Hearted Youtube comments on Mathologer (@Mathologer) channel.
-
17000
-
5700
-
2900
-
2000
-
1800
-
1700
-
1700
-
1700
-
1300
-
1300
-
1300
-
1300
-
1300
-
1200
-
1200
-
1100
-
1100
-
1100
-
1000
-
1000
-
1000
-
989
-
986
-
985
-
958
-
834
-
832
-
822
-
818
-
815
-
808
-
777
-
750
-
717
-
712
-
708
-
677
-
673
-
659
-
658
-
648
-
644
-
623
-
620
-
619
-
612
-
7:17 Actually, I would argue that neither you nor the young man in the clip missed out on that oversight, because there was no oversight to miss. The coach never stipulated that all of the cards had to end up being face-up, only that the sequence of moves had to terminate. In other words, the task was simply to show that there is no possible infinitely long sequence of moves, or to put it in yet another way, once you apply a move to any arrangement of cards, it is impossible to apply any combination of moves to the resulting arrangement such that the result is the original arrangement. Therefore, the answer to the question of what to do if only the right most card is face down is to do nothing. Such an arrangement simply represents a terminating condition.
Also, about the shorter video, I actually enjoy this format very much and would enjoy more shorter videos.
595
-
590
-
576
-
574
-
567
-
563
-
561
-
554
-
549
-
525
-
520
-
514
-
506
-
498
-
486
-
485
-
480
-
475
-
468
-
468
-
466
-
461
-
459
-
455
-
453
-
447
-
(28 March) Really bizarre, this video was basically invisible for almost two weeks with hardly any recommendations going out to fans. Only now YouTube has decided to actually show it to people. Who knows, maybe it was a mistake to mention cat videos in previous videos and the YouTube AI is now under the impression that the target audience for these videos has changed :)
The longest Mathologer video ever, just shy of an hour (eventually it's going to happen :) One video I've been meaning to make for a long, long time. A Mathologerization of the Law of Quadratic Reciprocity. This is another one of my MASTERCLASS videos. The slide show consists of 550 slides and the whole thing took forever to make. Just to give you an idea of the work involved in producing a video like this, preparing the subtitles for this video took me almost 4 hours. Why do anything as crazy as this? Well, just like many other mathematicians I consider the law of quadratic reciprocity as one of the most beautiful and surprising facts about prime numbers. While other mathematicians were inspired to come up with ingenious proofs of this theorem, over 200 different proofs so far and counting, I thought I contribute to it's illustrious history by actually trying me very best of getting one of those crazily complicated proofs within reach of non-mathematicians, to make the unaccessible accessible. Now let's see how many people are actually prepared to watch a (close to) one hour long math(s) video :). Have a look at the description for relevant links and more background info.
The first teaching semester at the university where I teach just started last week and all my teaching and lots of other stuff will happen this semester. This means I won't have much time for any more crazily time-consuming projects like this. Galois theory will definitely has to wait until the second half of this year :( Still, quite a bit of beautiful doable stuff coming up. So stay tuned.
445
-
440
-
429
-
428
-
426
-
426
-
425
-
422
-
419
-
416
-
383
-
379
-
379
-
376
-
375
-
372
-
371
-
367
-
366
-
361
-
358
-
357
-
352
-
348
-
348
-
343
-
340
-
339
-
331
-
329
-
329
-
326
-
320
-
318
-
314
-
309
-
304
-
303
-
298
-
297
-
291
-
288
-
286
-
283
-
280
-
279
-
276
-
271
-
269
-
268
-
265
-
262
-
262
-
261
-
259
-
253
-
253
-
253
-
245
-
242
-
242
-
240
-
239
-
238
-
234
-
233
-
231
-
221
-
220
-
220
-
220
-
219
-
218
-
218
-
218
-
217
-
217
-
215
-
Challenges:
1) We can guarantee at least 5 on one side. There are two ways to split the pigeons: 3-2 and 4-1. The max you can hit for both of them is 3 and 4 respectively. So we have at least 3 pigeons on one side. Add in the two pigeons on the equator and we have 5.
2) let the mystery value equal x. by multiplying by 10^some integer until the digits line up again, and then subtracting, we have 10^whatever*x - x = the non-repeating part you're left with. Factor out x on the left and that 10^whatever - 1 becomes a bumch of 9s. Then divide and we have the non-repeating remainder over a bunch of 9s. If ever the remainder isn't an integer just multiply both the numerator and denominator by 10 to knock out the decimal.
Example: 0.4318181818...
x = 0.4318181818...
100x = 43.18181818....
99x = 42.75 (all the 18s afterward cancel out)
x = 42.75/99
x = 4275/9900
x = 19/44 (simplify)
3) let's pretend we color the "didn't shake hands" white. We leave the "shook hands" black. So all we have to do is prove a triangle exists somewhere, and it doesn't matter what color. Note that each dot will always have at least three lines of the same color connecting to it, for the same reason the pigeons in challenge 1 can be split into a side with at least 3 and a side with less.
Look at a point, and follow three lines of the same color (I'll choose black) to three other points. Any line connecting two of these three points would either be black or white. If it were black, that would make a triangle with the two points and the starting point. If it weren't, then the three dots all have white lines between them, which makes another different triangle.
4) 315. And no, I didn't count manually. Here's how I found out:
The 7 edges UB, UR, RB, DB, DL, LB, and UL, all commute in one big cycle of length 7. The remaining 5 edges do the same. The 5 corners URF, ULF, DRF, DLF, and DRB go back in their positions but in the wrong orientation. So it takes 3 cycles to get back to their original orientation. The same thing happens with the other 3 corners in the back, only this time they take 9 cycles because they permute around each other too as well as reorient themselves.
All that's left is to compute the lcm of 7, 5, 3, and 9, which turns out to be 315.
5) 1, 2, 4, 8, 16, 32, and 64. Any collection just results in the binary representation of the number. Since every number can be written in binary in only one way, every single collection will add to a different number.
6) Queen of Hearts. The 9 of Hearts at the start signals the suit of the missing card is Hearts. The remaining three kings can be sorted as MBT (Diamond, Club, Spade), which is assigned 3. 3 spots after 9 is the Queen, and putting it all together we have Queen of Hearts.
Edit: I forgot one, challenge 5 at chapter 6
214
-
212
-
211
-
208
-
206
-
206
-
202
-
202
-
201
-
200
-
199
-
197
-
6:28 - The reason taxes come out first isn't some kind of greed on the part of government. It's because only paying the taxes first ensures that nobody will go bankrupt to their friends as a means of avoiding the payment of taxes to the government. If my assets are only just enough to cover my back taxes, I can call my friend and say "Do you remember that money I borrowed from you 20 years ago? We kinda forgot about it, but I think it's time to formalize it with a document because I'm about to go bankrupt". The friend comes over and I sign the note, which is for 99 times the amount of my assets. If taxes are not given priority, then when I go bankrupt the government gets its penny on the dollar in proportion to all debts, while my friend gets 99 cents on the dollar. When the dust settles, my friend simply gives me back that 99% of my assets. I end up losing 1% to back taxes instead of 100%. But if everyone knows that taxes come first, then everyone sees the fruitlessness of such schemes and they won't be attempted. Of course, I can run this same scheme on any NON-governmental entity to whom my debt equals my assets, and get 99% of everything back while the debt to that creditor is marked "paid in full". That creditor will go to court and challenge the validity of a promissory note signed one day before a bankruptcy-filing, but the government, by taking taxes first, doesn't have to go through that hassle.
197
-
196
-
194
-
193
-
192
-
192
-
191
-
190
-
187
-
187
-
186
-
186
-
186
-
185
-
185
-
183
-
183
-
183
-
180
-
180
-
178
-
178
-
177
-
176
-
176
-
176
-
176
-
175
-
174
-
171
-
170
-
170
-
168
-
168
-
168
-
168
-
167
-
166
-
165
-
164
-
163
-
163
-
163
-
161
-
160
-
158
-
156
-
156
-
155
-
155
-
154
-
154
-
154
-
First challenge: The chessboard cannot always be tiled after removing 2 black & 2 green. Suppose we remove the two black squares adjacent to the upper-left corner and the two green squares adjacent to the lower-left corner. Then the left corner squares are no longer adjacent to any other squares, so the board cannot be tiled.
EDIT:
Second challenge: if m and n are odd, then ⌈m/2⌉ = (m+1)/2 and ⌈n/2⌉ = (n+1)/2. Now the j=(m+1)/2, k=(n+1)/2 term of the product is 4cos²(π/2) + 4cos²(π/2) = 0, so the product is 0. Moreover, if m is not odd, then 0 ≤ j < (m+1)/2 in all terms of the product, hence 0 ≤ jπ/(m+1) < π/2 in all terms, so cos²(jπ/(m+1)) > 0 in all terms, so each term of the product is nonzero. This means the formula gives a nonzero answer whenever m is even -- symmetrically, the answer is nonzero whenever n is even. Thus, the formula returns 0 if and only if m and n are both odd.
154
-
151
-
149
-
148
-
147
-
147
-
146
-
146
-
145
-
145
-
144
-
142
-
142
-
141
-
141
-
140
-
138
-
138
-
135
-
134
-
133
-
133
-
131
-
131
-
130
-
130
-
129
-
128
-
127
-
127
-
127
-
127
-
126
-
126
-
124
-
124
-
124
-
123
-
122
-
121
-
121
-
120
-
119
-
118
-
118
-
118
-
118
-
117
-
117
-
116
-
116
-
114
-
113
-
113
-
112
-
112
-
111
-
110
-
110
-
109
-
109
-
109
-
107
-
107
-
107
-
107
-
107
-
107
-
106
-
106
-
105
-
103
-
103
-
103
-
102
-
101
-
101
-
101
-
101
-
100
-
100
-
Solution to the problem at 16:14
Start with this equation, which is true for every value of k:
5^k * 2^k = 10^k
The digital root of any power of 10 is 1, so
DR(5^k * 2^k) = 1
Using the multiplication rule you explained earlier,
DR(DR(5^k) * DR(2^k)) = 1
In other words, DR(5^k) and DR(2^k) have to be multiplicative inverses of each other.
Taking the “digital root” of an integer is equivalent to modding it by 9. (The only difference is that if DR(n) = 9, then n mod 9 = 0.) In mod-9 arithmetic, every number except for 0, 3, and 6 has a unique multiplicative inverse. Since the digital root of a power of 2 is never 3, 6, or 9, this means that DR(2^k) completely determines DR(5^k).
As k increases, the value of DR(2^k) cycles as follows:
2 4 8 7 5 1 2 4 8 7 5 1 …
Taking the multiplicative inverse of each number above gives the values of DR(5^k).
5 7 8 4 2 1 5 7 8 4 2 1 …
So DR(2^k) and DR(5^k) cycle through the same values, but in reverse.
100
-
99
-
99
-
98
-
98
-
98
-
98
-
97
-
97
-
97
-
97
-
96
-
96
-
95
-
95
-
94
-
94
-
94
-
93
-
92
-
91
-
91
-
91
-
91
-
89
-
88
-
88
-
87
-
86
-
86
-
86
-
85
-
85
-
85
-
84
-
84
-
84
-
84
-
84
-
84
-
83
-
83
-
82
-
82
-
82
-
81
-
For the demonstration of A(X) + A(Y) = A(XY)
Definitions :
- we define A’(x,y) as the area under the curve between x and y (we notice that A’(1,y) = A(y) and that A’(a,b) + A’(b,c) = A’(a,c))
- we define f_p(z) the transformation of an area z by a factor p
Let’s find an expression for f_p:
Given z=A’(a,b), the top left hand corner of the area is mapped to:
1/p * 1/a = 1/(ap) (squeezing)
which corresponds to the inverse function for a value ap (shifting). Same goes for the top right hand corner.
We can conclude that:
f_p(A’(a,b)) = A’(pa,pb) = A’(a,b)
Let now prove the theorem :
A(x) + A(y) = A’(1,x) + A’(1,y)
= A’(1,x) + f_x(A’(1,y))
= A’(1,x) + A’(x,xy)
= A’(1,xy)
= A(xy)
81
-
81
-
81
-
81
-
81
-
80
-
80
-
80
-
80
-
79
-
79
-
78
-
78
-
76
-
76
-
75
-
75
-
75
-
75
-
74
-
74
-
74
-
74
-
74
-
74
-
74
-
73
-
73
-
73
-
73
-
73
-
73
-
72
-
72
-
72
-
72
-
72
-
72
-
72
-
72
-
71
-
71
-
70
-
70
-
70
-
70
-
69
-
69
-
69
-
What makes Quadratic Reciprocity so special, and in particular, why was it so important to Gauss? Why is squaring integers so important in Number Theory? To understand the answers to these questions, you need to appreciate the work Gauss did in the theory of quadratic forms and their application to differential geometry. Gauss's "fundamental theorem" of differential geometry is about how the 2nd order partial derivates of a function that defines a "surface", e.g. `f(x,y)=height`, depends on a specified quadratic form called the "metric tensor" of the surface (and its first derivates), which gives a rule for how to measure distances "on the surface". This shows a profound and deep insight into the nature of the relationship between a large number of otherwise seemingly unrelated abstract concepts.
Squaring numbers plays a fundamental role in both Number Theory and Geometry. After watching this video twice, going on a few wild goose chases, and not being able to stop thinking about it: I finally think I understand why Gauss placed such a high degree of importance on this particular theorem, and feel like I have a lot of reading to do now!
69
-
69
-
69
-
68
-
68
-
68
-
67
-
67
-
67
-
66
-
66
-
66
-
65
-
65
-
65
-
65
-
64
-
64
-
64
-
64
-
64
-
64
-
63
-
63
-
63
-
63
-
63
-
63
-
63
-
62
-
62
-
62
-
62
-
62
-
62
-
So for that first puzzle... I think my mind just got blown
I started with 5 since I was basing on the fact that the first equation of both patterns end and start with 3 (1 + 2 = 3 and 3² + 4² = 5²), so why not do the same for the cube pattern. What I ended up is that 5³ + 6³ = 341, which is almost close to 7³ = 343. I was curious enough, so I tried to expand this by utilizing the same trick from the previous patterns and used 6*(1 + 2) for the next pattern and got 16³ + 17³ + 18³ ≈ 19³ + 20³ (with the difference being 18)
Expanding this pattern leaves me with a list of differences and... I don't know about you, but (2, 18, 72, 200, etc.) just screams "I have a pattern"... and it does!
Each difference is just twice a triangular number *squared*, and knowing that just blew my goddamn mind... because THAT'S LITERALLY HOW I STARTED working on this pattern. For each sum that uses 6(1 + 2 + ... + n) cubed as a starting point, there is a difference of 2(1 + 2 + ... + n)². How cool is that!?
In honor to document this amazing pattern, here's my christmas tree for the cube pattern, with the difference added in. You can think of them like they're ornaments or something xD
(the formatting might only work with monitors)
5³ + 6³ = 7³ - 2
16³ + 17³ + 18³ = 19³ + 20³ - 18
33³ + 34³ + 35³ + 36³ = 37³ + 38³ + 39³ - 72
56³ + 57³ + 58³ + 59³ + 60³ = 61³ + 62³ + 63³ + 64³ - 200
85³ + 86³ + 87³ + 88³ + 89³ + 90³ = 91³ + 92³ + 93³ + 94³ + 95³ - 450
...
and here's another one with the sum expanded, just to see the beauty :D
5³ + [6(1)]³ = 7³ - 2(1)²
16³ + 17³ + [6(1 + 2)]³ = 19³ + 20³ - 2(1 + 2)²
33³ + 34³ + 35³ + [6(1 + 2 + 3)]³ = 37³ + 38³ + 39³ - 2(1 + 2 + 3)²
56³ + 57³ + 58³ + 59³ + [6(1 + 2 + 3 + 4)]³ = 61³ + 62³ + 63³ + 64³ - 2(1 + 2 + 3 + 4)²
85³ + 86³ + 87³ + 88³ + 89³ + [6(1 + 2 + 3 + 4 + 5)]³ = 91³ + 92³ + 93³ + 94³ + 95³ - 2(1 + 2 + 3 + 4 + 5)²
...
61
-
61
-
61
-
61
-
60
-
60
-
60
-
60
-
59
-
59
-
59
-
59
-
59
-
59
-
59
-
58
-
58
-
57
-
57
-
57
-
56
-
56
-
56
-
56
-
55
-
55
-
55
-
55
-
55
-
54
-
54
-
54
-
54
-
54
-
54
-
54
-
53
-
53
-
53
-
53
-
52
-
52
-
52
-
52
-
52
-
52
-
52
-
52
-
51
-
51
-
51
-
51
-
51
-
51
-
51
-
51
-
51
-
51
-
51
-
50
-
For the grid problem, sum up every the possible 2*2 suqare exactly once. Notice that the total amount of times red, green, yellow, purple appeared MUST BE EQUAL. The corner squares are counted once, the edge suqares are counted twice (it is calculated in 2 of the 2*2 suqare), and the squares in the middle are calculated 4 times(it is in 4 of the 2*2 squares). Since the total amount of 2*2 squares is ODD, every color must appear an ODD number of times. Ans since only the corner squares are counted an ODD number of times, each color MUST appear in the corner square. QED.
By the same argument, we can show that every color appears on the corner of the 4n*4n*4n cube, if it is colored with 8 colors such that every 2*2*2 cube consist of 8 distinct colors.
We can generalize this further to m dimentional cube with 4n as its sides.
50
-
50
-
50
-
50
-
50
-
50
-
4:54 This actually isn't quite the origin of the name. From Wikipedia:
Dirichlet published his works in both French and German, using either the German Schubfach or the French tiroir. The strict original meaning of these terms corresponds to the English drawer, that is, an open-topped box that can be slid in and out of the cabinet that contains it. (Dirichlet wrote about distributing pearls among drawers.) These terms were morphed to the word pigeonhole in the sense of a small open space in a desk, cabinet, or wall for keeping letters or papers, metaphorically rooted in structures that house pigeons.
Because furniture with pigeonholes is commonly used for storing or sorting things into many categories (such as letters in a post office or room keys in a hotel), the translation pigeonhole may be a better rendering of Dirichlet's original drawer metaphor. That understanding of the term pigeonhole, referring to some furniture features, is fading—especially among those who do not speak English natively but as a lingua franca in the scientific world—in favour of the more pictorial interpretation, literally involving pigeons and holes. The suggestive (though not misleading) interpretation of "pigeonhole" as "dovecote" has lately found its way back to a German back-translation of the "pigeonhole principle" as the "Taubenschlagprinzip".
50
-
49
-
49
-
49
-
49
-
49
-
49
-
49
-
49
-
49
-
Burkard, you asked specifically for feedback on how successfully you made these explanations accessible.
I am an electronics engineer who has always been fascinated by mathematics, but struggled with the higher, more abstract elements of the subject. My way of learning is like constructing a building: laying foundations and fully understanding foundational material before laying the next layer of bricks. And I always felt uncomfortable just learning calculus identities without understanding their derivation. I think that's one of the big things that held me back.
Your videos follow a logical progression, building layers of bricks on top of understanding I already have, and at last I can trust in my gut that these identities are correct because you have shown me, even if I cannot repeat the derivation myself.
A few times in this video I had to stop, rewind and check some simplification or rearrangement you did, because it wasn't immediately clear to me, but I got there.
Thank you for making this fascinating subject accessible in a way I have never seen before. You have rekindled my love of higher math.
By the way, I much prefer the taurean version of Euler's identity: e to the i tau equals one, and I think your beautiful graphical demonstration could easily be adapted to demonstrate this form. I'd love to see your taurean conversion of other identities, which will be similar but I think lead to a far greater intuitive resonance than using pi.
48
-
48
-
48
-
48
-
48
-
48
-
48
-
48
-
48
-
48
-
48
-
47
-
47
-
47
-
47
-
47
-
47
-
47
-
47
-
47
-
47
-
47
-
46
-
46
-
46
-
46
-
46
-
46
-
46
-
46
-
46
-
46
-
45
-
45
-
45
-
45
-
45
-
45
-
45
-
45
-
45
-
45
-
45
-
45
-
@Mathologer It does rhyme! and it is deliberately written in the form of seven-word-poem (much like the solution poem to the Chinese remainder theorem).
On this same topic, there's another, much more ancient (and famous) "poem" on just the 3x3 magic square:
九宫之义,法以灵龟,二四为肩,六八为足,左三右七,戴九履一,五居中央
Translation:
"""
The way to fill a 9-square palace, is to imagine a turtle (back):
2 and 4 as the shoulder, 6 and 8 as feet, 3 on the left, 7 on the right, 9 as hat, 1 as shoe, and 5 in the middle
"""
This text came from an ancient manuscript, that says these numbers/pattern comes on a turtle, and it's a sign of miracle. I guess that's part of what you said in the video: some people think magic square is truly magical.
44
-
44
-
44
-
44
-
44
-
44
-
43
-
43
-
43
-
43
-
43
-
43
-
43
-
42
-
42
-
42
-
42
-
42
-
42
-
41
-
41
-
41
-
41
-
41
-
Thanks a lot, Mathaloger! I have studied some works of Madhava but didn't know about these beautiful correction terms.
Our textbooks in India still reflect a post-colonial mindset, resulting in a significant lack of awareness and ignorance. Unfortunately, many important contributions from Indian scholars are overlooked or not adequately highlighted. For instance, Baudhayana, who authored the world's first geometry textbook, including the Pythagorean theorem long before Pythagoras, is rarely mentioned. Similarly, Pingala's profound insights into permutations and combinations within the context of literature are often disregarded, despite their beauty and significance. Aryabhata's foundational work in trigonometry, which forms the basis of our modern understanding, also receives insufficient attention. Furthermore, there's no mention of Madhava, or any mathematician from the Nila school, in our great textbooks. They only perpetuate a subtle sense of inferiority complex by failing to acknowledge these remarkable achievements. It is high time for a change in our educational curriculum.
One aspect I particularly admire is India's significant contributions to philosophy and spirituality. These traditions are not only highly logical but also explore the limitations of logic itself, as evident in the concept of non-dualism. Additionally, architectural marvels like the Kailash temple exemplify India's profound knowledge in the field of architecture. Regrettably, we have lost a great deal of knowledge due to constant invasions, including the destruction of the precious Nalanda University, the world's first residential university.
There is so much more to be said about India's marvelous contributions to the world. It is my hope that our generation recognizes and appreciates the immense legacy left by our ancestors.
Thank you so much again.
Keep up the amazing work that you do!!
41
-
41
-
41
-
41
-
41
-
41
-
41
-
40
-
40
-
40
-
40
-
40
-
40
-
40
-
40
-
40
-
40
-
40
-
40
-
40
-
40
-
40
-
40
-
39
-
39
-
39
-
39
-
39
-
39
-
39
-
39
-
39
-
39
-
39
-
39
-
38
-
38
-
38
-
38
-
38
-
38
-
38
-
38
-
38
-
38
-
38
-
38
-
38
-
38
-
37
-
37
-
37
-
37
-
37
-
37
-
37
-
37
-
37
-
37
-
37
-
37
-
37
-
36
-
36
-
36
-
36
-
36
-
36
-
36
-
36
-
36
-
36
-
36
-
The general formula for the sum of 1/x^(2n) is hidden in here too, and only requires a few extra steps. Start from the chapter 2 formula and take logs and derivatives to get the formula at 10:40:
cot(x) = 1/x + 1/(x-pi) + 1/(x+pi) + 1/(x-2pi) + 1/(x+2pi) + ...
Move the 1/x term to the left side, and take 2n-1 more derivatives. The k'th derivative of 1/x is (-1)^k * k!/x^(k+1), so when k = 2n-1 this is -(2n-1)!/x^(2n). So we have
(d/dx)^(2n-1) (cot(x) - 1/x) = -(2n-1)! * (1/(x-pi)^(2n) + 1/(x+pi)^(2n) + ...)
Now divide by -(2n-1)! and take the limit as x -> 0.
1/(2n-1)! * lim x -> 0 [(d/dx)^(2n-1) (1/x - cot(x))] = 1/(-pi)^(2n) + 1/pi^(2n) + 1/(-2pi)^(2n) + 1/(2pi)^(2n) + ...
On the right side, all the negatives are squared away, and we end up with 2 copies of each term. So multiply by pi^(2n)/2, and we get this:
pi^(2n)/2 * 1/(2n-1)! * lim x -> 0 [(d/dx)^(2n-1) (1/x - cot(x))] = 1 + 1/2^(2n) + 1/3^(2n) + ... = zeta(2n)
This formula looks messy, but there's a trick: notice that on the left, we have something of the form 1/k! * k'th derivative of f(x) at x=0. These are just taylor series coefficients! The left side is really just pi^(2n)/2 times the coefficient of x^(2n-1) in the taylor series of 1/x - cot(x).
There are a few other things we can do to make the formula easier to read. We can multiply the function by x to make the powers line up nicely (otherwise the 1/k^8 sum will be related to the coefficient of x^7, instead of x^8). This gives:
zeta(2n) = pi^(2n)/2 * coefficient of x^(2n) in the taylor series of 1 - x cot(x)
The next thing we can do is move the pi^(2n) "inside" the taylor series, by replacing x with pi x. We can also move the factor of 1/2 into the function. Then we get:
zeta(2n) = coefficient of x^(2n) in the taylor series of (1 - pi x cot(pi x))/2, or equivalently,
(1 - pi x cot(pi x))/2 = sum n=1..inf, zeta(2n)x^(2n)
And indeed, if you ask wolframalpha to compute the taylor series of (1 - pi x cot(pi x))/2, you get pi^2/6 x^2 + pi^4/90 x^4 + pi^6/945 x^6 + pi^8/9450 x^8 + ...
Finally, comparing this series to the standard taylor series for cot in terms of Bernoulli numbers gives Euler's general formula for zeta(2n)
35
-
35
-
35
-
Since I made it to the end of this, and since you asked nicely: I'm an enthusiastic amateur, never took much beyond AP calculus formally but I read & study the subject pretty broadly as a hobby.
I was cruising along smoothly here right up until about chapter 4 or so, after which there was a lot of pausing, rewinding, workings-out on paper or Python, and a couple side trips into Wikipedia & Wolfram.
The Pascal/Bernoulli thing was mind-blowing to see in action. It definitely made sense on the surface, though as it got deeper i started to feel like it was going in circles? Something like, "We can easily derive S using B. But how do we get B? Well by deriving it from S of course." At least that was my initial impression.
Really enjoyed it even if it did start to outpace me towards the end (or more accurately, because it did). It'll be a few more viewings before everything clicks for sure, but I'm looking forward to the challenge!
35
-
35
-
35
-
Hi, I did find another way to "build" youre cubes. Instead of adding the hexagons, you can add 1+ (1+2+3), than add 7 + (3+4+5), and 19+ (5+6+7), ...
If you look in the video at 8:39 you see the second cube. To go to the next cube you need to add one side, so 3 pieces, lets say on the left, then you add 4 pieces to the right side (3 pieces for the original cube and one piece for the just added row), and finaly you add the bottom with 5 pieces to get back to the cube.
Now if you look at the numbers you add they are left and right from the yellow marked number in the top row (2 and 4 sandwidch the 3; 5 and 7 the 6,...)
And these yellow numbers are diviceble by 3 (by construction)
So you get a sequens of (3n-1)+(3n+1) and that is the same as 2(3n).
We could write 2(3n) also as 3(2n).
This we can convert in (2n-1)+(2n)+(2n+1).
And this is the secuence you are adding to youre cube.
(Note "n" is the base number from youre last cube, and the counting number of youre tripple (where 3 is the 1st; 6 the 2nd;9 the 3th;12 the 4th,..))
Notice that by every step you go up you start with the same number you stopt with the last time.
(2n+1)=(2n+2-1)=(2(n+1)-1)
So the cube you end up with has a sidelenght of 2n+1. And you start the next step just adding that side to youre cube
I hope you can see how i find this a bit easyer to construct the cube.
Also this is in a way the same you construct youre squares (if you start with "skipping" just one number at the top) if you notice that every odd number is the sum of its position + its position -1. (5 is the 3th odd number and it is 3+2; 9 is the 5th odd number and it is 5+4; ...). Also here you get that the last digit you added the last time you start to add the next time. (but you only get 2 numbers).
Note; this methode does not work in the fourth or higher dimensions. Why? Well I'm not a mathematitian. So I don't know. And I have trouble imagining in four dimentions.
Hope this is helpfull.
Sorry for my bad English, I speak Dutch.
Greetings from Belgium.
35
-
34
-
SPOILERS SPOILERS SPOILERS: It's uncountable; here's a funny algorithm: given a sequence (n_k) of natural numbers, we start out just as normal, adding positive terms until we overshoot. But rather than turning around immediately, we keep going for n_0 many positive terms, and only then turn around. Next, once we undershoot, we don't immediately turn around, but keep adding n_1 many negative turns before we turn around. Then once we overshoot again, we keep adding n_2 many positive terms before we turn around, etc. Note that the original algorithm is what you get when you start with the constant sequence (0,0,0,...).
Now of course, this isn't guaranteed to converge; you could end up over/undershooting too wildly. But - certainly if all the n_i are 0 or 1, it will converge. And there are already uncountably many sequences of just 0s and 1s.
34
-
34
-
34
-
34
-
34
-
34
-
34
-
34
-
34
-
34
-
34
-
34
-
34
-
34
-
34
-
34
-
33
-
33
-
33
-
33
-
33
-
33
-
33
-
@zacharyjoseph5522 It's a bit tricky to explain without diagrams, but I'll do my best. Let us call the whole shape G for glasses. First consider the pair of 2x3 rectangles at the extreme left and right of the glasses. If a domino is placed which crosses the border between the 2x3 and the rest of the glasses, then the 2x3 is left with a region of size 5, so cannot be covered.
We therefore know that the dominoes covering the pair of 2x3 areas are necessarily wholly within the 2x3 areas, so the tilings would be the same if the 2x3 regions were actually disconnected from the glasses. So if we let G' be the glasses without this pair of 2x3 regions and if we let N be the function counting the number of tilings of a shape, then we have N(G) = N(2x3)^2*N(G'), as the tilings of the pair of 2x3 regions and the tiling of G' are independent.
We'll need names for a few other things, so we'll call the boundary of a hole E for eye. So the eye is the 10 squares directly surrounding the hole. Consider the middle 2x2 in G'. Imagine a vertical line L cutting this 2x2 into two pieces. This line divides G' into two equal copies of the same shape, an eye with a 2x1 hanging off one side and a 2x2 hanging off the other. We'll call this shape E+2x2+2x1
Now if we consider tilings with no domino crossing L, then clearly the number is N(E+2x2+2x1)^2, as the tilings of the two shapes on either side of L are independent.
If instead we have a domino crossing L, then we must in fact have 2 dominoes crossing L. This is because otherwise we'd create a region of odd size, which couldn't be tiled. In this arrangement with 2 dominoes crossing L, we again have two identical regions to tile, but this time they are E+2x2, following the naming convention for the previous shape. In this case there are N(E+2x2)^2 tilings.
So we've shown that N(G') = N(E+2x2+2x1)^2+N(E+2x2)^2.
Next we consider E+2x2+2x1. If the 2x1 area is covered by one domino, then we're left with tiling E+2x2. Otherwise, then the tiling is forced all the way to a remaining 2x2 region. Therefore N(E+2x2+2x1) = N(E+2x2)+N(2x2).
Now we consider E+2x2. Imagine a line L' dividing the E from the 2x2. If no domino crosses L', then the E and 2x2 are tiled separately, so we get N(E)*N(2x2) tilings. If instead a domino crosses L', then the rest of tiling is forced, so we get 1 such tiling. Therefore N(E+2x2) = N(E)*N(2x2)+1.
The remainder is not especially hard to check. N(E) = N(2x2) = 2, so N(E+2x2) = 2*2+1 = 5, N(E+2x2+2x1) = 5+2 = 7, N(G') = 7^2+5^2 = 74. Finally N(2x3) = 3, so N(G) = 3^2*74 = 666.
33
-
33
-
32
-
32
-
32
-
32
-
32
-
32
-
32
-
32
-
32
-
32
-
32
-
32
-
13:46 I was so annoyed that the pattern was that simple. I had worked out a completely different, more complicated pattern. The sums of the differences were always factors of the double position number they surrounded. (e.g. the position numbers surrounding 2 were 1 and 3, which sum to 4, which is double the original position number). Furthermore, the number needed to multiply the sum to get to double the position number went 1, 1, 2, 2, 3, 3, 4, 4, 5, 5, etc. It was a pattern, just a way more complicated one. Now I'm going to have to try to prove that they are equivalent patterns, which may be quite difficult.
31
-
31
-
31
-
31
-
31
-
11:56 challenge question: The Pythagorean triple (153, 104, 185) corresponding to the box
[ 9, 4, 17, 13 ]. If you call the children A,B,C, the 153, 104, 185 is the A'th Child of 'CCC'
31
-
31
-
31
-
31
-
31
-
31
-
31
-
31
-
31
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
30
-
29
-
29
-
29
-
29
-
29
-
29
-
29
-
29
-
29
-
29
-
29
-
28
-
28
-
28
-
28
-
28
-
28
-
28
-
28
-
28
-
28
-
28
-
28
-
28
-
28
-
28
-
28
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
27
-
26
-
Merry Christmas!
6:11 - You could pick the two blacks in the top left corner. It would isolate the corner green square, so not every combination of 4 squares removed is tileable.
10:13 - say m = 2p-1 and n=2q-1. The denominators in the cosines will be 2p and 2q. Carrying out the product, when j=p and m=q, we will have a term (4cos²(π/2)+4cos²(π/2)), which is 0, cancelling out everything else
13:50 - Lets say T(n) is the number of ways to tile a 2xn rectangle. First two are obviously T(1)=1 and T(2)=2. For the nth one, lets look at it from left to right. We can start by placing a tile vertically, which will isolate a 2x(n-1) rect. - so T(n-1) ways of doing it in this case. If we instead place a tile horizontally on the top, we will be forced to place another one directly below, so we don't isolate the bottom left square, this then isolates a 2x(n-2) - so T(n-2) ways of doing it in this case.
----
We have T(n) = T(n-1) + T(n-2). Since 1 and 2 are fibonacci numbers, the sequence will keep spitting out fibonacci numbers
14:33 - It's 666. I did it by considering all possible ways the center square can be filled and carrying out the possibilities. It was also helpful to see that the 2x3 rectangles at the edges are always tiled independently. I was determined to do all the homework in this video, but hell no I wont calculate that determinant, sorry
30:12 - I'll leave this one in the back of my mind, but for now I'm not a real math master. I'm also not a programmer, but this feels like somehting fun to program
37:28 - Just look at the cube stack straight from one of the sides, all you'll see is an nxn wall, either blue, yellow, or gray
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
26
-
Your concerns are well founded. I was a math and science teacher at two private schools in the United States (each time for two years). The textbook selections never aligned well with my ideals of making math and science relevant, useful and entertaining. For 3 of the four years, I was able to accomplish that in spite of the books. During the 4th year the academic police state finally caught up with me, insisting that the reasons we teach anything are foremost to increase test scores and grow our market share through test-based reputation. That authority banned all non-standard curricula and forced me out of the profession I loved. All the texts from which I taught were ostensibly aligned with their goals. They were also filled with endless drills, BS examples, incomplete history, and frankly serpentine reasoning far more likely to confuse that to convey any valuable understanding.
I’ve lived 4 years in Germany (which was a bit better) and 4 more in South Korea (which seemed much better). Unfortunately, the Korean kids were under immense pressure to perform, making almost all of them profoundly depressed and/or stressed from perhaps 10 years of age.
Most of the great contributors to the betterment of our world recognize the importance of conquering fun challenges. Both mindless, meaningless repetition and idiotic complication turn beautifully curious and malleable children into miserable adults.
When teaching became primarily the indoctrination of future workers, it necessarily ceased teaching young people to think critically and creatively in favor of teaching them to do what those in authority tell them. This bodes terribly for the future and causes me to be deeply concerned for our posterity.
Thank you for making learning the entertaining challenge it’s meant to be!
26
-
26
-
25
-
25
-
25
-
25
-
25
-
25
-
25
-
25
-
25
-
25
-
25
-
25
-
25
-
25
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
24
-
23
-
23
-
23
-
23
-
23
-
23
-
23
-
23
-
23
-
Hey, I won gold because of this channel (long story lol), and have a suggestion for a small addition to the part two.
The correspondence between the map of Lambert and Kepert is done by taking circle segments between two points, and varying the "angle" of the circle segment. 0° is a line segment, and 90° is a halfcircle, like used in the video.
One of my own proofs, of the cylic quadrilateral angle theorem (that has undoubtedly been found by someone else as well) is that given a quadrilateral, you can look at the line segments like they are circle segments. The "circle segment quadrilaterals" have invariant α–β+γ–δ. Since you can merge two pairs of circle segments, you basically directly get the theorem.
This even works for hyperbolic geometry!
It is a nice proof, using unorthodox techniques, so it probably has to show up on this channel eventually, although it likely won't fit.
It also has a dual theorem, where the opposing sides of a quadrilateral sum to the same length if and only if the quadrilateral has an inscribed circle.
23
-
23
-
23
-
23
-
23
-
Just to go a step further, the perimeter of the shape described by the wiper arcs is exactly pi times the perimeter of the triangle i.e. (pi * the chord length), whereas the circumference of the circle is (pi * 2 * the radius) and, as noted above, it is easily seen that (2 * the radius) is greater than (the chord length).
To check that the perimeter of the SOCW is (pi * the chord length), consider each arc length. If we label the triangle sides A,B,C and their opposite angles a,b,c, then the arc lengths are given by: Aa + Bb + Cc + (B+C)a + (A+C)b + (A+B)c = (A+B+C)(a+b+c) = (chord length)(pi)
23
-
23
-
23
-
23
-
23
-
23
-
22
-
22
-
22
-
22
-
22
-
22
-
22
-
22
-
22
-
22
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
Fantastic video!
A more computationally-efficient method than the determinant method, which also easily and transparently deals with single and multiple zeros:
Step 1: After using k delayed copies of the original sequence, instead of repeatedly computing determinants (as in the video), take a rectangular “snapshot” — this is a rectangular matrix of size k-by-n, with n > k.
Step 2: Transpose this matrix (so now it’s n-by-k with n > k) and compute its “economy” SVD (singular value decomposition). There are many existing software libraries to do this. This step is the only computationally-intensive step, whose complexity is O(n k^2), which is much less than the determinant method as stated in the video, which is O(n k^3), or even more if care is not taken…
Step 3: Now, look at the singular values: If one or more of the smallest singular values is/are zero, then we’re sure this is what Burkard calls “Fibonacci-like sequence”, something usually called “linear recurrence with constant coefficients” (which he mentioned in the video). However, if none of the singular values is zero, then repeat with larger k, i.e., more delayed copies of the original sequence.
Step 4: The linear recurrence with constant coefficients is easily determined from the column of the SVD result corresponding to the zero singular value (because this is the vector which can zero-out any set of k consecutive elements of the sequence). Specifically, if the “economy” SVD result of the original n-by-k matrix is U, s, Vt, then the last row of the square k-by-k matrix Vt will be the desired set of linear constant recurrence values. This last row has size 1-by-k, of course. These values will probably need to be scaled by a common multiple if nice integer values are desired! Move the 1st element of this set of k values to the other side of the equation to use as a prediction recurrence equation, i.e., the next value is determined linearly from the previous k-1 values.
Remark: The idea of a singular value decomposition (SVD) is mathematically very closely related to a determinant, because it essentially also determines linear combinations of columns (and/or rows). But it is more computationally-efficient in this case, because it can be computed for a rectangular set of values simultaneously, rather than small squares of values one after the other. The second advantage of the SVD is that it can also give us the linear recurrence with constant coefficients “for free” (from the same computational result).
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
21
-
20
-
Let's collect some interesting comments and remarks as they come in (also check out the description of this video):
VincentvanderN
I don't know if this is already down here somewhere in the comments, but 2 x 2 = 2 + 2 can be used to show that there are infinitely many prime numbers. We generate a sequence a_1, a_2, a_3 etc by starting with a_1 = 3 and a_{n + 1} = a_n^2 - 2. Now we use our freak equation to show that if q is a prime divisor of some a_m, it cannot be a prime divisor of a_n for any n > m. (And then, as a result neither of an a_n with n < m, because otherwise we could repeat the proof with n in the role of m and get a contradiction.)
Here is the argument. Look at the sequence a_m, a_{m+1}, ... modulo q. We have a_m = 0 mod q, a_{m+1} = - 2 mod q, a_{m + 2} = 2 mod q, and then, by the freak equation 2 x 2 = 2 + 2 (in the form 2 x 2 - 2 = 2) we get that a_{m + k} = 2 for all k >= 2.
Neat, right? I believe I learned this from Proofs from the Book.
charlesstpierre9502
Use 2x2=2+2 to make a formula for generating Pythagorean triples. Start with two positive integers m > n. Then
(2×2)(mn)² = (2+2)(mn)²
(2×2)(mn)² = 2(mn)² + 2(mn)²
(2×2)(mn)² = 2(mn)² + 2(mn)² + (m⁴ - m⁴) + (n⁴ - n⁴)
(2×2)(mn)² = m⁴ + n⁴ + 2(mn)² - m⁴ - n⁴ + 2(mn)²
(2mn)² = (m² + n²)² - (m² - n²)²
Set:
a = m² - n²
b = 2mn
c = m² + n²
Then:
a² + b² = c²
This will not work for: aᵏ + bᵏ = cᵏ; k > 2
franknijhoff6009
Hi Burkhard, the 3-variable equation plays a role in integrable systems. In fact, it appeared in Sklyanin's work on quadratic Poisson algebras (around 1982) providing solutions in terms of elliptic functions, and (with an extra constant) as the equation for the monodromy manifold of the Painleve II equation.
Sklyanin's paper is: Some algebraic structures associated with the Yang-Baxter equation. Functional.Anal.i Prilozhen 1982 vol 16, issue 4,pp 27-34 ; look at equation (27).
Furthermore, in L.O. Chekhov etc al., Painleve Monodromy Manifolds, Decorated Character Varieties and Cluster Algebras , IMRN vol 2017, pp 7639--7691 you can find in Table 1 a close variant of the 3variable equation with extra parameters.
I think it would be interesting to explore this idea with exponents. For example 1+1+1+1+2+3=3^2^1^1^1^1
2¹×2¹=2¹+2¹
3½×3½×3½=3½+3½+3½
4⅓×4⅓×4⅓×4⅓=4⅓+4⅓+4⅓+4⅓
5¼×5¼×5¼×5¼×5¼=5¼+5¼+5¼+5¼+5¼
...
tanA + tanB + tanC = tanA x tanB x tanC (that's the identity that features prominently in the Heron's formula video)
For completeness sake and for fun let's mention
2⁴ = 4²
2 + 2 = 2 × 2 = 2²
log(1+2+3)=log(1)+log(2)+log(3)
log(2+2)=log(2)+log(2)=2log(2)
https://www.mtai.org.in/wp-content/uploads/2023/09/IOQM_Sep_2023_Question-paper-with-answer-key.pdf ... 29th question in an Indian maths olympiad problem
Check out this paper by Maciej Zakarczemny, On the equal sum and product problem (linked to in the description of this video)
http://www.iam.fmph.uniba.sk/amuc/ojs/index.php/amuc/article/view/1662
Theorem 5.1. says If there are exactly two sum-equals-product identitites , then one of the following two conditions is true:
1. n - 1 is a prime and 2n - 1 \in { p, p^2, p^3, pq } ,
2. 2n - 1 is a prime and n - 1 \in { p, p^2, p^3, pq } ,
where p, q denote prime numbers.
Why don't we start out the sequence of basic sum-equals-product identities with 1=1? Well, N=N for all N :) Very tempting to add the 1=1 at the top. To not do so is very much in line with not calling 1 a prime (saves you a lot of unnecessary heart ache further down the line :)
For this video I played around with generating some AI animated photos of Sophie Germain using https://www.myheritage.com/deep-nostalgia as well as with the ChatGPT generated rendering of Sophie Germain in the thumbnail. Here I fed the AI some of the existing pictures of Sophie Germain and asked it to play artist to come up with a picture of her in the same style used in this real picture of Leibniz https://www.alamy.com/gottfried-von-leibniz-image5071937.html). Lots of fun and quite eerie :) Not perfect, but will be interesting to find out what the machines can do in 6 months time (and at what point the machines can make better Mathologer videos than the Mathologer :(
In general quite a bit of AI hate in this comment section. Strangely I've never had people complain about me using any of those generic white men with beards images of ancient mathematicians that museums are full of.
Sophie Germain primes and safe primes (the 2n+1 primes) are important in cryptography, those involving prime fields. Because if the size field is a prime p, the size of its multiplicative group is p-1. For the group to be secure, the multiplicative group must have a large subgroup of prime size so p-1 must have a prime factor. So, having a prime q such that 2q+1 is also a prime is a good candidate. Even though it is not strictly required as long as p-1 has a big enough prime factor, this is what students are taught as a start.
Kaprekar's constant is among the lengths corresponding to exactly two sum-equals-product identities (discussed at the end of this video :) https://en.wikipedia.org/wiki/6174
Cunningham chains are an interesting generalisation of Sophie Germain primes (chains of primes such that the next prime in the chain is always the previous one times 2 plus 1). https://en.wikipedia.org/wiki/Cunningham_chain
If you are happy to also play this game with negative integers then you get more solutions. In particular, since (-1)+(-1)+1+1=0 and (-1)x(-1)x1x1=1, you can splice these two blocks into a/any sum-equals-product identity of length n to arrive at a sum-equals-product identity of length n+4.
And for complex integers we've also got things like this: (1 - i) + (1 + i) = 1 - i + 1 + i = 1 + 1 -i + i = 2 + 0 = 2 = 1 + 1 = 1 - (-1) = 1 - i^2 = (1 - i)(1 + i)
2xy - (2+x+y) +1 = N-2
<=> 2xy - 2 - x - y +1 = N-2
<=> 2xy - x - y +1 = N
<=> 4xy - 2x - 2y +2 = 2N
<=> 2x(2y - 1) - 2y +2 = 2N
<=> 2x(2y - 1) - (2y - 1) = 2N-1
<=> (2x - 1)(2y - 1) = 2N-1
Probably the easiest way to demonstrating this equivalence is to go backwards and start by expanding (2x-1)(2y-1)...
Relevant Project Euler problem: 88 https://projecteuler.net/problem=88
Relevant Online Encyclopedia of integer sequences: A033178
@Frafour
Tangentially related to 2x2 = 2+2: there is this video of Gromov (Gromov: 4 = 2 + 2 as "proof of Donaldson's theorem") https://youtu.be/bgePMb8wxv0?si=DmTf2wzbIr1F21f8 observing that 4 = 2+2 in 3 ways. That is, there are 3 ways to partition a 4-element set in two 2-element subsets. The fact that 3 is smaller than 4 is unique to 4, and produces a surjection from A_4 to A_3 - the alternating groups on 4 and 3 elements. This shows that A_4 is not simple, 4 is the only number where this happens, and this is responsible for many weird things in dimension 4.
In another direction, I first encoutered Sophie Germain primes accidentally when learning group theory in my undergrad. There is an elementary proof of simplicity of the Mathieu groups M_11 and M_23, which actually gives a general simplicity criterion for subgroups of S_p, p prime https://www.jstor.org/stable/2974771?origin=crossref. When I read this I wondered if this ever works for proving simplicity of the alternating group A_p. If I remember correctly, it works precisely when p is a Sophie Germain prime!
@YSCU261
The roots of a quadratic equation of the form x^2-bx+c=0 satisfy the following equations :
r1+r2=b
r1*r2=c
So we have that the solutions of xy = x+y can be expressed as the solutions to the quadratic equation x^2-bx+b for all b
this can be extended to x+y+z=xyz and so on with vieta's formulas
20
-
20
-
20
-
20
-
20
-
20
-
20
-
20
-
20
-
20
-
20
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
19
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
18
-
17
-
17
-
17
-
17
-
17
-
17
-
17
-
17
-
17
-
17
-
17
-
17
-
17
-
17
-
17
-
17
-
Hey Mathologer,
My solution to the Strand Problem has a much more intuitive way to find the x^2+x = 2y^2 identity.
It helps a lot to instead write in this way:
x(x+1)/2 = y^2
x is the largest number house, and 1 is the 1st house. The sum of all house numbers is the average of the highest house x and 1, so (x+1)/2, multiplied by the number of houses there are, which is x.
Ergo, sum of all houses is x(x+1)/2.
If the sum of all houses is equal to a perfect square, y^2, then the occupant lives in the house y.
After you reach this point in the puzzle, you can solve with wolfram to find all solutions.
(1,1) (6,8) (35,49) (204,288) (1189,1681)
Wolfram gives an equivalent formula to the one you reach at end of video. Would have done it myself but too lazy 🙂
17
-
17
-
17
-
17
-
17
-
I was introduced to the Steinbach ratios 8 years ago. When I am working with the heptagon ratios, I prefer using the three side lengths of a heptagon with the smallest side negated:
a=2sin(τ/7), b=2sin(2τ/7), c=2sin(4τ/7)
Using these values, any polynomial f(x,y,z) such that f(a,b,c)=0 also has f(b,c,a)=0 and f(c,a,b)=0
It also has the property that ab+ca+bc=a+b+c+abc=a²+b²+c²-7=0
Also, to expand on your formulas at 43:00, A(m,n)=C(m,n+1), B(m,n)=C(m+1,n),
and C(m,n)=(1*(r/1)^m*(s/1)^n+r²*(-s/r)^m*(1/r)^n+s²*(1/s)^m*(-r/s)^n)/(1+r²+s²)
17
-
17
-
17
-
17
-
17
-
17
-
17
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
16
-
Programming challenge (did this around 13:41): because all of the coins evenly divide a dollar, it must be the case that a sum of coins to a multiple of a dollar can be separated into some groups of coins in which no single denomination adds up to a dollar, and the rest, in which each single denomination adds up to a multiple of a dollar. The former cases can be enumerated by doing the product trick without the exponent of 100, and taking the coefficient of each exponent divisible by 100, including 0. There are ways to make change like this for zero through four dollars inclusive. This gives us one part of the solution. The other part is to determine how many ways the remainder could be made. But this is a simpler problem, because it's equivalent to asking "How many ways are there to add five (number of denominations - 1) boundaries to a set of a given size" which is equivalent to asking for (size of set + 5) choose 5, which can be hard-coded as a sequence of multiplications followed by a division. (I could have tried expanding out the polynomial explicitly, but that sounds like effort).
So, the final answer is to, for each amount of dollars that can be made without using a dollar's worth of a single denomination, multiply that by the number of combinations for the remaining dollars.
I coded this up in Python, and it's pretty fast. I just started added zeroes to the exponent on the ten, and it was pretty acceptable performance up to 2 * 10 ^ 100,000. I'll post the value for 2 * 10 ^ 1,000,000 when it finishes, because that's much slower. Okay, yeah, that took a few minutes. One sec... Oh geez, it overflowed the buffer.
I'm not pasting five million digits in here. See https://pastebin.com/MA2a9p3R
EDIT: At 23:02 I'm seeing the same coefficients in the center row that I had my program calculate for "ways to make dollars without a single denomination adding up to a dollar" So I guess this is going in the same direction.
16
-
16
-
16
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
@ravenecho2410 In fact some actuaries are fantastic at pure maths but most are what you would say “good at maths”. It’s prerequisite to study highest level at school to enter courses. As you’d know, but many wouldn’t, actuarial studies is a discipline of applied maths combining a diverse range of fields such as demography, finance, investment, commerce, marketing, accounting, risk analysis, modelling to problems in insurance, banking, health, pensions etc. So the emphasis is on applied maths to real world problem solving. Being brilliant at calculus or number theory is not of much use. if your friend chose another calling where such maths fields are useful, that’s fine. My son is an aeronautical engineer, who despite all the theoretical maths work at uni, discovered real world work generally didn’t require it.
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
15
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
Great video! Another way I noticed of describing why that (1st power) integer sums pattern works is that the left hand side always begins with a square number n^2, and apart from that term, there are n terms on each side, with each of the terms on the right side each having a term on the left side which it is exactly n larger than. For example, with 9+10+11+12 = 13+14+15, the terms 13, 14, and 15 each have a term on the left hand side they are 3 larger than (10, 11, and 12) making the difference between those be 3 times 3, which equals the square number 9 on the left.
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
Since 2020/4 = 505 which has prime factors 5 and 101, we just have the power set of {5,101} which is {1,5,101,505}. Since both 5 and 101 are congruent to 1 mod 4, they are both good and so are all their products. You definitely gave us an easy one. So 4 good - 0 bad gives us 16 ways of writing 2020 as the product of squares.
To find them, we'll first use the usual trick of dividing out the 4's. We'll find the ways to write 505 as squares and multiply each component by the square root of 4. True confession time: I'm adapting this from work by Dario Alejandro Alpern, whose fsquares program I ported to Gnu GMP.
We can find the ways to write 505 by finding the way to write its factors and using the fact that
(a^2+b^2) (A^2+B^2) = (aA+bB)^2 + (aB-bA)^2
We'll find the solutions in positive integers, and then convert each such solution (a,b) into 4 solutions, {(a,b)(-a,b)(a,-b)(-a,-b)}.
We know that every prime congruent to 1 mod 4 is the sum of two squares. For 5 this is easy: 5 = 1^1 + 2^2 and that's all. For 101 it's not hard either: 101=1^2+10^2, and this confirms that you are pitching us your softest softball. A quick check vs {49, 64, 81} confirms that this is the only way to write 101. Again, this is just the positive/positive solutions.
So we have:
(1^2+2^2)(1^2+10^2) = (1+20)^2 + (10-2)^2, which gives us:
505 = 21^2 + 8^2
2020 = 2*21^2 + 2*8^2 = 4*441 + 4*64 = 1764 + 256
and consequently 3 other solutions,
2020 = (-42)^2 + (16)^2 = (42)^2+(-16)^2 = (-42)^2+(-16)^2.
We also have:
(1^2+(-2)^2)(1^2+10^2) = (1-20)^2 + (10+2)^2 = (-19)^2 + 12^2 = 361 + 144
Thus we find 8 ways of writing 2020:
2020 = 42^2+16^2 = -42^2+16^2 = 42^2+-16^2 = -42^2+-16^2 = 19^2+12^2 = -19^2+12^2 = 19^2+-12^2 = -19^2+-12^2
According to the formula, there must be 8 other solutions out there, but I'm not seeing the permutation of these equations that gives them.
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
14
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
Interesting video! I ran into vortex maths videos some years ago. I drew the number circle in photoshop and played around with it awhile, trying to understand what I was doing. I'm not great at maths so in the end I didn't properly understand what you taught in this video, but I did come to the conclusion back then that the vortex diagram was just an emergent feature of our arbitrary numbering system and you could make any number of them and make them seem mystical by highlighting patterns and features. It does make pretty patterns.
Was Tesla really fixated to 3,6,9? What I was left wondering then and now is what kind of number theories Tesla was really thinking about! I don't think I've seen... objective video on the topic, and he must've had better stuff in his genius brain than what the youtube videos on vortex math are showing!
This is a topic for a different channel, but did Tesla really believe that 3,6 and 9 were the key to the universe, and if so, surely not because of the "vortex diagram"? He's a smart guy so he must've understood that our numbering system is just one of many. What do we actually know about his fixation with numbers? Is the whole vortex math thing even in anyway linked to Tesla, other than by coincidence through youtube conspiracy mythos?
Anyone know good sources that try to understand how Tesla thought by referencing his notes and such?
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
13
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
I did a bit of trigonometry to express the six angles with the coloured dots in terms of the angles of the given triangle. Here's what I figured out. (I'm sure this is known to the triangle experts.)
With the usual notation, let's call A, B, C the points of the triangle, a, b, c the edges and alf, bet, gam the angles. The median through A divides alf into the angles alf_b and alf_c (to the side of the edges b and c respectively). Similarly, bet=bet_c+bet_a and gam=gam_a+gam_b.
With this, one gets:
cot alf_b = 2 cot alf + cot gam,
cot alf_c = 2 cot alf + cot bet
and two similar pairs of equations. (The proof uses the law of sines and the addition formula for cot.) Btw., it can be checked that cot(alf_b+alf_c)=cot(alf).
Now the folded triangle has angles
alf_F = bet_a + gam_a,
bet_F = gam_b + alf_b,
gam_F = alf_c + bet_c,
and one obtains
cot alf_F=(-cot alf + 2 cot bet + 2 cot gam)/3
and two similar expressions for cot bet_F and cot gam_F, i.e. a linear relation between the cotangents of the angles!
So, if one forms a 3-vector from the cotangents of the angles, then the folding operation from the video is the multiplication of this vector with the 3×3-matrix M which has -1/3 on the diagonal and +2/3 in all other entries. This matrix satisfies M^2=1, reflecting the fact that folding twice reproduces the triangle up to size.
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
12
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
Professor Polster,
The Mathologer is THE best thing that happens to me in all of the media including Internet, blogs, youtube, social media, podcasts, TV, radio, movies, lectures, courses, and print. I can not describe how much pleasure I derive from watching Mathologer. At times, it's like a detective solving a crime (the crime being the difficult math problem at hand). Thank YOU, MARTY ROSS and OTHER GOOD PEOPLE who make Mathologer, such a good unusual program, possible. May you continue to produce your programs for at least another 50 years.
Behnam Ashjari, PhD EE
11
-
11
-
11
-
11
-
11
-
13:00 Spoilers below!
Using a=1, b=c=d=...=2, the highlighted sequence is
1, 2+2, 3+2+4, 4+2+4+6, ...
In each term of the sequence, we can split up the first term of the sum into k 1s and spread them out over the rest of the terms of the sum, giving
1, 1+(1+2), 1+(1+2)+(1+4), 1+(1+2)+(1+4)+(1+6), ...
Each term of the sequence is the sum of the first k odd numbers, which are the squares.
That means that the sequence generated by the squares is
1, 2*1^2, 3*2^2*1^2, 4*3^2*2^2*1^2, ...
In general, the kth term is k(k-1)!^2=k!(k-1)!
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
At first sight, there is something slightly unsatisfying about Aumann's game-theoretic solution, because it assumes a happiness function. However, standing a bit further back, among the uncertainties of the original lending decisions might be ignorance of where you stand in the list of creditors, ordered by size. And another might be uncertainty what your happiness function would be, at the time the dividend is being determined. From this perspective (originated, as far as I remember, by Harsanyi) there is an intuitive basis for wanting small creditors to be favoured when the dividend pool is small relative to the total liability of the estate, and for them to be subordinated to large creditors when the pool is larger, but still deficient. It could be argued that it is an even greater merit of the Talmudic solution that it responds to this intuition than that it is optimal given the Aumann utility function. It is also an attractive feature of the hydraulic model that it visually exhibits the ranking between creditors when there is a super-dividend, as well as generating the solution at every level (whereas the garment model generates the solution without showing how it works).
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
11
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
Ah, I did it. Here is the proof of the cube shadow theorem (with terms defined at 13:50):
Without loss of generality, assume S is the lowest point of the cube. The area of the shadow is then just the area of the parallelograms formed by the projections of the faces spanned by (u, v), (v, w), and (w, u) onto the x-y plane, respectively. These areas are given by the z components of the cross products u x v, v x w, and w x u. By construction, these cross products are just the vectors w, v, and u, so the area is given by the sum of the z components of all three vectors.
For the length of the line projection, assume (without loss of generality) that S is at height z=0. Since we also assumed it is the lowest point, the distance between S and the opposite vertex u + w + v (which is the highest vertex by symmetry) is the length of the z projection. Therefore, this length is also the sum of the z components of u, v, and w.
This completes the proof.
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
10
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
Extremely cool! Made me realize there’s a very practical strategy for handling these problems that, in my opinion, beats the pure programming AND pure math approach: programming this sort of thing correctly is nontrivial, and even with a smart implementation isn’t that fast (my quick and dirty python version takes about 20 seconds for $2500 - I just remembered I only bothered with 1, 5, 10, and 25 cent coins). The pure math approach to obtain a closed form expression through generating functions is mechanical and straightforward (and beautiful) but incredibly unwieldy for generating functions of this size. Even if you use Mathematica to do the algebra, you still face the problem of convincing yourself you didn’t make a mistake in writing the code! My approach, inspired by your video, is to simply observe that since the generating function is a rational function, its coefficients will be exponentials times polynomials in the index (and I think a bit more reasoning gets you to the fact that if you only extract coefficients with the same residue mod 100, they’ll just be polynomials (as you show for residue 0, of course)). For N coins, this polynomial will be of degree N-1. So pages and pages of computer algebra, in the end, are just to get N rational numbers! A much more direct way is to write a piece of code that is just barely fast enough to deliver N values of the function, and then solve the resulting linear system for the coefficients of the polynomial! No need for super slick iterative implementations, which will top out way before a million of dollars anyway; no need for huge amounts of algebra. Fast, easy, and delivers an essentially O(1) algorithm for a huge number of problems (ie. anything with a rational generating function, so anything produced by a finite state machine)!
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
I've been recently working on Tower of Hanoi animations for my channel and in the process discovered some of the stuff mentioned in the video, for example the algorithm to solve Tower of Hanoi without recursion by moving the smallest disk every second turn. I also proved why it works.
Answer to challenge questions:
Direction to move the smallest disk is: if (n is even) then Right else Left
That's because to move (N)-tower in some direction, you first have to move (N-1)-tower in the opposite direction, so the direction to move the smallest disk changes every time and it starts with Left for N=1 disk
The number of turns is (as mentioned in the video) equal to the number of reachable states, which is = 3^n, because:
there are 'n' disks, each disk can be placed in one of 3 towers, so there are 3^n valid states. Each valid state is reachable with a standard recursive algorithm, so there are 3^n reachable states.
Another interesting puzzle for you:
given a number of disks (N) and a number of turns (T), how do you quickly find the state of all disks in the optimal solution of N disks after making T turns?
I had to solve this puzzle while making my Tower of Hanoi animations, so I can instantly preview my animation at any moment in time (instead of re-playing it from the start), which becomes increasingly important with several-hour animations. Check out my channel for these animations: I've already published the shorter ones and longer once, including 10-hour 16-disk solution animation, coming soon:)
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
9
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
Hallo Burkard,
Ich schreibe Dir einfach mal auf Deutsch, weil mir das ein bisschen einfacher fällt.
Ich möchte Dir einfach mal Danke sagen! Dein Video zur quadratischen Reziprozität hat mich zum ersten Mal mit der modularen Arithmetik in Kontakt gebracht und, was soll ich sagen, jetzt bin ich im JugendForscht-Wettbewerb (in der Altersparte "Schüler Experimentieren", weil ich noch etwas jünger bin...) Landessieger in Berlin mit einem Projekt aus dem Fachbereich Mathematik, inspiriert durch Dich! Falls es Dich interessiert, die Preisverleihung ist auf dem Youtube-Kanal des innoCampus der TU Berlin zu finden. (Es ist ja alles digital gewesen; ich bin ab 40:24).
Es ist wirklich schön und auch irgendwie beeindruckend, was für ein Projekt aus dieser Inspiration und diesem tollen Video entstanden ist. Ich weiß noch, als ich diesen ersten "Wow!"-Moment hatte und bis nach Mitternacht wach war, um die ersten Ansätze auszuarbeiten. Ich finde es überwältigend, was die Mathematik bisher alles Großartiges bereitgehalten hat, was ich dann entdecken durfte, und das alles nur Dank Dir! Ich bin mittlerweile sogar soweit, dass wieder einen Bogen zur quadratischen Reziprozität selbst bekommen habe! ;)
Aber auch generell möchte ich Dir einfach mal danken für Deine tollen Videos, die mir durch diese schöne Mathematik immer wieder den Tag versüßen, wenn nicht sogar der Höhepunkt sind!
Sollte Dich vielleicht interessieren, was ich so gemacht habe, kannst Du mich gerne kontaktieren!
Viele Grüße und noch einen wunderschönen Tag, Tim
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
Hero's formula for the area of a triangle is one of those things introduced as a curiosity in a math textbook way back in middle school that I never really got a handle on. To a sixth grader, that is an impressively complex formula for an ancient to have discovered, and no proof was forthcoming. Through high school I saw it a couple more times, always in passing, a sort of neat oddity that seems compact but rarely gets used in practice. It was really neat to see an intuitive proof and motivation after all these years.
That said, it doesn't exactly seem useful. Even if you somehow do know the lengths of a triangle but not its angles, this formula is still not the fastest way to find the area. Typically, if you're doing this by hand, you will either have a table of square roots (for Hero's method) or of logs and logs of sines (for the law of sines method). That method is still faster, because you skip all the multiplication steps. If you want to compute the area of the triangle with a computer, you can use Newton's method to get the square root, and I assume Heron's formula really is faster. But the thing is, you basically never know all the side lengths of a triangle (and nothing else) before trying to find its area. Rather, you probably have coordinates, in which case the shoelace formula is by far the fastest.
So like, what is this formula actually good for? Is it just a novelty like the quartic formula? If it's never used, then no, I don't think it should be taught as part of a standard curriculum. The brief mentions in books for interested students are probably enough. There is so much I want to add to the math curriculum, and the curriculum is already packed as it is. It's hard to justify cramming in more random formulas to teach, prove, and memorize.
(BTW, although the phrase "Heron's formula" is seen pretty often in mathematical texts, in pretty much all other contexts in English, "Hero" is far more common than "Heron." Similarly, we say "Plato" rather than "Platon." The practice of Latinizing ancient Greek names is pretty standard in English. In classical Latin, the nominative singular would be "HERO," and the genitive singular would be "HERONIS." Since the Latin stem is still Heron-, the English adjective would be "Heronic" rather than "Heroic." Again, that's like the adjective "Platonic" rather than "Platoic. Other examples include "Pluto/Plutonic" and "Apollo/Apollonic." Admittedly, there are some exceptions, like the word "gnomon.")
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
8
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
Thank you for this great video. A long time ago I heard that Zagier did a one-sentence-proof without knowing what it was until two weeks ago. I did a bit of thinking on my own and want to share what I found (probably not as the first one) because it might be interesting.
In his original paper Zagier states that his proof is not constructive. In itself both involutions (the trivial t:(x,y,z) --> (x,z,y) and the zagier-involution z as discribed in the video) don't give many new solutions starting from a given one. But combined they lead from the trivial solution to the critical, from the fixpoint of the zagier-involution F := (1,1,k) to the fixpoint of the trivial involution t.
Proof (sry no latex here): Let n be the smallest integer with (z*t)^n(F) = F. So t*(z*t)^(n-1)(F) = F (multiply by z on both sides). And therefore (t*z)^m * t * (z*t)^m (F) = F with m = (n-1)/2. Bringing (t*z)^m to the other side proofs that (z*t)^m (F) is a (the) fixpoint of the trivial involution, ie a critical solution.
Note that n is always odd, assuming n is even results in a contradiction: If n is even we have t*(z*t)^k * z * (t*z)^k * t(F) = F with k=(n-2)/2. So again we see that (t*z)^k*t(F) is a fixpoint, this time of z, and therefore equals F. Multiplying by z gives us (z*t)^(k+1)(F) = F contradicting the choice of n.
7
-
7
-
7
-
7
-
7
-
Nice work! Gives me some flashbacks to Paul Lockhart's lovely book Measurement, which is like the one textbook that I, as a soon-to-be math teacher, actually really like; of course, I can't base my teaching on it because of these soul-crushing standardized requirements and tests, though (I guess I can at least use it for inspiration). One of my favorite related proofs when it comes to understanding the logarithm based on the graph of the function defined by f(x)=1/x that could have been mentioned, though, is a lovely visual argument that the alternating harmonic series sums to the natural logarithm of two. You restrict yourself to the little region from 1 to 2 and ponder the area under the graph – it's the natural logarithm of two, of course, but you can also approximate it with rectangles. If you use a 1x1 square, you're overshooting a lot, but you can remedy that by taking a rectangle out horizontally to remove the second half, where the overcounting is most egregious. You then add the smallest rectangle you can that still covers the desired area in that region, which has a height of 1/(3/2)=2/3 and a length of 1/2, so its area is a third. I think most of you will know where this is going, right? Our estimation, which still overcounts but is much more accurate, is 1 - 1/2 + 1/3. Now, we remove the second half of the first rectangle (area 1/4) and add a better fit back in that has an area of 1/5, and we also remove the second half of the second big rectangle we have (area 1/6) to get a better approximation with an area of 1/7. And so on and so forth – it's not too hard to convince yourself that this approximation strategy does indeed amount amount to evaluating partial sums of the alternating harmonic series, and the argument shows that they approach the logarithm of 2. Isn't that beautiful, even if it's tricky to describe everything just with words? This is exactly the kind of thing I want to teach eventually.
7
-
7
-
7
-
In fact, I've seen these diagrams as divisibility test. Make the multiplier equal to 10 (or whichever base you're working in), and the modulus equal to the divisor you want to test divisibility/remainder by. Then, with the arrows as you suggest in your third wish, start by moving clockwise from 0 as many steps as the first (most significant) digit, then follow exactly one arrow in the diagram, move clockwise as many steps as the second digit, follow one arrow, so on. Do not follow an arrow after your last digit. If you follow these instructions correctly, you'll end up at the remainder (and if and only if you end up at zero, the number is divisible by the modulus).
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
The primitive pythagorean triples are given by the general formula (a,b,c)=(u^2-v^2, 2uv, u^2+v^2) where u and v are positive integers, u>v, coprime, and of different parity (i.e. one is odd the other even). This shows that the 4k+1 primes are always also pythagorean triples' hypotenuses. The converse is not true though, for example (7, 24, 25) is a primitive Pythagorean triple but 25 is not prime.
7
-
Taking a shot at the "Multiplication Partition" problem at the end...
Empirically, the numbers seem to follow the pattern F(2n), where F(n) is the Fibonacci function. So the pattern is every-other Fibonacci number, henceforth called the "Skiponacci sequence". I will prove the hypothesis that the sum of products generated from partitions of the number n follows the Skiponacci sequence.
Here, S(n) is the Skiponacci function, and P(n) is the partition-product-sum function.
Firstly, analizing the stacks of equations, we can utilize the "recursion" mentioned early in the video. Looking at the example given for n=4, we see that the products are:
4
3 * 1
1 * 3
2 * 2
2 * 1 * 1
1 * 2 * 1
1 * 1 * 2
1 * 1 * 1 * 1
We focus on the products ending with "1", and removing the "* 1" we see:
3
2 * 1
1 * 2
1 * 1 * 1
Oh look, its the products for n=3 ! Looking at the products ending with "2" and removing it:
2
1 * 1
Its the products for n=2. The pattern is becoming clearer. Looking at the remaining products:
4
1 * 3
The "1 * 3" clearly follows the pattern, being 3 times the n=1 product. The 4 sticks out, but for now its easy to write it off as just "n". The final formula for this pattern is:
P(n) = P(n-1) + 2P(n-2) + ... + (n-1)P(1) + n
The reason for this formula makes sense. The "recursion" is because the partition products that are multiplied by 2 are made from partitions that are 2 less than n. Hence the "+2" in the partition list becoming a "*2", giving us the 2*P(n-2) part of the equation.
Now how does this fit into the Skiponacci sequence? It becomes clearer if we write the terms out into a pyramid. For instance, for n=5, the answer is the sum of these numbers. Each row has n copies of P(n-1), except the last row, which is written as n "1"s, for the "+n" term.
21
8 8
3 3 3
1 1 1 1
1 1 1 1 1
To aid in making sense of this, here is the pyramid for n=4:
8
3 3
1 1 1
1 1 1 1
Notice the recursion? The n=5 pyramid contains the n=4, just with the extra diagonal. This makes sense, since every time n increases by 1, each P(n-k) factor's coefficient increases by 1 (and the "+n" term increases by 1, naturally). All this means that this equation holds:
P(n) - P(n-1) = P(n-1) + P(n-2) + ... + P(1) + 1.
This gives us a neater equation for P(n) if you add P(n-1) to both sides, but for now lets test our hypothesis and replace P(n) with S(n).
S(n) - S(n-1) = S(n-1) + S(n-2) + ... + S(1) + 1
The left side is easy to simplify, because S(n) = F(2n)
S(n) - S(n-1)
F(2n) - F(2n-2)
F(2n-1)
For the right side, we can recursively replace the two right-most elements with another fibonacci number, until we are left with F(2n-1)
S(n-1) + S(n-2) + ... + S(1) + 1
F(2n-2) + F(2n-4) + ... + F(4) + F(2) + F(1)
F(2n-2) + F(2n-4) + ... + F(6) + F(4) + F(3)
F(2n-2) + F(2n-4) + ... + F(8) + F(6) + F(5)
...
F(2n-2) + F(2n-4) + F(2n-5)
F(2n-2) + F(2n-3)
F(2n-1)
This leaves us with this equation, which is obviously true:
F(2n-1) = F(2n-1)
Therefore, because we were able to replace P(n) with S(n) in our equation, we showed that P(n) = S(n). QED.
Also I did the math and found that the general equation for S(n) and P(n):
S(n) = 2/sqrt(5) * sinh(2 * ln((1+sqrt(5))/2) * n)
This is a long way of saying I think the next number is 55 :)
7
-
Congrats on 100 videos, mate. You really made it as a maths educator and content creator on YouTube, and I'm looking forward to seeing you do even better in the future. Hope you blow our minds again with each video you make
That said...
CHALLENGES!
11:18
I have no army of middle school minions but I am still ready to attack
Same reasoning as before. This time we start with the second paraboloid, the one carved from the cylinder. Makes the maths a little easier.
The paraboloid has radius R and height H. Cutting it at height h will leave a ring with outer radius R and inner radius r.
The paraboloid is modeled after a parabola y=ax^2, and so we should have H = aR^2 and h = ar^2. So it's possible to solve for r and get r = Rsqrt(h/H).
The ring thus has area pi * R^2 - pi * R^2 * h/H, or piR^2(1-h/H).
The first paraboloid should also have that area. Thus its radius should be Rsqrt(1-h/H).
Now the inverted paraboloid can be modeled by another quadratic, but the important takeaways are H = bR^2 and H - h = br^2. Solving for r this time gives Rsqrt(1-h/H), exactly the same as what was predicted by the circles area argument.
Or you can use integrals. Whatever floats your boat
17:42
The layers of the onion look almost like surface areas stuck together. That can be written as: V(R) = the integral from 0 to R of SA(r)dr
By FTC1, we can also write this as V'(R) = SA(R)
And so the derivative of the volume is the surface area. Even in 420 dimensions.
18:04
The base of the cylinder has area piR^2. The height is 2R, and the circumference is 2piR. In total, the surface area of the cylinder is 6piR.
With the surface area of the sphere being 4piR, the ratio of the surface areas of the two shapes really is 3:2.
7
-
I am absolutely trash at math but when I stumbled across a numericology video on 369 and they explained the theory I immediately commented that it was clearly because we use a base ten system. For instance base 14 was used in the Middle East for a long time. Based on counting the inner pads of the fingers, I am a behavioral scientist and Buddhist philosopher so I am well studied but I really am horrible at math. I can’t even do basic division if the numbers go into the hundreds without taking a long time. It was actually obvious because from a behavioral psychological point of view when you have something people can’t explain the culprit is most often a blind spot in your paradigm. Anytime we attempt to build a thought model of what we believe is true we create a type of tunnel vision because our imagination becomes constrained by your base assumptions. If you think a certain action is always bad you will be blind to anything good that comes from that action and when it’s pointed out you won’t be readily able to accept the evidence, it literally requires an individual to rewrite the most fundamental assumptions about life because our thoughts are like cards stacked upon each-other our conclusions come as a result of previous conclusions, as this happens our ability to recognize anything new or aberrant narrows and our ability to explain concepts becomes reliant upon what we have “figured out” thus creating a blindness in our personal philosophy. Most people assume that math is always a base ten system. But there are as many systems as there are numbers and historically societies that weren’t connected to the western tradition came up with novel calculation systems. Most notably the Indians who were the first to recognize the need for the number zero. Incendently this comes from Hindu philosophy which posited emptiness or void and singularity or oneness. Both of these concepts are represented by the number zero which like consciousness can be posited and observed in its function but cannot be qualified or countrified as it is immaterial like space. Anyways from a purely psychological perspective the answer was obvious I didn’t need to comprehend the math to see that this was merely a blind spot made by a false axiom.
Rant about Ether vs Space:
I will add that Tesla was correct about ether. When people talk about bending space they are talking about bending ether. Space is not a medium and cannot be acted upon it is a potentiality for expansion and position. If something bends in a given position that which is bent must rest within something. Without a medium the distance between all things would be zero, the fact that we can imagine folding space means that something which has qualities is being manipulated, however actual void ness, emptiness, the expanse of nothing allows for all laws of physics to be unobstructed and to function within their own constraints. If space was not completely void there would be no possibility for anything to bend, 0 the absence of all is what gives freedom to bend the ether. When you bend ether that fold is still a location in space and relies upon the void as a platform for existence. Nothing has freedom or capability to be anything without a complete void to appear within. This void exists outside of time and was present as the externality of the singularity but void or emptiness is not a thing and so it is completely logically coherent to say nothing existed outside the singularity because emptiness has no qualities and cannot be proven to exist. It can only be pointed out as being the function of unobstructed freedom of material phenomena. If space was void you could not bend it, it has qualities and so it cannot be said to be the expanse which allowed for the Big Bang. Without this freedom of nothingness and infinite indescribable void the singularity would never have been able to exspand as it’s size would be fixed.
Even the diagrams of bending space require the illustration of grid planes which automatically describes a medium which has properties. True space ie the infinite void has no properties it is the absence of properties and can only be proven to exist because the bending of space happens within a free unobstructing openness. It a real shame because I suspect this distinction is likely an incredible resource for computation. Again if there was no ether than all of existence would not have any measurable distance, distance in space is a fixed property which is measurable as space/time, but something has to exist that undergirds and existed before space time in order to allow it to come into being and to exist in an exspansive way. Nothingness is that thing. Without the absence of all nothing can expand out into infinity. Without void everything would be permanently constrained. Space itself is a constraint as it is governed by the speed of light.
Space what I would call ether acts on objects because it is a medium with properties subject to the qualities of other objects like planets whose gravity actually warps it. True void is inert and has no properties it cannot be affected or measured it is merely the absence of all qualities which allows for qualities to come into being without being constrained.
Can you comprehend emptiness or is your paradigm constrained by preconceived notions of what your science books have told you. I have made a coherent argument that we are missing a fundamental piece of reality in our current scientific paradigm. Exists in math as zero it is already proven to exist in math and many higher calculations cannot even be done without zero proving that as a model of reality zero is acknowledged as indispensable to model our reality. ... yet we ignore emptiness. We already know zero cannot be affected by any other number zero cannot be divided subtracted or multiplied because it is an absence of qualities. It can be added to because it is inert and unobstructed and so it is a vessel which is truly and absolutely infinite in its allowance for qualities to exist within it. We already know all this yet our current models of physics refuse to acknowledge that what we call space has qualities that occupy the infinite void as a basis for all. Without void we cannot posit existence we understand that two things cannot occupy the same place which means for anything to occupy any space there must be a void which does not obstruct position, characteristics or behavior. As I said if what we call space can bend it is bending within emptiness. There must be a basis for reality which is without qualities or existence would be constrained. If what we call space was truly the basis for reality it wouldn’t have qualities. If space was not a medium there would be no time or distance to travel between objects, you could move from earth to Mars in an instant because there would be nothing to travel through. Traveling through something means it is a medium especially if it has coherent properties, it’s qualities are consistent not chaotic meaning like atoms it has a particular measurable property which can be counted on to act and behave in a uniform manner. Having inherent measurable qualities means it exists as an object.
Please don’t reply if you’re just going to quote text book physics as my entire point is to point out something missing in the current assumed model of reality. We know the number zero is real functional and vital yet we deny it’s actual existence in the world despite our model of reality requiring zero to be calculated. The absence of everything came before anything this should be obvious as only an absence of things allows for an unobstructed exspansion. Without an absence of qualities qualities could not come into being without being affected by constraints. If there was no external void the singularity would have nothing to expand to.
This is important for positing the big crush as there may be a limit to this ether or space time, it may have an elastic property where it can only stretch so far, space time may be capable of thinning metaphorically like a gas perhaps creating a gravitational vacuum. Just like as gas thins it creates a vacuum which can literally implode or crush objects. Once we see space as a medium this new paradigm means measuring the qualities of this ether may actually allow for incredible breakthroughs. I believe humanity is constraining it’s progress by resting on a preconceived notion that space/time is the undergirding foundation of reality when In my mind you always must have nothing before you have something. If
7
-
Very interesting! This reminds me of a result I came across recently. Take a regular polygon positioned anywhere in a circle. Extend the sides to form two sets of segments with lengths x_1,x_2,...,x_n and y_1,y_2,...,y_n, so that each y_i is counterclockwise from x_i. Then applying power of point to each vertex, we get
x_k(L+y_(k+1))=y_k(L+x_(k-1)),
where L is the side length of the polygon. Adding all these equations and cancelling the x_ky_(k+1) terms, we get
x_1+x_2+...+x_n=y_1+y_2+...+y_n.
So the two groups of segments have equal sums.
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
7
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
A relatively easy proof:
Any time you finish adding positive numbers (i.e. any time you go above the limit), it doesn't do any harm to add up to a fixed number of extra positive numbers. Let's say add an arbitrary amount of extra positive numbers that is between 0 and 9.
Because the sequence of positive numbers converges to 0, the distance this gets you away from the target real number will be arbitrarily small (namely, in total 10 times the epsilon from the definition of being a zero-sequence), and so, if you do this, your rearrangement will still converge to the same real number.
But this means that any real number R on [0, 1) can be turned into a new rearrangement:
Take the decimal expansion of R. Then when creating your rearrangement and deciding how many extra positive numbers you want to add at the end of one consecutive sequence of positive numbers, just choose the next decimal digit of R, and add as many extra positive numbers as that digit is high.
Because the set of real numbers on [0, 1) is uncountably infinite, and each number gives a distinct rearrangement, the set of all rearrangements resulting in the same sum must also be uncountable.
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
You can also use the golden ratio to drastically improve the prime number theorem.
Number of primes less than (n/phi)^2=((n-n/phi)x(n/Ln(n-1))/2.
For example: n=641,894
Pi(157,380,656,251)
=6,361,970,514
n/Ln(n-1)=6,350,670,612
Using the golden ratio=6,360,264,553
Not sure if this shows some connection between phi, e and primes or just a neat trick that works really well.
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
2:34, check of the calculations up to that point:
For the equilateral triangle, the height is the length of one of its legs times the cosine of π/6, so the height equals 2*sqrt(3) times sqrt(3)/2 .
Thus, the area is half its base, i. e. sqrt(3), times 2 times sqrt(3) times sqrt(3) times 1/2 which amounts to 3*sqrt(3).
For the isosceles triangle with legs of the length 2+φ with a base of 2*φ, Pythagoras's theorem gives sqrt( (2+φ)^2 - φ^2 ) = sqrt(4 + 4φ + φ^2 - φ^2) = sqrt( 4(1+φ) ) .
The golden ratio is a solution to the the equation x^2 = x+1 , so sqrt( 4(1+φ) ) = sqrt( 4φ^2 ) = 2φ .
(Since we're talking about lengths, we can ignore negative results.)
Finally, the area of this isosceles triangle is then half its base times its height: φ times 2φ = 2φ^2 .
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
I've got a proof for the last problem distinct from ilonachan's very elegant sudoku-style proof, inasfar as grid-puzzle proofs can be distinct, giving the same fully general 2n*2m result.
I) Any solution must have an "Alternating Orientation" (equivalent solution under reflection over a diagonal or 90° rotation) where each row alternates between two colors, and the sequence of rows alternates between two such pairs of colors.
II)
a. For an even number of columns, an alternating orientation of a solution must have different colors for the first and last columns in each row.
b. For an even number of rows, an alternating orientation trivially has different colors between the first and last rows.
c. The colors in each of the four corners of an alternating orientation must be different.
III) Reflection along the diagonal or rotation by 90° preserves both the property of being a solution, and the span of the colors in the corners. (Trivial)
Proof of I:
Ia) If the first row alternated over a pair, the second row must contain the complement pair; and that row must therefore alternate, as, trivially, two adjacent cells cannot share a color, and there are only two colors in the row. Inductively, each successive row must also contain the complement pair to the previous row, and must itself alternate. This satisfies the definition of an alternating orientation.
Suppose there is a solution which does not have an alternating orientation.
By Ia, the first row does not alternate over a pair.
Without loss of generality, it can be shown that the following configurations remain:
i. abcd[...]
ii. abcb[...]
iii. abac[...]
Where [...] represents any sequence of successive terms, for grids of more than 4 columns.
We shall first assume configurations i and ii, together.
Both configuration i and ii start with the same three colors, and we may say they both satisfy of configuration 'abcx[...],' where 'x' is undefined.
As no steps of the remaining steps do not depend on cells in columns 3 or later for this orientation, configurations will be denoted with only the first three color values.
Cell (2, 2) is part of both the top-left 2x2 square, and the top-center 2x2 square. It must therefore not be the same color as the first two or second two cells in row 1. As all three colors are represented in those three cells, cell (2, 2) must be the 4th color.
Cells (1, 2) and (3, 2) are now determined, as they each have 3 colored neighbors.
In particular, row two is now in a configuration of 'cda.'
This configuration is isomorphic to configurations i. or ii. of row one. Therefore, each row can be determined inductively.
In particular, row three is also in a configuration 'abc,' which is the same as row one.
If we look at column 1, we find it must alternate between colors 'a' and 'c.' This means that if we reflect over the diagonal, we must have a solution where the first row, equivalent to the original solution's first column, must alternate over the pair of colors 'a' and 'c.'
This satisfies the condition of Ia), and so this reflected solution must be an alternating orientation.
This means the original solution had an alternating orientation, contradicting the premise for configurations i and ii.
We may use a similar argument for configuration iii, but shifting our perspective over by 1 column. We first remove column one from the solution, and observe that we now have a configuration isomorphic to configuration i or ii. As removing a column does not alter the property of being a solution, as it does not affect 2x2 grids not intersecting that column, the argumentation for configurations i/ii follow.
If we take the resultant alternating orientation, and then reflect this solution vertically so that the top row is now at the bottom, we may notice that our new top row is also alternating, as given by Ia. If we try to replace an appropriately reoriented row at the bottom corresponding to the column we removed here, the solution must still satisfy Ia, as the new top row is not changed.
The property of having an alternating orientation is trivially preserved under horizontal and vertical reflections. As such, this re-appended row must still be alternating, and our overall solution still has some alternating orientation.
I: QED
QED
(Sorry for the haphazard alterations to the proof of I, but I realized I needed to generalize it to get the full 2n*2m result, but that proved non-trivial.)
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
Excellent work! And yes, you should be wary of discontinuous "logarithms" lurking in the dark :) The axiom of choice guarantees their existence (but that also explains why you can't find any of them - because they're non-constructive).
Essentially, the multiplicative group of positive real numbers forms a vector space over Q (taking vector addition to be multiplication and scalar multiplication to be exponentiation). So the axiom of choice guarantees that there is a Hamel basis for the multiplicative group of positive real numbers over Q, and we can assume e is part of that basis. Similarly, the additive group of all real numbers is a vector space over Q, so the axiom of choice guarantees us a Hamel basis here, too, which we can assume contains 1. (Note that we can prove the dimension of both of these spaces over Q is |R|.)
Then we can create a function which maps the Hamel basis from the multiplicative positive real numbers to the Hamel basis of the additive real numbers, choosing e to get mapped to 1, and choosing it so that some basis element d of the multiplicative positive reals is mapped to something other than ln(d). But a function defined from a basis of one vector space to elements of another vector space defines a unique linear transformation between the two vector spaces. Let's call this linear transformation L. We can then check by the definition of these vector spaces and the definition of linear transformations that it has all the algebraic properties of the natural logarithm.
If a and b are two positive real numbers, then L(ab) = L(a)+L(b) because vector addition in the domain is actual multiplication but vector addition in the codomain is actual addition. For any positive real number a and any rational number c, we have L(a^c) = cL(a) since a^c is scalar multiplication of a by c in the domain but scalar multiplication in the codomain is actual multiplication. We also have L(1) = 0 since 1 and 0 are the identities respectively, and L(e) = 1 by choice. So we have the "properties" of a logarithm that you were able to pin down without continuity. But we have L(d) = something other than ln(d), so L is not the same thing as the natural log.
(As an added bonus, I made sure L is bijective too!)
But your argument pretty much shows that ln(x) is completely pinned down by
ln(e) = 1, ln(ab) = ln(a)+ln(b) for all real positive numbers a and b, and continuity.
So the only conclusion here is that this nasty L function cannot be continuous. And since it has the other properties of logarithms, it is in some sense a discontinuous "natural log".
6
-
6
-
6
-
6
-
6
-
6
-
6
-
6
-
I was thinking about that Pythagorean triple tree. If one associates going straight with a 1, going left with a 2 and going right with a 3, then one can encode every primitive Pythagorean triple (PPT) with a rational number between 0 and 1 in base 4 as follows:
0 corresponds to (3,4,5), 0.1 to (21,20,29), 0.2 to (5,8,17), 0.3 to (5,12,13), 0.12 to (77,36,85) and so forth. That way one gets a 1-1 correspondence with the finitely representable base 4 numbers between 0 (included) and 1 that do not contain the digit 0. The three families of PPTs correspond to the numbers 0.11111..., 0.22222 and 0.3333...
I wonder if one can get anything geometrically meaningful with that correspondence.
Can one interpret the periodic or irrational numbers as interesting infinite paths/families of PTTs in the tree?
What about interpreting numerical manipulations like multiplication of 0.1111 with to to get 0.2222 in terms of the associated PTTs?
6
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
I came up with a poem of my own for the cubic formula:
x^3 + n*x = m
x = cbrt(m/2 + sqrt(D)) + cbrt(m/2 - sqrt(D))
D = (m/2)^2 + (n/3)^3
When x cubed and x times n,
Are added and equal to m.
The values of x,
The goals of our quest,
Here's how to calculate them.
Cube roots to add,
Square roots they had,
Both of a term we'll call D.
Square half of m,
Cube third of n,
Add together and see.
Half of m, adds to the root,
First with a sign of plus.
Its little brother,
Is just like the other,
Except with a sign of minus.
Cube rooting time, of both the brothers,
Add up the roots with glee.
We found our first x,
But where is the next?
I know there have to be three.
With help from DeMoivre,
Who's theorem, we love ya,
There's cube roots all over the plane
Yes, they're complex,
But do not perplex,
A new kind of numbers we gain.
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
15:23, I arrived at the same answer as the Talmud, by the following reasoning: the contested child ("red") who has a possible claim to the 1st son's estate takes first pick, and is given half-way between what he would receive in the two scenarios: (1/2 + 1/3) / 2 = 5/12. What remains (7/12) is given to his two siblings, and is shared equally between them, so they each take 7/24.
My reason for giving the "red" child precedence was more of a practical than a legal one: it's messy to modify the fractions for all the children at once, and you can easily end up with something that doesn't add up to 1. By first settling the "red" child's claim, it's then trivial to divide what remains between the "green" children.
Interestingly, we can go about it the other way, settling the "green" claim first: (1/2 + 2/3) / 2 = 7/12, then 5/12 remain for the "red" child. So, the solution is the same regardless of who takes first pick, which is definitely a plus.
A more modern approach would be to assign the two scenarios some probability, then divide between them in proportion to how likely they are; but given that no information is available about the paternity of the "red" child, a 50-50 divide is kind of forced.
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
In hindsight, I saw my math teacher was showing us some of the math-related to famous anecdotes, like the sum 1..100 and also Collatz conjecture. Without ever mentioning it. As a planted memory, that might be a reminder. I like my math teacher very much. He was an Egyptian, and though he wasn't our main "Tutor" teacher, he invited the Leistuingskurs home and showed us a way to make a sweet (Mr. Tom) with peanuts and sugar sirup ourselves, without going to a shop in the school breaks and leave the school area, which actually was forbidden.
I got the idea to group 1+100,2,+99, too. It's not that far fetched to get this idea. But powers is another story. But thanks for another reminder of a great time in school.
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
I'm actually surprised and pleased with myself for being able to understand the solution from the clip alone, haha.
Of course it is such a simple that anyone should be able to understand it, was is not for he using a binary number representation, and most people being unfamiliar with binary. If, for instance, he decided for no particular reason to use 0 and 9, everyone who isn't familiar with binary would immediately have an easier time understanding it.
You can also describe the solution without using any numbers at all. You can just use the letters D and U, and show that the only two things that can happen to the sequence of letters is two Ds turn into Us, or a D swaps places with a U to its right.
As a result you either end up with two Ds next to each other, or they are separated by some run of Us. In the latter case you can move the rightmost one until its all the way to the right, at which point the only remaining option is to move the leftmost D closer to the rightmost one until you're back to the former case, meaning you must turn both Ds into Us.
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
I have a special relationship to Ptomelys Inequality because it's the first theorem where I came up with an original and beautiful proof of it. And it was actually because of you :D Back then, I was re-watching some video about proofs of Pythagoras and one of the proofs involved scaling the original triangle by the factors a, b and c and then rearranging the parts in a clever way. You can do the same thing with Ptolemys Inquality!
Take any quadrilateral with side lengths a, A, b, B, c, C like in your video but make sure that c is "in between" a and b. Take three copies of the quadrilateral and scale it by a, b and c respectively. Notice that two of your copies now have a side with length ab and diagonals ac and bc respectively. Join them on the side ab. Now notice that you can fit the third copy perfectly on the diagonals of the other two, as the third copy has two side lengths ac and bc too and the angle also matches. Put it in place and look carefully. You will have formed a triangle with side lengths aA, bB and cC. Ptolemys Inquality therefore follows. And the equality will hold exactly when the point lands exactly on the cC side. Which will happen exactly when the opposite angles of the original quadrilateral add up to 180 degrees which is exactly the case when it is circular. qed
I stand by my opinion that this proof is absolutely marvelous and the best proof of Ptolemys Inequality ever. And it's perfect for an animation so if you ever want to show it, I absolutely allow you to do so :)
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
Very nice geometric proofs. It might be also interesting to look at it a bit more algebraically. Let's say a',b',c' are the big parts (as in 2/3) of the respective medians. From Steiner theorem (or law of cosines) we have a'^2=-a^2/6+b^2/3+c^2/3 and similar equalities hold for b', c'. So if we represent the triangle by the vector (a^2,b^2,c^2), "folding" is just a matrix multiplication
(a'^2,b'^2,c'^2)=M*(a^2,b^2,c^2),
where
M=[[-1/6,1/3,1/3],[1/3,-1/6,1/3],[1/3,1/3,-1/6]].
Since M^2=I/4 (where I is the identity matrix), folding twice means making the squares of sides 4x smaller, i.e. scaling the triangle down by a factor of 2.
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
11:48
The box for 153^2 + 104^2 = 185^2 is
9 4
17 13.
The upper two number can be calculated via: b = sqrt((z - x) / 2), a = sqrt((z + x) / 2) - b, where x^2+y^2=z^2.
Then c = a+2b, d=a+b, in the form of:
a b
c d.
The path there is: right right right left
The algorithm to find the path to a given pythagorean triple in the tree: (x^2+y^2=z^2 , x,y,z ∈ N
(0) Reduce triple to primitive triple by dividing by their greatest common divisor gcd(x, y, z)
x = x / gcd(x, y, z)
y = y / gcd(x, y, z)
z = z / gcd(x, y, z)
(1) let a, b:
b = sqrt((z - x) / 2), a = sqrt((z + x) / 2) - b
if a is even or either of them is not an integer or either of them is 0, repeat the calculation with y instead of x!
(2) Calculate the Category of a and b:
C: N^2\{1,1} -> N
C(a,b) = 2 - floor(2 / ceil(2 * b / a)) for a != 1 or b != 1
if C(a, b) is undefined, jump to step (6)
(3) Save C(a, b) in memory
(4) Calculate a' and b'
let c = C(a, b)
{ a-2b for c = 0
a' = { 2b-a for c = 1
{ a for c = 2
{ b for c = 0
b' = { a-b for c = 1
{ b-a for c = 2
(5) Repeat from (2) with a = a' and b = b' as long as a != b and a * b != 0
(6) Read the saved C(a, b) from (3) in reverse order
0 corresponds to going left,
1 corresponds to going straigt,
2 corresponds to going right!
DONE!
Example code and Algorithm here:
https://pastebin.com/T71NP8Z9
JS-Applet here:
https://jsfiddle.net/ptu6f5cr/4/
5
-
5
-
Herr Dr Polster: That was lovely. My chemistry teacher was like you, so I became a chemist. My maths teachers were terrible; but I like (love?) math anyway. I did have the calculus, some algebra which included Power series (which I never got), some differentials, partials, numerical methods, and so on. But I never got power Series at all until I just watched the Power Series Master Class. I got through all of it in two nights; I will have to watch it again and get out a pencil and paper to get more (Erdos: "if your fingers don't hurt, you're not working enough"). I got most of it, except the animation at the end, and I'm sure I must have missed something as I'm not that smart. But, thank you so much!
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
Always, fascinating!
I followed, surprisingly, until the last example. That's when my brain went flat and would inflate no more, until patched with a second cup of tea.
Why, at the end of each video, is there a smile on my face, and a chuckle in my heart? Perhaps they have much to do with your own gentle encouragement along the way with a chuckle here and a giggle there, however, I'm also quite fond of that sweet tune, pleasingly harmonized on a collection of folksy strings, that serves as a segway and ending to your videos.
As always, thank you and I'll look forward to the next time. 😊
5
-
5
-
5
-
5
-
5
-
5
-
5
-
5
-
4
-
4:50 already every power of 9 plus one is special, because the triangles can be collapsed in like so: every 9th hexagon is sufficient to account for 9 generations in the future, so ignore the eight between that and the previous one. 10 is good, and a group of 82 would behave just like a group of 10 (with 8 between each one, 8*9 = 72, which is 82-10), and generation 10 of 82 hexagons would behave just like generation 1 of 10 hexagons. likewise generation 82 (assuming the first row is generation 1) is determined only by the top corners in the same way. Therefore 82 behaves like 10, and so 9^3 + 1 will behave to 82 the way 82 behaves to 10, and so on. My first expectation is that since there are 9 possible rules that this is the root behind why 9 is so special, but it may be that 4 generations also obey this property (powers of 3, plus 1 instead)
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
I like the Pigeons at the Olympiad proof best.
For the challenge at the end of chapter 3:
Because every repeating decimal can be written in the form
(x+y)/10^n + x/10^2n + x/10^3n ...
for some integers x,n,y.
We can factor out x, and we can rewrite:
1/10^n + 1/10^2n + 1/10^3n ...
as 1/(10^n - 1)
Noticing that when n is an integer, 10^n and (10^n)-1 are integers, we have the sum of two rationals,
y/10^n + x/(10^n - 1)
which is therefore rational, QED.
For the challenge at the end of chapter 6:
7 integers from 1 to 100 such that no 2 different collections from that set have the same sum:
{1,2,4,8,16,32,64}
This is easily seen by putting the sum in binary form. Each bit in the sum is set by exactly one number from the collection, so if collection A sets a given bit, disjoint collection B lacks the only element that can set it.
For the think-about-it pause at the beginning of chapter 7, I found a much harder and clunkier solution:
There are 2*3*4 = 48 permutation of 4 cards. That's almost enough to pick out 1 card among 52 - if only somebody removed 4 cards from the deck. But somebody did! The mystery card can't be any of the 4 cards you passed to your assistant, which leaves 48.
So we place some natural ordering on the 4 cards. Say, suit is the primary sort key and number/jack/queen/king is the secondary.
We also number the permutations of 4 items - so we have a table that gives us a bijection from a permutation like C D A B to/from a unique index.
And we number the remaining cards in the deck. Same idea: force an ordering, make a table.
Then we look up that card in the deck table, look up that index in the permutations, and arrange the cards in that permutation.
Our assistant does the reverse: Sees what permutation we handed him, looks up its index, constructs a deck table for the ordered deck minus the 4 cards he sees, looks up the permutation index in that, and announces that card as his answer.
Ta daah, except that we need to build large tables on the fly. On the plus side, we can do it for any mystery card, we don't have to choose it ourselves from among 5 cards.
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
I'm from Brazil, and I was taught Heron's formula very early on, around 3rd or 4th grade. It always puzzled me, because it seemed to have been dropped on my lap out of nowhere (as far as my education was concerned, it really was). Still, I'm personally thankful it happened, because my first memory of doing maths by myself, out of sheer curiosity and without any obligation to do so, was trying to verify it from the usual formula for the area of a triangle, equating the two and working through the algebra. I wasn't successful, of course, but it got me here eventually, so that's nice.
4
-
4
-
4
-
4
-
4
-
4
-
4
-
30:07 Minimum number of moves for 9 disk and 4 pegs (I think): 41 (^.^). You explained how to divide the disks into 3 superdisks for 8 disks but not for 9. So, just for fun, I tried every possible division in a very inefficient c++ code. I found that you can solve optimally in 41 moves if you divide the disks into 3 superdisks of 2, 3 and 4 disks (from top to down); also dividing them into 4 superdisks of 1, 1, 3 and 4; 1, 2, 2 and 4; AND 1, 2, 3 and 3. I didn't check if these divisions provide different solutions (Challenge for anybody?), but my guess is that, at least, the 3 and 4 superdisk divisions are different optimal solutions.
Calculation (hopefully) explained: Assume that you got to solve for 8 disks and 4 pegs, and you divide the 8 disks into 3 superdisks of 1, 3 and 4 disks (from top to down, like Mathologer did). Also note that f(n)=2^n-1 is the optimal number of moves to solve n disks and 3 pegs. Then, (hopefully) it's easy to see that the number of moves, following the Mathologer path, is equal to f(1)+f(3)+f(1)+f(4)+f(1)+f(3)+f(1)=4f(1)+2f(3)+f(4)=2^2 f(1)+2^1 f(3)+2^0 f(4). Note that the arguments of the f's add up to the 8 disks and also note the decreasing powers of 2. (hopefully) it's easy to see that, for n disks, 4 pegs and m superdisks of a1, a2, ..., am disks, such that a1+a2+...+am=n, the resulting number of moves is equal to 2^(m-1)f(a1)+2^(m-2)f(a2)+...+2^0 f(am). For the code, I simply apply this formula to every possible division of the 9 disks into superdisks. I cycle through the possible divisions in a very inefficient and lazy way: I go though all possible lists of between 1 and 9 elements with all possible values between 1 and 9, checking if the accumulation of the elements add up to 9. Seeing the formula, it's easy to see that it's not worth it to check superdisk divisions that are not in decreasing amounts of disks, but whatever.
Here's the code for anyone interested:
#include <iostream>
#include <vector>
#include <numeric>//accumulate algorithm
using namespace std;
int pow(int b,int e){//b to the power of e with e being a non negative integer
int ans=1;
for(int i=0;i<e;i++)ans*=b;
return ans;
}
int h3(int n){//optimal number of moves with 3 pegs
return pow(2,n)-1;
}
bool operator!=(vector<int>v0,vector<int>v1){//how to compare vectors
if(v0.size()!=v1.size())return 1;
for(size_t i=0;i<v0.size();i++)
if(v0[i]!=v1[i])return 1;
return 0;
}
int main(int argc, char *argv[]) {
int N=9;//total amount of disks
for(int i=1;i<=N;i++){//i is the number of superdisks used in this attempts
vector<int>v(i,1),v0=v;//v holds how many disk in each superdisk, v0 is the initial attempt
do{
if(accumulate(v.begin(),v.end(),0)==N){//the total amount of disks has to be N
int m=0;//total number of moves in this attempt
for(size_t i=0;i<v.size();i++){
m+=pow(2,i)*h3(v[i]);
cout<<"2^"<<i<<" f("<<v[i]<<")+";
}
cout<<'\b'<<'='<<m<<endl;
}
for(size_t i=0;i<v.size();i++)//cycle through all possible v's
if(v[i]==9)v[i]=1;
else{++v[i];break;}
} while(v!=v0);//all possible attempts has been searched when v loops back to v0
}
return 0;
}
Enjoy. (^.^)
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
The whole idea of "summoning" satan with pentagrams/pentacles comes from horror movies heavily misrepresentating thing.
First, the pentacle/pentagram, if pointed up, is a protection symbol, and has been used by many religions, including Christianity, Pagans, and others.
Second, if pointed down, the symbol is linked to Satanist, which doesn't really have the right name since we don't even believe in Satan, and most Satanist are basically atheists with a few more rules (which are mostly common sence, ex. dont hurt kids; dont hurt innocents; if someone invites you, respect them; and so on) and the few that do believe in an entity believe in either Lucifer (which is NOT the one from the Bible) or Baphomet, which represents balance as they are both female and male, light and dark, good and bad, and so on.
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
I asked ChatGPT to flesh out your idea Title: Pentagon’s Shadow
Synopsis:
In the heart of a bustling city, a series of seemingly random crimes baffle the local police. Each crime scene forms a perfect pentagon, its vertices marked by the locations of heinous acts: arson, robbery, assault, kidnapping, and murder. The pattern is unmistakable, but its origin and purpose remain elusive.
Main Characters:
1. Detective Laura Mathis: A seasoned detective with a sharp mind for patterns and an unconventional approach to solving crimes. She has a background in mathematics, which becomes pivotal in cracking the case.
2. Dr. Petr Novák: A reclusive mathematician known for his work on complex geometric theorems. His breakthrough, known as “Petr’s Miracle,” describes the transformation and center-shifting of pentagons, a theorem that holds the key to the mystery.
3. Lucas Grey: A brilliant but troubled mathematician turned criminal mastermind. Obsessed with proving his intellectual superiority, he uses Petr’s Miracle to orchestrate his crimes.
4. Captain James O’Neil: Laura’s superior, a pragmatic and seasoned officer who struggles to understand Laura’s mathematical approach but trusts her instincts.
Plot:
Act 1: The Initial Crime Scenes
• Scene 1: Detective Mathis is called to a grisly murder scene, the fifth in a series of crimes that seem unrelated except for one peculiar detail: each crime scene location, when connected, forms a perfect pentagon.
• Scene 2: Laura revisits each crime scene, mapping out the pentagon. She notices subtle clues suggesting the involvement of higher mathematics.
• Scene 3: Captain O’Neil is skeptical but authorizes Laura to pursue her hunch. Laura reaches out to her old professor, Dr. Petr Novák, who explains his theorem and how it could apply to the crimes.
Act 2: Uncovering the Pattern
• Scene 4: Laura discovers that each pentagon transforms according to Petr’s Miracle, with the center point and vertices shifting in a predictable manner.
• Scene 5: The team realizes that each new pentagon is formed based on the previous crime’s locations, creating a series of nested pentagons.
• Scene 6: Laura and Dr. Novák work together to predict the next vertices. They identify a pattern and pinpoint the next potential crime scene.
Act 3: The Chase
• Scene 7: As they rush to the predicted location, they narrowly miss catching Lucas Grey, who leaves behind a taunting message: “Catch me if you can.”
• Scene 8: With the stakes higher, Laura delves into Lucas’s background and discovers his motivation: a twisted game to prove his genius.
• Scene 9: A cat-and-mouse game ensues, with Lucas always one step ahead, using Petr’s theorem to stay elusive.
Act 4: The Final Showdown
• Scene 10: Laura finally deciphers the ultimate goal of Lucas’s geometric game – the final pentagon points to a significant, symbolic location in the city.
• Scene 11: A high-stakes confrontation at the final predicted crime scene. Laura uses her knowledge of the theorem to outmaneuver Lucas, predicting his moves.
• Scene 12: Lucas is captured, but not before a tense standoff where he reveals his admiration for Laura’s intellect, acknowledging her as a worthy opponent.
Epilogue: Resolution
• Scene 13: The city breathes a sigh of relief as Lucas is put behind bars. Laura reflects on the case, realizing that her unique skills in mathematics have not only caught a criminal but also saved lives.
• Scene 14: Dr. Novák publishes a paper on the practical application of his theorem, crediting Laura for her intuitive leaps and deductive prowess.
Themes:
• Intellectual Duel: The battle of wits between Laura and Lucas showcases the power of intellect and strategy.
• Math in the Real World: The plot highlights how abstract mathematical concepts can have real-world applications, even in crime-solving.
• Moral Ambiguity: Lucas’s genius is both his strength and downfall, raising questions about the ethics of using intellect for harmful purposes.
Style:
The narrative is a blend of crime drama and intellectual thriller, with detailed explanations of the mathematical concepts woven seamlessly into the plot. The tension escalates with each new crime, keeping readers on the edge of their seats, while the unique use of geometry adds a fresh twist to the genre.
Potential for Adaptation:
“Pentagon’s Shadow” is ripe for adaptation into a TV series or film, with its intricate plot, compelling characters, and the unique blend of crime-solving and mathematics. Each episode could explore a new facet of the theorem and its application, culminating in a thrilling final showdown.
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
That table at 37:31 really shows how outstanding 355/113 is as an approximation of pi. It is astonishingly good for a three digit denominator.
.
Here's a question: Is there any other famous irrational number that has a rational approximation with a three digit denominator that is anywhere near as good as this one?
To decide this, we need to think about making some measure of the "goodness" of a rational approximation (q) of an irrational number (r). To start with, we should calculate the proportional difference, (r-q)/r.
We then realize that we must take absolute values to ignore the sign of the difference.
We could then take the reciprocal so we can express this amount as being accurate to within 1 part in N, where N=|r/(r-q)|. (I use N here to indicate nearness of the approximation.)
If we care only for how accurate the approximation is, this is a good measure and the higher the value of N, the better the approximation.
But, that would have us just choosing bigger and bigger denominators as "better" when really some preference should be made for smaller denominators as these are generally found sooner and with less effort than bigger denominators (and perhaps smaller denominators are easier to remember).
So, we could apply a penalty for larger denominators by dividing the score, N, by the denominator, d. But is this a sufficient penalty? Perhaps a better measure is to divide by the square of the denominator. I will settle on having a "rational approximation goodness score" calculated as:
RAGS = | r / (r-q) | / d²
Here are some scores (N and RAGS) for approximations of π
(I have used the fractions given in the table at 37:31)
Rational N-score RAGS
3/1 22 22.188
13/4 29 1.811
16/5 54 2.152
19/6 125 3.480
22/7 2,484 50.704
179/57 2,530 0.779
201/64 3,247 0.793
223/71 4,202 0.834
245/78 5,541 0.911
267/85 7,549 1.045
289/92 10,897 1.287
311/99 17,599 1.796
333/106 37,751 3.360
355/113 11,776,666 922.286
52,163/16,604 11,801,038 0.043
Note: The fractions given by the convergents from the continued fraction representation of π are
3/1, 22/7, 333/106, 355/113 and then the following:
Rational N-score RAGS
103,993/33,102 5,436,310,128 4.961
104,348/33,215 9,473,241,406 8.587
208,341/66,317 25,675,763,649 5.838
312,689/99,532 107,797,908,602 10.881
Here are some scores for rational approximations of √2
Note: The sequence is given by a/b → (a+2b)/(a+b)
Rational N-score RAGS
3/2 16 4.121
7/5 99 3.980
17/12 576 4.003
41/29 3,363 3.999
99/70 19,600 4.000
239/169 114,243 4.000
577/408 665,857 4.000
1,393/985 3,880,899 4.000
3,363/2,378 22,619,537 4.000
8,119/5,741 131,836,323 4.000
19,601/13,860 768,398,423 4.000
It is interesting that while the nearness (N-score) increases at a rate that converges towards 3+2√2, the denominator increases at a rate that converges towards 1+√2. Since (1+√2)² = 3+2√2, the adjusted score (RAGS) converges to a constant (=4) as the sequence progresses.
Here are some scores (N and RAGS) for approximations of the golden ratio φ:
Rational N-score RAGS-score
3/2 14 3.427
5/3 33 3.697
8/5 90 3.589
13/8 232 3.629
21/13 611 3.614
34/21 1,596 3.620
55/34 4,182 3.617
89/55 10,945 3.618
144/89 28,658 3.618
233/144 75,024 3.618
377/233 196,419 3.618
610/377 514,228 3.618
987/610 1,346,270 3.618
1,597/987 3,524,577 3.618
2,584/1,597 9,227,466 3.618
4,181/2,584 24,157,816 3.618
6,765/4,181 63,245,986 3.618
10,946/6,765 165,580,143 3.618
In the case of approximations of φ the improvement in nearness (N-score) occurs at a rate that converges towards 1+φ = φ². Then, as the denominator increases at a rate that converges to φ, the RAGS-score also converges to a constant which happens to be 2+φ = 1+ φ².
The approximations of φ and √2 are generated by formulae that create convergence to fixed multiples for increases in the denominator. There is also a fixed rate of convergence towards the rational, as evidenced by the continued fraction representations of φ and √2 being 1+1/(1+1/(1+1/(1+1/… and 1+1/(2+1/(2+1/(2+1/… respectively.
I think it is fair that this scoring system gives these formulaic fractions equal standing. It is also worth noting that the RAGS for φ is less than for √2, which is consistent with φ being the "more irrational" number that is harder to approximate.
I have also calculated the N-scores and RAGS for √3 = 1+1/(1+1/(2+1/(1+1/(2+1/(1+1/(2+1/…
The continued fraction pattern is 1,2,1,2,… It shouldn’t be a surprise that the N-scores increase at a rate that converges to 2+√3 while the RAGS converges to an alternating sequence of 3 and 6, with the higher RAGS coinciding with the approximations that are slightly greater than √3. This is because the approximations that are greater than √3 have a proportionately smaller increase in denominator than those that are less that √3 – i.e. if you go from above √3 to below √3, the denominator has increased by more than the numerator to obtain a smaller fraction.
The scoring for the approximations of π is certainly more interesting since the continued fractions are not in a fixed pattern and so the quality of the approximations relative to denominator as indicated by the RAGS varies considerably. And just to come back to it: How good is the 355/113 approximation for π? It is the absolute stand-out amongst those shown here.
[Note - these calculations were done "quick and dirty" in a spreadsheet and so the values are likely inaccurate past 10 digits.]
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
24:30 For a while, I struggled with figuring out, why the ”>”-sign can’t become a ”<”-sign, straight away, as the quadrilateral becomes flat. But, I’ve finally got it: The transition between 3D and 2D is continuous; and, as such, the difference (Aa+Bb)-Cc must either stay positive (corresponding to the ”Strictly greater, than” -case), or, on its way to negative (corresponding to ”Aa+Bb < Cc”), it must pass through 0 (Aa+Bb = Cc), as you indicated, in your video on the ”Turning the wobbly table” -theorem (in a slightly different context). But; because, in a continuous transition, there is only 1 instant, where the quadrilateral is flat (and, therefore, the ”strictly greater than” -inequality doesn’t necessarily hold), the difference is always either positive or 0; and therefore, for all quadrilaterals, Aa+Bb >/= Cc. Very satisfying, I must say. 😌
4
-
4
-
4
-
4
-
4
-
4
-
Big fan of Heron's formula here, so I may go on a bit. Here in the US, New York State, under the "Regents" math curriculum, in the late 1960's, we did learn Heron's formula. We were not required to memorize the derivation, but it was in the textbook. Decades later, I decided to test my algebraic chops, and try to derive it with the only clue I remember from the book, that it relied on factoring the difference of two squares. If you drop an altitude on one of the sides, you can solve for the altitude and one of the unknown segments on the base simultaneously, using the Pythagorean Theorem. Mathologer takes a shortcut by using the law of cosines, but my method is how you get the law of cosines as well (and you can get rid of the fraction in the Mathologer's version with some convenient cancellations). But once I was there, and I got the formula knowing what I was looking for, I thought of four justifications for "discovering" Heron's formula instead of calling it a day having a formula for the area in terms of the sides: 1) it lacks symmetry. There is nothing special about any side, other than that you chose one to drop an altitude on; 2) it's pretty nasty to calculate from. Many of us have probably calculated nastier ones, but we can do better; 3) it is badly scaled. You end up raising numbers to the fourth power, which usually results in something large, then subtracting them, leading to truncation or round-off errors if you don't keep a lot of decimal places (I realize Heron wasn't thinking about floating point calculations. Heck, he didn't even have a decimal system) 4) it's not obvious from the original formula, at least to me, that your area isn't going to turn out to be the square root of a negative number.
If you expand the trinomial and collect terms, you do get something symmetrical, but all the other objections remain. That might tell you that, since expanding didn't work out very well, maybe the opposite--factoring--is the solution. In addition to factoring the difference of two squares, twice, at one point you have to collect some terms and recognize it as the square of a binomial. It's all quite pretty. But it's 100%, algebra, none of the geometric insight of this video.
With the final formula, in addition to being much prettier and easier to calculate from (if the need arises which, truthfully, it rarely does) you can see at once that for a legitimate triangle, no one side being longer than the sum of the other two (or equivalently, no side being longer than the semi-perimeter) you'll never get a negative number and moreover, if a "triangle"s one side is exactly equal to the sum of the other two it has zero area, as expected.
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
Loved this Video! Always wondered whether there was any research on the addition of odd numbers to get squares. I knew how to get the cubes, but was pleasantly surprised to know that the pattern continues! Made my day, (the week actually). Loved the fact that some of the approaches in the video share a big similarity to the ones in your power sum video, grids and various others (And pascal's triangle!). Also, how you can use geometry to not only visualize, but also solve what is, essentially, a number theory and algebra problem. Gently makes you appreciate how interconnected everything in mathematics is. I hope you can keep making these videos, and that they reach an even larger audience, so everyone can fully enjoy and learn advanced mathematics.
PS : Is your geometric proof the first one?
PPS : I think, the output sequence might be n!*(n-1)! ? if a =1 and b, c, ... =2?, Since then each term would become n + 2(n(n-1)/2) = n^2.
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
I'm a brazilian olympiad student, the time I traveled to where I was going to participate in Conesul (South America Olympiad) there was a training during 3 days before the test. One of the problems that my professors showed was this one, pretty fun problem, by the way. My solution at the time was, without loss of generality fix the point that's gonna be reached by the soldiers, and say it has an energy of 1. And then, the natural ideia that came in my mind was to say that any square that was K squares apart from the square we're going to has an enery of (phi)^{-K}. Like this, as well as your solution, the energy of the whole table is always, at least, preserved, and summing the energy of every square below the red line equals something less than 1 (which you can easily compute by some infinite power sum tricks). It's always great to watch your videos! thanks🇧🇷
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
By bloopers, do you refer to the formatting inconsistency?
For example, 二百五 is "two hundred five", so 205. Next to it is 九八七, being "nine eight seven", which is contexts like this would refer to 987.
In these settings, you'd just list the numbers off one by one. You wouldn't give the full name. A bus numbered 64 would be read as 六四, which is "six four", rather than 六十四, which is "six ten four", so "sixty four".
Basically, you wouldn't have things like 白 (hundred) in this sort of thing. Just the base characters listed like a phone number.
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
This is one of the many things I enjoy about traditional geometric patchwork design quilt patterns , (like Flying Geese, Irish Chain, Feathered Star, Grandmother's Flower Basket, Disappearing 9-patch,
Bear's Paw, Log Cabin and all it's variations, Pinwheels, etc.) with 1/2 square or equilateral triangles, rectangles, squares, etc. and shuffling them around to create different patterns. Now I have a name for them, and may design quilt patterns with the names of Pythagoras or Trithagoros, or the long ago Chinese mathematician's name! Great visual explorations of these relationships, thank you. (of course I enjoy even more, the random sewing together of odd shaped and sized pieces, called, "crumb quilting"! lol)
(see also, the 'golden ratio' seen so much in nature and used from antiquity in architecture to produce unconsciously, naturally pleasing, building facades.)
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
For the puzzle at 11:37, I was too tired yesterday to get it, but now that I had time today I gave it another shot.
My square was defined as:
┌──┬──┐ I defined the layout of any pythag.. triple as:.
│B1│B2│ X^2 + Y^2 = Z^2 .
├──┼──┤ .
│B4│B3│ .
└──┴──┘ .
I started out pretty standard with the main three equations of:
B1*B4 = X
2*B2*B3 = Y
B1*B3 + B2*B4 = Z
I also included these:
B1 + B2 = B3
B2 + B3 = B4
To define the innercircle, I used r_i to denote its radius. The equation is then:
r_i = B1*B2
I started by creating an equation to find r_i. To do this, I used the slope of the hypotenuse (Y/X) to solve for an angle, divide it by two, and convert it back to a slope. This gives a slope which bisects the angle XZ. The equation is:
y=tan(1/2 * arctan(Y/X)) * (x+X) (my origin is placed at the corner XY so x+X places the lines origin at the corner XZ)
Using trig identities, namley half angle identity and inverse trig identities, I could simplify the expression:
tan(1/2 * arctan(Y/X)) =>
(sin(arctan(Y/X))) / (1 + cos(arctan(Y/X))) =>
(((Y/X)/sqrt[1 + (Y/X)^2]) / (1 + (1/sqrt[1 + (Y/X)^2])) =>
(Y/X) / (sqrt[1 + (Y/X)^2] + 1) =>
Y / (sqrt[X^2 + Y^2] + X)
This means that the equation is:
y = (Y / (sqrt[X^2 + Y^2] + X))(x+X)
Which is our equation for the first bisector line.
Since the angle XY is 90°, it has a bisector with a slope of -1. This results in the equations:
y=-x
Setting them equal to find the centre of the incircle results in a simplified equation of:
x = -XY / (X + Y + sqrt[X^2 + Y^2]))
Since X^2 + Y^2 = Z^2, this becomes:
x = -XY / (X + Y + Z))
OR:
r_i = (XY) / (X+Y+Z)
Plugging in the values, we get:
r_i = 36
since r_i = B1*B2 = 36, this means B1 = 36/B2.
plugging into the equation: B1 + B2 = B3 yields:
36/B2 + B2 = B3 =>
36 = B2*B3 - B2^2
using Y = 2*B2*B3 = 104, we know B2*B3 = 52.
36 = 52 - B2^2 =>
B2^2 = 52-36 =>
B2 = 4.
So, using B2*B3 = 52 we find:
B3 = 13
Since B1 + B2 = B3:
B1 + 4 = 13
Or:
B1 = 9
Since B2 + B3 = B4:
4 + 13 = B4
Or:
B4 = 17
To Conclude, my guess as to the numbers in the square are:
┌──┬──┐
│ 9│ 4│
├──┼──┤
│17│13│
└──┴──┘
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
10:46
For negative integers, A(x^n x^-n) = A(x^n) + A(x^-n) = nA(x) + A(x^-n), so A(x^-n) = -nA(x).
For rationals, qA(x^(p/q)) = A(x^p) = pA(x), so A(x^(p/q)) = (p/q)A(x).
For reals, I'm pretty sure we have to assume A is continuous? For any real number c, we get a sequence of rationals q_1, q_2, ... approaching it. By continuity, the limit of A(x^(q_i)) is A(x^c). To compute this, we want the limit of q_iA(x), which is cA(x).
If we don't assume A is continuous, I'm not 100% certain. My real analysis class has traumatized me to the point that I'm pretty sure there's a weird discontinuous counterexample lurking around the corner, but I can't find it.
4
-
4
-
4
-
4
-
4
-
4
-
4
-
4
-
This presentation remind me a experience. At high school my math is bad, at exam my teacher give one question, how much summary number from 1 until 1000, i slove this one with draw a diagram from 1-1000 like stair case , then divide that draw to be a large triangle and many minor triangle, and calculate that's area.
And next day, she call me and said to me how dumb i'm, she said i'm not even use right formula and get right answer, said that i'm cheated and give me 0 score i just laugh at thats time.
But my economic teacher pass by on righ time and right place like superman, hearing her bit anger he said to me "what have you done?".
She explain to him and give my paper to him, then suddenly he said "ok... i get it, leave it to me, you can back now."
I saved by economy teacher at math problem, truly i cant forget that. 😂😂😂
4
-
4
-
4
-
Hi there, Mathologer. Really nice explanations -- thank you.
One method that I've used to convince my calculus students that expressions are indeterminate is to give them, or have them try to construct, functions with the specified limits which show that these indeterminate forms could be "reasonably" assigned any number that they like. For example, for a given nonzero complex number w, the functions f(z)=wz and g(z)=z give rise to a quotient f(z)/g(z) which takes on the indeterminate form 0/0 as z approaches 0. But of course the limit is w, so this gives a "convincing" argument that 0/0=w. Note that f(z) and g(z) can be taken to be holomorphic functions, and one can run similar arguments for most of the other indeterminate forms.
Edit 12/6 : Incorrect reasoning in the following paragraph. Please see my reply to this comment.
One curious thing I've encountered recently is that I run into problems in applying this process to the indeterminate 0^0. For example, it is easy to make an argument that 0^0=1 -- simply take f(z)=z and g(z)=0. Then f(z)^g(z) clearly approaches 1 as z approaches 0. But suppose I sought to make a convincing argument that 0^0=w for a given complex w using this method. I seek two (holomorphic, say) functions f(z) and g(z) such that f(z) and g(z) both approach 0 as z approaches 0, but f(z)^g(z) approaches w. By L'Hopital's rule, I determine that a necessary condition for this to happen is that the limit as z approaches 0 of the expression [(f'(z)/g'(z))(g(z)/f(z))](-g(z)) should equal ln(w). But by L'Hopital's rule the two quotients f'(z)/g'(z) and g(z)/f(z) must be the reciprocals of each other and therefore the limit of their product must be 1, which of course forces w=1 (since g(z) approaches 0 as z approaches 0). Thus, in order to have f(z) and g(z) behave like I want them to, at least one of them must not be differentiable at 0. Can you comment to what extent the indeterminate form 0^0 is different from the others?
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
2:25 - "Why? Who knows..." and that's why there's this expression "talmudic reasoning" or "talmudic problems" - meaning obscure and convoluted (and often illogical) way of thinking using all sort of arbitrary "laws" and "precedents" - or bizarre/ weird interpretations. Why so? Who knows, or as Tevye the Milkman has said:
You may ask, how did this tradition start?
I'll tell you! - I don't know...
;-)
11:00 - the creditors may suffer "the same loss" in absolute numbers, but certainly not in terms of "relative loss". Someone who can lend $300 (i.e. is capable of giving such a loan) must, obviously, have MORE than that amount (say/ assume $600 of cash) - while someone who loaned $100 might be much poorer (say, $200 of ready money at most) - so the same AMOUNT of lost money (say, $50) would hurt the "$200 guy" much more than this "$600 chap".
But then you never really know what is their financial situation/ status - it could be just that this "$300 sucker" loaned all he had, while this "$100 scrooge" holds $10,000 - but he's a scrooge and a penny pincher, and that's why he shelled out a mere $100.
And since this part of equation is beyond any reasonable judgement of any "inheritance court" it should be left out and never taken into consideration, as this "talmudic do-gooders" approach/ reasoning can actually create worse and less fair outcome for the debtors.
Which only goes to show "why this talmudic reasoning expression exists".
When it's the time to light up Shabbat candle? "When one can see stars on the night sky" (i.e. they become visible) goes one (of probably gazillion other interpretations. So when a first star appears, it's only ONE star, not "stars", hence it doesn't count ("einmal ist keinmal" as Germans say - or "once/ single occurrence is no occurrence/ it doesn't count").
So, when the SECOND star appears it's still only a "single" star, as the first one "doesn't count, so only after a third star appears you can say there are TWO stars, so it is "stars" (not "a star"), so there you go. I mean, "there you can light up a Shabbat candle".
Simple, no? ;--)
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
13:00 Answer: n! * (n-1)!
Edit: I was a few hours too late apparently.
We start by using what you showed at 11:56 and choose a = 1 and b = c = d = ... = 2, because from that follows that
1^2 = 1a
2^2 = 2a + 1b
3^2 = 3a + 2b + c
--- and so on.
The Moessner sequence is then 1^a, 2^a * 1^b, 3^a * 2^b * 1^c, which we can simplify since a is 1 and all the other variables are 2.
The general formula is: n * 1^2 * 2^2 * ... * (n-1)^2 = n *((n-1)!)^2 = n! * (n-1)!
1, 2, 12, 144, 2880, 86400, ...
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
@Mathologer Concerning the higher dimensions: I always assumed that, say, in 5D you can construct a regular pentagon by taking points (1,0,0,0,0) and (0,1,0,0,0) ... and (0,0,0,0,1). Arguably, these five points are very regularly distributed, forming (seemingly) something 5-fold rotationally symmetric around axis (1,1,1,1,1). But your proof suggests, that this is not a pentagon. So what is this? Do these five points not lie in a 2D plane? (Very nice video btw, I am huge fan!)
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
solution to the ending puzzle (49:52)
(SPOILERS!!!)
the solution as indicated by others is the odd-indexed fibonnaci numbers if you count f(0) = f(1) = 1. with f(n) being the n-th fibonacci number.
here is my proof:
first ill mark p(n) = the n-th term in this series, p(1) = 1. and for convenience, p(0) = 1 as well.
as shown in the beginning of the video, every way to break up n into a sum of integers can be expressed as n-1 "spaces" between n boxes, when each way of closing some of the spaces is a unique integer sum of n.
this means that you can express every integer sum of n with a vector with n-1 terms, all either 0 or 1 .
example: take n=5 and the vector 1101. this represents 3 + 2 = 5.
now i will organize the vectors in a row by their values in binary.
meaning that for example for n=3, the 4 vectors will be ordered as such:
00
01
10
11
and the corresponding sums like this:
1+1+1 = 3
1+2 = 3
2+1 = 3
3 = 3
now for n, there will be 2^(n-1) rows in this kind of ordering.
the first half of the rows all have 0 as their left most term, meaning all sums look like 1+ integer sum of n-1.
and since the rest of the vector goes from 00...0 to 11...1 this covers every integer sum of n-1.
meaning that in the first half of the rows there are exactly, all integer sums of n-1 with an added 1 on the left.
that 1 does'nt change the product of the rest of the numbers, hence the sum of the product of each row is just p(n-1).
now ill look at the second part, and again at its first half. this time all vectors start with a 1,0 and the rest corresponds to an integer sum of n-2.
now every row looks like 2+ integer sum of n-2. meaning that the sum of the product of those rows is 2*p(n-2).
continuing like this by induction we get:
P(n) = sum from k=1 to n-1 of k*p(n-k)+n , p(1) = 1.
first, f(2*1-1) = f(1) = 1.
ill now show that p(n) = f(2n-1) satisfies this recurrence relation.
meaning that i need to show that:
f(2n-1) = [sum from k=1 to n-1 of k*f(2(n-k)-1)] +n
proof by induction:
(n=2) base:
[sum from k=1 to 1 of k*f(2(2-k)-1) ]+2 = f(1) +2 = 1+2 = 3 = f(3).
step : ill assume the reccurence holds up to n.
now ill show it for n+1:
first, f(2n) = f(2n-1) + f(2n-2) = f(2n-1) + f(2n-3) +f(2n-4) = ... = f(2n-1) + f(2n-3) +...+ f(1) +f(0). (*)
now :
f(2n+1) = f(2n) + f(2n-1) =
= f(2n-1) + f(2n-3) + f(2n-5) +...+ f(1) + 1 by (*)
+ f(2n-3) + 2f(2n-5) +...+ (n-1)*f(1) + n by induction.
= f(2n-1) + 2f(2n-3) + 3f(2n-5) +...+ n*f(1) +n+1
= [sum from k = 1 to n of k*f(2(n+1-k)-1)] + n+1.
hence for every n, p(n) = f(2n-1).
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
17:18 I rather think this construction works for all n ≠ 3, 4, 6. As I see it, we are constructing a non-degenerate star n-gon with side length equal to that of the original polygon. Taking the convex hull of the star gives a smaller n-gon, and we are done.
A non-degenerate star has Schläfli symbol {n/q}, where 2 ≤ q ≤ n-2 and gcd(n, q)=1, so the construction can work as long as such a q exists. (To do the octagon for example, we construct an {8/3} star.) If such a q does not exist, we must have (phi() being the Euler totient):
phi(n) = 2 <=> n = 3, 4, 6.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
It heartening, hilarious, and humbling to see the various people reporting "I discovered this when I was much younger" (or a variation on it). I am fascinated by math and terrible at it. So I was 26, did everything by hand without a calculator (not because I'm hardcore, but because that was all I had), and twice ran into worsening and worsening polynomials to try to deal with, before I finally came up with a one-line formula that spat out the polynomial for a sequence of numbers. In this attempt, I started over twice and I don't know what I did differently the third time (when I say I started over twice, I mean over months of hand calculations). I certainly had no names for the sums of sums, but I did see Pascal's triangle buried in it, was multiplying the falling "n choose k" statements by coefficients derived from differences of differences. I was using horizontal rows, rather than diagonals though.
Watching the video, I can't put together the relationship between what's happening here and what I did exactly; this is obviously much more straightforward and sensible, but both end with elegantly simple statements. I have a really ugly Excel spread sheet (the "function finder") that crunches a list of numbers into a polynomial. The background calculations are just as ugly as my stumbling attempts, but it gives you the correct polynomial.
Just to show the prettiness, if you have a sequence of four numbers [A, B, C, D] that generates a degree 3 polynomial (variable n>3), the not-simplified formula that kicks out is (I think):
[-(n)(n-1)(n-2)(n-3)]/3! * [A/(n) - 3B/(n-1) + 3C(n-2) - D/(n-3)]
3
-
3
-
4:30 I'd say that the mistake is that chinese numerals are not directly positional, you need to add the appropiate symbol to mark the tens, hundreds, thousands and beyond, but given that the 1080 at the top of the picture is perfectly correct (it does have a 1 before the 1,000, which is not done usually but is done in places where there is a need for accuracy) and knowing that there are indeed times where they are used with no position markers in places that look to keep a "traditional" ambient (mainly in restaurant menus, at least outside of China, although I don't know how modern the practice is), they may be using a positional system as a kind of shorthand. (EDIT: 7:45 shows the tens with the marker, so the theory is quite likely, although I don't know if it's anachronistical or not.)
3
-
3
-
3
-
3
-
The magic square proof is brilliant!
To generalize it, i turned the tiles' numbers into coordinates
for a n*n square, T=x+(y-1)*n, 1<=x, y<=n, T->(x,y) is a one-to-one map
therefor we have (1,1) to (n,n). 1->(1,1), 2->(2,1) etc.
arrange the tiles into the big tilted square as the video shows, for every tile T, the only tiles that share a common x or y coordinate are the ones that share the same diagonal with T
Thus, every tile on the same horizontal/vertical with T must have different x and y coordinate (that is the first part)
Now we define horizontal and vertical distance, which is the number of horizontal and vertical steps required to move a tile to another square (empty or another tile)
As the video says, when creating the magic square, all tiles outside moves exactly n steps to it's destination, which gives us a distance (horizontal or vertical) of n
However, as there are only n squares on a diagonal, the maximum (horizontal or vertical) distance between T and any tiles that share the same x or y coordinate <=n-1 (that is the second part)
Thus, on the final magic square, all tiles on the same horizontal (vertical) with T must have different x and y coordinates
this holds for every tile T
so every horizontal and vertical sum= n(n+1)/2+(n(n+1)/2-1)*n=n(n^2+1)/2 which is correct
the diagonal sums are n*(1+n)/2 (same x coordinate (1+n)/2) + (n(n+1)/2-1)*n (y coordinate 1-n) and n(n+1)/2 (x coordinate 1-n) + n*((1+n)/2-1)*n) (same y coordinate (1+n)/2)
this proves that it is a magic square
My wording is really bad, but this is the proof in the video
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
Top shelf stuff Burkard.
Extending the Fib S "leftwards" we obtain:
...-8,5,-3,2,-1,1,0,1,1,2,3,5,8...
and note that just as F(n)/F(n-1)->φ for n->+∞ it also->-1/φ for n->-∞.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
The "windscreen wiper" method can be used to prove the same result on a sphere, which by extension also proves the 2D case as well.
1)Draw a great circle around a sphere, marking two points at random on this circle.
2) Mark a third point at random, but not on the same great circle, and connect this to the first two points around two new great circles. This creates a spherical triangle.
3)Rotate the sphere so that the "north pole" lies within the triangle in such a way that the northernmost points on the three great circles all lie at the same latitude.
4)Draw the "whiskers" by extending the sides of the triangle as shown in the video, following the great circles in each case.
5)The great circles all have exactly the same radius of curvature at every point, so the windscreen wiper method can be used to rotate and tilt any one of the extended sides around each vertex to bring it into superposition over the adjacent side. Three rotation and tilts brings a side back to its original position, but pointing the opposite direction, just as in the video. This proves that the three extended sides are the same length, and the northernmost points are the exact midpoints of each arc.
6)Starting at the northernmost point of any of the three great circles and measuring the same arc in each direction takes you to two points which lie at the same latitude. Since the three arcs are all the same length, all six end points must lie at the same latitude.
Ta dah.
7)Let the radius of the sphere tend to infinity. The surface of the sphere tends to a plane surface, and the original Euclidean result follows directly.
Ta dah!
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
Fantastic! I would love to see a mathologerisation of the pattern for the continued fraction of e.
[2; 1, 2, 1, 1, 4, 1, 1, 6, 1, 1, 8, 1, . . . ]
It's so beautiful and so hard to wrap my head around...
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
By the way, the second method you demonstrated makes use of the Pell number, which can be defined as P(0) = 0, P(1) = 1, P(k) = 2P(k-1) + P(k-2), The limit of the sequence of P(n+1)/P(n) = 1+root2 which is the silver ratio. In fact this is related to the Metallic Ratio. Metallic ratio is the root of the equation x^2 - nx - 1 = 0, if n = 1, you get the golden ratio. If n = 2, you get the silver ratio. If n = 3, you get the Bronze ratio, and so on. And the sequence of number you can obtain for each n = {1, 2, 3, 4, 5, ...} can be defined by f(0) = 0, f(1) = 1, f(k) = n.f(k-1) + f(k-2). So if n=1, then you get the Fibonacci number. If n=2, you get the Pell Number, and so on. So to get the approximation of root2, you go like this: Since x^2=nx+1, then x=n+1/x, then 1+root2 ≈ 2 + P(k-1)/P(k) or root2 ≈ 1+P(k-1)/P(k) = (P(k)+P(k-1))/P(k). For instance, the Pell sequence: 1, 2, 5, 12, 29, 70, 169, 408, ... root2 ≈ (P(7)+P(8))/P(8) = (169+408)/408 = 577/408
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
to check how many sum=prod identities are there for a given length n, you first check what is the maximal number of non-one numbers in your identity. you do this by taking a list of n ones and seeing how many ones can you replace by twos s.t. the product is less than the sum. now you manually check cases for each number of non-1 numbers below the suprimum. so for 10 you've got 2 identities: 4,4,1,1,1,... and the trivial case 2,10,1,1,1,...
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
30:15 phi can be written like (1+sqrt(5))/2, which is what you get from the positive solution. for the negative solution the square root is subtracted, so we end up with (1-sqrt(5))/2, and doing some algebra leads to:
-(1+sqrt(5))/2 + 1 which is -phi + 1. at this point i would try find a correlation between the two solutions (or try finding a mistake in my calculations), but i've known for quite some time now that 1/phi = phi-1, and i will try prooving it now: if you have a golden rectangle with sidelengths 1 and phi, and then remove a square, the new sidelengths are going to be 1 and phi-1. but since the ratio is the same, we conclude that 1/phi-1 = phi/1, which (hopefully) proves phi-1 is the reciprocal of phi.
31:50 for puting real numbers there is already a video on youtube online, it basically just results in complex numbers because of the exponents. for the negatives, it just goes backwards and switches between positive and negative, which, now that i think about it, makes the sequence even better! when taking the ratio of two consecutive positive fibonacci numbers, their ratio approaches phi as you go bigger and bigger.
if you go in the opposite direction, the ratios approach -1/phi, which is the second solution to the previous equation! i find that funny as it seems as such a clear fact, the equation we use to get the fibonacci numbers uses -1/phi afterall.
41:35 my guess is that it will produce the mixed r and s identities.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
I’m pretty sure (for the three color game) that is it MULTIPLES of powers of three plus one. (m*3^n)+1. For example, 19 works, 2* 3^2 + 1. You may have said this in the video and I just missed it, but if not I think this is very interesting because it means that 2,4, 7, 10, 13, 16, 19, 28, 37, ect. all work.
Lmk if I’m wrong
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
Calculating the area of a triangle is part of the Mathematics Olympiad program (IMO training), because counting things in multiple ways can reveal complicated identities.
The proof used the cosine rule. Ew.
I actually discovered (and proved) the formula myself, when I was about ~13, just by using the right-angle theorem (Pythagoras).
More recently, I found out (thanks to Wikipedia) that the maximum area of a quadrilateral happens with a cyclic quadrilateral. This is beautiful, because we can extend this for any polygon.
I used some ugly math to prove this, but I did found out about some nice things. I also found out about the correction term. I think that many people discovered this, even before 1800.
Your video's encourage me to seek for simple proves myself.
The method for the proof of p=a²+b², can be extended for all natural numbers.
You can prove that numbers that shouldn't be written as sum of two squares, have an even number of representations.
Next, you parameterise n=a²+b²=c²+d², to show that n is the product of two numbers (greater than one), as the sum of two squares.
Next, you can say that we had the lowest number that works (with some extra things), to finish the prove.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
Yes; you can just pick a direction and it will work out, because you always have a choice of at least 2 cards to take out.
Say the magician and assistant have agreed always to add the encoded number to the first card value. Say the 5 cards include the ace of spades, and there is one other spade in the group. If it's the 2, 3, 4, 5, 6 or 7, you take out the larger card and use the ordering of the others to encode a 1 ... 6 and all is well. But what if it's the 8? Well, we're using modular arithmetic, and (8+6) mod 13 is 1. So you take out the ace and encode a 6. Similarly for all cards larger than an 8, and for all other combination of cards; just choose whichever card makes the gap between 1 and 6.
3
-
The Wikipedia article I have seen starts with integer pairs (2,1) and (3,1). Call these pairs (m,n). It then applies the following algorithm separately to each pair: Branch 1 = (2m-n, n); Branch 2 = (2m+n, n); Branch 3 = (m+2n, m). The algorithm in the video starts with the same seed pairs, arranged in a 2X2 array, which you can alternatively view as rational pairs (1/3,1/2)
The Branch 2 algorithm reproduces the same “middle children” as those derived in the video. The other two algorithms reproduce the left and right children but swap one column in each for one column in the other. So in the video left child (3/5,1/4) and right child (1/5,2/3) become in the Wikipedia algorithm (3/5,2/3) and (1/5,1/4), starting with an initial pair (1/3,1/2) as in the video.
So the two algorithms are equivalent.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
the magic method in 18:47 is simply drawing diagonals but using modular arithmatic or if you perfer imaging the board goes on forever and then collapsing in back to a single board so using (row,column) notation : we have (5,3) draw a diagonal thats (6,4)->(1,4) then (2,5) diagonal (3,6) ->(3,1) then (4,2) . once the first 5 diagonal has ended go down 1 and draw the second 5 diagonal thats (3,2) then (4,3)
MARTY : you've made a mistake no (4,3) green dot!
(5,4) , (6,5)->(1,5) (2,6) ->(2,1) go down (1,1) , (2,2) and so on
this ensures every column has all (x+5*y) sequences (x being the sequence in the 1-5 drawn by the y'th diagonal)
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
I wonder what the 3 dimensional equivalent. Like if we work in reverse, we would have all our linesrly independent eigen polyhedra(whatever that entails), and we can have some linear transforms that map each of the miscellaneous polyhedra to 0. I havent finished watching the video, but if im to guess, the eigenpolygons are all the possible orderings of vertices that can be attained from connecting the points of the regular polygon by skipping n of the vertices(i.e original clockwise pentagram attained by connecting without skipping, clocksise pentagram attained by skipping 1 at a time, anticlockwise pentagram skips 2 at a time, anticlockwise pentagon skips 3 at a time). This doesnt translate well to 3d, but I can guess that this specific process of forming eigenpolygons is just one of many possible basis polygons, and that there might be another basis that translates well to 3d. If we dont care much for clenliness, we can just have a non linearly independent basis by having every single permutation of our vertices, modulo rotations, which certainly spans all of our possible polygons just the same, but I dont think this will work. As in the original basis was isotropic(i think thats the right word), we could send eigenpolygon components to 0 by adding ears on, which was a process that had no preference of starting vertice, and the action was identical for each pair of vertices, but if we wish to send arbitrary permutations of vertices to 0, I'd have to guess this isnt going to turn out the same. Which makes me wonder, is there a convenient basis for polyhedra, where an isotropic process will send each eigenpolyhedron to 0, and the basis spans all of our polyehdra. I think i would be comfortable in saying it isnt possible.
But now that I've taken another step back im beginning to think its possible again. Instead of thinking of a process that relied on an explicit ordering of the vertices, why dont we think of all the possible choices faces of polyhedra that are rotationally invariant, that way an action on a single surface is the same for all surfaces. I still dont think there is enough of them to form a basis here. Additionally, the process of adding ears was linear in the 2d case, but would it be in the 3d case? I suppose we can always modify our action so that it must be linear, we have plenty of degrees of freedo. there(takes 3 vertices as input, must be isotropic, must be linear, and must send a specific face to 0, feels like we have the freedom here). And now that i think about it the eigenpolyhedra dont even have to be nice whole surfaces, we only care that all the faces are identical, and its rotationally invaiant, and that it forms a basis. The more I think about it the more I feel its possible, though whether our 3d analogue of adding ears is natural in any way, thats debatable. But now im doubting myself again and thinking its impossible, and my reasoning is now impossible to explain.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
Interestingly (and perhaps not too surprisingly), while there is an infinite solution to Conway's Soldiers in a certain sense, there is no well-founded solution. That is, every solution involves sequences of steps which are forwards-infinite (no final move) and also ones which are backwards-infinite (no initial move). You can't do it in the intuitive way of having some function f(α) on the ordinals α≤κ (where κ is some large countable ordinal), where f returns a legal position for each ordinal in its domain, each proceeding from the last by a legal move, where f(0) is the starting position and f(κ) is the winning position.
If we do allow arbitrary forwards- and backwards-infinite sequences in a solution, we can actually do better. We can make a soldier appear anywhere on an empty board! That's clearly not acceptable, so an additional requirement is necessary. Specifically, there is some number N such that no space on the board flips from empty to occupied more than N times. With that extra condition, the best we can do is reach row 5, using up every soldier in the process.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
Christmas glasses:
The squares numbered 20r and 20g must be part of their outer 2x3 "arms", or they'd leave a 5-square board that can't be tiled. These two 2x3 sections can be tiled 3 ways each.
So far: 3 x 3 = 9 ways to tile the outer two rectangles.
Then, the red and green squares numbered '8' and '15' can only be part of 4 different pairings (e.g. 8r and 15g must both be paired to their left or their right), otherwise they'd isolate a section with an odd number of squares. Call these 4 pairings First, Outside, Inside, Last (like FOIL). EG: 'First' = both sets of pairs made with the left neighbour.
Using 'First': Working left-to-right, we get a 2x3 grid (3 ways), then 13g-9r and 14g-10r must be paired, then another 2x3 grid (3 ways), then 17r-7g and 16r-6g must be paired, then a 2x2 grid (2 ways). This multiplies to a total of 3x3x2 = 18 ways.
Using 'Outside': Working left-to-right, we get the same 2x3 grid (3 ways), then 13g-9r and 14g-10r must be paired, leaving a 2x2 grid in the middle (2 ways) and a 2x3 grid on the right (3 ways). This makes a total of 3x2x3 = 18 ways.
Using 'Inside': 17g-7r and 16g-6r must be paired, leaving a 2x2 grid to the left (2 ways); likewise, 17r-7g and 16r-6g must pair, leaving a 2x2 grid to their right (2 ways). This all leaves a 2x4 grid in the middle (5 ways). Total: 2x2x5 = 20 ways.
Using 'Last': Same as 'First', but mirrored. Total: 18 ways.
Multiplying all this together gives:
3 x 3 x (18 + 18 + 20 + 18)
= 9 x 74
= 666
Truly a 'beast' of a solution.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
I liked the way you saw it as an ordinary number rather than a binary number. It removes one layer of obscurity. Also, yes, short problems between your larger more involved ones would be great.
Regarding the incomplete rule problem:
There are three options:
- You can't flip the last card.
- If you flip the last card, you don't flip any other card.
- If you flip the last card, you flip the first card.
The first two options work but using the last option, it is possible that it may not terminate though I'm not sure if I can prove it except that the upturned cards can no longer accumulate to the left.
Also, if you reverse the labels (0 for down-turned card, and 1 for up-turned card) then, obviously, the number will progressively increase till you reach the maximum number of 11111111 (or, in non-digital form: 11,111,111).
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
WOW!!! Bravo again! This is another indisputably perfect example that supports my theory, metatheory, proofs & metaproofs that show how and why nature's astrophysical geometry, geometry, numbers, maths, and logic are enabled & sustained by the natural metalogical principles of being (i.e., the "cosmos"). I will definitely cite (& link) this episode in my next draft of "Astronomy, Geometry, and Logic" (and formally request permission to use a pic or 2 from the video). Dear Burkard & Marty, thanks again for doing the best, most useful maths series on Youtube.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
I have known Ian's website for ages, being myself a kind of knots geek. I have used Ian's knot for my shoelaces for almost 20 years (more than 15 for sure), and it really drives me crazy when I see someone who ties their laces with a granny knot. Lastly, my 4yo niece was proud that she could tie her shoes, alas, her mother learned her, and she always ties them the wrong way. My niece was so disappointed when I told her, as gently as I could, that she learned it incorrectly and that it would be better for her to learn it again... I think I will try to teach her Ian's knot. It's really easy and fast, once you have the hang of it.
By the way, am I right to suppose that the boat illustration comes from the Ashley's book of knots?
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
Challenge time! I will write out the answers, but I don't have time to write the whole proofs, so I should conjure Fermat and only write the main points for each challenge.
1. 5. Big circle through 2 points leaves 5 points to be split, so 2+3=5.
2. Trivial, as "a"."b"_"c"_, with a, b, c length l, m, n can be turned into a fraction by multiplying 10^m and 10^(m+n) respectively and subtracting. "abc"-"ab"/10^(m+n)-10^m is the result.
3. Ramsay theory. For each point there is either 3 segments or 3 non-existing segments. Focus on the three segment/non-segments and the three end points. If there exists a same type interaction between the three, a triangle is created. If there are none of that type, the three automatically make a triangle.
4. 315? Not really sure, as I do not do Rubik's cubes, but each part seems to have 7, 5, 3, 9 cycles. Lcm is 315.
5. {1, 2, 4, 8, 16, 32, 64}. Binary represantion, trivial as 1, 2.
6. Queen of Hearts. First card shows that the suit is hearts. Dictionary order is 3rd, so 3 spaces away means queen.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
4:26 If the mini-blooper is (probably spoiler below)
is that there are no character for zero (零/〇), that's not a blooper. Ancient Chinese numerals do omit these zeroes when written out. The notation of "〇" as a zero digit actually comes from Hindu-Arabic numerals, and the usage of "零" for zero digit comes from its original meaning "remainder". During calculation, zero digits are represented by empty space with counting rods, or no beads moved in a position in abacus.
For example, "二百五" is "two hundred (and) five" or 205; in comparison, modern Chinese write 205 as "二百零五" ("two hundred (and) remainder five"). To disambiguate the position of zero digit, the position value of the present digits will write out if it is not at 1s place: "一千八十" is "one thousand (and) eight -ty", the presence of "十" (ten) indicates the eight is in 10s place, so this number is 1080, not 1008.
And actually, ancient Chinese numerals also likes to omit ones if they are not in 1s place; the position value will indicate the presence of digit, and defaults to one on that position. So 1080 may also omit the "one" and written as "千八十".
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
#1: Nine disks need forty-one (41) moves. My solution is the "minimum of (two times a smaller number of disks plus the minimum of the remaining disk moves) method.
#2: An observation: if d (# of disks) < p (# of pegs), then you ALWAYS only need 2*d-1 moves. Basically, having more pegs than disks means you can move every disk to its own peg, including the base disk (to the designated ending peg), then move all of the other disks back on to the next bigger disk, which you should have already moved to that ending peg: (d - 1) + 1 + (d - 1) = 2d - 1.
#3: Likewise, if d = p, then m = 2d + 1, since you will have to move whichever disk is on the ending peg (or move another disk, if the ending peg is the only one available to the biggest disk), move the biggie, move the disk you had to move beforehand, and then move all the other disks onto the biggie: (d - 1) + 1(clear off the ending peg) + 1(biggie) + 1(split the double stack up with the starting peg) + (d - 1) = 2d + 1.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
Due to the PHP, you'll always have a pair with the same suit. Put one of those two cards in the front, so you get the suit. The same does not work for the value of the card, though, so we need another plan... At this point, you have 3 cards of encoding to figure out which of the 12 remaining cards (same digit, not the card that encodes the suit).
My first idea for encoding the digit of the card was to use the order of the cards. Since the first card is necessarily at the front, you only have three cards, or 3x2x1 permutations to work with, which isn't enough. My second idea was to use high-low order of the cards, which could use the first card for encoding. This gets to 2^3 or 8 different values, still not enough.
In the second idea, though, an edge case popped out where the card indicating the suit is the highest or the lowest value, and can't be "higher or lower" than the next. That can be solved by giving different suits higher values (i.e. diamonds are "lower" than hearts, etc. However, the edge case does point out that we can choose which of the cards to take, which gives us another "digit" to encode with. For clarity, we'll call them the numbers 1 through 13.
So how do we use those digits? If you always take the lower card, you get a worst case of 1, which still gives you 12 possibilities, which is not less than the 8 we need. What if we always take the card closer to 7, though? Well, 7 is the worst case, and give us 12 cards to dig through still. 6 would give you 1-5, 8-13, so 11
I gave up here and watched the video. Makes sense.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
Aww, this is sweet!... I knew it, that this clever way of dealing with the coin change problem must have been done already!
I've played with this coin change problem recently, deliberetely not knowing about backtracking approach or dynamic programming - I just wanted to see what can I discover on my own (later on, I found that I did not do anything radically new, but I was pleased anyway). I'd loved, of course, to find some formula for it, but theoretical approach was not my goal back then, I just wanted to make my computer to calculate the damn thing.
I was not particularly interested in calculating the minimum number of coins (of certain denominations) that add up to a given amount of money, but only in the number of ways such amount of money can be presented. I was also not very keen in writing computer program to do the job, just to make a table, filled with formulae that are calculating everything. The reason for all this is that for many years dealing with various problems I developed for myself a lazy approach to programming - part of the work (a huge chunk of it, if possible) can be done in table forms, another part - with much simpler program, if writing such a program is inevitable, otherwise custom functions are good enough in many cases (I mean, in Excel spreadsheets). So, practising such "programming" style helped me to learn how to compartmentalise any heuristic problem in smaller, more "edible" problems.
Firstly, I made a counter that gives me the answer, just for verifying the results (for relatively small amount of money, of course; for bigger ones it has to count for hours). Easy-peasy.
The analysis of the problem resulted in following table:
1. The number of columns = the number of 'n' distinct positive integer values (whole numbers), arranged in increasing order as c(1) through c(n); oddly, the results do appear in the leftmost column, associated with the smaller coin in the set, c(1).
The set of coins may be whichever one wants {1; 2; 5; 10; 20; 50; 100;...}, {1; 2; 5; 10; 25; 50; 100;...}, or some exotic: {3; 7; 12; 19}, {1; 2; 3; 5; 7; 11; 13; 17; 19;...}. One even can play dumb and use a set as {6; 9; 18; 30}, but the table will give back zeroes for all odd and not divisible by 3 sums anyway.
2. The number of rows is not limited above, with increment of 1 cent per row. Naturally, you should have 100 rows if you want to calculate the function for 1 dollar, or 250 rows if the target is 2 dollars and 50 cents.
3. In every cell [x, y] there is a formula that does one simple thing (for the leftmost column is slightly different) - it calculates the ways 'x' dollars can be presented by coins with nominal value >= c(y): cell[x, y] = sum( cell[x - c(y), j] ) for j = y to n / for 1st column: cell[x, 1] = cell[x-1, 1] + sum( cell[x, j] ) for j = 2 to n.
This is all, now every row will give you the corresponding number, for example, in the sequence A000008 (it is only for set [1, 2, 5, 10], doesn't include coins of 20, 50, 100 cents). If we use all coins and banknotes up to 20$ , {1; 2; 5; 10; 20; 50; 100; 200; 500; 1000; 2000}, the sum $20,21 will be presented in 30,399,653,516
ways. Clearly my counter will outlive me here ;-).
Now on, the hypothetical program is easy to be written and if one looks closer a little bit on the method, will see that this program will have to memorise maximum n*c(n) numbers to do its job (or even less, if you are greedy), for arbitrary large amount of money.
I was delighted that this problem has such simple practical solution, but thankfully, here I learned about the heavy, theoretical stuff, and they are also so elegant. The practical approach has some similarity with Pascal's triangle, so, no wonder that binomial coefficients do appear here and there.
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
3
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Mr. Burkard, first of all thank you so much for your Mathologer channel. It is wonderful. I enjoy each and every video and keep waiting for the next one :) I would like to share one thought I've got after watching the latest episode about the Ptolemy's theorem - suspecting it is quite trivial and seems different to me only - for which I apologize in advance. Here is what I thought: the Ptolemy's imparity can be generalized a little bit. Let say, we have a quadrilateral - any in fact. Add the diagonals, i.e. get all vertices mutually connected. We will have a number of line segments (original shape's sides and diagonals). Now split those in pairs such that the segments in each pair would not have common vertices. Then apply that naming convention you've been using in your video, where one pair would have A and a segment, the next - B and b, and C and c. Then it seems that the rule Aa+Bb>=Cc will hold despite which pair is a''s, b's or c's, that is c-pair should not necessarily be of the diagonals. And it is pretty obvious based on the proof you have presented. Then, the case of equality follows with all vertices located on a circle, and c-pair being the diagonals. Thank you so much for your time - and again sorry for being not enough educated to recognize a trivial result. Am already waiting for your next video! Best regards, Mike Faynberg
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Such puzzles and mathematical "mysteries" are like a concentrated food for always hungry young mathematical intellects - they're teaching them to mathematical beauty, elegance and aesthetics, and hone their logic, heuristic ability, pattern recognition, intuition (I saw the shortcut too, yesss!), making seemingly impossible connections between semantically distant objects or facts.
For not such young minds it is pure pleasure and recreational activity to tackle such problems.
So, thank you, sir, for making my day.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
So as shown in the proof, we know what the (x + 2)^n formula actually represents: we start with a 0-d "cube", which is just a point, so it gets assigned a polynomial p_0 = 1*x^0. Then every time we go up a dimension, we make two copies of the previous object (multiply the previous polynomial by 2), then another copy where all the elements are lifted one dimension higher - vertices become edges, edges become faces, etc - so we multiply the previous polynomial by x. That "lifted" copy connects the two normal copies together, and you get an (n+1)-dimensional hypercube from an n-dimensional one. Or in other words, p_(n+1) = (x + 2)*p_n.
Now you can think of what objects you could represent if you took a different "base" instead of x + 2. x + 3 doesn't seem to make much sense, as you're making 3 copies of the lower-dimensional objects, then connecting them together with only 1 of the higher-dimensional copies. Seems like one of the pieces will be left by itself, just floating out there. What if you took 2x + 3 as the base instead? It's easy to imagine getting three points connected by two edges after the first step, then you get a 3x3 square grid after the second step, and so on. In fact, you can take any k and get a k×k×k×... grid, then count its features with the polynomial ((k - 1)x + k)^n.
These seem like the most obvious options, maybe you could create other shapes by picking different numbers, but I'm not sure if they could be expressed with such a simple formula.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
squares or 'cubes'? THough as that honestly brings an impressively complicated line of thinking if applied towards dimensional logics.
Though as a for a gride, the number of squares is for a flat grid. is for each to points where it's height & length are the same,
I am sure this is poorly worded, but assuming a grid of 6x6, there are 36 points, & a shrinking list of pairs, & of that list x&y must be the same. (though a rough concept is in (0,0) (x,y) where y is 6x, 6y and the limit of grid size 6 squared, & the number of squares would be [Grid size=>]6 - Z[<=Largest value of X or Y] = Squares of (x,y) pair [where Z the number of highest value of the number pairs for that point, that is valid] & is repeated for each valid number pair.
-GS = Grid size (6x6 = 6, etc)
-- that said... the list of valid points would be [GS -1]^2 = VL (valid list of pairs that'd produce atleast 1 square)
y = [1, .... GS-1] (independent from) x = [1, .... GS-1] (treated as a range/array, that's lowest value is 1& no repeating numbers, sharing the same numbers - being a square)
- The number of squares for that grid would be
---- initial Z value is 0
---- for each x, do for each y, [GS - {X or Y, Highest value} = tmp] & z = tmp + z
Though for a cuboid structure.. it would that on the total assumption of cubic forms (GS^3)*z = TZ (total Z)
appologizes for the half programmatic thought on how it'd work. Not entirely sure how it's be written down, but that's how I thought.
This is of course, assuming 'squares' equal sides & flat.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Thanks for another fine work of videomath! Some of my thoughts going into this:
There are methods for accelerating the convergence of certain series, especially ones that are slow to begin with.
Sometimes they can be applied multiple times, in order to really streamline convergence.
Sometimes this will work at first, but then make matters worse after a repetition or two.
For this starting series, the "alternating odd harmonic (AOH) series," a very simple acceleration method is to add half of the first omitted term, as you showed early on.
This amounts to averaging each pair of consecutive partial sums.
BTW, another fairly good computational series for π is the Taylor Series for arcsin of one-half:
sin(⅙π) = ½
π = 6sin⁻¹(½)
It's not nearly as simple as the AOH series, but it converges a lot faster, with odd powers of ½. It would be most interesting to apply the techniques in this video to that series, I expect.
Of course, the Machin formula,
¼π = 4tan⁻¹(⅕) – tan⁻¹(1/239)
is much faster still, but it has those pesky powers of 1/239.
And then there was Ramanujan's series and its refinements...
Fred
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
About the parametrization of the unit circle using Pythagorean triples scaled down to have a hypothenuse of unit length:
Take the first box, about the 3-4-5 or 4-3-5 triangle. The fractions 1/2 and 1/3 are the Y-axis intercept (the parameter t of the parametrization) of the line between the point (-1,0) and the points (3/5, 4/5) or the (4/5, 3/5) respectively that exist on the unit circle. Each box is for one triangle, representing the two possibilities of which side lies on x axis.
This is a very rich subject!
This rational, pure math parametrization is explained beautifully by Norman J. Wildberger on his YouTube channel. Also there's another subject worthy of Mathloger exploration& perspective perhaps one day, the Miller Protractor and the core circle, x²+y²=x as key between projective (2 variants, an eighth of a turn of rotation away of each other, basically the x=y asymptote of the hyperbola of the red projective quadratic norm x²-y² rotated on the coordinate axis to yield what Norman calls the the green geometry based on the projective quadratic norm 2xy, (x²+y²)²=(x²-y²)²+(2xy)² is the among founding identities of chromogeometry) and the Euclidean, x²+y² -based blue geometry, it is marvelous. 😁
Thanks again for the great video I have been waiting for a long time! 😁🙏🙋
2
-
i think the F_n*F_(n+2)+F_(n+1)*F_(n+3)=F_(2n+3) thing would allow us to calculate a fibonacci number in a logarithmic amount of time, if multiplication and addition took a constant amount of time (but sadly, it takes longer)
edit: still, the naive way of calculating F_n would take O(n^2) steps, while this way takes only O(n*log(n)^2), so i'll go into more detail.
you first calculate some 5 consecutive fibonacci numbers, let's say 2,3,5,8,13 because that's the first time it skips anything.
then you do 2*5+3*8=34, skipping 21, and 3*8+5*13=89. calculate the 89-34=55 between 89 and 34, and then the next two numbers 89+55=144 and 144+89=233, and then you have 5 more numbers, so you can repeat:
34*89+55*144=10946 and 55*144+89*233=28657, and then fill in the numbers around that with simple subtraction and addition.
now, this on its own doesn't help calculate all fibonacci numbers, but you can adjust by shifting around the 5 numbers. so for example instead of using 34,55,89,144,233, we could calculate 144+233=377 and use 55,89,144,233,377, and get to different parts of the sequence. but you can figure out when to do that based on the index of the fibonacci number in something like binary (or maybe even exactly in binary)
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Also, I really like the falling factorial notation. I learned that combination C(n, k) = n! / (n-k)! k!, but I've always disliked that formula. First, if n is big enough, this makes it uncomputable as n! is too big. Also, it hides the fact that terms cancel out between n! and (n-k)!. If I note the falling factorial like this: n^k, then n^k = n! / (n-k)!, which seems like a much more useful notation, also easier to compute, and yields C(n, k) = n^k / k!, nice.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
14:23 I got 666. Hahaha, what's the obsession with the number of the Beast? (^.^) I'm not 100% sure I got this right however. Here are the matrices I got:
Arbitrary incremental number assignment of the 2 sets of colored tiles.
1 1 2 0 0 2 3 3 0 0 4 4 5 0 0 5 6 6
7 7 8 8 9 9 0 10 10 11 11 0 12 12 13 13 14 14
0 0 0 15 15 16 0 17 16 18 17 0 18 19 19 0 0 0
0 0 0 0 0 20 20 21 0 0 21 22 22 0 0 0 0 0
Magical 22x22 matrix (rows are green and columns are red).
1 1 0 0 0 0 i 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 1 0 0 0 0 0 i 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 1 0 0 0 0 0 0 i 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 1 0 0 0 0 0 0 i 0 0 0 0 0 0 0 0 0
0 0 0 0 0 1 0 0 0 0 0 0 0 i 0 0 0 0 0 0 0 0
i 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 i 0 0 0 0 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 1 1 0 0 0 0 0 i 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 1 1 0 0 0 0 i 0 0 0 0 0 0
0 0 0 i 0 0 0 0 0 0 1 0 0 0 0 0 i 0 0 0 0 0
0 0 0 0 i 0 0 0 0 0 0 1 0 0 0 0 0 i 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 1 1 0 0 0 0 0 i 0 0 0
0 0 0 0 0 i 0 0 0 0 0 0 1 1 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 i 0 0 0 0 0 0 1 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 i 0 0 0 0 0 1 0 0 0 0 i 0 0
0 0 0 0 0 0 0 0 0 i 0 0 0 0 0 1 0 0 0 0 i 0
0 0 0 0 0 0 0 0 0 0 i 0 0 0 0 1 1 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 i 0 0 0 0 0 1 1 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 0 0 0 0 1
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 0 0 0 1
Determinant of last matrix: 666 (positive real)
I actually calculated everything with this c++ code I wrote:
#include <iostream>
#include <vector>
using namespace std;
struct Ci{
int re,im;
Ci():re(0),im(0){}
Ci(const Ci&z):re(z.re),im(z.im){}
Ci(int n):re(n),im(0){}
Ci(int a,int b):re(a),im(b){}
Ci operator+(Ci z){return Ci(re+z.re,im+z.im);}
Ci operator*(Ci z){return Ci(re*z.re-im*z.im,re*z.im+im*z.re);}
bool operator!=(Ci z){return re!=z.re||im!=z.im;}
};/*complex integer*/
ostream&operator<<(ostream&out,Ci z){
if(z.im)
if(z.re)
if(z.im==1)out<<z.re<<'+'<<'i';
else out<<z.re<<'+'<<z.im<<'i';
else
if(z.im==1)out<<'i';
else out<<z.im<<'i';
else out<<z.re;
return out;
}/*complex integer console print*/
static Ci i=Ci(0,1);/*i constant*/
int pow(int b,int e){
int ans=1;
for(;e>0;--e)ans*=b;
return ans;
}/*integer power*/
void minor(vector<vector<Ci>>&ans,const vector<vector<Ci>>&m,int r,int c){
int h=m.size(),w=m[0].size();/*height and width of matrix m*/
ans.resize(h-1,vector<Ci>(w-1));
for(int k=0;k<h-1;k++)
for(int j=0;j<w-1;j++)
if(k<r)
if(j<c)ans[k][j]=m[k][j];
else ans[k][j]=m[k][j+1];
else
if(j<c)ans[k][j]=m[k+1][j];
else ans[k][j]=m[k+1][j+1];
}/*minor matrix*/
Ci det(vector<vector<Ci>>m){
if(m.size()==1)return m[0][0];
Ci ans=0;
for(size_t k=0;k<m.size();k++){
vector<vector<Ci>>m0;
minor(m0,m,0,k);
if(m[0][k]!=0)ans=ans+m[0][k]*pow(-1,k)*det(m0);
}
return ans;
}/*determinant*/
int main(int argc, char *argv[]) {
vector<vector<char>>m0={
{1,1,1,0,0,1,1,1,0,0,1,1,1,0,0,1,1,1},
{1,1,1,1,1,1,0,1,1,1,1,0,1,1,1,1,1,1},
{0,0,0,1,1,1,0,1,1,1,1,0,1,1,1,0,0,0},
{0,0,0,0,0,1,1,1,0,0,1,1,1,0,0,0,0,0}
};/*sunglass*/
char r=0,g=0;/*numbers to assign to tiles*/
for(int k=0;k<4;k++)
for(int j=0;j<18;j++)
if(m0[k][j]){
if((k+j)%2)m0[k][j]=++g;
else m0[k][j]=++r;
}/*number assignation to red and green tiles independently of each other.*/
vector<vector<Ci>>m1(r,vector<Ci>(g,0));/*final matrix*/
for(int k=0;k<4;k++)
for(int j=0;j<17;j++)
if(m0[k][j]&&m0[k][j+1]){
if((k+j)%2)m1[m0[k][j]-1][m0[k][j+1]-1]=1;
else m1[m0[k][j+1]-1][m0[k][j]-1]=1;
}/*1's of the final matrix*/
for(int k=0;k<3;k++)
for(int j=0;j<18;j++)
if(m0[k][j]&&m0[k+1][j]){
if((k+j)%2)m1[m0[k][j]-1][m0[k+1][j]-1]=i;
else m1[m0[k+1][j]-1][m0[k][j]-1]=i;
}/*i's of the final matrix*/
for(int k=0;k<4;k++){
for(int j=0;j<18;j++){
cout<<int(m0[k][j])<<' ';
if(m0[k][j]<10)cout<<" ";
}
cout<<endl;
}cout<<endl;/*glass print*/
for(int k=0;k<r;k++){
for(int j=0;j<g;j++)
cout<<m1[k][j]<<" ";
cout<<endl;
}cout<<endl;/*final matrix print*/
cout<<det(m1)<<endl;
return 0;
}
(^.^)
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
30:38
This is really poorly written so please don't judge but I threw this javascript code together. It runs in nodejs and contains all of the steps, but only shows the completed result in the console at the moment. If I have the time I might try to make a webpage around it and make it a lot nicer, or if anyone else wants to feel free to steal any of this code if it helps
function getRandom(num = 2){
return Math.floor(Math.random() * num);
}
class ArcticCircle {
constructor(num = 1){
this.size = num;
this.grid = new Array(num * num * 4);
for(let i = 0; i < this.size*2; i++){
let offset = this.getLineOffset(i);
let length = this.getLineLength(i);
for(let j = 0; j < this.size*2; j++){
if((j >= offset) && (j < offset + length)){
this.grid[this.getIndex(j, i)] = "_";
}else{
this.grid[this.getIndex(j, i)] = " ";
}
}
}
}
getIndex(x, y){
if(x >= (this.size*2) || y >= (this.size*2)) return -1;
if(x < 0 || y < 0) return -1;
return y * (this.size * 2) + x;
}
toGridCoords(line, column){
let nl = (line < this.size) ? line : (this.size*2) - line - 1;
let offset = this.size - nl - 1;
return [line, offset + column];
}
getLineOffset(line){
let nl = (line < this.size) ? line : (this.size*2) - line - 1;
return this.size - nl - 1;
}
getLineLength(line){
let nl = (line < this.size) ? line : (this.size*2) - line - 1;
return nl*2 + 2;
}
print(offset){
for(let i = 0; i < this.size*2; i++){
let s = "";
let padding = "";
for(let j = 0; j < offset; j++)padding+=" ";
for(let j = 0; j < this.size*2; j++){
let tile = this.grid[this.getIndex(j, i)];
let c = "\x1b[0m";
if(tile == '>') c = "\x1b[41m\x1b[36m";
if(tile == '<') c = "\x1b[43m\x1b[34m";
if(tile == '^') c = "\x1b[44m\x1b[33m";
if(tile == 'v') c = "\x1b[42m\x1b[35m";
s += c + tile + " ";
}
console.log(padding + s + "\x1b[0m" + padding + ".");
}
//console.log(this);
}
fill(){
for(let i = 0; i < this.size*2; i++){
for(let j = 0; j < this.size*2; j++){
if(this.grid[this.getIndex(i, j)] == "_"){
let tiles = (getRandom(2)) ? ['^', '^', 'v', 'v'] : ['<', '>', '<', '>'];
this.grid[this.getIndex(i+0,j+0)] = tiles[0];
this.grid[this.getIndex(i+1,j+0)] = tiles[1];
this.grid[this.getIndex(i+0,j+1)] = tiles[2];
this.grid[this.getIndex(i+1,j+1)] = tiles[3];
}
}
}
}
static GenerateNew(){
let circle = new ArcticCircle();
circle.fill();
return circle;
}
static EmbiggenCircle(oldCircle = this.GenerateNew()){
let circle = new ArcticCircle(oldCircle.size + 1);
for(let i = 0; i < oldCircle.size*2; i++){
for(let j = 0; j < oldCircle.size*2; j++){
let newI = i+1;
let newJ = j+1;
let tile = oldCircle.grid[oldCircle.getIndex(i, j)];
if(tile == "v"){
if(oldCircle.grid[oldCircle.getIndex(i, j+1)] != "^"){
circle.grid[circle.getIndex(newI, newJ+1)] = "v";
}
}
if(tile == "^"){
if(oldCircle.grid[oldCircle.getIndex(i, j-1)] != "v"){
circle.grid[circle.getIndex(newI, newJ-1)] = "^";
}
}
if(tile == ">"){
if(oldCircle.grid[oldCircle.getIndex(i+1, j)] != "<"){
circle.grid[circle.getIndex(newI+1, newJ)] = ">";
}
}
if(tile == "<"){
if(oldCircle.grid[oldCircle.getIndex(i-1, j)] != ">"){
circle.grid[circle.getIndex(newI-1, newJ)] = "<";
}
}
}
}
circle.fill();
return circle;
}
}
let circle = ArcticCircle.GenerateNew();
let iterations = 32;
for(let i = 0; i < iterations; i++){
circle = ArcticCircle.EmbiggenCircle(circle);
console.log();
circle.print(iterations-i);
}
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Great video! :)
Your Simple is King remark reminds me of how the divide-and-average method for calculating square roots is occasionally explained to a layman audience via incorrect reasoning in educational websites or modules. The very first “explanation” I came across in 7th Grade is if the initial guess x you pick to calculate √N is too small, N/x would be too large & vise versa. Then they erroneously infer that since √N is between x & N/x, the average must be closer to √N than both x & N/x [this doesn't immediately follow because 2 is between 1 & 1000 but I think we can all agree that (1+1000)/2 = 500.5 is a worse approximation for 2 than 1]
This lead me to grow a dislike for the method especially since the first legit explanation I heard of later on involved the Newton-Raphson method which I think is rather overkill for most smart 7th Graders. So, eventually I found an explanation that I think is way better using the AM-GM inequality & basic algebra.
Essentially, after applying the procedure of divide & average once, we end up with a new guess which is greater than √N which would then get closer and closer while stay greater than √N. So, it must converge to some limit L which, once we know it exists, we can solve for algebraically & get that L does in fact equal √N :)
The claims in the previous paragraph's first sentence can be proven with the AM-GM inequality & algebra, while the only bit that requires calculus to rigorously justify would be the rather intuitive fact used in the next statement that a decreasing sequence with a finite lower bound must have a limit.
I'm a bit saddened that the simple & valid mostly algebraic explanation is still way less popular than the simplified but blatantly flawed “explanation” that I heard in 7th Grade. Still, I'm glad people like you show simplified explanations without having to risk their validity! :)
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Today, I was writing a solution for the question in Michael Penn’s video by using this method and I did not know that you did this video (great video by the way), what a nice coincidence! If I were doing a youtube channel on math, this (something similar to that) could have been one of my first videos. I have been using this for more than 15 years already in action. So I love using Lagrange Interpolation/ Newton’s Divided Differences and use it in places where people don’t expect to use, as a number theorist, it’s brother Chinese Remainder Theorem is something I like too. Once, when I was in high school I wrote down first 25 primes in a big poster to see their pattern looking at the consecutive differences of primes and tryed to find a pattern for prime numbers. Later on, I figured that primes do not satisfy any polynomial. By the way, there is Gilbreath Conjecture on primes (considering consecutive absolute differences), as a number theorist, I am still interested to see if there is a connection between Newton’s Finite Differences and Gilbreath Conjecture. This is a very hard problem in professional math level…
Also I once trolled a facebook post, a what comes next question like the following:
The question was asking
9->90
8->72
7->56
6->42
3->?
I just thought of the sequence 9,90,8,72,7,56,6,42,3,19. In my post I said that the answer is 19 because I like 19!!! And explained that the terms of the huge 9th degree formula/polynomial (I used Wolfram Alpha to get the expanded polynomial of course! ) of the sequence exactly match with 9,90 etc… 😀. Because you can literally continue any finite sequence in any way, Lagrange Interpolation Theorem guarantees unique polynomials for each value of ?
2
-
2
-
2
-
2
-
2
-
2
-
2
-
So, I learned something today. That thing that my brain does with numbers sometimes, is called the digital root.
What I figured out when I was in my teens made me more interested in maths than any of my maths teachers could do.
(Note, Im only at 5 minutes in, so this might even be talked about.. who knows).
This is probably really obvious to most, but if you have a string of numbers where each double adds up to a multiple of 9 (eg 728145 [7+2=9 etc]) then that number is divisible by 9. So, 7+2=9, 8+1=9, 4+5=9, giving a root of 999 which is then rooted to 9+9+9=27, 2+7=9. Obviously the example looks divisible by 9 at first sight, but if the numbers were re-arranged to something less obvious like 785124, due to the fact that we can root sum to 9 means this number is also divisible by 9.
Now, you may be asking, in what situation would you need to know if a number is divisible by 9, where you wouldnt simply just divide the number by 9 to check. And youve got me there.
My brain just likes to point out useless weird facts to me every so often. :)
Another side note. I dont think its esoteric or anything. Its just a strange interaction between base 10 and the weirder numbers (3, 6 and 9).
Im now at 8m30s... wait... you were taught this in school? >>> Ahh, ok, so the first root is maybe taught, but the rooting to a single number isnt. OK, fair :)
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
5:12
Not that good in writing down the math here. But....
Please look at the start of the sequences.
The 1, 4 and 9 start as 1^2, 2^2 and 3^2. Of course the next one is 4^2 to begin with.
And of course, you can keep counting in between.
Either way, the intervals are 1. 1+1=2 2+1=3
Then X^2
The 3, 10 and 21 are the sums of 2, 4 and 6. Or in other words. (X^2 + X)/2, with intervals of 2. 2+2=4 4+2=6 etc.
The next start is at 6+2=8, then (8^2+8)/2=36
Thus 36^2+37^2+38^2+39^2+40^2 = 41^2+42^2+43^2+44^2
Now to look at the end of the left side of the equations.
We see 2, 6 and 12. Then for the power of 2 sequences we see 4, 12 and 24. That just so happens to be the double.
The 2, 6 and 12 follow the sumation again. What I mean is (X^2 + X)/2, but then times 2. Or just X^2 + X
The power of 2 is 2* (X^2 + X)
And you can see this at the 4th of both sequences as well. 20 and 40^2
Not sure what logic we can use for the cubic numbers though.....
2
-
2
-
2
-
2
-
2
-
I’m not sure, if I’m missing something; but that doubling of the product (like the 1 * 2, in the first example) seems, to me, a bit forced, and ”ad hoc”; like they were thinking: ”Damn! Can’t connect 1 * 2 to the ”3-4-5” -triangle. Let’s double it, and get 4; and if that didn’t work, we could have tripled it.”. Like, there’s no clear reason for, why you should double it; other, than getting it to work. Seems very random. The same kind of goes for the adding of the products, to get the area; like, now we suddenly went from multiplication to addition? What gives? 🤔😮🤷🏼♂️
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Disclaimer : I only watched up to around 14:00 before making this comment.
I had a thought for my program : This looks like the infinite series for 1/(1-x), but composed with the function x^n (for a n-cent coin), namely 1/(1-x^n).
So, instead of doing this giant product, which would get extremely hard extremely quickly, I just take my "vocabulary" of coins, find the product of (1-x^k) for k in my set of n coins, then do a partial fractions decomposition which gives me easy terms like 1/(1-x) and 1/(1+x)^2, which can easily be further expressed into their power series. Finally, I add them all up to get a general formula for the number of ways to make a certain amount of money using my coin vocabulary.
I'll admit, I didn't do the whole thing myself, most of the heavy lifting is done by SymPy, but I'm pretty happy about my program, especially after spending the last 4 hours on it.
Now I can finally watch the rest of the video.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
This is not quite the same as sum = product, but it might be related:
Consider the pair of pairs of numbers (1, 5), (2, 3). The sum of the first pair equals the product of the second, and the sum of the second equals the product of the first. Aside from the obvious (2, 2), (2, 2) and the trivial (0, 0), (0, 0), I can't find any other pairs of pairs of positive integers for which this relationship (ab = c+d and a+b = cd) holds.
If we allow negative numbers, then (-1, n), (-1, -n+1) is a general solution. I have no idea what non-integer solutions exist.
I have no idea if this is an already-explored topic, or whether it's of any significance. But I find it quite interesting.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
If we look at the finished magic square, we can imagine that it has been filled using the following process:
Put the average number 13 in the center.
Fill the anti-diagonal using consecutive numbers with 13 in the center, that is, with 10+1, 10+2, 10+3, 10+4 and 10+5.
These sum up to 65 as was explained in the video.
Now we fill the main diagonal with the numbers 3+0, 3+5, 3+10, 3+15 and 3+20, again with 13 in the center, also summing up to 65.
Now start at any remaining number on the main diagonal, e.g., 3+5, and move diagonally down/left counting up and diagonally up/right counting down
until all numbers of the respective group, i.e., 1+5, 2+5, 3+5, 4+5, 5+5 (or 1+k.5 to 5+k.5 in general) have been filled in.
Do this for each number on the main diagonal.
When you reach the edge of the magic square, cycle around!
Just imagine a second square adjacent to where you reached the edge and see which field of the adjacent square the next diagonal step would end up in.
Filling up the square diagonally this way ensures that all numbers of a group 1+k.5 to 5+k.5 are all in different rows and columns.
All numbers of the form k+0.5 to k+4.5 are also all in different rows and columns, because that is true for the main diagonal and the filling up process starts from there.
This ensures that all rows and columns add up to 65.
The cycling around when you reach the edge of the board is equivalent to the royal shifting method.
This is not quite rigorous but you get the idea.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
For the 3rd assignment at 9:40, I first looked in what ways 2020/4=505 can be written as a sum of two squares.
Mod 5, the only ways that two quadratic residues (0,1,4) add up to 505=0 are 0+0, or 1+4.
[Here observe that (0,0) is not a possibility, because it would imply that 25 would have to divide 505]
Mod 7, the only ways that two quadratic residues (0,1,2,4) add up to 505=1 are 4+4 and 0+1
Mod 9, the only way that two quadratic residues (0,1,4,7) add up to 505=1 are 0+1
Mod 11, the only way that two quadratic residues (0,1,3,4,5,9) add up to 505=10 are 1+9 and 5+5
Listing the candidates 0 through 22, these conditions each exclude some possibilities, leaving only the squares 1^2, 8^2, 12^2, 19^2, 21^2. By multiplying everything by 2, the 16 (ordered) ways to write 2020 as a sum of two squares are easily found now:
(+/-2×8)^2+(+/-2×21)^2=(+/-2×12)^2+(+/-2×19)^2=(+/-2×19)^2+(+/-2×12)^2=(+/-2×21)^2+(+/-2×8)^2=2020
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
About the coding competition: Is there a preferred coding language (example HTML5 with javascript, Processing, Unity, Scratch, Python+Pygame, C# windows application, Python Turtle, other)? keyboard interaction?, mouse interaction?, webbrowser?, windows/mac/linux application?, android/iphone?, zooming?, slow motion buildup?
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
The hyperbola in the spinning cube intrigued me.
Suppose the unit cube rotates about its main diagonal (from vertex (0,0,0) to vertex (1,1,1))
The largest distance to the main diagonal is at an edge, the interesting edges are those not connected to the rotation axis, for example x=0, y=1, 0<=z<=1.
The squared distance from a given point (0,1,z) to some point (t,t,t) on the main diagonal is
(0-t)^2+(1-t)^2+(z-t)^2. The distance to the diagonal has its derivative to t equal to 0, so that z=3t-1
We find that the curve is d = sqrt(6t^2-6t+2).
The interval for t that is featured by the cube is then [1/3 , 2/3].
My quadrics / cone sections are a bit rusty, who can show that this is a hyperbola?
2
-
2
-
2
-
2
-
2
-
It took me a long time to get to what the Mathologer pulls out of his hat after the first minute (following "whoa! What just happened?") Part of the problem was I steadfastly refused to accept that the three angle bisectors of the triangle meet at the center of the inscribed circle, without proving it according to Euclid--perhaps because of it's similarity to the flawed proposition that leads to the "theorem" that all triangles are isosceles. Getting past that hurdle, though, it's pretty straightforward to derive a formula for any triangle with a unit circle inscribed in it. Each colored segment (BTW of course you wouldn't use "blue" because it's name would be the same as side B) is simply equal to the cotangent of the half angle of each vertex. For the 3-4-5 you can apply the half-angle formulas to get the cotangents in terms of the cosines (sqrt(1+cosine/1-cosine) ), which for the 3-4-5 are 4/5/ 3/5 and 0, and you miraculously get whole numbers for each side. Time to unpause the video. It follows that the sum and product of the cotangents of any 3 angles that add up to 90 degrees must be the same. I guess the only hope of proving that is expressing them in terms of imaginary exponentials, and grinding through a lot of algebra. No geometric insight in that though. For another time.
BTW while Heron's formula in terms of the sides instead of the semi-perimeters might be preferable in some ways, if you're actually calculating it you annoyingly have to form 4 sum-difference combinations of the sides, whereas in terms of the semi, you can just stick it in a memory on your pocket calculator and re-use it. Did Heron have a pocket calculator? He was an engineer, wasn't he?
In all seriousness, Heron being an engineer clears up a question I've had for years: were this Heron of Alexandria and the one credited with inventing the first steam engine-which consisted of a steam boiler with jets on the side that set it spinning- one and the same? Apparently so. It's inefficiency, and other logistical problems problems, probably explains why the chariots of the day weren't powered by Heron engines.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
@ragnkja Yes! I had forgotten that. There's a logarithmic scale from 1 to 10 on the sliding part, and a mirror of the same scale on the frame of the ruler - this is for multiplication. And, there's another line of numbers on the frame of the ruler. Where the second scale has '1', the third has e, Euler's constant. Where the second scale has '2', the third has e^2. Where the second scale has '3', the third has e^3, and so on. So, if you slide the '1' of the slider to '2' on the second scale, the same place as e^2 on the third scale, then the slider numbers '2', '3', '4', and '5' match up with the second scale at 4, 6, 8, and 10, and match up with the third scale at e^4 (or, e^2, the starting number, squared), e^6(=(e^2)^3), e^8(=(e^2)^4), and (e^2)^5. Putting the '1' of the sliding scale at any value 'x' on the third scale means '2' will match up with x^2, '1.5' will match up with x^1.5; putting the '10' of the sliding scale at any value 'x' on the third scale means '5' will match up with x^0.5, '3' will match up with x^0.3, etc..
But the third scale only works for numbers between e and e^10. There's a fourth scale for numbers between e^0.1 and e, and depending on the slide rule, a fifth for e^0.01 and e^0.1, a sixth for e^0.001 and e^0.01, a seventh for e^10 and e^100, and an eighth for e^100 and e^1000. For multiplication, the second scale works for all powers of 10, but this method for exponentiation needs another printed scale for each power of 10.
I'm curious if there's a way to make a (finite) slide rule with just two scales that will compute one number to the power of another, generally.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
When you handshake someone, they also handshake you. Therefore, the total amount of handshakes having orcured, appears to be even. 0 + 1 + 2 + 3 + 4 + 5 + 6 = 21, and 21 isn't even. So seven people can't hold one by one distinct values from the set: (0, 1, 2, 3, 4, 5, 6). So there must have been a handshake twin.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
An n-cube is the n-fold product of (1 line + 2 vertices). In the formula, the 1 line becomes 1 x^1 = x and the 2 vertices become 2 x^0=2. The geometric shape described by this (x + 3)^n is the n-fold product of 1 line and 3 vertices. Ie. just draw a line with two end points and then another point on its own, next to that line:
o-o o
Now "multiply" that with itself n times. (A line times a line gives a (square) face, a vertex times a line gives a line, a vertex times a vertex gives a vertex.) The result is described by the formula (x+3)^n as in the Mathologer video.
For example:
(o-o o) x (o-o o) =
o-o o
| | |
o-o o
o-o o
You can check that the number of vertices, lines, and faces of this are described by (x+3)^2.
This formula generalises - start with V vertices, E edges, F square faces, C cubical cells, ... and represent them as a polynomial V x^0 + E x^1 + F x^2 + C x^3 + ...
Powers of this polynomial will describe powers of the associated geometric shape.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Here's my proof (I did not watch beyond 2:22). Clearly the three line segments have the same length, being each composed of one segment of each colour. Then any two of them are interchanged by the reflection in the inner angle bisector of the vertex where they intersect. It is well known that the inner angle bisectors intersect at the center C of the inscribed circle of the triangle. Then the mentioned reflection implies two equalities of distance from C between a pair of end points of the reflected segments (one of them being the ends of a pair of same colour whiskers, the other the remaining pair). One gets that for each of the three inner bisectors, and by transitivity these equalities imply equal distance of each of the six end points to C.
Viewing a bit more of the video, I can add that for the successive application of the three mentioned reflections in some order, there is in each case one of the segments S that is mapped successively to each of the other two segments and then back to S again, but with its end points interchanged. The combined action is the colour-reversing one in your "swiveling proof" that the midpoint of S is it point of tangency with the inscribed circle; you realised each swivel as a rotation, but a reflection does the same with less drama. Now the composition of any three reflections in lines concurrent in C is again a reflection with respect to some line passing through C (since the only isometries fixing C are rotations and such reflections, and the composition reverses orientation). Here that composition coincides with the reflection with respect to the perpendicular bisector of the segment S. The fact that this axis passes through C means that the midpoint of S (on this axis) is the point of tangency of S to the inscribed circle of the triangle.
When composing reflections through three concurrent lines, the result is the reflection in the line obtained by rotation the axis of the first reflection by the angle turning from the axis of the second reflection to that of the third. Applying that general fact in this situation gives a relation between the angles formed, at C, between the angle bisectors and the perpendiculars to the sides, and interesting triangle fact. Finally the fact that any reflection is its own inverse menas that if we apply the three reflections and then all three again in the same order, the composition will be the identity. The "orbit" of any point of the plane under these successive reflections gives six points, clearly all at equal distances from C; that's your windscreen wiper theorem.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
PUZZLES!
14:57 How many 1-S-S isosceles triangles are in the diagram?
7 * something, that's for sure
I'm gonna count the ones where the tip of the triangle is at the top most corner. There are 9 of those.
7 * 9 = 63
Or alternatively you can count by base.
On a 1 line, there's only 1 triangle.
On an R line, there are 3 triangles positioned like a W.
On an S line, there are 5 triangles positioned again in a zigzag pattern.
There are 7 of each of those lines, so there are 7 * (1 + 3 + 5) = 63 triangles.
30:12 Show that -1/phi is a solution to x^2 = x + 1 just like phi itself.
-1/phi + 1 = 1 - phi^-1
= (1/phi^2) * (phi^2 - phi)
= (1/phi^2) * (phi + 1 - phi)
= (1/phi^2) * 1
= (1/phi)^2 * (-1)^2
= (-1/phi)^2
31:51 What happens when you try other indices of fibonacci numbers?
Negative numbers are easy. You can just as easily extend the fibonacci numbers backwards as forwards.
2nd: 1
1st: 1
0th: 0
-1st: 1
-2nd: -1
-3rd: 2
-4th: -3
-5th: 5
It seems to be the case that the magnitude is the same but the sign alternates. Let's see:
F(-n) = (phi^-n - (-1/phi)^-n) / (phi - (-1/phi))
= ((1/phi)^n - (-phi)^n) / (phi - (-1/phi))
= (-1)^n * -F(n)
For non-integers, you have a (-1/phi)^n term in there. You're gonna get a complex number. As n gets larger the complex part disappears for the most part but generally it's... a bit weird
41:20 What happens when you do the R maneuver on a 1x1x1, then an S maneuver on that?
Algebraically, the R maneuver does this:
As + Br + C => (A+B)s + (A+C)r + B
Notice that the R term becomes the constant in the term below. In the pascalian grid, the same applies: the R term in one tile is the same as the constant term in the next one below in the R direction. So there are only TWO pascalian grids, not three.
Sorry I got sidetracked let's apply the R maneuver, then the S maneuver
1, 1, 1 => 2, 2, 1 => 5, 4, 2
Because RS = SR, we should get the same result if we do the S maneuver first, then the R maneuver
1, 1, 1 => 3, 2, 1 => 5, 4, 2
43:43
PF: Everything we've been doing with all these maneuvers is basically one big linear transformation. The matrices correspond exactly to the reverse R and S maneuvers earlier. Their inverses should be the forwards R and S maneuvers, then. The basis is (s,r,1) btw
Then it follows quite obviously that since we modeled these matrices exactly from the properties of the numbers themselves, then the matrix multiplication should obey those exact same rules.
The bottom row corresponds to what would happen if we multiplied this component matrix to the column vector (0,0,1), interpreted as a constant 1.
44:51 The chase is on for finding rho and sigma in nature!
Idk, look for anything heptagon-shaped. Pretty rare in nature because it's almost always 2's, 3's, or 5's
Obviously it's not gonna show up with something as simple as ruler and compass, you can't construct a regular heptagon using ruler and compass
47:14 What's next?
depends on how I'm feeling, really.
feeling with the theme of the video? 14. It's a fibonacci sequence.
feeling like one of those crazies you mentioned at the beginning? 2, it's the next digit of pi. close enough
feeling overly rigorous? literally anything, the operation isn't given to us. it could be the square root of R for all that is holy
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
I did your euler machine challenge.
Heres the C++ code: https://github.com/Bynar0b11001001/mathologer_challenges
I wanted to use String Arithmetics instead of something like bigint, but the functions for the string arithmetics turned out very messy
idk if by 666th u mean partition(665) or partition(666) so here are both
665: 11,393,868,451,739,000,294,452,939
666: 11,956,824,258,286,445,517,629,485
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Prime Three
1x1=, 1+1=2, 2+1=3,
5=1x2=2, 1+2=3+2=5,
7 1x3=3, 1+3=4, 4+3=7 now this will also create 19 as 3x4+3+4=7+12=19, how to get to....
11 is created from 2x3=6, 2+3=5, 5+6=11, as well 1x5=5, 1+5=6, So 23 and 15 or 32 and 51 create 11
27 and 35 will crteate 23, as 7x2=14+7+2=9+14=23,
29 is created from 14 as 14 then creates 54 then creates 20+9 as I multiply and add you can also subtract...
Crazy numeric messages and the link to the geometic dots and lines. Where are the lines , HOW to do..Quadrilium paper
I have 5000 videos some with sanity
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Back in 2016 when I saw Beck present the proof on Numberphile, I looked it up - I remember reading it in 1990 when it came out. Since I was teaching a number theory course at the time, I considered inlcuding it in my course, and so I studied it. Zagier says that it was not constructive , but I discovered that it is easy to use Zagier's involutions to give a constructive proof. Of course, I should say I rediscovered it for the ?th time. Letting S represent the switching involution and Z represent Zagier's windmill involution, take the fixed point of Z, namely x=1, y=1,
z = k=(p-1)/4 as the start, compute SZ((1,1,k)), (SZ)^2((1,1,k)), etc. That is, the orbit of the solution (1,1,k) under the mapping SZ.
Then RIGHT IN THE MIDDLE OF THAT ORBIT is found a fixed point of the switch involution S, that is, the sum of squares solution. Try it!!
Also, this is quite general:
TWO INVOLUTION LEMMA. If S and Z are two involutions on a finite set and v is a fixed point of Z, the the orbit of v under the mappint SZ will contain another fixed point of Z or a fixed point of S RIGHT IN THE MIDDLE! WHICH OCCURS DEPENDS ON WHETHER THE LENGHT OF THE ORBIT OF v is even or odd.
In Zagier's case, Z has a unique fixed point so the fixed point found must be a fixed point of S, which is the sum of squares solution. (This is the constructive version of Zagieer's proof.)
As Zagier points point, many of his statements are true when p is just an integer of the form 4k+1. What happens when you start with p = a product of two primes of the form 4k+1 OR with two primes of the form 4k+3 (so their product is of the form 4k+1)?
I know some answers. Try it!
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
I am uniquely placed to authoritatively answer your question, Ser Polster. It's because YouTubers use the English Wikipedia.
On 2021-09-24, Jade Tan-Holmes on the YouTube channel Up And Atom covered this, with a notable "What is this? Physics?" aside in the middle of the video. In response the next day I improved the English Wikipedia article vastly, based upon several university press mathematical textbooks. You will find me in its edit history. Until that point, the English Wikipedia article had only laid out the calculus approach. YouTubers, working from it, took its approach. I added all of the things that you mentioned to the English Wikipedia article, which at that point hadn't mentioned Oresme, hadn't properly laid out Torricelli's original proof, hadn't included Torricelli's original shape with the cylinder, or even supplied Torricelli's original name. I also added the explanation of the differences between "mathematical" and "physical" paint, and the contemporary 17th century philosophical and mathematical debates that ensued.
The French Wikipedia had already gained a graphic for Torricelli's shape, fortunately. Although its coverage at the time was better than the English Wikipedia's, it too didn't address Torricelli's proof in detail or go into things like Hobbes and Wallis and Barrow. I like to think that the French Wikipedia was spurred, by my efforts, into attempting to leapfrog the English one and be ahead again, because it has gained a lot more in 2022 and 2023.
I wrote in 2021 in a comment on the Up And Atom video that "One day, video makers referencing Torricelli will actually show the shape that Torricelli used in his proof. It's not that day yet, though. Hint: There's a cylinder on the end." It has taken over 2 years for that day to finally come. Well done.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
A little late to the party, but I loved this video and accepted the challenge to write a domino-dance program, stretching my modest Python skills a little bit. It doesn't use a square array but instead a dictionary to represent the grid. Keys are grid coordinate pairs; to represent the state at each coordinate pair, values are one of five vectors (four directions, 0 for empty). Domino flow patterns occur in the values of the dictionary. Probably this is quirky, but it makes a form-fitting (diamond-shaped) grid that extends naturally and performs decently, thanks to fast dictionary lookups. I don't know anything about graphics or web coding, so I used Processing and Trinket to share it. It has a couple of niceties for the user.
Here's a link: https://trinket.io/library/trinkets/5b574f6671
Version without code window: https://christopherphelps.trinket.io/sites/arctic_circle_theorem_simulator
Signed,
a Mathologer fan, math tutor & enthusiast, first-time commenter
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
In reference to the problem at (8:24), the magic constant of the King's 33x33 magic square is 17,985.
In general, the Magic Constant for the King's Magic Square is n*(n^2+1)/2 where n is odd.
As can be seen at (11:46), the magic constant is given by a sum natural numbers, and a sum of multiples. Specifically, it is the sum of the natural numbers up to and including n [1+2+3...+n], and the sum of the first n multiples of n [0n+1n+2n+3n...+(n-1)n]. We can be sure that this will always be the case, since when labeling the tiles, we always begin by filling a diagonal (as seen at 6:50) and then the squareness of the setup guarantees that the other diagonal will always also have the same number of tiles. Important to this proof, we know that 1+2+3...+n = n(n+1)/2. So, the King's magic constant (M) is:
M = (1+2+3...+n) + (0n+1n+2n+3n...+(n-1)n)
M = (1+2+3...+n) + n*(1+2+3...+(n-1))
M = n(n+1)/2 + n*n(n-1)/2
M = n*(n^2+1)/2
2
-
2
-
Do people have difficulty seeing 3+ dimensions?
1: Linear line of Points, going along One Axis.
2: Square Grid of Points, going along Two Axis.
3: Cubic Array of Points, going along Three Axis.
And continue...
4: A Linear line of Cubes, where each Cube has Three Axis of Measurement. Picture Multiple "Rubik's Cubes" laid out in a line. The 4th Dimension refers to which 3 Dimensional cube we are looking at.
5: A Square Grid of Cubes, where each Cube has Three Axis of Measurement. Picture Multiple "Rubik's Cubes" placed in a Grid. The 4th and 5th Dimensions refer to which 3 Dimensional cube we are looking at.
6: A Cubic Grid of Cubes, where each "Internal Cube" has Three Axis of Measurement. Picture Multiple "Rubik's Cubes" stacked into a Cube. The 4th, 5th, and 6th Dimensions refer to which 3 Dimensional cube we are looking at.
So, a 4 Dimensional scenario, such as the Location of a moving object, could have a set of 3 Dimensional Coordinates, with the 4th representing different times.
A 5 Dimensional Scenario, could be the Location of a moving object through different Mediums, could be 3 Dimensional Coordinates, with the 4th representing different Times, and the 5th includes all previous information, but each "Array" would represent a different Medium.
A 6 Dimensional Scenario, could be the Location of a moving object through different Mediums with different Engines.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
13:27 I would guess there is a O(N^2 * M) dynamic programming approach for that (N = the amount we want to partition, M = the number of distinct coins). I will assume we have a sufficient amount of every coin type. Let's define count(i, j) as being the number of ways to partition i dollars with the first j coins (coin values are sorted in ascending order). Then count(0, j) = 1 for all j and count(i, 0) = 0 for all i. The recursion comes out to be count(i, j) = count(i, j-1) + count(i - c[j], j-1) + count(i - 2*c[j], j-1) + ... The answer will then be stored in count(N, M). If you do some clever stuff with the memoization, the time complexity can also be reduced to O(N * M), I think.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
In the beginning of the video, Mathologer mentions the Gaussian integral as ∫[−∞,∞] e^(−t^2) dt = √π.
However, at 10:30 ish, he's not using the exact same integral anymore; instead, he's using ∫[−∞,∞] e^(−t^2/2) dt.
Starting with the original Gaussian integral ∫[−∞,∞] e^(−t^2) dt = √π, we can perform a change of variables u/√(2) = t to obtain
1/√(2) · ∫[−∞,∞] e^(−u^2/2) du = √π , from which we conclude
∫[−∞,∞] e^(−t^2/2) dt = √(2π).
2
-
2
-
I found the adjacency matrix for the glasses and the determinant is 666. Maybe I did something wrong because you said the determinant doesn't give the answer straight away?
Matrix that you can input to a CAS:
[0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,1,i],
[0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,1,i,1],
[0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,i,1,0],
[0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,1,i,1,0,0],
[0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,1,0,i,1,0,0,0],
[0,0,0,0,0,0,0,0,0,0,0,0,0,0,1,i,0,0,0,0,0,0],
[0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,i,i,1,0,0,0,0],
[0,0,0,0,0,0,0,0,0,0,0,0,1,0,0,0,1,0,0,0,0,0],
[0,0,0,0,0,0,0,0,0,0,0,1,i,i,0,0,0,0,0,0,0,0],
[0,0,0,0,0,0,0,0,0,0,0,0,0,i,1,0,0,0,0,0,0,0],
[0,0,0,0,0,0,0,0,0,0,1,i,0,1,0,0,0,0,0,0,0,0],
[0,0,0,0,0,0,0,0,1,0,i,1,0,0,0,0,0,0,0,0,0,0],
[0,0,0,0,0,0,0,1,i,0,0,0,0,0,0,0,0,0,0,0,0,0],
[0,0,0,0,0,0,0,0,i,i,1,0,0,0,0,0,0,0,0,0,0,0],
[0,0,0,0,0,1,0,0,0,1,0,0,0,0,0,0,0,0,0,0,0,0],
[0,0,0,0,1,i,i,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0],
[0,0,0,0,0,0,i,1,0,0,0,0,0,0,0,0,0,0,0,0,0,0],
[0,0,0,1,i,0,1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0],
[0,0,0,i,1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0],
[0,1,i,1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0],
[1,i,1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0],
[i,1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0]
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
I'm a little sad you left out the hyper triangles: they can be categorized in a similar fashion by Pascal's Triangle, which of course contains the coefficients to (x+1)^n. That is,
1
1 1 (if we truncate the final one in each row, this marks 1 vertex)
1 2 1 (again truncating the final one, 1 edge and 2 vertices)
1 3 3 1 (similarly, a triangle has 1 face, 3 edges, 3 vertices)
1 4 6 4 1 (the tetrahedron has 1 cell, 4 faces, 6 edges, and 4 vertices)
&c.
2
-
This video has in it the first thing that I ever saw that was interesting (to me) as mathematics. As far as I knew before then, mathematics functioned only to accomplish something else, in particular far-out physics and miraculous inventions. I therefor took algebra (and got all A's) because that was what I was told Einstein was writing on the blackboard in those nostalgic, old photographs. I was disillusioned and frustrated (as teenagers are) to find by the second semester of algebra that there was nothing yet that could honestly be claimed to be useful or practical whatsoever, and no clues to how or when that could ever happen. Then a thermonuclear device was set off in my brain, by an incidental, optional section of the textbook, just there for its own sake, unrelated to the rest of the chapter. The well-known observation that the differences between successive squares are the odd numbers, also means, they said, that the second differences are all 2. Likewise with cubes, the third differences are all 2*3. Likewise with fourth powers, the fourth differences are 2*3*4. Mind totally blown, I had to calculate this by hand (no personal calculators in 1959 and I hate arithmetic), for larger and larger numbers, and higher and higher powers, over the next few weeks, just to see it (not that I doubted it), just like watching Mr. Wizard (Don Herbert) freeze a hot dog in liquid air and break it on the floor never gets old. I thought I should be able (since I got A's) to figure out why this works from what had been in the book, so I spent the next couple of months working on it (while higher priority things needed doing) without success. (But not without finding chains of things that were not in the book.) Of course the cause of the phenomenon would be in some book somewhere, but that's not the point, anymore than some one else looking at the sunset is the same thing as you looking at it. Blessing to the textbook author, who probably had to lawyer to keep it in, for putting that incidental aside in the book. And blessings to mathologer.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Proof of the rainbow square corners:
I'll be referring to the example at 13:00 a lot, as "the canonical example".
All right, we need some lemmas.
Lemma 1 ("alternating stripes"): In any 1xN subregion in which no run of three squares have three different colors, the region alternates between two colors.
Example: the top row of the canonical example.
Proof: consider the two leftmost squares (rotate WLOG an Nx1 region into 1xN). Since they are inside a 2x2 subgrid, they must have different colors; call them A and B. Consider now the third leftmost square; since no three run of squares have three different colors, it must also have color A or B. But since it is adjacent to the second leftmost square, which has color B (WLOG), it cannot also have color B; hence it must have color A. Consider now the second, third and fourth squares; they must have colors B, A, B by a similar argument, and in general every square must have the same color as the square two squares to its left.
Corollary 1: if the board has no 1x3 region (in either orientation) with three different squares, no two corners have the same color.
Example: swap yellow and green in the bottom row of the canonical example; it now conforms to this condition.
Proof: by the same argument as above, every square must be the same color as two squares to the left, and two squares up. Hence if you divide the grid into disjoint 2x2 squares (this is possible since the side length is even), each square has the same color pattern as every other square. But then the bottom left corner of the whole square has the same color as the bottom left of the top left 2x2 square; and likewise, the corners of the top left 2x2 square has the same color as the corners of the whole square. But the corners of every 2x2 square are all different.
Lemma 2 ("zig-zag bands"): if there is a vertical run of three squares of three different colors, this determines the complete coloring of the containing three rows. Likewise for horizontal runs.
Example: the lower three cells in the leftmost column in the canonical example.
Proof: consider three squares with coordinates (r-1, c) and (r, c) and (r+1, c), colored A, B and C respectively (example: yellow, red, green). Consider now (r, c+1). It is in a 2x2 square with (r, c) and (r+1, c), so it cannot be color B or C. It is in a 2x2 square with (r-1, c) and (r, c) so it can not be color A or B. But then it must be color D (e.g. blue). Consider now (r-1, c+1). It is in a 2x2 square with (r-1, c) and (r, c) and (r, c+1), colored A, B and C, so it must be color C. Likewise, (r+1, c+1) is in a 2x2 square with (r, c) and (r+1, c) and (r, c+1), colored B, C and D, so it must be color A. But note that (r-1, c+1) and (r, c+1) and (r+1, c+1) are themselves of three different colors. Repeat this analysis, extending to the right edge. Now mirror this analysis and also extend it to the left edge. The pattern you will see is that the middle row (row r) alternates between colors B and D, and the two other rows alternate between A and C, one row starting with A and the other row starting with C. Every intersection of this three-row band with a column contains three squares of three distinct colors.
[I call them zig-zag bands, as the yellow color zig-zags in knight's moves around the alternating red-blue central row; so does the green color.]
Corollary 2: there cannot both be a vertical and a horizontal zig-zag band (= run of three squares of different colors).
Example: none, they don't exist ;-)
Proof: the middle column of every vertical zig-zag band is an alternation between two colors. Every column intersects a horizontal zig-zag band at three squares of three different colors; this includes the middle columns if vertical zig-zag bands. Three squares cannot simultaneously have exactly two and exactly three distinct colors.
Corollary 3: if there is a horizontal (vertical) zig-zag band, every row (column) is an alternation of two colors; and each row (column) uses the opposite pair of its two neighbouring rows (columns).
Example: the rows of the canonical example.
Proof: let a horizontal zig-zag band be given, and consider its central row, colored A and B in an alternating pattern. Every square in an adjacent row is in a 2x2 with two squares from the central row, colored A and B, hence the two adjacent rows must be colored C and D. But then by lemma 1 they must alternate. But then by a similar analysis their neighbouring rows must alternate, using colors A and B. Apply this reasoning inductively until you get to the edges of the board (in both directions).
Theorem: in every rectangle with even side lengths, with each tile colored in one of four colors such that every 2x2 square of tiles has all four colors present, the corners have different colors.
Example: the canonical example.
Proof: if there is no horizontal (vertical) zig-zag band, this is corollary 1. If there is a zig-zag band, the top row alternates between colors A and B, and since the vertical side length is even the bottom row alternates between colors C and D. Since the horizontal side length is even, any row in which two colors alternate (that's all of them), the leftmost and rightmost squares have opposite colors. This applies in particular to the top and bottom rows, and so the four corners have colors A, B, C and D, all different.
(4n-by-4n squares are all examples of 2m-by-2k rectangles, if we choose m = k = 2n.)
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
In the Doctor Who episode, even the toymaker got bored of watching the Doctor moving disks, so he accelerated the pattern, making hundreds of moves at once and then the Doctor had to continue. What is the best way for the Doctor to proceed? He either needs to move the smallest disk or make the only other valid move.... but which one? If he knew the move number, then it would be trivial, but here's the question: Given an state within the optimal move sequence, how do you continue?
I once decided to see what the Towers would sound like as music. I gave each disk a note in C major, 1=C, 2=D, 3=E and played each note for each moved disk. So CDCECDCFCDCECDCG... etc. Played as simple 8th notes, it gives a very mechanical feel with the CDC making a continuous pattern and the other notes seeming sort of random. It gives a very mechanical vibe, like something you could play over Metropolis.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
CHALLENGES!
3:20
Let y = 1/x.
Squishing vertically by a factor of c corresponds to replacing y with cy.
Stretching horizontally by a factor of c corresponds to replacing x with x/c.
Now substitute: cy = 1/(x/c)
cy = c/x
y = 1/x
And we recover our original shape.
Everything in the interior stays in the interior for obvious reasons.
8:53
A(X) ranges from 1 to X, and has width X - 1 and height 1/X on the X end.
A(Y) ranges from 1 to Y, and has width Y - 1 and height 1 on the Y end.
The A(Y) shape needs to squish by a factor of X to fit. And it should stretch by that same factor of X.
The total width should be (X - 1) + X(Y - 1)
= X - 1 + XY - X = XY - 1, exactly the width of the shape from 1 to XY.
Or you can use FTC and notice that the antiderivative of 1/x is ln(x) + C, and that ln(X)+ ln(Y) = ln(XY). Provable with the integral definition but that's for later [°1].
10:48
Just use the subtraction formula!
N = 0: log(X) - log(X) = log(X/X)
0 = log(X^0) = log(1)
N = -1: log(1) - log(X) = log(1/X) = log(X^(-1))
N = -M: Nlog(X) = log(1) - Mlog(X) = log(1) - log(X^M) = log(1/(X^-M)) = log(X^N)
Rationals is slightly trickier.
R = P/Q: log(X^(P/Q)) = P * log(X^(1/Q)). Now multiply and divide by Q (the analytic version of shape-shifting I guess)
P * (Q / Q) * log(X^1/Q) = (P/Q) * log(X^(Q/Q) = R * log(X).
For reals, I'm just gonna say log is continuous. I'll leave a non-rigorous "proof" for later [°2].
[°1] Integral definition
Suppose ln(x) was defined by this integral:
Integral[1/t dt; from 1 to x]
We want to prove ln(xy) = ln(x) + ln(y).
Integral[1/t dt; from 1 to xy] can be split into two:
Integral[1/t dt; from 1 to x] + Integral[1/t dt; from x to xy]
The first half is ln(x) by our definition. The second requires a little work.
I want to take x to 1 and xy to y. Setting u = t/x will do the trick.
u = t/x
du = dt/x
lower bound = 1
upper bound = y
Integral[x/t * dt/x; from x to xy]
Integral[1/u du; from 1 to y]
which is exactly equal to ln(y).
Notice what we did here. We split the integral into two, and rescaled one of them to fit.
That's the reverse of the shape-shifting magic! And the u-substitution is that magic shape-shifting squash and stretch in the geometric proof!
[°2] Probably Continuity of the Logarithm
We know the area function is defined for each rational number. Clearly the area function is also monotonic since if A(x) < A(y) if and only if x < y.
I can approximate any real number using a sequence of strictly increasing or strictly decreasing rationals. For example, I can approximate sqrt(2) with the sequence 1, 4/3, 7/5, 41/29, 239/169, ... from below, or 2, 3/2, 17/12, 99/70, 577/408, ... from above. Since A(x) is monotonic, we expect the first increasing sequence to output areas that increase, and the second to output areas that decrease. The LH limits and RH limits exist, and they can't explode because LH limit <= RH limit.
Also, the integral of a function is usually continuous, especially of the integrand itself is continuous. We know quite easily 1/x is continuous everywhere ln(x) is defined (the interval (0, +inf]) so we expect ln(x) to also be continuous.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Today has been something else for me. My wife and I had a break through in our relationship. We came to terms with something incredibly taxing on us.
Second, I discovered a video on the Richat structure in Africa. Stirring a curiosity in my mind I haven't felt in... maybe ever.
Third, through that curiosity I found you. I have never sat through a lecture, not without fidgeting, smoking, going for a snack. Nothing would keep my interest. I barely took a breath watching your explanations of 3,6 and 9 and now this video.
As someone who has felt like he was awful at math, only ever memorizing material. Even in college. I now understand why. The way I learn was never catered to, in the least it seems. I perhaps could've spent my whole life not understanding why we use Pythagoras for things, other than it just works. But this has shown me the beauty in mathematics. No wonder so many strive to this. Wow. I feel like an empty cup, eager to be filled. Thank you for making these, genuinely from the bottom of my curious little heart.
2
-
2
-
Well, here's my "ahaa" recorded:
[Spoilers ahead]
The numbers of tilings with identical 2×1 pieces of 2×n plains seem to match up with the Fiblnacci numbers, such that for a plain of 2×n the number of possible tillings is F(n+1).
So let's see why.
Before proceeding, note that any 2×n board can be covered with "vertical" tiles, whose height is 2 and fills completly the 2 unit long side of the board. Also, note that any 2 parallel tiles can be rotated over their joined centre (so that vertical tiles now become horizontal). These are the only two ways of tilling a 2×n board.
For the 2×0 board the one possibility is to use 2 tiles. Duh.
For the 2×1 board the one possibility is to use one vertical tile. Duh.
For the 2×2, things start to grt interesting. Now, you can use 2 vertical tiles to make the all-vertical tilling, or rotate the 2 to make a tilling with 2 horizontal tiles.
2×3 allowa for the all-vertical tilling and for 2 other tillings: tiles 1 and 2 may be rotated, or tiles 2 and 3 may be rotated intead. There still isn't enough space to rotate more that one pair of tiles. Vertical: 1; 1 horizontal: 2; 2 horizontal: 0.
With a 2×4 board, the results are as follow: vertical: 1; 1h: 3 (=4–1); 2h: 1.
K, we need formulas. For the all-vertical tilling, it is allways possible, and thus v=1.
For 1 horizontal (one pair rotated, h1), since the pair itself is 2 units whide, it is n–1 for n greater that 1 (and I guess it also works for 1, as it just spits 0), where n is one side of the board (2×n is the board).
For more than one pair of horizontal tiles, this gets complicated. For 2×4, there's one possibility. For 2×5 there's 3. I will further subdivide these into groups defined by how much space there is to the left of the leftmost horizontal tiles.
If they are right touching the side, then there's as many possibilities as for covering a 2×(n–2) board, due to the space they take up. If they were 1 unit away, there would be as many possibilities as for a 2×(n–3) board. Since I cannot type epsilon, I will just explain the infinite sum that is about to take place. Say t(2×n) is the amounth of ways to tile a 2×n board; then t(2×n) = t(2×n–2) + t(2×n–3) + t(2×n–4) + t(2×n–5) + t(2×n–6) + t(2×n–7) + … + t(2×3) + t(2×2) + t(2×1) + t(2×0) + 1
The last one is for the all-vertical tilling.
But don't you worry, we can simplify this a bit by noticing that t(2×n–1) is the sum of all of the above except for t(2×n–2), and thus t(2×n) = t(2×n–1) + t(2×n–2)
Beautifull.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
@Mathologer
Sorry I'm not sure what happened to the other comment, I think I accidentally deleted it. But I was informing of Season 2 Episode 12. It's a math geometry problem that went like this:
"In the figure on the right, cubes with side length "a" are stacked and arranged repeatedly, with an atom placed at each vertex and in the center of each individual cube as shown, in the form of a body-centered cubic crystalline lattice. Assume that the structure is composed of sodium, potassium, or some other alkali metal. Within the body-centered cubic crystalline lattice, the atom at the center is point A{0}. Inside the cube, the region that is closer to A{0} than any other atom is D{0}. Find the volume of D{0},"
You're gonna need the figure but I haven't found a good clip. If you see the episode you can see how two students arrive at the answer using similar means. A channel called CyclicSquares did an explanation (Video: Weird Geometry) but I'd like to see people try a "simpler" explanation I guess...
I was going to link a writeup but I guess youtube doesn't like links in the comments, google it and you'll find it. Good Luck!
:D And thanks for the videos!
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
That was fun!
For power = 1 or 2 the differences are zero, as shown in the video.
For power = 3, the differences are 2* triangle(n)^2.
2 18 72 200 450
For power = 4, the differences are 64* triangle(n)^3.
64 1728 13824 64000 216000
For power = 5, the differences are
2002 162066 2592552 20002600 101258850
.... and I'm stumped!
(2002 = 2*7*11*13, not that THAT is any help!)
Here some R code for it.
multipliers = c(2, 4, 6, 8, 10)
pow = 5
results =
sapply(1:5, function(n) {
tri_pow = triangle(n) * multipliers[pow]
a1 = sum( ( (tri_pow - n) : tri_pow)^pow )
a2 = sum( ((tri_pow + 1) : (tri_pow + n) ) ^pow )
diff=a2-a1
guess3 = 2* triangle(n)^2
guess4 = 64 * triangle(n)^3
return(c(n=n, pow=pow, tri_pow=tri_pow, a1=a1, a2=a2, diff=diff, guess3=guess3, guess4=guess4))
})
results
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
I think the answer to the initial puzzle is off summing i**10 from 1 to 1000. Double checking with Ruby I get 90,409,924,241,424,243,424,241,924,242,500 whereas the dear professor got 91,409,924,241,424,243,424,241,924,242,500?
My program was merely:
def sum
a=0
for i in 1...1000
a+=i**10
end
return a
end
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
I have a question?
If with-out removing any numbers the series goes to infinity (initial case in the video); and if removing the “9’s” the series does not go to infinity (special case you had us wager). Does that mean if you only sum up the series with “9’s” does it grow to infinity?
A is the set with out removal (1,2,3,4,5,6,7,8,9,0*)
B is the set with removal of 9’s (1,2,3,4,5,6,7,8,0*)
C is the set with just 9’s (9)
•The set A = set B + set C )
The set A grows to infinity
Then (set A - set C); are the ones with the 9’s removed and Converges (not infinity)
So then does set C which is (set A - set B) diverge grow to infinity
???? Thank you in advance for your response
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
≈ 5½ min – Lucas (pron: LOO-kah) numbers, are sums of pairs of Fibonacci numbers that are next-but-one:
0 + 1 = 1
1 + 2 = 3
1 + 3 = 4
2 + 5 = 7
3 + 8 = 11 . . . Lᵢ = Fᵢ₊₁ + Fᵢ₋₁ from which their adherence to the same recursion rule as the Fibonacci numbers, follows.
But I wonder – is there a relation of these to the Suzanne Vega song, "Luka"? It's pronounced the same!
I know we've touched on this in previous comments, that you too are a Martin Gardner fan; so I'm curious whether you got the idea for this from his early MG column on Fibonacci numbers? In it he mentions extending the Fibonacci recursion rule to k'th order, still with all 1's as coefficients, and that as k → ∞, the limiting ratio of consecutive terms → 2⁻⁻.
That limiting ratio, R(k), is always the root of the characteristic polynomial, that's between 1 and 2.
R(2) = φ; φ² – φ – 1 = 0 (Fibonacci)
R(3) = x; x³ – x² – x – 1 = 0 (Tribonacci)
R(4) = y; y⁴ – y³ – y² – y – 1 = 0 (Tetranacci)
. . . etc.
The "Infininacci" numbers (k=∞) are just the powers of 2.
Another curiosity – the discarded solution, "phi-red", –1/φ, is also the consecutive-term limit of the Fibonacci sequence –– but in the negative direction, where the signs alternate!
..., 13, –8, 5, –3, 2, –1, 1, 0, 1, 1, 2, 3, 5, 8, 13, ...
So the ratio F(n)/F(n-1) → –1/φ as n → –∞; and it → φ as n → ∞.
Question: Is there an analogous rule when the recursion order, k > 2?
Danke schön, Herr Polster, for a splendid presentation!
2
-
2
-
2
-
2
-
2
-
focus on one intersection of two supposed chords and notice that it has the same power wrt one pair of endpoints as the other (the colors show it) and this means the ends of those two chords lie on the same circle, now if we show that an endpoint of horizontally oriented chord lies on the smame circle as the circle formed by endpoints of other two chords, we are done. Label P,Q;R,S;U,V the endpoints in the positive angular direction starting with horizontal red endpoint and A the red vertex of triangle, B blue vertex and C green vertex. Since PBV is isosceles, angle <BPV is pi/2-beta/2, <APQ is pi/2-alpha/2, so <VPQ is pi-(alpha+beta)/2 and because <VUQ is equal to <VUC which is pi/2-gamma/2, it follows that pi-<VPQ=<VUQ which means VPQU is a cyclic quadrilateral, which finishes the proof
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
5:27 Iron Man proves the beard man formula at the bottom of the left page!
Let E(m,n) be the number of m-cubes in an n-cube. An n-cube is made up of two (n-1)-cubes whose corresponding parts are all connected. So, for example, E(2,4) (the number of 2d faces in the 4-cube) equals 2*E(2,3) (the number of 2d faces in the two 3-cubes) + E(1,3) (since connecting a pair of corresponding 1d edges creates a 2d face).
In general we get E(m,n) = 2 E(m,n-1) + E(m-1,n-1), which is the recurrence relation on the page there. From here you can prove the beard man formula by induction. Or, if you want to be fancy, you can use generating functions to prove the (x+2)^n expansion thing directly :)
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Firstly: brilliant video, explanation. Consider my mind blown.
1. There are many examples of where processes and decisions expressed in flow can be used to solve problems in an ideal way. Traffic jams are one example, where most people would assume it would help to have higher throughput on the main roads by having longer traffic light phases of go, when in actuality, a "proportionally" bigger share of green light should be given to side roads leading to and from main arteries. I would love to see if communicating vessels could be used in a similar way here. Possibly these issues are equivalent?
In the traffic jam example, some solutions are found by simulating the traffic as a directed graph and "water" is pressed into the "pipelines" and you check where the water leaks :)
2. In a lot of evolution simulations there is a significant issue with predation. The mere existence of predation in a simulation affects survivors, survival strategies and life expectancy much more than energy density, regeneration and most other factors, one would assume to have a significant impact. I am wondering if the idea of using a solution like this, that prefers a happy outcome, could be used to in simulations to "curb" predation. Without dragging this into a political discussion, these patterns can conversely be seen in the real world. It would be fascinating to see if such an approach on a fundamental level could lead to an overall "happier" outcome for everyone.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
7:30 Hmm, at first I was wondering if this could be described as Pascal's triangle mod 3, but it quickly became apparent that it couldn't, and also that this algebra isn't even associative.
(R+B)+Y = Y+Y = Y but R+(B+Y) = R+R = R
All that idempotency must be getting in the way.
Maybe there's still a mod 3 isomorphism going from the tip (like Pascal's actual triangle), but using colour shifts for different bands of hexagons or something. With the rotational symmetry of this "algebra", any tip would do.
18:22 Well, I guess it's going to be clearer soon!
20:50 !
Guess I wasn't too far off with that thought about colour shifts, in the end. Flip red for blue every other row (but not starting from the tip).
28:30 If C = -A-B, then -A-C = -A-(-A-B) = -A+A+B = B. Hmm, yep, that works out.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
There are 10 essentially different triangles of width 4: BBBB, BBBY, BBYB, BBYY, BBYR, BYBY, BYYB, BYYR, BYRY, BYRR. Now, I have no idea why...
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
You've just asked (24:27) what happens if you apply some ears repeatedly, in a combination of ears. Looking at the "eigenpolygons", I notice they're not exactly axis-aligned, although three of them are close and the other is almost diagonal-aligned for one of its points, but I'm guessing that's a distraction and what we're doing is, for an original p-gon with corners represented by complex numbers z(n) for n in p = {0, ..., p-1}, a decomposition of form z(n) = sum(j in p: q(j)*cycle(j*n/p)) where cycle(t) = exp(2*pi*i*t) is the unit-period traversal of the unit circle; q(0) is the centroid (cycle(0) = 1 so j = 0 just contributes q(0) to the sum), the phase and scale of other q(j) configure the orientation of the p-gon traversed j steps at a time. The real fun must be in proving that, for any z, there is a q that makes this work; but, give or take normalisation (my guess is 1/p), I'm guessing q(j) = sum(j in p: z(n)*cycle(j*n/p))/p will do the job. Then application of each of the "ears" is linear in the two end-points of the edge to which it's applying the ears, so maps onto doing the same to each cycle(j*n/p) polygon contributing to the sum. The effect on these regular polygons is a simple scaling, scaling each q(j) by a j-dependent (complex) factor that's zero when the ear is either j/p or 1 -j/p turns (depending on the direction of our traversal, I suspect). If you apply k/p turn ears for each k from 1 through p-1 you'll thus annul all q(j) except q(0) and be left with the centroid. Repeated application of a given "ear" angle thus just repeats that scaling of each q(j) by the same j-dependent factor, so doesn't annihilate any q(j) that the first application didn't. So if you apply k/p turn ears for all but one of the k from 1 through p-1, you'll get a regular p-gon; and repeating some of those ears will only change the scaling. Now to resume watching and see if I'm wrong ...
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Great video as always (I know this comment comes a bit late)! What is even more mind-blowing (and would require another Mathologer video!) is that even the set of numbers that we can describe with a finite number of words (and symbols) is countable. This implies that the vast majority of all real numbers remains for us a complete mystery beyond our grasp.
To be more precise, let us decide a set of symbols, for example S=English Alphabet U Digits U parentheses U logical connectives (and, or, not, imply) U {symbol for belonging} U quantifiers. We may want to add some more special symbols (for example +, /, etc.) so in the end we may agree that S has at most - say - 200 symbols (or in general N symbols). You can list all the possible finite "sentences" made with the symbols from S starting from length 1 (each individual symbol), then length 2 etc. So, all the set of finite sequences of symbols from S is countable. Some of these sequences will not make much sense (e.g. "2+"). Some others will represent a number (for example "pi" will be pi, "square root of 2" will be the square root of 2 and so on). Well, the numbers represented by these sequences will constitute a countable set, thus a negligible part of all the real numbers.
Notice that the same applies if instead of a finite set S we use a countably infinite set S. The resulting set of sequences is the union of S, SxS, SxSxS (Cartesian Products) etc. which is a countable infinity of countably infinite sets, thus it is countable as well and as a consequence it covers 0% of all real numbers!
I bet that this will beat the 0.999... = 1 incredulity...
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Shalom Mathologer.
I have said this several times before, and it is very interesting to hear what is your opinion about it.
There is a small problem with the GOOGOL number
And I will explain:
Only numbers with an amount of zeros that is multiple of 3 can have a "name".
The number 1,000 has a name. Thousand.
The number 1,000,000 has a name. one million. now, this million, with one more zero will be called ten million. 10,000,000 (7 zeros)
Another zero and it will be one hundred million. 100,000,000 (11 zeros)
Only after we add another zero (12 zeros in total) will we get a number with a new name. Billion.
Therefore the number GOOGOL, 1 with 100 zeros, can only be ten "something" ...
Assuming there is a number like 1 with 999 zeros, suppose its name is Boomboom
So only if we multiply it by ten we get one and a thousand zeros.
And it will be called
Ten Boomboom.
GOOGOL can't have a name.
It can only be 10 "something"
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
This is all quite fascinating, as I implied in my earlier comment. As for the relation of circle (disk), to surface of sphere, to volume of sphere, I discovered a little gem in '65, just after graduating HS, which makes me wonder whether it has been known for longer.
It relates, using Pappus' Theorem, the capacity of an N-ball to the surface capacity of an (N+1)-sphere, which then gives the capacity of an (N+2)-ball using the
" surface=d(volume)/dr " relation [the (N+1)-sphere being the surface of the (N+2)-ball.]
Then, starting with the formulae for N=1 and N=2, the formulae for all greater N's can be obtained. And all of them fit neatly into the formula
V(r;N) = (πr²)^(N/2) / (N/2)!
where the factorial definition is extended, using the gamma function: x! == Γ(x+1); more formally, replace x! with ∏(x) == Γ(x+1).
If you think this could be a good theme for a Mathologer video, contact me for details; I don't have the expertise to make the wonderful animations you & your team make.
Fred
YT seems to prevent posts that contain emails, so if there's a way of contacting me through my channel...?
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Steinbach's ratios are extremely pretty, thanks for showcasing them! I'm going to have to look at the ratios α, β, and γ for the enneagon. I'm surprised you didn't mention that A, B and C, the Fibonacci-like recurrent functions are actually only a single function. In particular, taking their definition r^m s^n = A(m,n)s+B(m,n)r+C(m,n) and multiplying through by r and applying the relations for r and s give recurrences for A(m+1,n), B(m+1,n) and C(m+1,n), and similarly multiplying through by s gives recurrences for A(m,n+1), B(m,n+1), and C(m,n+1). And one obtains that C(m+1,n) = B(m,n) and C(m,n+1) = A(m,n). Therefore not only is C a shifted copy of A, but so is B, as B(m,n) = A(m+1,n-1).
So r^m s^n = A(m,n)s+A(m+1,n-1)r+A(m,n-1). That can make building the values up by recurrence slightly awkward, but you need to build them from a set of initial values anyway, so it's not any less practical. And this means you only really need a Binet type formula for A(m,n), and the formula for B(m,n) is just a re-indexing of that.
2
-
@Mathologer Most things work out similarly nicely for the enneagon, with one exception which is particularly curious. There's still a little I haven't quite figured out, so I'll explain what I do know. In the paper Steinbach uses α,β and γ for the ratios in increasing size, but I'll be using a, b and c. Applying Ptolemy's theorem we obtain the following 6 relations, after simplifying a few things:
a^2 = b+1
ab = a+c
ac = b+c
b^2 = b+c+1
bc = a+b+c
c^2 = a+b+c+1
Just as in the case with the pentagon and heptagon, we can use these relations to reduce any non-negative integer powers of a, b and c to an integer linear combination of 1, a, b and c. However, unlike the pentagon and heptagon, we don't get all integer powers in general. The relations alone can be used to obtain 1/a = a+b-c-1 and 1/c = c-b, however 1/b cannot be obtained from the relations alone, for reasons we shall see later.
We can then define maps P, Q, R, S: ZxNxZ->Z, where Z and N are the integers and the non-negative integers respectively, such that:
a^i b^j c^k = P(i,j,k)c+Q(i,j,k)b+R(i,j,k)a+S(i,j,k) for all i, k in Z and j in N.
Multiplying through by a, b and c gives forward recurrences for P,Q,R and R in the i, j and k indices, and dividing by a and c gives reverse recurrences for P, Q, R and S in the i and k indices. Like with the pentagon and heptagon, we see that Q, R and S are just shifted copies of P. In particular:
Q(i,j,k) = P(i,j+1,k-1), R(i,j,k) = P(i+1,j,k-1) and S(i,j,k) = P(i,j,k-1) for all i, k in Z, j in N.
I do wonder if it's perhaps better to write P, Q and R in terms of S, as then P, Q and R are positive shifted versions of S in a fairly symmetric way.
To understand why we don't get 1/b from the relations, consider that there are multiple solutions to the relations. Any triple (x,y,z) satisfying the relations of (a,b,c), must also have:
x^i y^j z^k = P(i,j,k)z+Q(i,j,k)y+R(i,j,k)x+S(i,j,k) for all i, k in Z and j in N.
What are these solutions? Well we can use the relations to show that if (x,y,z) is a solution, x, y and z must be roots of the polynomials x^4-x^3-3x^2+2x+1, y^4-3y^3+3y and z^4-2z^3-3z^2+z+1 respectively. Each has four real roots which match up to form 4 distinct solutions and can be expressed as (a,b,c), (1,0,-1), (-c/a,b/a,-1/a) and (-1/c,-b/c,a/c).
We can now see the problem with deriving 1/y from the relations; the solution (1,0,-1). If 1/y could be derived from the relations, we'd be able to express negative powers of y as a linear combination of 1, x, y and z. But negative powers of y in the solution (1,0,-1) correspond to dividing by 0. We shall see this isn't all to the story though.
Now that we have our solutions to the relations, we can form a matrix equation A = MB, where:
A = {a^i b^j c^k, (-1)^k δ_j, (-c/a)^i (b/a)^j (-1/a)^k, (-1/c)^i (-b/c)^j (a/c)^k}
B = {P(i,j,k), Q(i,j,k), R(i,j,k), S(i,j,k)}
M = {{c, b, a, 1},
{-1, 0, 1, 1},
{-1/a, b/a, -c/a, 1},
{a/c, -b/c, -1/c, 1}}
Note that 0^j for j in N is just δ_j the Kronecker Delta at j, where we're using the combinatorial convention that 0^0 = 1.
We can then invert M and obtain B = M^(-1)A, where:
M^(-1) = 1/9 {{c-b+1, -3, -a+1, b},
{-c+2b-1, 0, c-b+a, -a-b+1},
{a-1, 3, -b, -c+b-1},
{-b+3, 3, -a+b+1, c+1}}
So we do indeed get Binet-like formulae for P,Q,R and S, for i, k in Z and j in N. But the story doesn't quite end there. What would happen if you try to use these formulae for negative values of j? Do we just get nonsense? We no longer get integer outputs for P, Q, R and S, we get 3-adic rationals in general, but they do indeed allow you to express negative powers of y as a linear combination of 1, x, y and z for three of the four solutions (x,y,z). And the formula actually works correctly! I don't currently have a complete explanation as to why, but I did observe one thing.
If y =/= 0, we can obtain a new relation z = x+1. We see this relation is indeed satisfied by three of the four solutions, but is not satisfied by the solution (1,0,-1), so perhaps this additional relation is enough to show that the Binet formulae indeed works for three of the four solutions for negative values of j.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Most memorable part for me is the grader giving your proof a zero. I got a few A's in math at IU generating "original" proofs for homework or on tests. Not that I ever found a proof that wasn't already known somewhere else, but just not the one given in class. At least the professors or students grading my work went to the trouble of checking if my proofs were valid. It reminded me of my method of getting A's in all my math classes in College: In the summer before the next school year, I simply spent my time studying all the old textbooks in the school's Math library that covered my upcoming classes for the Fall and Spring semesters. I mean, they don't make any secret about what the upcoming classes will cover, and, for undergrad classes, most of the Math is not exactly new, cutting-edge stuff. And as a full-fledged Math nerd, I didn't have anything else to do all summer that was half as fun. So, by the time I took the classes, I already knew the material through and through. (I also discovered that some lazy professor's hand-out tests contained problems that were culled from other older textbook's worked examples. Tsk, tsk.) Strangely enough, in all the hours I spent in the Math library getting ahead of my studies, I never met even one other Math Geek doing the same preparatory research.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
You can paint an infinitely long fence as far as you like with just one (finite) tin of paint. Assume the fence consists of an infinite number of identical finite sections. Use half the tin to paint the first section, then half of the remaining paint to paint the second section, then half of what remains to paint the third section, and so on for as many sections as you like. The thickness of paint on any section will be half the thickness of that on the previous section, so every section you painted will be covered in paint, yet there will always be some paint left in the tin at the end.
It is similar to painting the inside of Gabriel's horn by introducing paint into it. The thickness of the paint at any point on the inner surface is equal to the radius of the horn at that point, being ⅟ₓ at point x on the axis. Starting with a pot of π volume units of paint, fill the section from x=1 to x=2. This will use half the pot, the paint thickness varying from 1 to ½. Then fill the section from x=2 to x=4. This will take half of the remaining paint at a thickness from ½ to ¼. Continue with the next section from x=4 to x=8, and so on. Each section is twice as long as the previous one, but only requires half the quantity of paint, so you can fill/paint as far as you like but there will always be some paint left in the pot. Since the paint thickness varies inversely with length from the "mouth" there will always be some paint on the inner surface of the horn that you have painted so far, and you can carry on "painting" the horn like this indefinitely from the same pot of paint.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Another grand video brimming with Mathologer's characteristic good cheer and containing several playful challenges for the audience. The content is quite accessible, despite coming from relatively recent mathematical literature. "This is not the Discovery Channel" was my favorite quote. Kudos to Bjarne Fich for rising to Burkard's challenge and creating such a professional Arctic Circle animation.
Here is my response to Mathologer's request for feedback.
Challenge 1: No, removing 2 black & 2 green squares will not always yield a tile-able board. For example, removing (2,1) and (1,2) leaves (1,1) isolated.
Challenge 2: If m & n are odd, then ( j, k ) = ( (m+1)/2, (n+1)/2 ) yields a zero term in the product.
Challenge 3: 2xn squares yield (1,2,3,5,8,...) tilings for n=(1,2,3,4,5). This sequence follows the Fibonacci rule. To see why, observe that T(n+1), the number of tilings for n+1, can be computed by adding the number of tilings with the last domino vertical, which is T(n), and the number of tilings with the last two dominoes horizontal, which is T(n-1). Aha!
Challenge 4: The determinant for Tristan's glasses gives 666. By drawing the 4 ways of tiling around the holes I get the same result, but have yet to see why the determinant formula should work when the holes themselves can be tiled.
Challenge 5: Previous comments allowed me to see why a hexagon tiling must be split evenly into the 3 tile orientations. Unfortunately, I did not see this for myself.
Most of the video was clear. However, I am mystified as to why the dance yields all possible tilings, and why, for example, a pair of adjacent 2x2 squares don't count as a 2x4 rectangle. Mathologer dropped some clues, but I guess I need to consult the references.
Looking forward to more great content in 2021!
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
When I was a child, my dad bought a bunch of stuff from an auction, and among it was a large-diameter (maybe three or four inches; I don't know my artillery history) shell casing, such as from an artillery gun. But not an intact shell casing, quite -- nope, this one was half crushed, into precisely the type of lantern pattern you show here, but crushed vertically, parallel to the straight sides of the cylinder, to the very extreme limit of the strength of the material. It was made of. Steel? Bronze? Brass? I don't know my artillery composition, either.
Someone once expressed the opinion that this casing had been crushed by getting caught in the ejection mechanism when fired, which I suppose is possible (though I don't know my artillery mechanisms), but It seems to me that the crush pattern was much more uniform than I would think an accident would produce. Surely wouldn't an accidental crushing produce some bit of bias, a tilt to one side or some other non-uniformity in the bucklings? Or does the math prohibit that and force an all-or-nothing symmetry? So someone else suggested it was a deliberate artwork, such as you have begun to produce in your final footage of that soft drink can which you score with the piece of wood. (Thank you for that, incidentally, I've been looking for a start on that technique most of my life. Score the can as you have done, then press it from above, and you can make quite an attractive little vase for some pretty flowers, exactly the way my mother used that shell casing in my youth.)
I'm also surprised you can't compensate for the buckling inward of the triangles by simply calculating and applying a correction factor in determining their contributions to the surface area of the cylinder. You know the normal vector along which the surface area must point, and you should be able to calculate the normal vectors to each triangle in the lantern pattern, and from the supply a (sine theta?) correction factor to each off-normal triangle's surface area contribution. Is it simply that that technique was outside the scope of this video? I find it hard to believe you wouldn't at least mention it in passing. Or is there some other mathematical dragon that arises to defeat the valiant knight who attempts this correction?
One could even envision cutting the surface of the cylinder as though to unwrap it, then examining the zigzag/seesaw pattern formed by the cross-sections of the lantern pattern at each point in its back and forth cycle, then figure out some way to correct for that, as a whole, and apply that correction to the summation over all the triangles. Now, damn it, either you need to make a video discussing this, or I'm going to have to try it myself.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
When you have your list at 13:43, instead of throwing away every other combination, a solution that uses all of the pairs: house number is top times bottom. To get the last house number: min of [(top^2), (bottom^2)*2], which will be two numbers in the form of (x) * (x+1), with x being the last house number. So, for 7|5 - 7*5 is the house number 35. min(7*7,5*5*2) = min(49,50), so x is 49.
For the answer to the puzzle, 17*12 = 204. min(17*17=289, 12*12*2=288) = 288 for the number of the last house.
2
-
2
-
2
-
2
-
2
-
3:00
Bottom row: a_0 = 1, a_n = 1
4th row: b_0 = 1, b_n = b_0 + \sum_(k=0)^(n-1) a_k = 1 + n
3rd row: c_0 = 1, c_n = c_0 + \sum_(k=0)^(n-1) b_k = 1 + n + 1/2 n(n-1) = 1 + 1/2 n + 1/2 n^2
2nd row: d_0 = 1, d_n = d_0 + \sum_(k=0)^(n-1) c_k = 1 + n + 1/4 n(n-1) + 1/12 (n-1)n(2n-1) = 1 + 5/6 n + 1/6 n^3
Top row: e_0 = 1, e_n = e_0 + \sum_(k=0)^(n-1) d_k = 1 + n + 5/12 n(n-1) + 1/24 (n-1)^2 (n^2) = 1 + 7/12 n + 11/24 n^2 - 1/12 n^3 + 1/24 n^4
We already see a few powers of 2 when n = 0, 1, 2, 3, 4, 9. But I'm not sure if there are others.
May I ask which pattern are you expecting us to find out? Thank you!
P.S: After watching the video, I realized I was doing quasi-sequence integral using {\sum x^n} formulas. Now I know G-N formula is the way!
26:04
^(_m) denotes " to the falling power m".
x^(_m) - (x-1)^(_m)
= 1/(m!) [x(x-1)...(x-m+1) - (x-1)(x-2)...(x-m)]
= 1/(m!) (x-1)(x-2)...(x-m+1) [x - (x-m)]
= 1/[(m-1)!] (x-1)(x-2)...(x-m+1) = (x-1)^(_m)
36:30
Isn't this true given the definition of "difference"? :P
44:15
A quote from Mathologer:"The infinite sum is just a finite sum in disguise." :)
Great video as always! Among all these aha moments, it's the rotation into Pascal triangle that impressed me most.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
24:13 proof:
let us form a sequence of integers that are used to generate the sequence 1, 2, 5, 7, 12, ..... (let this be sequence 1) using the formula f(n) = n(3n-1)/2 in same specific order.
which is: 1, -1, 2, -2, 3, -3, ....n, -n, (n+1), -(n+1), .....(let this be sequence 2).
now take a note of the n, -n, (n+1), -(n-1) (let this be genSeq1) part cause that gives us the generalization of the sequence we need for every 'n'. we can deduct the result for the difference between 2 consecutive numbers from sequence 1 by breaking the problem into 2 cases.
case 1: difference between 2 numbers from sequence 1, where those 2 numbers are obtained by a +ve number followed by a -ve number respectively. now, the difference can be expressed as f(-n) - f(n) (n and -n are taken according to genSeq1) .
which gives us: diff(+ve followed by - ve) = (-n(3(-n) - 1)/2 - n(3n-1)/2)
=> diff(+ve followed by - ve) = ((3n^2 +n) /2 - (3n^2-n)/2)
=> diff(+ve followed by - ve) = (3n^2 + n - 3n^2 + n))/2
=> diff(+ve followed by - ve) = n
which means that the difference between the two numbers that are obtained by plugging +n and -n into f(n) is 'n' itself.
case 2: difference between 2 numbers from sequence 1, where those 2 numbers are obtained by a -ve number followed by a +ve number respectively. now, the difference can be expressed as f(n+1) - f(-n) (-n and n+1 are taken according to genSeq1) .
which gives us: diff(-ve followed by +ve) = ((n+1)(3(n+1) - 1)/2 - (-n)(3(-n)-1)/2)
=> diff(-ve followed by +ve) = ((3(n+1)^2 - (n+1)) /2 - (3n^2+n)/2)
=> diff(-ve followed by +ve) = (3n^2 + 6n + 3 -n -1 - 3n^2 - n))/2
=> diff(-ve followed by +ve) = (4n +2)/2
=> diff(-ve followed by +ve) = 2n + 1
which means that the difference between the two numbers that are obtained by plugging - n and (n+1) into f(n) is '2n +1'.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
11:51 - (right, right, right, left)
going right the pattern is (?,1,?,3), (?,2,?,5), (?,3,?,7), next we get our (?,4,?,9) and go left to take the 4 and 9 as our seed to create (9, 4, 13, 17)
2
-
2
-
There's actually a nice formula for the partial sums that can be derived as follows:
We want to find 1/1 + 1/2 + 1/3 + ... + 1/n. Replacing the numerators with x^(denominator) gives us the function
g(x) = x/1 + (x^2)/2 + (x^3)/3 + ... + (x^n)/n.
So g(1) is our desired value and g(0)=0. Then, we may take the derivative of this polynomial giving
g'(x) = 1 + x + x^2 + ... + x^(n-1).
We're going to exploit that this has a nice closed form expression. To see this, multiply each side by (x-1):
(x - 1)g'(x) = (x - 1)(1 + x + x^2 + ... + x^n) = (x + x^2 + x^3 + ... + x^n) - (1 + x + x^2 + ... + x^(n-1)) = x^n - 1,
so g'(x) = (x^n - 1)/(x - 1). Now, by the fundamental theorem of calculus, we may integrate this function to get our result:
∫[0,1] (x^n - 1)/(x - 1) dx = ∫[0,1] g'(x) dx = g(1) - g(0) = g(1).
Thus this integral is our final formula. I think Euler was the first one to find this, though I only know it because I happened to rediscover it in high school. This method can also be extended to the partial sums of this series for higher powers. For instance,
1/1 + 1/4 + 1/9 + ... + 1/n^2 = ∫[0,1] (1/y) * ∫[0,y] (x^n - 1)/(x - 1) dxdy.
The interesting thing about these higher power examples is that they actually converge, and that the integral forms reflect this:
1/1 + 1/4 + 1/9 + ... = ∫[0,1] (1/y) * ∫[0,y] -1/(x - 1) dxdy = ∫[0,1] (1/y) * (-ln(1-y)) dy = pi^2/6.
Obviously proving that last integral equality isn't trivial, but I think this comment is long enough so I'll leave it there.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
p(666)=11,956,824,258,286,445,517,629,485
Python solution: (Runtime is a few ms)
import itertools
_p = [1, 1]
def p(n: int) -> int:
if n < len(_p):
return _p[n] # answer is known; return it (avoids recalculation; otherwise this'll take forever)
value = sum(
(
sign * p(_n)
for sign, _n in reversed( # reversed to avoid recursion limit
list(
zip(
itertools.cycle([1, 1, -1, -1]),
itertools.takewhile(
lambda e: 0 <= e,
map(
lambda e: n - e,
itertools.accumulate((e for _e in zip(range(1, n, 2), range(1, n)) for e in _e))
)
)
)
)
)
)
)
_p.append(value) # save value for reuse later
return value
print(p(666))
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Every odd number can be written as a difference of two squares in at least one way.
Trivial proof:
n² equals the sum of the first n odd numbers 1, 3, 5, 7, …, (2n-1), which is easily shown by induction and very easy to visualize too. Just try squared paper and see. 😎
So it is easy to see 2n-1 equals n² - (n-1)².
For any odd prime number p = 2n-1 this is the only way to equal a difference of two squares.
Proof by contradiction:
Let's assume p is a difference of two squares n² - (n-m)² with m > 1 and m < n
Then
p = n² - (n-m)²
p = (n + (n - m)) * (n - (n - m))
p = (2n - m) * m
p is prime and p | m means m = p or m = 1, and the last one contradicting m > 1.
But if m = p, then 2n - m = 1 and therefore m = 2n - 1, contradicting m < n.
So there is exactly one way to express any odd prime number p as a difference of two squares.
This was so much easier than Fermat's theorem. Even possible to me. 😉
(Excuse my English, but it is better than most people's German.)
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
For the programming challenge for dollar amounts, this is how I might try to calculate it: if a_n is the number of ways to make change for a value n coin using 1, 5, 10, 25, 50, and 100 value coins, I have the recurrence a_n = a_(n-1) + a_(n-5) + a_(n-10) + a_(n-25) + a_(n-50) + a_(n-100). (a_k for k < 0 is 0 and 1 for k=0). You could then just use the recurrence relation to calculate the first however many terms in the sequence. If you wanted a closed formula for a_n, Because each a_n depends linearly on the last 100, you can find a (sparse) 100x100 matrix of 0s and 1s that when applied to the vector [a_n, a_(n-1), a_(n-2),..., a_(n-99)] results in the vector [a_(n+1), a_n,...,a_(n-98)]. Then you could convert this to Jordan canonical form and find the change of basis matrix. In principle, you could figure out the formula for raising each of the Jordan block matrices to the n-th power, then use this to find a closed form expression for a_n. I have a feeling that the latter method might be very inefficient, and susceptible to round off error or something like that.
2
-
2
-
2
-
2
-
@Mathologer Bob Newhart was an American comedian, now passed on, who got his start as a stand-up act, ca 1960, with the release that year of two hit LPs, The Button-Down Mind of Bob Newhart, and The Button-Down Mind Strikes Back, then got starring roles in 2 or 3 TV sitcoms in the following decades. His brand of humor was what could fairly be called "deadpan."
In his penultimate sitcom he played a psychiatrist (in Seattle?), and Suzanne Pleshette co-starred as his wife.
In his final sitcom, titled simply, Newhart, he and his wife Emily, played by Mary Frann, ran a small hotel (the Stratford Inn) in Vermont, and it co-starred Tom Poston, among others. The final episode got a lot of buzz with its closing scene which started in darkness, then a light switched on showing Bob and Suzanne Pleshette in their roles in the earlier sitcom, with Bob describing his "strange dream" in which he ran a hotel in Vermont. Suzanne tells him to go back to sleep, the light is quenched, and the credits roll.
IMDB will undoubtedly have more info on him. Wikipedia also has a good bio. He was quite famous in the US.
2
-
His proof didn't rely on the 10 numbers at all (besides that they were between 10 and 99, inclusive), so it works for any collection of 10 numbers in the range.
The way this works is that he way overcounts the "possible sums". This is essentially like overcounting the number of pigeonholes. This is not a problem, though! Even overcounting the number of pigeonholes, we still have fewer pigeonholes than pigeons, so we know that at least one of the pigeonholes has to have at least two pigeons in it.
How does this work? He considers the lowest possible sum you could have. If your set of 10 numbers includes the number 10, then the lowest possible sum you could have is 10 itself. If your set of 10 numbers does not include the number 10, then the smallest possible sum will be larger than 10, since 10 is the smallest number possible. So 10 is the lowest possible sum you could have. With most sets of 10 numbers, the true smallest sum will be larger than 10.
Similarly, he considers the largest possible sum you could have. If your set of 10 numbers includes the numbers 91-99, then the largest possible sum you could have is 91+92+...+99 = 855. If your set of 10 numbers does not include every number from 91 to 99, inclusive, then the largest possible sum you could have will be smaller than 855, since 91-99 are the 9 largest numbers possible. With most sets of 10 numbers, the true largest sum will be smaller than 855.
So, we know that, no matter what 10 numbers you get, there is no way to have a sum smaller than 10 and no way to have a sum larger than 855. So, no matter what 10 numbers you get, any possible sum is between 10 and 855, inclusive. So there are 846 possible values of a sum we are considering here. With a fixed set of 10 numbers, there are way fewer options, but here, these 846 options account for any possible set of 10 numbers you could have.
Now, when you have 10 different numbers, there are exactly 2^10 = 1024 subsets of these 10 different numbers, including having none of them and having all of them. Since we only want to consider nonempty (and hence also non-full) subsets, there are 1022 possible subsets of a fixed collection of 10 different numbers. Counting the number of nonempty non-full subsets does not at all depend on what the numbers are - just how many of them there are. So his count is accurate.
So, no matter what 10 numbers you get, there are only 1022 possible subsets of them. Also, no matter what 10 numbers you get, there are only 846 possible values you could have as a sum. So at least one of those sum values must have two subsets corresponding to it. There are, therefore, at least two subsets with the same sum.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Great to see!
A few avenues not explored:
The four triangles and their trigonometry. Including THE heptagonal triangle (with sides 1:r:s or angles pi/7, 2pi/7, 4pi/7)
r satisfies r^3-r^2-2r+1=0, and s is a root of the reciprocal polynomial.
Recurrence relations, similar to those relating to the Tribonacci constant.
Rational approximations to r (and s), starting with 1,2,9/5,182/101 (and 2,9/4,182/101)
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
Favorite part: Chapter 5, and the proof that the total excess over the logarithm is less than 1.
That visual proof practically hands us the answer to why gamma is greater than 0.5: Each of the blue regions is larger than a triangle. It has the area of a triangle plus an extra belly. Perfect triangles would cover half the 1x1 area; area=base*height/2.
I found the odd/even thing in chapter 4 perhaps less surprising. The thing is, all your denominators are going as some factor of n!. It's as if there is some formula where the "raw" denominator is exactly n! while the "raw" numerator has no direct factorial pattern. Thus putting the fraction in lowest terms cancels a few prime factors between the raw numerator and denominator. But since the raw denominators are n!, they have plenty of factors of 2, always more than the few 2's in the relatively rough raw numerators. Thus always odd/even.
The same pattern emerges more weakly when you look at it mod 3: After a while (everything past 63/40, as far as I looked) all the numerators are 0 mod 3, for similar reasons. For instance, for lurkers:
7433/3960, 673/360, 8719/4680, 63373/32760, 65557/32760
This also tends to explain why the series misses all the integers, or at least make it less surprising: there are too many factors in the denominator to expect them to all be cancelled by chance, since no simple factorial pattern arises in the numerator.
2
-
2
-
2
-
2
-
Homework from 11:56
Part 1. For the triangle given (153, 104, 185). The four numbers A B
C D are as follows.
A=9, B=4, C=17, D=19.
To answer the bonus question. The sequence from triangle (3,4,5) to triangle (153,104,185) is (right, right, right, left).
2
-
2
-
2
-
If you play with the first 20 integers for √a+√b=√c you get that a+b can be 32.00%, 37.50%, 44.44%*, 48.00%, 48.98%* or 50.00% of c but never more. The best, ie 50% is where a=b, so a+b=½c.
[Within the arbitrary limit I set] The squares 1, 4, 9 & 16 have the most integer solutions, 4 each, and the odd primes have the fewest, one only where a=b≡p because the square root of any prime is irrational.
* not exactly, the others are precise.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
@MonstraG55 So, I once heard someone describe common ways to solve a hard mathematics problem. Sometimes we can add structure to a problem in order to make it easier. For example, if you had an 8x8 grid and you removed two opposite corners and wanted to tile the remaining grid with 2x1 or 1x2 pieces, could you do this? One way to solve the problem is to add structure to it - imagine the 8x8 grid is a chess board, where squares alternate colors. Then the two opposite corners are the same color, so it's not possible because 1x2 or 2x1 pieces always take up 1 of each color (this is essentially the argument used at the beginning of this Mathologer video: watch?v=Yy7Q8IWNfHM). So adding structure (color) to the grid helped us solve the problem.
But sometimes removing structure from a problem can help us solve the problem. The example I heard for this possible way of solving a problem is to imagine you have a log which is 1 meter long. And you drop a bunch of ants onto this log. The ants are all on a single line on this log (lined up with the log, so 1 meter long) and they always move along this line. All ants move at a constant speed of 1 meter per minute. Suppose that if two ants run into each other (going opposite directions), then both ants turn around and move in the opposite direction. What is the maximum amount of time it will take until all of the ants walk off the log? It's a hard problem, but there's a nice insight. If two ants are walking toward each other, collide, and both turn around, this is the same situation as if the ants are able to pass through each other. (The ant moving left switches roles with the ant that was originally moving right but turned around to move left, and vice versa.) So we can remove this whole requirement about ants turning around when they collide and can solve a new problem where ants pass through each other. Then the answer becomes easy. The longest it can possibly take is 1 minute, since that's the longest it can take an ant to get from one end of the log to the other.
That's what's happening in this scenario with chapter 6 in this video. We can remove structure from the problem. We want to find two non-overlapping sets with the same sum. That seems hard to do, so let's remove part of the problem. Remove the requirement that they're non-overlapping. If we are able to find any two different sets that have the same sum, then we can just remove their overlap (removing the same amount preserves the equalities of the sum). So really, all we have to do is fine two different subsets of the same sum. We don't have to do everything at once. We don't have to use the pigeonhole principle to guarantee that there are two non-overlapping sets with the same sum - we just have to use the pigeonhole principle to guarantee two sets with the same sum. The existence of two sets with the same sum guarantees the existence of two non-overlapping sets with the same sum. So we're done once we put these two facts together.
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
2
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
First thing I would think of is the number fields Q[r, s] (and Q[phi]). Q[phi] is the splitting field of x^2-x-1, in particular Q[phi] is isomorphic to Q[x]/(x^2-x-1). Q[r, s] is isomorphic to Q[x, y]/(xy-x-y, x^2-y-1, y^2-x-y-1).
We can look at the ring of algebraic integers inside these number fields. For Q[phi], that would be Z[phi]. If I recall correctly, this is a unique factorisation domain. Perhaps it would be interesting to check if the ring of algebraic integers in Q[r, s] is a unique factorisation domain (else we may consider computing class numbers, it might be interesting too).
Since we get them from polygons, I think these lie in cyclotomic extensions, so Q[phi] is inside Q[z] where z^5=1. It should be clear that Q[r, s] is inside Q[z] where z^7=1, and for the infinite family of polygons I would expect something similar. (Proof: identify the vertices of the polygon with the complex numbers 1 to z^6. Then r^2, s^2 are clearly |1-z^2|^2/|1-z|^2=|1+z|^2 and |1-z^3|^2/|1-z|^2=|1+z+z^2|^2 respectively. Since s^2=s+r+1 and r^2=s+1, we can solve for r and s from there.) I believe it should be clear that [Q[r, s]:Q]=3 since all powers of r and s can be written as Ar + Bs + C.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Thanks much for another wonderful, excellent video. Equality and inequality are powerful subjects, and can have quantitative (applied math?) as well as qualitative (pure math?) aspects. For an arbitrary set of 4 planar vertexes, choose any subset of 3 which then specify a circle. The 4th vertex is either on the circle, if Ptolemy's formula gives zero, or non-cyclic if non-zero. My conjecture (wish I had time to work on it) is that the non-zero result gives the "degree of non-cyclicness" and maybe inside/outside measure of distance to the circle.
This is related to what may be the simplest geometric equality, the standard form of the line equation. I think that early students don't get the more exciting introduction; the focus tends to be on the slope-intercept form and graphing of functions, helpful but rather boring. Strangely, they short-shrift the general form, and even get students to derive it from the slope-intercept form! But, the general form is much more useful, and is derived more simply by a cross-product type formula of the endpoint coordinates. The resulting Ax + By + C = 0 is then a formula -- if you plug in a point it gives zero if it's on the line, but if >0 then the point is on the right side of the line, on the left if <0, and the magnitude of the result gives the distance from the line, if "normalized" by the length of the line segment. This is extremely useful but students don't hear about it.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Circular sliderule. Its only magical to the kids, who have not grown up with sliderules, and those of us, who had to use it. Works with the more common linear ones, too. I still have a couple.
Fun Fact: In electronics, this logarithmic relationship is used to establish component values. If you have ever wondered why resistor & capacitor values are multiples of: 1.0, 1.5, 2.2, 3.3, 4.7, 6.8, and so on. Its that the successive value is a fixed multiple, of the prior one, within a given tolerance.
1
-
1
-
I have a possible improvement to suggest for the card trick. It would be somewhat more impressive, would it not, to have the 'mark' not only randomly select the 5 cards, but then to select one of them, when they are spread face down on the table, as the 'selected' card for your assistant to guess. You are then free to arrange the remaining 4 cards to hand to your assistant, passing info using just their ordering in the same manner as you've described. The twist is that you are going form a number from 1 to 48 using the card ordering to point to the 1 of 48 position of the 'selected' card. I say 1 of 48, not 1 of 52 because obviously the selected card cannot be one of the 4 that are passed. Now this calls for not a strict mapping of cards to numbers, but one that maps a bit dynamically to the 48 available positions, taking the 4 unavailable cards into account. So a bit of mental gymnastics for you and your assistant to perform. But, I hear you object, there are only 4! = 24 ways to arrange the 4 cards, so insufficient pigeons here. Ah ha, but you can pass them to either the assistant's left or right hand, yes, to make 48 pigeons!
1
-
1
-
Regarding the box at 11:50, the answer is [9,4,13,17] and it's a grandchild of [1,3,4,7], via [1,4,5,9].
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
@7:45 I would propose to adjust the connected dots slightly to emphasize the proof.
For the first chain of each new shell (2, 5, 8 etc), it should go from the lower right corner of the hexagon, via the lower middle and the lower left corners, ending at the upper left corner, but excluding the start and end corners. For the second chain (4, 7, 10 etc), it should also go from the bottom right corner of the hexagon, but instead via the upper right, the upper middle and end at the upper left corner, and this time including the start and end corners.
This is slightly different from the visualisation in the video, where the chains don't always start and end exactly like this pattern.
By following this pattern, it is obvious that each chain will extend it's length by three for each new shell, since it always runs the entire length of exactly three sides (excluding corners, which are always the same for each shell - two for the first chain and four for the second chain) and each side extends by one for each new shell. The extension by three is of course directly related to the skipping of each third number in the number line.
The chains in the video does not give this visual proof of why each new shell is a combination of the two new numbers, but with my proposed change, this will be more obvious.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
About all the details that were skipped over in the visual sum = ln(2) proof:
First of all, to show that the sum up to the n-th positive number is equal to the sum of rectangle n up to rectangle 2n-1, a simple inductive argument:
The sum up to the first positive number is just 1 (so rectangle 1 up to 1).
Then, to get from the n-th positive term to the (n+1)-th positive term, you're adding (1/(2n) - 1/n) + 1/(2n+1).
Subtracting -1/n removes rectangle n, so the sum now starts at rectangle n+1, and adding 1/(2n)+1/(2n+1) adds rectangles 2n and 2n+1, so the sum now goes up to 2(n+1)-1.
Now let's say the left-most edge of a given rectangle starts at x. It will therefore have a height of y=1/x. If you squish the rectangle by a factor of n and stretch it horizontally by that same factor, and instead of chosing the corner of the rectangle as the fix-point of the scaling, you choose the origin x=0, then the rectangle will move to x'=x/n, and the new height will be y'=n y. With this, the new rectangle also follows the law y' = n 1/x = 1/(x/n) = 1/x'. This transformation will also move all the rectangles to horizontally fill the interval [1,2], as the first rectangle with x=n will be moved to x'=n/n=1, and the last rectangle with x=2n-1 will be moved to x'=(2n-1)/n=2-1/n, so its right-most edge will be at 2-1/n+1/n=2 because its original width of 1 was squished to 1/n.
Therefore, the new rectangles fill the area under the curve 1/x on the interval from 1 to 2, and because all we did was move them around a bit, and squish them by the same factor as we stretched them, the total area didn't change.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
In the case of the four colour problem:
Consider the grid as a set of nodes with coordinates A1 through B2 for the 2x2 case, A1 through D8 for the 4x4 case, et cetera.
Choose an arbitrary colour for A1, and a different one for A2
Once A1 and A2 have colours, B1 and B2 must have the two other colours to satisfy the criterium for the A1 through B2 square. Since no distinction has been made between these two colours yet, choose one arbitrarily for B1 and the other for B2.
Once B1 and B2 have their colours, C1 and C2 must have the other two colours to satisfy the criterium for the B1 through C2 square. This means that C1 and C2 are now confined to the same colours as A1 and A2, though not necessarily in that order.
This pattern of altering pairs continues, with every subsequent pair of squares in a column along the 1- and 2- rows being confined to one of two pairs: the same two colours as A1 and A2 for "odd" letters, and the same two colours as B1 and B2 for "even" letters.
Equivalently, once A1 and B1 have their colours, A2 and B2 must have the other two colours to satisfy the criterium for the A1 through B2 square. This logic continues equivalently along the rows exactly equivalent to the case along the rows, and thus ever subsequent pair of squares in a row along the A- and B- columns is confined to one of two pairs: the same two colours as A1 and B1 for odd numbers, and the same two colours as A2 and B2 for even numbers.
Therefore, for any given corner alpha in an x by x grid where x is an even number, the corner squares beta and gamma which it shares sides of the grid with must be differently coloured to it, as beta and gamma will each differ from alpha in the evenness of either the row number or of the column number. Thus if we can prove that alpha also differs in colour from its opposite corner delta, then it must differ from all other corners; and if it is true for any given corner that it differs in colour from all other corners, then each corner is uniquely coloured.
That part of the proof, however, keeps eluding my 11pm mind and should probably wait until morning
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Einst soll ein französischer Philosoph gemeint haben "am Grunde eines jeden Problems sitzt ein Deutscher", doch mit Burkart Polster "haben wir" anscheinend eine art anti-shapeshifter, der auch die Lösung eines jeden (echten) Problems von einem Deutschen auf den Tisch, die Strasse oder einfach nur zu den Interessierten trägt oder bringt. Oder wäre Lob hier zuviel des Guten, das er sich nicht auf Lorbeeren ausruhte, so er nicht ein Rom-o-philer sein wird? Solch "mathematisierte Unterhaltung" ist stichhaltig, nicht nur weil man das so nicht mit "infotainment" vergleichen kann, meinte ich.
Euclid wurde vor einigen Tagen am Lagrange Punkt L2 , weitgehend hinter dem Mond "parkiert", doch die "probe" teilt sich den "Punkt" mit dem JWST und GAIA, wie "kriegen" die das hin, in Dynamik mit geometrischen und gravitativen Aspekten?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
When you asked what they'd have in common, I didn't guess, I thought: (i) each of these three extended lines has the same length; (ii) each pair, crossing at a vertex of our triangle, is split in the same lengths, so has the same products of its parts' lengths, matching the intersecting chords theorem; each pair of the lines thus puts four point on a circle; so I concluded you were about to tell us it was a circle. That isn't quite a proof, though, and I can't immediately turn it into one, but it smells like it's relevant: we have three cyclic quadrilaterals (albeit we don't know they share a single circle), each of which shares each of its diagonals with one of the others; and those three diagonals are all of the same length. So Ptolemy's theorem's product of diagonals is the same for all three, implying their sums of products of opposite edges are the same. But after scratching my head for a bit about that surely being a clue, I think I'll watch your video and find out what JHC came up with ;^>
1
-
OK, so you show the circle shrink and it looks suspiciously like its centre is also the incentre, where the angle bisectors meet. And, of course, each of our three extended lines is, by construction, the mirror image of the one it meets at a corner of the triangle in the bisector of that corner's angle. So it suffices to show one of these has its two ends equidistant from the incentre, so that there is a circle centred on it through those two ends and reflections in lines through the incentre will ensure the other two lines also have ends on the same circle. Let the length of each edge be denoted r, g, b by the first letter of its colour, with the semi-perimeter s being half their sum; each of our extended lines has length 2.s. The perpendicular from the incentre to (say) the blue edge splits it into an s-r on the end we extend by r and s-g on the end we extend by g so this is the mid-point of the extended edge and a radius is its perpendicular bisector; we are done ;^>
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Peg number formula: if we relabel pegs (A,B,C) as (0,1,2), then for peg, p; move m, and disk, d (where the smallest disk, d=1), then I get that the peg number is
p=2^(d-1)*floor((m+2^(d-1))/2^d) mod 3, for any number of disks. At move zero, all disks start on peg zero.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Full comment has challenge answers (attempts really) down below
OOOH I UNDERSTAND THE DETERMINANT TRICK
The determinant can be described as taking the product of all terms which don't share a row or a column.
If you decide that one particular square is paired to another square (oppositely colored) using a domino, then you can't tile it again, which by the determinant process, disallows you to pair that square/row with another term in your product.
Two i's represent two vertical dominoes. Twisting them, involves turning them into horizontal dominoes (two 1s). The product turns from -1 to 1, and the minus sign is then compensated for by the change in the sign of the permutation.
Now, to understanding why twisting gives all possible tilings 🤔.
1) (removal of 2 black + 2 green squares) Not generally, because the two black squares connected to a corner green square can be removed, isolating it and making it un-tile-able
2) (how the crazy formula spits out a 0 for oddXodd boards) the denominators of the cos input are (n+1) and (m+1), so for odd m and n, they are even. Since the products upper bounds round up, inputting the upper bounds gives the input of both the cos functions as (π/2), making them 0, and making the entire product 0.
(The last puzzle of why the number of blue, orange and grey tilings) Rotational symmetry!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I don't unbderstand the part that starts around 20:00. It shows that the isolated solution, that is the cross solution, requires X to be equal to 1. But in the case of 13, the solution is actually 2 squared and 3 squared. So for 13, x is NOT equal to 1. So when factored as p = x(x + 4z), x is not one, so... p is not prime? So, I don't see how crosses, that are the unique solutions, which should always have x = 1 can correspond to the 13 case.
Maybe there is soemthing else going on. If so, then choosing a prime other than 37, that is a prime whose decomposition into two squares does not contains 1 squared, would have been a better illustration.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Wonderful video as always!
23:38 Challenge 1: Because all the "attaching ears" operations preserve centroids. And centroids are additive, i.e. C(polygon #1 + polygon #2) = C(polygon #1)+C(polygon #2).
Challenge 2: The "attaching ears" operations are commutative. After one round of each, we already have a regular pentagon. So the result will still be a regular pentagon after another round of 144 and two rounds of 216.
27:34 Is this a special case of Vandermonde matrix?
28:36 Proof using complex numbers: Since the two original triangles are similar, there exists a complex number z=(X(2)-Y(2))/((X(1)-Y(1)), where X and Y are any two colors among {Black, Brown, Red}. As a result, ([X(2)+X(1)]-[Y(2)+Y(1)])/((X(1)-Y(1))=z+1, thus the sum is still similar.
One minute into this video, I was already thinking about complex numbers since I know the Napoleon's theorem. :P
Then I tried to use the same method on the Petr-D-N theorem:
Let the five vertices of the original pentagon be complex numbers u, v, w, x, y.
First consider the operation of attaching 72° ears, the first new vertex becomes
u' = u+ {e^(3πi/10)/[2cos(3π/10)]}(v-u) = (e^(-3πi/10)u+e^(3πi/10)v)/[2cos(3π/10)], a linear combination of u and v.
Then comes the 144° ears, and the first new vertex becomes u'' = (e^(-πi/10)u'+e^(πi/10)v')/[2cos(π/10)] = some much more complicated linear combination of u, v and w.
That's why I really appreciate the discrete Fourier transform point of view. Thank you!
1
-
Full disclosure - the referenced work is mine, but this is not a mindless plug.
Intimate Fibonacci/binomial coefficients/partitions connections with generalizations over 2 integer parameters are detailed with code in Sloane's OEIS A347187 and detailed in videos on the Feral Mathematician channel. This one is a summary from earlier videos https://youtu.be/zdhdiO4uUl0
The Fibonacci numbers are generalized to F(a,b,n) =F(a,b,n-a)+F(a,b,n-b) and shown to be the first 2 terms of the recurrence for the partitions function, similarly generalized and expressible in terms of binomial coefficients, which carries over to the sigma sum of divisors function. I have no doubt the work detailed in this video fits right into this emerging framework, just needs to be sussed out.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
DDDDDAAAANNNGGIIIITTTT I wanted to make a video about this!
This is something that I had explored myself like three years ago. I came up with many of the insights you presented here. My original motivation was trying to figure out how I can come up with a formula for the sum of squares (up to some point), which had a known formula and you could obviously prove it inductively, but how do get the formula to begin with right?
Eventually I thought of the idea to take differences on top of differences... then do linear combinations of the difference sequences you see for 1, x, x^2, x^3, etc to get a formula for an arbitrary sequence (such as the sum of squares up to some term), which only works perfectly if it fizzles out to a constant row of course(or at least, if there is a finite number of interesting rows, you don't have to do any further analysis). There were problems I've come across where I knew the end result should be a polynomial and I needed to do polynomial interpolation, so this was a handy tool for that. At some point I also discovered you can use starting values/diffs of 0,0,0,0,...,1 as your basis instead which turned out to be (n choose h) for "height" h.
This also gives you a handy way to do polynomial evaluation without using multiplication, and given this tool you can figure out how to come up with formulas for programs that involve loops where variables just keep adding to other variables in a non-cyclic manner.
I will say though, that I probably wouldn't have made this video for quite a while longer(procrastinating...).
Some things that I had thought to explore but didn't(because I got distracted by other things)... what if we have infinitely many rows of these difference sequences (which you touched upon). What about trying to interpolate in between the terms, between the rows? What about extending these difference sequences into the real and imaginary axes, can this lead to a reasonable interpretation of summations where the indices are complex numbers, at least for polynomials? What if our input sequence isn't a linear function?
I gotta start making videos on my ideas before I get scooped as a result of my own laziness...
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
12:00 RRRL: {{1,1},{3,2}} > {{1,2},{5,3}} > {{1,3},{7,4}} > {{1,4},{5,9}} > {{9,4},{17,13}}
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
You only need 10 "essentially different" combinations of 4 for a simple test.
AAAA = [RRRR, BBBB, YYYY]
AAAB = [RRRB, RRRY, RBBB, RYYY, BRRR, BBBR, BBBY, BYYY, YRRR, YBBB, YYYR, YYYB]
AABA = [RRBR, RRYR, RBRR, RYRR, BRBB, BBRB, BBYB, BYBB, YRYY, YBYY, YYRY, YYBY]
AABB = [RRBB, RRYY, BBRR, BBYY, YYRR, YYBB]
ABAB = [RBRB, RYRY, BRBR, BYBY, YRYR, YBYB]
ABBA = [RBBR, RYYR, BRRB, BYYB, YRRY, YBBY]
AABC = [RRBY, RRYB, RBYY, RYBB, BRYY, BBRY, BBYR, BYRR, YRBB, YBRR, YYRB, YYBR]
ABAC = [RBRY, RBYB, RYRB, RYBY, BRBY, BRYR, BYBR, BYRY, YRYB, YRBR, YBYR, YBRB]
ABCA = [RBYR, RYBR, BRYB, BYRB, YRBY, YBRY]
ABBC = [RBBY, RYYB, BRRY, BYYR, YRRB, YBBR]
3+12+12+6+6+6+12+12+6+6=81, so every RBY combination is used.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
One thing I didn't see pointed out (sorry if I missed it) - if you look at the highlighted diagonals at, e.g. 14:25, but read them across, there's a pattern. The third row, for example, shows 10, 40, 90... That is, 10, 10*2^2, 10*3^2,... And in general every row r whose first highlighted digit (generated by Pascal's Triangle) is a_r (where the 0th row is all 1s) generates the sequence n^r * a_r for the nth highlighted number.
Since the last row in any such diagonal will always have a 1 as the first highlighted number, naturally the sequence of the bottom row collapses to n^r.
I don't think that counts as a proof (at least not without a lot of additional work) but thought it was an interesting angle. So to speak.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
3:09 Proof that the shape defined as "the curve y=1/x stays the same curve when you both stretch vertically and squish horizontally by a factor s":
A curve y=f(x) stretched vertically by a factor v is just y=v f(x), and squished horizontally by a factor h is just y=f(h x).
With that, we can show that in our case, y=s f(s x) still lies on the curve defined by y=f(x), because with f(x)=1/x, y=s 1/(s x) = 1/x = f(x). So the transformed curve is the same as the original curve, making it resistant to shape-shifting, and any section of the curve will turn into another section still on the curve when applying the transformation.
9:00 So you have A(X) and A(Y), and want to stretch A(Y) so that it fits on the right edge of A(X). You know that A(Y) has a height of 1 on the left and A(X) has a height of 1/x on the right, so for them to fit together, you need to vertically stretch A(Y) by a factor of 1/X (aka vertically squish by a factor of X). The bottom edge of A(Y) originally had the length Y-1 (because the edge goes from 1 to Y), but because we want the new shape to fit the 1/x curve, we also need to stretch it horizontally by X after having squished it by that amount. So the bottom length of the new A(Y) will be X(Y-1) = XY - X. The resulting total shape after adding A(X) and the transformed A(Y) together will fill the curve until the point where A(X) ends plus the length of the transformed A(Y), so in total, X + XY - X = XY. Per definition of A, this area is A(XY). But we also know that this area is just A(X) + A(Y), because the transformed A(Y) has the same area as the original.
10:47 Proof that it works for all integers: If log(XY)=log(X)+log(Y) for all (X, Y), then 0=log(1)=log(X/X)=log(X) + log(1/X), and therefore, log(1/X)=-log(X), and by extension, log(X^-N)=-log(X^N)= -N log(X). For the exponent 0, you can just use log(X^0)=log(1)=0=0 log(X).
For rational powers of X, so for the rule p/q log(X)=log(X^(p/q)), we can take log(X^p)=(q log(X^p))/q =(pq log(X))/q=p/q log(X^q). Now in that equation, substitute Y=X^q (which implies X=Y^1/q), and you get log(Y^p/q) = p/q log(Y). This equation must hold for all Y, because any Y can be expressed in terms of X using the earlier substitution rule.
For irrational exponents, you can argue that the logarithm must be continuous on any compact interval, as 1/x is continuous as well, and because the rationals are dense in the reals, that shows that the rule must also be true for any real number.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I went to high school in the 1960s and we had to use slide rules for physics problems. Everyone else had straight slide rules, but I had a circular slide rule — my father (an engineer) had always favored the circular ones. Two advantages: (1) you didn't have to figure out which direction to slide it, and (2) my 8" circular model had a circumference of over 25", compared to the 8" or 10" straight rules of my classmates, so it was more accurate. In addition (no pun intended) it had a log-log scale, so I could work with exponents, eg, raising pi to the power of e, as well as trig functions to work with sin, cos, tan, etc.
I used that thing almost daily for a long time. I kept it for many years, then finally threw it away, which I now sorely regret.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Solution to finding the square for the pythagorean triple at 11:45 (153^2+104^2=185^2):
Take prime factors of 153 and 104/2
3, 3, 17 and 2, 2, 13
Test possible combinations of two factor for each until one fits the rules of the square.
9, 17 and 4, 13 result in
9, 4, 13, 17
9+4=13
13+4=17
9*13+4*17=117+68=185
Therefore the square is composed of 9, 4, 13, 17.
Some extra facts I noticed while thinking about the problem:
The bottom parts of the square must each be greater than the sum of the top parts, and therefore the factorizations of the two pythagorean triples being summed must be in increasing order respectively within the square; this means that there is only 1 order that must be tested for any combination of factorizations.
Every triple must contain both an odd and an even number, where the odd corresponds to the left of the square, and the even corresponds to the right.
The top numbers in the squares as you follow the tree can never decrease.
The position in the tree for the solution to this triple, following the pattern for expanding the tree as shown in the video by this point, is:
Start, right, right, right, left.
EDIT: I realized that I accidentally answered the bonus question for the problem when resuming the video after posting this comment; hooray for my curiosity!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
If you have an n-dimensional cube of side x, made up of C = x * x * x * ... * x unit cubes, add an additional cube on every outside "face", an additional cube on every outside edge, etc. The remaining shape is an (x+2) * ... * (x+2) cube, with (x+2)^n unit cubes; at the same time, it consists of C + H + .... + F + E + V cubes.
Or, using coordinates and combinatorics: you can define an n-dimensional cube with vertices at (±1, ..., ±1). Every edge corresponds to a pair of vertices that differs in only one coordinate, (±1, ..., ±1, ?, ±1, ..., ±1). Every face corresponds to four vertices that differ only in two coordinates, (..., ?, ..., ?, ...), and so on. For an m-dimensional "bit", we have nCm choices for placing the question marks, and 2^(n–m) choices of sign, showing immediately that there are 2^(n–m) nCm of them.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Split-complex numbers relate to the diagonality (like how it's expressed on Anakin's lightsaber) of ring/cylindrical singularities and to why the 6 corner/cusp singularities in dark matter must alternate.
Dual numbers relate to Euler's Identity, where the thin mass is cancelling most of the attractive and repulsive forces. The imaginary number is mass in stable particles of any conformation. In Big Bounce physics, dual numbers relate to how the attractive and repulsive forces work together to turn the matter that we normally think of into dark matter.
Complex numbers = vertical asymptote. Split-complex numbers = vertical tangent. Dual numbers = vertical line. These algebras can be simply thought of as tensors. Delanges sectrices can be thought of as opposites of vertical asymptotes. Ceva sectrices as opposites of vertical tangents, and Maclaurin sectrices as opposites of vertical lines.
The natural logarithm of the imaginary number is pi divided by 2 radians times i. This means that, at whatever point of stable matter other than at a singularity, the attractive or repulsive force being emitted is perpendicular to the "plane" of mass.
In Big Bounce physics, this corresponds to how particles "crystalize" into stacks where a central particle is greatly pressured to degenerate by another particle that is in front, another behind, another to the left, another to the right, another on top, and another below. Dark matter is formed quickly afterwards.
Ramanujan Infinite Sum (of the natural numbers): during a Big Crunch, the smaller, central black holes, not the dominating black holes, are about a twelfth of the total mass involved. Dark matter has its singularities pressed into existence, while baryonic matter is formed by its singularities. This also relates to 12 stacked surrounding universes that are similar to our own "observable universe" - an infinite number of stacked universes that bleed into each other and maintain an equilibrium of Big Bounce events.
i to the i power: the "Big Bang mass", somewhat reminiscent of Swiss cheese, has dark matter flaking off, exerting a spin that mostly cancels out, leaving potential energy, and necessarily in a tangential fashion. This is closely related to what the natural logarithm of the imaginary number represents.
Mediants are important to understanding the Big Crunch side of a Big Bounce event. Black holes have locked up, with these "particles" surrounding and pressuring each other. Black holes get flattened into unstable conformations that can be considered fractions, to form the dark matter known from our Inflationary Epoch. Sectrices are inversely related, as they deal with dark matter being broken up, not added like the implosive, flattened "black hole shrapnel" of mediants.
Ford circles relate to mediants. Tangential circles, tethered to a line.
Sectrices: the families of curves deal with black holes and dark matter. (The Fibonacci spiral deals with how dark matter is degenerated/broken up, with supernovae, and forming black holes. The Golden spiral deals with black holes being flattened into dark matter during a Big Bounce event.) The Archimedean spiral deals with black holes and their spins before and after a reshuffling from cubic to the most dense arrangement, during a Big Crunch. The Dinostratus quadratrix deals with the dark matter being broken up by ripples of energy imparted by outer (of the central mass) black holes, allowing the dark matter to unstack, and the laminar flow of dark matter (the Inflationary Epoch) and dark matter itself being broken up by lingering black holes.
Delanges sectrices (family of curves): dark matter has its "bubbles" force a rapid flaking off - the main driving force of the Big Bang.
Ceva sectrices (family of curves): spun up dark matter breaks into primordial black holes and smaller, galactic-sized dark matter and other, typically thought of matter.
Maclaurin sectrices (family of curves): dark matter gets slowed down, unstable, and broken up by black holes.
Jimi Hendrix's "Little Wing". Little wing = Maclaurin sectrix. Butterflies = Ceva sectrix. Zebras = Dinostratus quadratrix. Moonbeams = Delanges sectrix. Jimi was experienced and "tricky".
Jimi was commenting on dark matter. How it could be destabilized by being slowed down, spun up, broken up by lingering black holes, or flaked off. (The Delanges trisectrix also corresponds to stable atomic nuclei.)
Dark matter, on the stellar scale, are broken up by supernovae. Our solar system was seeded with the heavier elements from a supernova.
I'm happily surprised to figure out sectrices. Trisectrices are another thing. More complex (algebras) and I don't know if I have all the curves available to use in analyzing them. I have made some progress, but have more to discern. I can see Fibonacci spirals relating to the trisectrices.
The Clausen function of order 2: black holes and rarified singularities are becoming more and more commonplace.
Doyle's constant for the potential energy of a Big Bounce event: 21.892876
Also known as e to the (e + 1/e) power.
At the eth root of e, the black holes are stacked as densely as possible. I suspect Ramanujan's Infinite Sum connects a reshuffling from the solution to the Basel problem and a transfer of mass to centralized black holes. Other than the relatively small amount of kinetic energy of black holes being flattened into dark matter, the only energy is potential energy, then: 1 (squared)/(e to the e power), dark matter singularities have formed and thus with the help of Ramanujan, again, create "bubbles", leading to the Big Bang part of the Big Bounce event.
My constant is the chronological ratio of these events. This ratio applies to potential energy over kinetic energy just before a Big Bang event.
Methods of arbitrary angle trisection: Neusis construction relates to how dark matter has its corner/cusp singularities create "bubbles", driving a Big Bang event. Repetitious bisection relates to dark matter spinning so violently that it breaks, leaving smaller dark matter, primordial black holes, and other more familiar matter, and to how black holes can orbit other black holes and then merge. It also relates to how dark matter can be slowed down. Belows method (similar to Sylvester's Link Fan) relates to black holes being locked up in a cubic arrangement just before a positional jostling fitting with Ramanujan's Infinite Sum.
General relativity: 8 shapes, as dictated by the equation? 4 general shapes, but with a variation of membranous or a filament? Dark matter mostly flat, with its 6 alternating corner/cusp edge singularities. Neutrons like if a balloon had two ends, for blowing it up. Protons with aligned singularities, and electrons with just a lone cylindrical singularity?
Prime numbers in polar coordinates: note the missing arms and the missing radials. Matter spiraling in, degenerating? Matter radiating out - the laminar flow of dark matter in an Inflationary Epoch? Corner/cusp and ring/cylinder types of singularities. Connection to Big Bounce theory?
"Operation -- Annihilate!", from the first season of the original Star Trek: was that all about dark matter and the cosmic microwave background radiation? Anakin Skywalker connection?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
For odd n (say 7, a heptagon), this means that you use the permutation (1, -2, 3, -4, 5, -6). But that is just (1, 5, 3, 3, 5, 1). And that doesn't give you a regular heptagon yet.
For even n (say 6, a hexagon), this means that you use the permutation (1, -2, 3, -4, 5), which is (1, 4, 3, 2, 5) and is just another permutation giving you the same regular n-gon after the 4th cycle (and giving you the single vertex center after the last cycle).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
As some one familiar in computer science, the 1, 2, 4, 8 question comes easy answer as all the binary representation of numbers from 1 to 15. Using the (1+x), (1+x^2), (1+x^4), (1+x^8), it can be seen by multiplying all with (1-x), we will get (1-x^16), and divide that of (1-x), we will get (1+x+x^2+x^3+...+x^15), aka the same answer. This is such a very interesting connection between polynomials and binary numbers!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
18 is 3^2 + 3^2 with either, both, or neither of those threes changing to negative.
2020's odd factors are 1, 5, 101, and 505 and all those are good, so this indicates 16 ways. Let's see...
16^2 + 42^2, 24^2 + 38^2... Ah, wait! I can switch the order, so either of those give eight ways, making 16 total. I was worried for a while!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The L1 metric (taxicab metric) is an interesting thing to bring up - I’ve thought about it too, as part of preparation of this video.
Since (R³,||.||₁) is not an inner product space, there is no natural way to define “normal”.
Also, the set of isometries (or rigid motions) of this space is a lot less interesting than for (R³,||.||₂) - basically only translations, reflections and rotations that preserve the “grid” - no arbitrary rotations.
Due to both of these it is a lot harder to get a sense of what the area of an arbitrary parallelogram is - you can’t compare it to a length x width rectangle in the same plane because right angles rely on inner products, and you can’t compare it to a rectangle in the coordinate plane because there is no guarantee of an isometry that carries the parallelogram into a congruent parallelogram in the xy-plane.
Without this fundamental idea settled, measuring surface area of a curved surface (using some sort of integral of areas of infinitesimal parallelograms) has a hard time getting started.
I think the only natural way to measure surface area in this space is via the “how much paint” method discussed in some recent mathologer videos in normal Euclidean space. We can define volume in (R³,||.||₁) geometrically, and for a surface S, the set T of points within ε of S is a volume we can measure. So consider the limit of T/(2ε) as ε -> 0. I haven’t investigated this myself beyond this, and I don’t know if anyone has.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Triangle lover here.. This is marvellous. Thank you so much
I am very surprised this was not 'known'.
It seems that it must have been known at some level or other by people making things - in cloth or clay - in tiles, especially in China, Japan, Korea, India and especially Hindu Buddhism or Islam where geometric repetition in huge number and scale are central to sacred geometric patterns, architectural relationships and texture.
Cut and fold with paper cloth wood, clay or fired ceramic.. glass tiles/windows ?
Embroidery, patchwork, quilting
String and stick tools used as templates and jigs.
Musical instrument makers workshop ?
All the natural patterns that derive from simple folding, repeats or recursive actions, or string straight edge and compass. Angle dividers...
My feeling is It would be hard not to know or have met this this in many ways
But might never have been presented, or kept and translated across the slik roads, libraries Alexandria Istanbul
Italy Germany France and England when renaissance printing took off.
Hope we get a superfast Visual-Math AI tool to search across imagery via time culture place pattern/
and perhaps even in number system counting ?
scaling tricks for people who make things ?
Thank you for this channel and your videos
inspiring!
I've seen on TikTok neat triangular vertex-hinged shutters transform from door to an open gateway.
This pattern is very scifi.
Now I want to try to use the geometry you present here to make a human sized set of experiments here at my studio
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I'm a layman and know almost nothing about math and haven't studied math in decades. But my answer for 36:17 is that the area of the square in the middle is 0.2 (or 1/5). I came to that conclusion by taking that 1/2 length of side "a", but then because those lines are parallel, using that 1/2 as the hypotenuse of a smaller triangle inside that middle square. If that's the case, then side "b" of that inner triangle would be the side of that inner square. And if we multiply that scaling factor (I don't know what to call it) to get the side "b", and then square that (to get the area), it comes to 0.2 (or 1/5). The actual calculations are as follows. We have (1/2)*2 + 1^2 = c*2. So 1/4 + 1 = 1 1/4 (= 1.25). That's c^2. The square root of 1.25 is 1.11803398... So that's "c". Now, we make 0.5 the hypotenuse, so what is "b"? 1/1.11803398... = 0.89442719... . So it's that times the side that's 0.5 units in length. 0.89442719... * 0.5 = 0.44721359. That gives the length of one side of the center square. And when you square that, the answer is 0.2 (or 1/5) for the area of the center square. Okay, well, that's the answer I came up with, at least!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
There is a potentially serious issue with this method of distributing claims, and that is the concept of splitting or merging claims. If the total amount of available funds is less than half of the total amount claimed, there is an incentive for a claimant to sell off partial claims in order to have a larger number of smaller claims filed, thus paying out a proportionally larger award to each of the smaller claimants than the one large claimant could receive if they filed individually. And if the total funding is more than half the total amount claimed, there is an incentive for smaller claimants to collectively file a single claim for their combined amount, then redistribute the amount awarded proportionally. This was actually alluded to in the 'three sons' example, as if each of the three sons filed a claim independently, the 'red son' would be awarded 4/9ths of the estate (more than 5/12ths), while each of the 'green sons' would be awarded 5/18ths of the estate (less than 7/24ths). The primary ways out of this situation is to either have very complex rules on how claims can be spread, thereby refusing the ability to split or merge claims, or switch to a proportionality-based claiming system.
1
-
1
-
1
-
1
-
1
-
Comforting part: how Nicole Oresme proved the divergence of the harmonic series, because it feels nice to find in your video something, I am already familiar with.
The best chapter: the fifth, about the gamma constant and its usefulness in defining how much terms you need in the harmonic series for it to be equal to some particular number, very nice stuff.
Answer to your question: why gamma is greater than 0,5?
The gamma constant as you say equals the totality of the blue area over the curve f(x)=1/x .
This area (gamma) is definitely greater than the sum of quarters of areas of blue squares, because looking from the bottom the curve is convex.
So, gamma > 1/4 + 1/12 + 1/24 + 1/40 + …
gamma > (1/2 *1/2) + (1/2 * 1/6) + (1/2 * 1/12) + (1/2 * 1/20) + …
gamma > 1/2 * [ (1 -1/2) + (1/2 – 1/3) + (1/3 - 1/4) + (1/4 – 1/5) + …]
Therefore, gamma > 1/2, because everything cancels out except for the first product (1/2 *1)
You can also see 1/2 coming from the formula for the area of a triangle.
Is there a simpler way to do it?
Oh, and it’s fun to replay the video and see in the introduction the parabolic tower on top of the letter L.
Your content is great!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
3:40 i started by trying to out the distance between the centre and 2 of the whiskers growing out of a single vertex, and that was easy because the centre is given by halving the triangle's angles - i just watched a 3B1B video proving that so that part of the proof isn't mine - the, we can deduce that the line that is obtained by splitting the angle can be extended in the other direction, giving us 2 angles, both equal to the other 2 angles we've gotten inside the triangle, therefore the points, which are the same distance from the vertex by definition, are also the same distance from the centre. then just noticed all the chords are the same length, and quickly proved that if 2 points on the ends of a pair of lines of equal length (correct me if thats not the right term) are the same distance to a point, then the other ends of the lines are the same distance from that point as well, using the same angle argument. this is the proof, i probably didnt get all the math terminology right so correct me please! thanks for making this!
also, i think they have the same diameter. the corners of the wierd conway's curve are more pointed and bulge outwards, while the edges bulge inwards, although its not a proof or anything
edit 2: i was wrong
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Regarding the first problem you postulated, the answer I got is that, for N cards, then N*(N+1)/2 , (the triangular numbers!) is the maximum number of steps to finish the game. I'm trying to construct a rigurous proof, but I have failed so far. But, at least, I want to present an intuitive argument of why I think this should be the case.
First, for N=1, it's easy, just one step, and the formula holds. :)
Now, let us assume that for N cards, this result is valid. What happens when we have N+1 cards?. An intuitive idea would be that we leave the first card (the one at the extreme left) alone for the moment, and we take as much steps as possible before we are forced to flip that card. So, with the rest of the N cards, we take the most possible number of steps to flip those, which is, by inductive hypothesis N*(N+1)/2. So, we end with the following scenario:
1 0 ... 0 0
So, in this case, it is straight forward what happens next: we just have one card, and the rest of the moves are just
1 0 ... 0 0
0 1 ... 0 0
...
0 0 ... 1 0
0 0 ... 0 1
0 0 ... 0 0
So, we have N+1 steps to finish the game from this state. And, of course (N*(N+1)/2) + (N+1) = (N+1)(N+2)/2, which yields the desired result.
Of course, this proof is NOT complete, because the obvious question would be: "Well, what happens if you flip the first card earlier?", and I've noticed that indeed, taking other strategies, we can also obtain (N+1)(N+2)/2 steps. The argument I'm currently working on is proving that, if you make a flip in any other point of the maximal sequence, you would end up in some corresponding step of the first maximal sequence of steps at best, or in, the worst case scenario, in a faster sequence of moves. For example, let us consider the following flip:
1 1 ... 1 1
0 0 ... 1 1
In this case, the first two card would be, effectively, out of the game, and you couldn't optimize the number of moves for N cards, because then you would only have N-1 "active" cards as maximum.
So, I think my answer is correct, but I'm currently working in the rigurous proof of why this formula should work for all possible strategies.
Thanks for the video, BTW!. Great format, is always a pleasure to watch your content.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The first numbers of each row in the linear pattern are the square numbers (1, 4, 9, 16...), while the first numbers of each row in the square pattern are every other triangular number (3, 10, 21, 36...) and the last numbers in each row are one less than the other triangular numbers (6, 15, 28, 45, ...).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
import numpy as np
from functools import reduce
def change(n, coins):
# Initialise all our polynomial coefficients to 0
coeffs = [np.zeros([n+1]) for _ in range(len(coins))]
# Place a 1 at the positions corresponding to our coin values
for (coeff, coin) in zip(coeffs, coins):
coeff[0:n+1:coin] = 1
# Convolve the coefficients to multiply the polynomials ... maybe there's a fast trick here ;)
coeffs_conv = reduce(np.convolve, coeffs)
# Return the coefficient at the desired position
return coeffs_conv[n]
# Example from the video
change(100, [1, 5, 10, 25, 50, 100])
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Your cardioids, late in this video, are epicycloids, which are what you would get if you played with Spirograph but rolled the little wheel around the outside of the big one, rather than the more typical hypocycloids, which are what you get when you roll the little wheel around the inside of the big one. Therefore, any mathematical articles and theorems regarding Spirograph patterns -- how to know how many cusps you're going to get, where they're going to lie, etc. -- ought to be equally applicable to your cardioids here -- if that's of any help to you.
(What continues to defy me, though, despite your having just now explained it, is how cardioids arise here, in the Vortex patterns, in the first place, as a consequence of the cumulative straight lines of the Vortex happening to collectively lie tangent to it. It's the same thing as getting a quarter-circle by joining with a straight line the points (K, 0) and (0, K), for K = 0...N, about which I am also still waiting, after forty years, to hear a reasonable explanation. This also applies to the question of "what is the shape of the curve you see along the edge of a cube, when you spin it at high RPM around an axis that passes through opposing vertices?", another one I've been waiting since the 1980s to even find out *how to approach*....)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
There's something extra special going on with those approximations. With pi, for example, Wolfram lists the following approximations.
3/1, 22/7, 333/106, 355/113, 103993/33102, 104348/33215,...
First: if you check the spirals corresponding to these fractions, you'll notice they alternate between spinning left and spinning right. (I'm too lazy to say CW or CCW/ACW/whatever) This is why 333/106 is listed even though 355/113 gets even closer for pretty much the same bang for your buck: the spirals must alternate in direction.
The reason for this will be clear by the end of this post.
Next: these fractions also correspond to evaluating the continued fractions.
Related to that point is how they appear on the microscope:
3/1 is the first good approximation, and at the beginning the spiral just looks like one arm.
Then as you zoom in, the approximation loops over itself until a better approximation appears: 22/7. The seven arms.
Then the arms continue to loop over and twist among themselves until the next one over lines up nicely. This becomes 333/106 and 355/113.
Notice what this means on the algebraic side:
22/7 = (7 * 3 + 1) / (7 * 1 + 0)
333/106 = (15 * 22 + 3) / (15 * 7 + 1)
355/113 = (1 * 333 + 22) / (1 * 106 + 7)
103993/33102 = (292 * 355 + 333) / (292 * 113 + 106)
104348/33215 = (1 * 103993 + 355) / (1 * 33102 + 113)
For starters, this comes directly from the continued fraction. But now the connection between the microscope and the continued fraction is more obvious.
Given the sequence A1/B1, A2/B2, the next term A3/B3 would be: (M * A2 + A1) / (M * B2 + B1)
The logic for the denominator is this: B2 represents how many spirals there were on the last iteration, and B1 how many spirals before that. When the spirals loop over, the spiral that approaches is labelled B1, and it comes over in tiny steps of B2.
This also explains the alternating nature of where the spirals curve. B1 must be opposite to B2's spin.
Idk it's midnight I'm yapping and idk if this is correct
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
transcribed problem: The inhabitants of little Wurzelfold have been considerably roused out of their
lethargy by the great war. Like the rest of the world, they talk of little else,
so it is natural that the particular conversation which we have to relate should
have a distinct material flavor. The little group of villagers gathered together
in the red lion Inn where all respectable neighbor whose ages may safely be said
to range from 50 to 70 years. All fir young men of the village had enlisted, and
the opinion was lead by everyone that if the English Navy and Army did not in
consequence "knock the Germans into a cocked hat", it was not the fault of little
Wurzelfold. There is no necessity to introduce the reader individually to these
estimable men, drawn together by bonds of mutual sympathy and desire to forget
their anxieties in pleasant intercourse. We are entirely concerned with the
curious riddles and proses that, without any premeditation, happened to be
forthcoming on the particular occasion of which we write.
"I was talking the other day" said William Rogers to the other villagers gathered
round the Inn fire, "to a gentleman about that place called Louvain, what well–
used to visit a Belgian friend there. He said the house of his friend was in a
long street numbered on his side one, two, three, and so on, and that all the
numbers of him added up exactly the same as all the numbers on the other side
of him. Funny thing that! He said he knew there was more than fifty houses on that
side of the street, but not so many as five hundred. I made mention of the matter
to our parson, and he took a pencil and worked out the number of the house where
the Belgian lived. I don't know how he done it."
Perhaps the reader may like to discover the number of that house.
"That reminds me," said James Woodhouse " of Miss Wilkinson, down Bradford way.
She promised to make a lot of red crosses, and to stitch 'em on to white bands
for hospital nurses at the war. She bought a lot o material and started cuttin'
out all the crosses. When she had done it she found that by mistake she'd made
em just twice as big as they ought for to be. She is very economical, as you
may say, is so she went to the schoolmaster and asked him how she was to cut each
cross into two crosses of the same shape. He showed her an artful way of doing
it in five pieces so that both crosses should be of one size, and no waste.
He's a sharp man, I reckon."
It is true that the red cross can be cut into five pieces that will form the two
crosses shown in our first diagram.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
22:07: How can the maximum cycle length be 2260? What if your algorithm has more than 2260 steps? And OBTW, what if it's prime? How can (for example) a non-trivial (do nothing) 2267* step algorithm possibly repeat in fewer than 2260 steps?
And OBTW, how can an algorithm of 43,252,003,274,489,856,001 steps (clearly mutually prime with 43,252,003,274,489,856,000) repeat in a factor of 43,252,003,274,489,856,000 steps?
* 2267 is prime.
1
-
1
-
Challenge 1:
292(there is only 1 way for 100cents to make 1 dollar) Or in the expansion of the series, multiplying the product before with (1+x^100) will only add 1 new term of x^100
For the 1,2,4,8 the expansion of (1+x)(1+x2)(1+x4)(1+x8) is 1+x+x2+....x15 (simply multiply and divide by term(1-x))
In other point of view, as all coins are in power of 2, write the values in binary, which is 1,10,100,1000. It is clear that the digits won't interfere each other no matter how we add them. Their combination is simply 2^4-1=15 (excluding 0)
Conclusion: 15 values can be made with 1 way each, if 0 is not considered as a value
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
2 is the sum of the reciprocals of [1, 2, 3, 6]
3 is the sum of the reciprocals of [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 15, 230, 57960]
Had to bust out python, but the algorithm is quite simple.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I actually calculated the (signed) area of the triangle with side lengths aA, bB, and cC, which I call the Ptolemian, in an attempt to prove the inscribed square theorem and because I thought it would be an efficient way to check if four points lie on a common circles.
It turns out to be a constant time the determinant of the 4×4 matrix, with row entries of 1, x, y and x²+y², for every point (x,y) of the original quadrilateral.
This is the algebraic way of saying that the points satisfy a common equation of the form
a+bx+cy+d(x²+y²), which is obvious in hindsight.
It also satifies some interesting identities using area, and given a quadrilateral ABCD, the Ptolemian of this quadrilateral is equal to a constant times the area of ABC times the power of D to ⊙ABC.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Mathologer, if you see this and want to know my background, it's as an actuary. Like almost everything in corporate employment, though, learning the science (or earning any degree) just gets you a ticket into the door. Unfortunately in the real world, theoretical fun rides like this are nothing but theoretical fun rides that I do on my own time. Lots of actuaries enjoy puzzle solving, telescoping equations and relationships, tricks for quickly doing amazing demos to others ("What's the 4th root of 4096?"), and elegant proofs. But it's, in my observation, strictly for theoretical enjoyable musings. It doesn't pay the bills. Or the Woos.
Mathematical skills (memorizations, tricks, quick mental reductions, etc.) however do aid tremendously in programming. And very much for zen worthy programming "with style".
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Let me see if I got the One-Sentence Proof. Working backwards, if (x,y,z)->(x,z,y) on S has a fixed point, then y=z, so x^2 + 4yz = p can be written x^2 + (2y)^2 = p, and we have a sum of two squares. If |S| is odd, then any involution must have at least one fixed point (more generally an odd number of them), because the involution (being its own inverse) swaps all the non-fixed points with each application, so they must come in pairs. The three-line involution can only have a fixed point if either x+2z=x, so z=0 (meaning x^2=p, which isn't possible if p is prime); 2y-x=x, so y=x=1 (otherwise, p has a factor other than itself and 1); or x-2y = x, so y=0 (not possible for the same reason z!=0). So x=y=1 is the only possibility for x and y. Substituting into x^2+4yz=p, we get 1+4z=p, which has one solution for z and also requires that p is congruent to 1 mod 4. So there is only one fixed point to this involution, so |S| is odd, so any other involution over S has a fixed point, so in particular (x,y,z)->(x,z,y) has one, so y=z for some element of S, so x^2+(2y)^2 = p has a solution. How did I do?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I first ran into this in 1st year university. Needless to say it has provoked endless conversations and disagreements, some of which get somewhat heated (all in good fun though). So when I asked my Calculus prof about this (back in the 70's) he gave me an answer which made me seriously re-think about what we really mean by area when applied to a curved surface. He simply said the we may not understand area on a curved surface as well as we think we do. Because, as we all know, area is defined - square meter- on a FLAT surface like the rectangle. But when we apply that unit to a curved surface we have to distort it (stretch it) to make it lie on the curved surface - and then all heck breaks loose. And, as you correctly point out, we are dealing with an "ideal" paint so squeezing some molecule into the horn just doesn't apply here.
On the other hand, the volume of the horn is based on the flat definition of volume (3 dimensional - but still "flat" space). And that means the volume concept does NOT have to be distorted or stretched when measuring the space inside of the horn. The inside of the horn simple lives in flat space. Whereas the surface area lives in curved space.
And, yes, I know that as you make a small area on the curved surface shrink and shrink the more and more it starts to look and behave like a flat surface - but it really never does become flat - certainly not in the macro world.
BTW, I know Mathematics Teachers who are still convinced that: .9999...... = 1 is a "paradox". But it really isn't. Infinity is not a number nor a place, it's just weird. Thanks so much for all your work. I really do like your approach.
1
-
1
-
1
-
1
-
1
-
1
-
The number of integer squares for 2020 by the heavenly formula is: 16
The possible factors are:
-42 & -16, -42 & 16, -38 & -24, -38 & 24, -24 & -38, -24 & 38, -16 & -42, -16 & 42, 16 & -42, 16 & 42, 24 & -38, 24 & 38, 38 & -24, 38 & 24, 42 & -16, 42 & 16.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
@Mathologer Yes, true that. Well; both the area and the SSSSSemiperimeter, of the 3–4–5-triangle, are 6; and, if we find a 7, we can divide other things, with it; and we’ll get the rest of the digits covered, except 9 and 0. But 2, 7, and 8 will be taken care of :). Plus; the fact that it’s precisely the LEAD digits of π that are so prevalent, in the 3–4–5-triangle, makes it all EXTRA special, to me; because they’re the ones that usually get represented, when π is displayed, in decimal form; and because it’s already an unreasonably good approximation (to 4 decimal places). 🙂
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I'm not sure if I would consider this as a proof or a postulate. The Pythagorean Theorem is Embedded within both the addition (+) and assignment, equality, identity operator (=). I'll start with the later.
We can take the simplest expression y = x. Depending on context this expression or statement is stating that either two quantities are equivalent, or that we are assigning x to y. I also call it identity since nothing within the expression has changed. Here's a list of our pairs {..., (-1,-1), (0,0), (1,1), ...}
We know that y = x is also linear from the slope-intercept y = mx+b where b is the y-intercept and m is the slope defined by (y2-y1)/(x2-x1) = dy/dx. The graph of the line bisects the XY plane in the 1st and 3rd quadrants. Since the x and y-axes are perpendicular or orthogonal to each other making a right angle of 90 degrees or PI/2 radians at the origin (0,0) we know that this line has a 45 degree or PI/4 radian angle that is between the line y = x and the +x-axis giving us an angle in standard form with respect to the origin. We also know that the slope of this line is 1. If we set the length of this line to 1 going from the origin (0,0) to some point (x,y) and draw a vertical line down from (x,y) to the +x-axis we end up with the coordinate ( sqrt(2)/2, sqrt(2)/2 ) and the lengths of x and y are sqrt(2)/2. We can take the slope formula dy/dx and simply use (x,y) since (x1,y1) = (0,0). We can see that (sqrt(2)/sqrt(2)) = 1 since sin(45)/cos(45) = tan(45) = 1. Here the angle theta (t) is the angle with respect to the line y = mx+b, the origin and the +x-axis. With this we can substitute the slope m = dy/dx with sin(t)/cos(t) = tan(t). This then would give us y = sin(t)/cos(t)x + b and y = tan(t)x + b. Why does this derived substitution from y = x work? This is where the addition operator comes into context.
Here we will use the very first and simplest of all mathematical operations 1+1 = 2. With this I'll demonstrate how this expression, equation indirectly satisfies and solves the Pythagorean Theorem and I will also show how it relates to the Equation of the Circle. We know that the Pythagorean Theorem is A^2 + B^2 = C^2 with respect to Right Triangles. We know that the General Equation to the Circle is (x-h)^2 + (y-k)^2 = r^2 where (x,y) is any point on the circle, (h,k) is the center, and r is the circle's radius.
Now let's see this in action. We are going to start from the left and work our way to the right. We are going to define our starting point as the origin (0,0) we will then move to the right 1 unit of arbitrary length to give us a Unit Vector at the coordinate (1,0). Now we are going to horizontally translate this vector in a linear progression towards the right by the magnitude of the original unit vector. This gives us a second unit vector that goes from the points (1,0) to (2,0). Here we have v1 = P1-P0 = (1,0) = (1,0) - (0,0) and v2 = P2-P1 = (2,0) - (1,0) where v1 = v2 = 1 and their combined length = 2. What does this have to do with a circle and the Pythagorean Theorem?
It's quite simple. Here we the basis for a unit circle with its center (h,k) located at the point (1,0). We can now use the equation (x-h)^2 + (y-k)^2 = r^2 and use the center location. (x-1)^2 + (y-0)^2 = 1^2. Rearrange and simplify to y^2 = 1 - (x-1)^2. Expand, y^2 = 1 - (x^2 + 2x) + 1 and simplify then gives y^2 = x^2 - 2x. Solving for y gives us y = sqrt(x^2 - 2x). Here is a sequence of some of the first few values of (x,y): { (0,0), (1,i), (2,0), (3, sqrt(3)), (4, 2*sqrt(2)), (5, sqrt(3)*sqrt(5)), ... } okay this doesn't seem to justify anything. Yet if you noticed, when we plug 1 into the expression we end up with i the sqrt(-1). Why is this happening? This is because the unit circle is located at (1,0) as it is translated to the right by 1 unit.
Let's move it to the left by 1 unit placing its center at the origin (0,0). Once we do this the Equation of the Unit Circle then becomes (x-0)^2 + (y-k)^2 = 1^2 which then becomes x^2 + y^2 = 1. Here this is a well known function and this is why the trigonometric functions have a Pythagorean Identity. Let's solve for y: y^2 = 1-x^2 then y = sqrt(1-x^2). We can now generate an equivalent sequence as before: { (0,1), (1,0), (2, i), (3, 2i), (4, 3i), ... (N, (N-1)i) }
This is also gives all Real and Complex Numbers and this is why we have the expression e^(PI*i) + 1 = 0 as well as the Taylor series and so on. Also if you look closely at the equation of the circle (x-h)^2 + (y-k)^2 = r^2 this is in fact The Pythagorean Theorem. We know that a circle with a radius of 1 has a diameter of 2 hence 1+1 = 2 and we know that the angle between the point (1,0) to (-1,0) is 180 degrees which equals PI radians.
And there you have, basic arithmetic{addition, multiplication, exponentiation, logarithms, etc.}, basic polynomial algebra, linear algebra with vectors & matrices, geometry and trigonometry all expanding from y = x and 1+1 = 2. And with all of these being related, we can extrapolate from them and their various functions and properties that gives us dy/dx the Derivative and with its inverse Integration. So the next time you add 1 and 1 to get 2 or the next time you set a and b equal to each other, remember that there is always a unit circle embedded within it waiting to be expanded and the limits of that expansion reaches +/-infinity!
But why Triangles, right angles to be specific? For this we need to jump over to a little bit of Physics. We typically use sine and cosine functions to map or graph wave motion, rotations, oscillations, cycles, etc... We know that both Sound and Light are forms of Energy. In essence the sine and cosine functions are basically Sound and Light waves in motion. We also use them to map or graph rotational energy and or angular momentum. So everytime you speak and sound emits from your mouth, your tongue, and your lips, the energy or vibration moving through the air (medium) is a Sine or Cosine Wave. And both of these functions are defined from both the Right Triangle and the Unit Circle again 1+1 = 2.
Numbers don't exist in nature as they are abstract concepts. They are a product of the mind. Yet the equations, formulas, functions, and laws of the sciences require and depend on them and nature adheres to the laws of physics and chemistry. The laws of nature obey Sound and Light, the properties of Circles and Right Triangles and things that wave, rotate and or oscillate. A clock goes tick tock with even intervals and can be represented as 1,0,1,0,1,0... and now we have a binary system. Log 2 mathematics. And it has been proven that Log 2 mathematics with an infinite amount of binary digits (bits) is infact Turing Complete. Math is Grand!
Math is a Product of the Mind and the Laws of Nature adhere to the properties of Math. Food for thought! So as you can see, The Pythagorean Theorem has always been long before any human thought that they have discovered it, long before any human existed. Since the properties of such things existed long before Humans, and considering that numbers are conceptual, abstract ideas, products of the mind, then the real question becomes: Whose mind did they originally come from? I can give a hit and it succeeds 1+1 = 2 which pertains to and rhymes with 3. The Divine Trinity! "Let There Be Light!"
1
-
1
-
1
-
1
-
1
-
That was a very nice, and simple, proof that the Kempner
series is finite. It occurred to me that the first part of the harmonic series might provide a better bound for the first part of the Kempner series. Ie, if Harm(a,b) means the harmonic series from a to b and Kemp(a,b) is similar for the Kempner series then:
Kemp(1,∞) = Kemp(1,10^k-1) + Kemp(10^k,∞) < Harm(1,10^k-1) +
Kemp(10^k,∞) < ln(10^k) + 1 + 80(9/10)^k = k*ln(10) + 1 + 80(9/10)^k
for all k. The RHS has a minimum when k = 12, giving a bound of about 51.2 on the Kempner series. .
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The most memorable part was guessing that Kepner's (spelling?) series had a finite sum even before the counting down started, and for much the same reason explained later. I recalled having encountered the idea that for any given digit, "almost all" natural numbers contain that digit, with the proof involving that as the number of digits increase, the proportion of numbers containing that digit tends to one. I think that idea may have been my first exposure, way back when, of the term "almost all" in a discussion of an infinite set. Anyway, in this case the complement is considered, and the fraction of numbers not containing the digit tends to zero. Not a proof, at least not without a whole lot more work, but definitely a way to see that finite summation is at least reasonable on its face.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
24:50 Looking at the shape of the individual slices, it is clear that when, instead of taking the "bow" shaped path form the bottom right to the top left corners, one closes the shape with a straight line, the resulting triangle has a smaller area (rigorously, this due the the fact that 1/x is convex, therefore a connecting line between ( n, 1/n ) and ( n+1, 1/(n+1) ) does not cross below the function). So the total blue area has to be smaller than the total area of such triangles.
Looking at the right side of the blue shapes, the total length of these sides add up to 1 (since 1/x -> 0 and the lengths are the differences between consecutive values of 1/n, starting at 1). And since the area of each triangle is one half times the length of it's right side, this means that the total area of these triangles is one half.
This proves the desired fact that the sum total of the areas of the blue slices is less than one half.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I learned about casting out 9s on my own reading math books as a kid, they never taught it in school. (Along with how to calculate square roots by hand and other interesting things.)
A few quick comments
- When making these sorts of diagrams, I think it's nice to, instead of using the numbers 0 through 8, use the numbers -4 through 4. (Or if you're working in something other than modulo 9, it goes from roughly -n/2 through n/2, give or take the parity). The reason is that if you number the diagram that way, with 0 at the top as usual, then you immediately see that 8 = -1 mod 9, 7 = -2 mod 9, 6 = -3 mod 9 and 5 = -4 mod 9, so the left side of the circle is just the negatives of the right side. But since ab mod 9 = (a mod 9) (b mod 9), you get exactly the same diagram as if you used 0 through 8.
However, now with those negatives, it's instantly obvious why the diagram is symmetrical about the axis - once you hit a negative number then multiplying it by 2 (or whatever the multiplier is) is going to give you exactly the same result as the corresponding positive number but with the sign flipped. So if you start at 1, make a couple of connections, then hit -1, the following connections are always just the mirrored versions of the ones you just drew. That's why all these diagrams, no matter what multiplier and modulus you use, are symmetrical about the y axis.
- Regarding the proof that "casting out nines" works, an alternative quick proof is to notice that 10 = 1 mod 9. So 2578 = 2*10³ + 5*10² + 7*10 + 8 , which modulo 9 is just 2*1 + 5*1 + 7*1 + 8 = 2+5+7+8.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
My favorite, fairly succinct proof of Fermat’s two-squares theorem goes a little like this:
(1): A whole number c can be written as the sum of two squares c=a^2+b^2 if and only if it can be written as (a+bi)*(a-bi), for integers a and b. Complex numbers a+bi with a and b both integers are called Gaussian integers.
(2): A number is called a Gaussian prime if its only Gaussian integer factors are itself and 1. Just as every integer can be written uniquely as a product of primes, a Gaussian integer can be written uniquely as a product of Gaussian primes. (I’ll reply to my own comment to prove this later this evening)
(3): If a prime number is not a Gaussian prime, it must factor into some (a+bi)*(a-bi), and so it must be a sum of two squares.
(4): For a prime p, if we can find some number m such that m^2+1=kp for some integer k, then we have (m+i)(m-i)=kp.
(5): But if we multiply p*(x+yi), we always get px+pyi. That means that p is a factor of a Gaussian integer c+di if and only if it is a factor of both c and d. But in (4), p is not a factor m, 1, OR -1, so p is not a factor of m+i or m-i
(6) This implies that, if we find m such that m^2+1=kp for any integer k, that means that p is not a Gaussian prime, and therefore can be written as the sum of two squares.
(7) Under what conditions can we find such an m? Well, we know that m^2 divided by p must give a remainder of (p-1), so we can answer this question by asking “What remainders can a square number give when divided by p?”
(8) Let’s assume we have some whole number r<p such that m^2/p never gives remainder r. For every number s<p, there must be some other number t<p such that st gives remainder r when divided by p. This follows from the fact that p is prime and the pigeonhole principle, since there are (p-1) whole numbers less than p to multiply s by, and (p-1) possible remainders, and if there were some t1 and t2 such that st1 had the same remainder as st2, then if we considered st1-st2, the remainders would cancel and we’d be left with a multiple of p, even though neither s nor t1-t2 is a multiple of p. Furthermore, t=/=s, since we know s^2 does not give a remainder of r.
(9)By Wilson’s theorem (which was proved on numberphile) (p-1)!+1 is divisible by p. But by (8), we can pair up the factors of (p-1)! Into (p-1)/2 pairs such that the product of each pair has remainder r. So r^((p-1)/2) gives the same remainder as (p-1)! when divided by p, that is, a remainder of p-1.
(10) We now consider p=4n+1 and r=p-1. In this case, r^((p-1)/2)= (p-1)^2n=((p-1)^2)^n. Since (p-1)^2=p^2-2p+1, the whole expression will give a remainder of 1 when divided by p.
(11) But by (9), if there were no m such that m^2 gave remainder p-1 when divided by p, the remainder of the whole expression in (10) would be p-1. So we conclude that there is such an m. But by (6), this means p is a sum of two squares. QED.
1
-
Now the proof that we really do get unique factorization in the Gaussian integers. It’s easiest to see the logic side-by-side with the logic that shows we get unique factorization in the integers.
(1): First, we show that there is a concept of “division with remainder” in the Gaussian integers. That is, for any two Gaussian integers z=a+bi and w=c+di, we have two more Gaussian integers q (the “quotient”) and r (the “remainder.) such that z=wq+r and such that the |r|<|w|.
(2): To show that this is true for the integers is easy. For two integers a and b, we can say q=a/b rounded to the nearest integer, and r=a-qb. Dividing both sides by b gives r/b=a/b-q. It’s easy to see that |r|<|b| if and only if |r/b|<1. And what is r/b? The distance from a/b to the nearest integer. But on the number line, the furthest you can be from an integer is ½. So r/b is certainly less than 1.
(3) Now we carry out a similar proof for the Gaussian integers. z/w is a point on the complex plane, so we can declare q to be the nearest (in terms of absolute value) Gaussian integer to z/w, and say r=z-wq. To show |r|<|w|, we need only show |r/w|<1. But |r/w| is just the distance from z/w to the nearest Gaussian integer. Since the Gaussian integers form the corners of a grid of 1-by-1 squares on the complex plane, and the furthest a point inside a 1-by-1 square can be from a corner is sqrt(2)/2<1, we can surely say |r/w|<1.
(4) We can now use the Euclidean algorithm,which is described on wikipedia, to find the least common divisor of w and z. Since this algorithm constructs the least common divisor through a series of steps which each involve multiplying the inputs w and z, as well as the outputs of the previous steps, by (Gaussian) integers and adding them together. If follows that the least common divisor of w and z can be expressed in the form LCD(w,z)=jz+kw, for (Gaussian) integers j and k.
(5) Using this fact, we can prove the following: if a (Gaussian) prime p divides a product of (Gaussian) integers a*b, it must divide either a or b.
(6) Assume p divides a*b, that is, ab=pq for some q, but p does not divide a. Since p is prime, this means LCD(p,a)=1. It follows that 1 can be expressed by 1=jp+ka for some Gaussian integers j and k. Multiplying both sides by b, we get b=jpb+kab=jpb+kpq=p(jb+kq). So p must divide b.
(7) Now let’s take c to be an arbitrary Gaussian integer. If it cannot be written as the product of two Gaussian integers of strictly smaller absolute value, then by definition it is a Gaussian prime. If it can, say, c=ab, then either one or both of them are Gaussian primes, or we can break down one or both of them further until we have expressed c as a product of Gaussian primes.
(8)But we can use (6) to show that we could not get a factorization that ends up having different primes. So factorization in the (Gaussian) integers is unique
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
So I take it (adding a few bits about fractals that I know) that when using rectangular corner-cutting to approximate a circle, you approximate the area of the circle, but the perimeter of your approximation is a fractal curve that has infinite right-angled discontinuities which are not present in the smooth circular curve. Fractal curves have "fractal dimension" of at least 1 - they can e.g. be area-filling curves that have an area rather than a length, or space-filling curves that have a volume. Fractal surfaces have fractal dimension of at least 2 and can be space-filling surfaces. Either way (and including all the cases with non-integer dimension), arguments about measures being preserved/increased/whatever as the approximation is improved break down when the measure (length or area) ceases to apply, so you can't then take those measures from fractal approximations and apply them to the non-fractal thing being approximated.
On the other hand when subdivision forms smaller and smaller in-between angles, in the limit you have infinitely many zero degree angles - the discontinuities have disappeared so that the infinite pieces (lines/triangles/whatever) form a smooth, continuous curve/surface.
There's a different non-convergence problem (Runge's phenomenon) which I find interesting when approximating smooth curves using high-order polynomial curves, or when approximating smooth surfaces using high-order polynomial surfaces. The normal solution is using piecewise curves/surfaces using low degree polynomials for each piece, though usually based on a single specification for the full curve (e.g. B-spline) rather than explicitly working out the polynomials for each piece. Converting into a connected sequence of simple polynomial curves (e.g. Bezier curves) is easy enough but unnecessary. Subdividing into linear B-splines into lines is the linear-polynomial-pieces special case that (like linear B-splines themselves) no-one uses. I mention this because I'm now wondering if the Runge's phenomenon convergence failure indicates another kind of fractal, with infinite extreme "oscillations" rather than infinite discontinuities.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Wonderful video as always!
9:20 Length 10:
Case 1: 8 copies of 1, a and b (2≤a≤b), 8+a+b=ab, (a-1)(b-1)=9, (a,b)=(2,10) or (4,4).
Case 2: 7 copies of 1, a, b and c (2≤a≤b≤c), 7+a+b+c=abc. c is not an integer when a=b=2. Thus ab≥6 and c≥3, then 7+3c≥7+a+b+c=abc≥6c, contradiction!
Case 3: 6 copies of 1, a, b, c and d (2≤a≤b≤c≤d), 6+4d≥6+a+b+c+d=abcd≥8d, contradiction!
By the same argument as in case 3, we know there are no solutions with fewer than 6 copies of 1.
More than 10 identities:
Consider the equation (a-1)(b-1)=2³·3²·5 i.e. 359+a+b=ab where 1≤a≤b, there are (3+1)(2+1)(1+1)/2=12 integer solutions. Length 361 is one solution.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
12:04: 153^2+104^2=185^2 is represented in the tree as 9 4 13 17 (read clockwise from top left.) From 1 1 2 3, you can get there by going right, right, right, left.
I got this by factorizing the first two numbers: 153=9*17, 104=2*4*13, testing that 9 4 13 17 fulfils the fibonacci formula and then working my way backwards by trying to complete 9 ? 4 ?, ? ? 4 9 or ? 4 ? 9, of which only the last one works (as 1 4 5 9), which corresponds to going left at the last step, and so on until I reached 1 1 2 3.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
First puzzle: just going off the pattern that the 1D starts with 1, 2 -> 3, and 2D goes 3,4 -> 5, I'd say 5^3 + 6^3 = 7^3. Obviously it doesn't (5^3 + 6^3 = 341, 7^3 = 343), it's off by 2. But that jives with what would happen if you took the 6 slices of the 6x6x6 cube, and tried to fit them around the 5x5x5 cube - you'd be missing two corner pieces!
Second puzzle: 10^2 + 11^2 + 12^2 = 100 + 121 + 144 = 365 (this one I know simply because I have those memorized). From the square pattern though, this also equals 13^2 + 14^2, so that's 2 * 365 in the numerator, the result is 2. Don't know of a quicker trick for this one though.
Final puzzle: I couldn't find any integer solutions for the extension past 3rd powers (3^3 + 4^3 + 5^3 = 6^3), so probably similar to how 5^3 + 6^3 ~= 7^3, I'll guess ~7 for the 4th powers, ~8 for the 5th powers, and so on.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
this is weird, because this is the only proof i knew about. it may not be exact proof, it was scaling sides, and then matching the two copies. i recall the "scale and match" part, since when someone showed me this proof, i had a small debate, how can you scale a side with another side? you could scale a side with a scalar, but how could you do with a measurement? the resolution, i set at that time in mind, is that you could either divide it by Unit length; or start measuring the length of the scaled version in Unit^2, where Unit was the original metric.
math is built on the shoulders of the giants. while this proof is beautiful visually, but i think there were shoulders of scale-n-match type proofs. of course, thanks for pointing out many of its applications and making a very beautiful educational video.
1
-
1
-
1
-
Hello Burkard,
Today swiping across Youtube, I saw that article where you asked why almost no one watched this video about logarithms the first weeks.
Earlier, I estimate some months ago, I watched several of the Mathologer videos, for example that one about the "do nothing grinder". And Youtube gave me recommendations for the other videos.
Then some time no more of the recommendations appeared.
As the information below told, all the the videos seemed to be released at least several years ago. I had already assumed, that you didn't produce new videos anymore.
In the meantime I watched lots of other math channels, as there are many Youtube members also producing content.
Sometimes I am not ready calculating an answer to a math puzzle, then there are already three new recommendations.
I don't know how the algorithm decides, which recommendations the users get, but perhaps channels you didn't choose recently, slide down in the list and are replaced by other similar contents, perhaps to avoid staying in a "bubble".
But I didn't forget you and Mathologer.
The videos are great 👍, even though I have to admit, that sometimes I'm not able to understand everything at once.
English to me (German) is a foreign language and the videos proceed very quickly.
And school didn't show some of the things, and school is in my case long time ago (I am 55). And I have not always as much time as I want, busy with work and so on.
But I don't want to miss Mathologer.
Greetings and best wishes!
1
-
1
-
1
-
1
-
Here's my slightly scuffed proof
Note that all the final lines are the same length because they're made up of red, blue, and green. Since they're all from the triangle, they all touch the inner circle of that triangle (inscribed? It's been a while since I've done Geometry), and are tangent to it.
Take one of these lines and imagine it is a chord of a circle. The circle that it is a chord of must have a center on the line that intersects at the midpoint, perpendicular to the chord. If you can show that the midpoint of the line makes a radius of the inscribed circle (I'm pretty sure that's true but I forgot how to prove it), then the center of the circle where that line is a chord could be the center of the inscribed circle.
Since each line is not parallel to the others, if they share a circle, it must have the same center as the inscribed circle.
Finally, note that you can calculate the radius by the inner circle and half the chord, since they're at right angles, Pythagoras can be used, and so you get that all the circles where the lines are chords and the center is the center of the smaller circle all have the same radius. Since the radius and centers are the same, the circles are the same
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The Moessner Miracle is the Golden Mean/Fibonacci Sequence in 3D with the potential to prove the existence of interdimensionality. 0 + 1 = 1, 1, 2, 3, 5, 8, 13, 21, 34, 55....INFINITY. Furthermore, it proves the existence of a FIRST CAUSE....THE LAW OF ONE within the VOID. The place of 'matter' in the void is that of an exquisitely beautiful precipitate established by the First Cause-Law of One through the unified field interface as determined by a geometrical necessity. (Restatement of John G. MacVicar) In other words, the universe is a holographic projection from an unquantifiable (in mechanistic terms) source, inadequately expressed (mechanistically) as ONE (1) IN CONTRADICTION TO NONE (0).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
If the largest angle between triangles approaches 0 or 180 depending on what angle you count the different planes or the angle at the intersection, and the surface approximated is differentiable and smooth and nice everywhere then it works. If the angles do not change, the area does not changes if the angle grows the area grows, if it shrinks then the area shrinks and approches the area of the shape. Talking about the angle between the planes the neighbour triangles are in, two triangles in the samenplane having 0 degrees between them, while the intersection spans 180 degrees, you cannspit the different definitions and how to convert the statement i am sure, this has to be true, because you can show it is so, take the initiam clue example meme, there you preserve the angle at 90 degrees, so the estimate will not change, of you clipped away edges such that angles keep getting sharper then the lenght will increase, and so on, andnthis generalizes to any dimension. :)
1
-
Amazing video as always, and although I am little late and I used Thales to solve it, I still got the solution.
The square is formed of various triangles, but we only need to pay attention to four of them, the one that gave us the units. we know that the hypotenuse of the biggest triangle is sqrt(5)/2, and that this hypotenuse is formed by segments, lets label them AB, BC and BD, and lets label the bottom of the triangle (a full side of the original square) FG and GD.
AB is a leg of ABF, which is a similar triangle to AFD, so we can extract that AB : 1/2 = 1/2 : sqrt(5)/2, and so AB = sqrt(5)/10. Then, we can extract that the BFD is similar to AFD and to GCD, so GCD is similar to AFD by transitivity, from this we can extract that sqrt(5)/2 : 1/2 = 1 : CD, and so CD = sqrt(5)/5.
Now, we know that AB + BC + CD = AD, and so, sqrt(5)/10 + BC + sqrt(5)/5 = sqrt(5)/2. Turning on the proverbial algebra autopilot, we get that BC = sqrt(5)/5. Returning to the original problem, we need to get AB squared, and so (sqrt(5)/5)² = 5/25 = 1/5.
Although solved via a different method than the one expected, we get the same result, but well, that is one of the many beauties of math I guess :).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
21:09 I try to avoid commenting before the end of the video, but (since I'm watching on a cycle of: watch 5 minutes, rewind 6 minutes) it might take a while to get to the end, here it is.
I noticed something that I found interesting and that hadn't been pointed out in the next 5 minutes (yes 'next': see rewind strategy above). But is probably explained later.
The coefficients of the top right diagonal's terms are 1, 2, 3, 4, 5 ... (adding one each time) which is 1x(1,2,3...)
Of the next diagonal down are: Minus 1, 3, 6, 10 (adding 2, 3, 4....) (haven't figured out how the even rows work, yet)
Next: Plus 1, 4, 10 (Adding 3, 6, which is 3x(1, 2...)
The next section is Pasqual's Triangle, which has to do with adding things, so I may find my answer later.
I could cook up a "pattern", differing on odd and even row, but I'd need to calculate more data. Probably I'm looking for something that either isn't there or is explained in the video.
Sorry for the ramble: so far, so great! Keep up the good work
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Top row: 1, 2, 3, 4, 5, 10th place, ... done?
Row below: 1, 2, 3, 4, 8, ... 24th place, thats it?Middle row: Powers of two on 1, 2, 3, 6, ... 91st place, done?
No more columns in open office ;-)
The only pattern i found was the doubling (5-10, 4-8, 3-6), but no idea, why that is so.
1
-
1
-
1
-
Here's my (relatively non-rigorous) go at proving that log-exponent property beyond the positive integers.
Log-exponent property for rationals:
----------------------------------------------------------
For any positive integer q, (1/q)A(x) added to itself q times is (q/q)A(x) = A(x).
Likewise, A(x^(1/q) * x^(1/q) * x^(1/q) * x^(1/q) ... ), where we multiply q times before applying A, is the same as A(x^(q/q)) = A(x). Therefore, (1/q)A(x) = A(x^(1/q)) for all positive integers q.
It's already known from the video that for any positive integer p, p*A(x) = A(x^p). Combining with our above proposition, we can conclude (p/q)A(x) = A(x^(p/q)) for any positive integers p and q, and thus this log-exponent property holds for the positive rationals.
For (positive) real numbers:
----------------------------------------------------------
By definition, any real number can be written as an infinite convergent sum of some arbitrary sequence of rational numbers (for our real number in question, r, let's call it q1, q2, q3...). We know that our log-exponent property holds for all the rationals, and so q_i * A(x) = A(x^q_i) at all indices i >= 1.
Consider A(x) * r = A(x) * lim(sum(q)). The log-exponent property holds at every finitely indexed step of this summation, so as our number of steps approaches infinity, the whole expression's limit must inevitably approach A(x^sum(q)) = A(x^r), i.e. A(x) * r = A(x^r) for any positive real number.
Probably not 100% adequate but I think it makes sense?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Lets label the block ((a,b),(d,c)) where (a,b) is the top row and (d,c) the bottom row. For the triplet 153, 104, 185, we need ad=153, 2bc=104, ac+bd=185. By definition of the Fib block, a+b=c and b+c=d. Then, our block is ((a,b),(a+2b,a+b)). We eventually find a=9, b=4, c=17, d=13. For the path, notice that the only possible parent is ((a,4),(9,c)). Then c=5 and a=1. The grandparent is ((1,b),(d,4)). So b=3 and d=7. Its parent is ((1,b),(d,3)), with b=2 and d=5. Finally, we reach the bottom, ((1,1),(3,2)). We then start there and go right three times, then left.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
BTW One formula (that I found myself) for generating infinitely many unique Pythagorean triplets is:
For any integer n > 0:
A = 2n+1
B = ((A^2)-1)/2
C = B+1
It outputs: [3,4,5], [5,12,13], [7,24,25], [9,40,41], and so on...
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
36:01 I thought about this one for a long time… And I came to a similar “sparseness” conclusion, though I wasn’t able to definitively prove that it converges
One thing I came up with is that for large N, the expected distance between the integers used for this series should be N^0.0458, which is somewhere between the 21st and 22nd root of N
(To be more precise, the exponent is exactly Log_10(10/9) = 0.04575749… Where the inverse of 0.9 shows up here, precisely because we are removing another 10% of the integers every time we approach a new power of 10
The way I’ve worded that might make it sound like I’m over counting by removing things that would already be removed… But it’s easier to think of it in terms of a weighted average of what has already happened. For instance, every 10 integers before 90 keeps exactly 9 integers, which is 90%. But then the 90s keep 0%, so we have 0.9(90%) + 0.1(0%) = 81% of all integers kept up to 100
Speaking of which… This whole pattern of removals reminds me very strongly of “negative space” fractals such as the Sierpinski Triangle, or the Cantor Set. I bet the fact that this series forms a fractal ultimately proves to be a huge factor in the reason why it converges, despite the fact that the “generation 0” series diverges
Edit: oh, you actually drew a representation of said fractal! That’s awesome. Thanks Tristan ❤)
My inspiration for this idea came from the prime number theorem, where we know the expected distance between primes is about log(N). As you mentioned earlier, the inverse prime series actually diverges, so I knew the integers for this Kempner series needed to be more sparse than the primes (if we want it to converge)
And of course, that turns out to be true! N^a grows faster than log(N) for all a > 0, even if it leads to something “very slow” like the 22nd root of N 😅
But unfortunately, I haven’t been able to see how I would prove that it actually converges. (I even went back and reviewed all the series convergence tests that I might have forgotten about, but nothing came to mind…)
I guess I’ll see how it works at the end of the video 🙂
Edit: Man, I totally could have come up with that proof! Kinda disappointed I didn’t figure it out, but oh well… That’s what I get for trying to do most of this in my head at the computer, without writing much down lol
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Had some fun with Python.
For the first puzzle, I used the formula 6*(1+2+3+...+n).
The first difference was 2. The second was 18. Third was 72...
Turns out the difference is always 2*(1+2+3+...+n)^2. I wonder where does that comes from?
Tried to do the same for the fourth power, and the difference is always 64*(1+2+3+...+n)^3. I don't know enough to figure out where do these come from. I sadly couldn't find a pattern for the fifth power. (2002, 162066, 2592552, 20002600, 101258850, .... Do these ring any bells?).
For the last puzzle, it seems that it goes 5,6,7,8,9... when you round the number, but the pattern breaks at power 39 where the sum rounds to the same number as power 38.
1
-
1
-
1
-
1
-
1
-
1
-
This is also how Charles Babbage's Difference Engine was supposed to work - the method of finite differences. I mean his first, limited design, not the more ambitious Analytical Engine, which was a general-purpose, Turing complete computer (or would have been, if it had ever actually been constructed).
Given the first few values of any polynomial of degree < 7*, it could calculate the rest - and therefore tabulate the polynomial's values as densely as one would care to. In practice, of course, it would not be used to make tables of polynomials, but more likely tables of transcendental functions that could be approximated by polynomials.
* 7 was just the "depth" it could go into, due to its mechanical design.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I was thinking more like "in any situation when from an eyelet there is more than one step the lace takes to reach another eyelet, you can rearrange the lacing in a way to put the lace just one or zero steps apart from the original eyelet. That closest eyelet now has three lace parts, rearrange one of them to be to that second eyelet we just took the lace from. Now the local lacing is the shortest for that eyelet. Take a random lacing. Let's rearrange the lacing to make it locally the shortest across the whole lacing. Let's start at the bottom, rearrange the lacing to make it so two bottom ones are connected, every other change happens above the part we have already, because there is no below, making it impossible to screw up the optimized part. Now rearrange the lacing to make to make it connect just one step above the lowest row, it naturally creates the cross lacing. Second row, if we try to connect them with step 0, we cannot connect all the other eyelets, thus another rearrangement to create cross lacing occurs naturally, and since no eyelet underneath is screwed up, the change only happens with the eyelets above, this will be true to every step. Thus no matter the amount of steps the cross lacing will naturally occur to locally optimize every eyelet. And since there is no way to create any line that isn't 1 or 0 step lines, but being shorter than 1 or 0 step lines, at all and specifically by creating longer lines elsewhere, which comes from the fact that eyelet columns are at an exact distance from each other (which doesn't necessarily mean that the solution would not be cross lacing for the inconsistent distance, but it means we don't have to think about those cases), if every eyelet is optimized, the whole system is optimized. The only way we can try optimizing the system to trouble individual eyelets, is to create more 0 step connections, which geometrically can be proven to only make the total length worse"
Or something like that
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Interestingly, when choosing U and V, the constraints on choosing the numbers have very little meaning. If you choose V<U, you still get a Pythagorean triple, you just reflect the triangle about the y axis. If you choose U and V such that they're both odd or both even, you get a primitive triple that is scaled by a factor of at least two. there seems to be a pattern for determining the scaling factor, it would seem to involve powers of 2 and powers of the greatest common factor, although I only tested with small numbers so it could be more complicated than that.
Still, interesting to know its possible to get EVERY triple, even those that aren't primitive, if you allow yourself to choose any U,V such that V>U. I wonder if that still only makes each triple show up exactly once?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
As soon as you pointed out that the number of vertices, edges, faces, etc. of an n-cube sum to 3^n, I had a total "aha moment", where I thought, "of course they do, just as a Rubik's cube has 1 mini-cube for each vertex, edge, face, and (hypothetically) an interior mini-cube for the 1 cell to make 27".
And then I paused the video and worked out a visual proof sketch (of the (x+2)^n coefficients counting the elements of an n-cube). I haven't kept records, but maybe once in 4 videos or so, something early will give me an "aha, I see where he's going" moment and that's always fun. And usually my aha moment pans out later in the video.
But this time you never went in my direction, so I'll present my proof concept (EDIT: Well, Tristan's proof is the same concept, but it treats the vertices and edges directly as algebraic objects instead of using n-volumes like my concept below. I thought my idea was missed because of the Euler tangent.)
If you have a 1+x+1 line segment partitioned into segments of length 1, x, and 1, you can raise it the nth power and get an n-cube with n-volume (1+x+1)^n that is partitioned into 3^n n-cuboids. Each vertex is adjacent to exactly one n-cuboid with n-volume 1^n*x^0=1. Each edge is adjacent to two vertex-cuboids, but also exactly one n-cuboid with n-volume 1^(n-1)*x^1=x. Each face is adjacent to vertex and edge n-cuboids, but also exactly one additional n-cuboid with n-volume x^2. And so forth. (And finally there is one with n-volume x^n in the center.)
It is clearer to start by illustrating for n=1,2,3:
https://imgur.com/a/VIajEwu
(1+x+1)^2 creates a square with 9 pieces, 4 of area 1 at each vertex, 4 of area x along each edge, and 1 of area x^2 in the center.
(1+x+1)^3 creates a cube with 27 pieces, 8 of volume 1 at each vertex, 12 of volume x along each edge, 6 of volume x^2 in the middle of each face, and 1 of volume x^3 in the center.
Now, I'm not an expert at turning visual proof sketches into proofs, but I think it's very pretty.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
My answer for 11:42 is the following
9, 4
17, 13
My logic? We have a function that takes four whole numbers as input a,b,c,d and outputs three whole numbers ac, 2bd, and ad+bc.
In this case ac = 153, bd = 104/2 = 52, and ad+bc = 185.
Next, we can determine the factors of ac & bd to guess & check possible solutions. ac = 153 = 3 x 51 or 9 x 17.
Setting a = 9 and c = 17,
my remaining equations become bd = 52 and 9d + 17b = 185.
You can either do algebraic substituion or continue guessing & checking using the factors of bd; I did the latter. bd = 52 = 13 x 4 or 26 x 2.
I set b = 4 and d = 13,
And all I had to do was check my final equation was true, 9*13+17*4 = 185, which it is!
----
Now I wish I could solve the next part but I ran out of steam.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
9:23 i hate this "paradox" and this is NOT a full rotation!!! From the rotating coin's perspective, it is currently upside down!!! Consider if you were a ball the size of earth, and you rolled around the equator of earth from 0 degrees to 0 degrees. You are inside the ball, and you start with your feet pointing at the ground, at 0 degrees on the equator. You start rolling around, and when you reach 90 degrees, your HEAD IS POINTING TO THE GROUND! You would never call this the "first rotation"!!!!!!!!!! Only when you reach 0 degrees on the equator again would you return to your starting position of feet down, making your SINGLE FULL ROTATION. The coin is exactly the same! From the coins frame of reference, it only makes one rotation! It is only when you place both coins in a different frame of reference and observe them arbitrarily from the side than it appears to be two rotations.
Really what you are asking with the coin "paradox" is if you rotate a coin about a point that is 2 times its radius away from its center such that the rate of rotation of the coin about its own center is twice that of the rotation rate about the point 2 times the radius, how many times will the coin rotate about its center per each rotation about the further point? This is obviously a stupid and trivial question because you are literally asking "if a coin rotates twice around its center for each time it rotates around an arbitrary point, how many times does the coin rotate for each time it rotates about its arbitrary point?" This is exactly the same as saying "if 2x = y, then what is y?"
In math we often toss out the trivial solution, or at least set it aside, but it often TELLS US NOTHING. The coin paradox is a case of this. The "it rotates twice" answer is not only wrong, its the trivial solution to an even more trivial question.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Amazing. I love all of your videos.
School maths generally depressed me for most of my life. Although the maths involved is great, the presentation of it was awful. For me, I felt they didn't spend long enough trying to be like Einstein:
"I never teach my pupils, I only provide the conditions in which they can learn."
It feels to me, schools just never answer children satisfactorily when they ask:
"What's the point of learning this stuff? It looks boring."
I think a perfectly natural response to a poor answer to this question is the very saddening things we hear: "I'm just not a maths person", "my brain doesn't work that way" and so many children get a bad relationship with maths from the beginning sadly.
Now Einstein's quote clearly went to an extreme and suggests he spent all his time only trying to give a strong answer to that oh so common question above. In reality, it would likely be split between teaching and inspiring but I'd say right now schools are 5% inspiration and 95% teaching (perspiration), when we really need to get to the exact reverse.
If we develop interest and show people where maths can take them and why it's fun, so much would be better in this world I feel, as a purely natural consequence.
I think many teachers in other fields are closer to the extreme of Einstein's quote and we need to get there with Maths too.
Thanks so much for your efforts, always love the little touches like the shirt also :)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
27:00 “Aumann and Maschler explain the Talmudic approach this way: they suggest that half of the total claim, representing half a container, serves as a psychological threshold between two perspectives. If a creditor receives less than half of their claim, they focus on the loss—seeing the situation as a complete loss with only a small portion salvaged. If a creditor receives more than half, they focus on what they recovered, perceiving it as a full repayment with a minor shortfall.”
It’s almost the reverse of counterfactual thinking in Olympic medal winners. The bronze medal winners are, supposedly, happier than the silver medal winners. If you win the bronze medal, you engage in downward counterfactual thinking, looking at all the competitors who won nothing at all, i.e., focusing on your win; if you get the silver, so the theory goes, you engage in upward counterfactual thinking, i.e., looking up at the gold medalist and thinking “That could have been me!” focusing on the loss.
The algorithm given by Presh Talwalkar in his 2008 blog post, “How Game Theory Solved a Religious Mystery,” (which is how I first learned of this distribution problem and its Talmudic solution), makes the logic clear:
(1) Order the creditors from lowest to highest claims.
(2) Divide the estate equally among all parties until the lowest creditor receives one half of the claim.
(3) Divide the estate equally among all parties except the lowest creditor until the next lowest creditor receives one half of the claim.
(4) Proceed until each creditor has reached one-half of the original claim.
(5) Now, work in reverse. Start giving the highest-claim money from the estate until the loss, the difference between the claim and the award, equals the loss for the next highest creditor.
(6) Then divide the estate equally among the highest creditors until the loss of the highest creditors equals the loss of the next highest.
(7) Continue until all money has been awarded.
That algorithm ensures that, to the extent possible, each creditor, from lowest to highest, gets at least half of the claim, and only after that is satisfied, the losses are equalized in absolute terms among the remaining creditors.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Many many years ago I read in a book by Ian Stewart ("Another Fine Math You've Got Me Into. . ") about the well known puzzle of crossing the river with a goat, a lion and a cabbage. If I remember correctly - or pause to think - this is isomorphic to Hanoi with three discs. Ian Stewart discusses Hanoi in that chapter and quotes Dr Andreas Hinz's wonderful article about the average path length of 3-peg n-disc Hanoi. I could not resist the temptation to try to prove the sqrt(17) infested formula and thought I discovered a mistake. I seem to remember there was actually a printing error in Stewart's book. Anyway, I spent many an evening with our 9-year old daughter on the balcony of our rented Swiss holiday chalet, enjoying the balmy evening weather and doing maths with a little girl, trying things out, sketching out attempts, thinking out loud, in short, something like real research mathematics. I wrote out my results and sent it to Dr Hinz who kindly sent me a reprint of his article with a friendly letter discussing my "paper" and gently pointing out one error of reasoning, in the form of a question, in the best Socratic fashion.
I will never forget the joy of doing real maths - never mind the fatal error. I learned so much. Best of all was the wonderful interaction with Dr Hinz and his generosity. I cannot help but feeling sorry for anyone who has not had the experience of immersing oneself in the beautiful world of mathematical ideas.
In my Covid-induced isolation I spend one hour each week with a high school student who wants to study mathematics. My aim is to cure her from high school maths and, after a fashion, show what maths is really about. After introducing her to the rationals and the notion of countable infinity, I showed her Cantor's diagonal proof of the uncountability of the reals. She exclaimed "But there are no holes between those fractions! Where do all those decimal numbers go????" Next, I will have to introduce her to Dedekind cuts and similar ideas. If I can convince her that the real line is very mysterious - and she needs at this stage little convincing - I believe she will be hooked on maths forever. Like me.
High school teachers who never showed the young minds in their class something like Euclid's proof of the infinity of the primes, should be lined up against the wall and shot 😎. How many people end up mathematophobes without ever having seen real maths? @Mathologer , you are one of my heroes - with Ian Stewart, 1B3B and the other soldiers fighting the good fight.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I like in depth mathematics topics. Not having formally done anything more complicated than differential equations, I often wonder about analysis, functionals, topology and a bunch of things I lack the adequate background to understand.
I still remember with amazement a linear algebra teacher generalizing linear span from vectors into polynomials to show how you could use this to represent a function. I often wonder (or maybe he explained and I don't remember) if you can generalize this, and have a set of infinite "vectors" (that represents polynomial coefficients) and with those generate "any" function.
In any case, I guess I miss the math that I didn't get the chance to learn in undergrad or gradschool. What worked and what not worked for me: I guess you did some jumps there in the end and there it required trust (he did the algebra, it probably works out); otherwise great work! If you are wondering what to show next, I guess I personally would like to know a bit about non-linear spaces and why it was hard for Einstein to come up with the general part of relativity theory; I am not particularly familiar with the difficulty of it and why, for instance it would be hard to adapt anything to a non-linear space, for instance, why is it hard to adapt quantum field theory to a non-euclidean space (in my head there is just some sort of jacobian somewhere and then you're done).
tl;dr. great video. Got to the end, didn't even notice it was 50 minutes long. Cheers
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
This is amazing in its simplicity and apparent obviousness... how it went so long unspotted (before the dot proof, of course... see what I did there?) is a testament to how the human brain works (or, often, doesn't work). My undergraduate degree is in cognitive science, and I never went to graduate school due to a combination of the effects of late-diagnosed neurodivergence (autism diagnosed in 2005, five years after graduating; ADHD diagnosed earlier this month) and shifting gears from academia to tech work the year after I graduated (which, again, was probably the ADHD talking). Mathematics, especially geometry and number theory, have always been fascinating to me, and your videos and those of 3Blue1Brown have done more to maintain that fascination than anything else. Thank you.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
13:57 I plugged the sequence 7, 15, 26, 42 into OEIS, and only got two results:
"[ n(n-1)(n-2)/8 ]" (sequence A011890)
and
"Row sums of triangular array T: T(j,1) = 1 for ((j-1) mod 8) < 4, else 0; T(j,k) = T(j-1,k-1) + T(j,k-1) for 2 <= k <= j" (sequence A131076)
No mention of Schwarz or infinite area in either one, which makes me wonder if you've found a new sequence.
1
-
31:48 "What does this formula say if you plug in a negative number?"
If I take the recurrence relation F(n) = F(n-1) + F(n-2), run it thru a change of variables n -> n+2, and rearrange terms to re-isolate F(n) on the left, I get F(n) = F(n+2) - F(n+1). Together with the base cases F(0) = 0, F(1) = 1, in this way I can iterate the sequence in the other direction. That is, F(-1) = F(1) - F(0) = 1, F(-2) = F(0) - F(-1) = -1, F(-3) = F(-1) - F(-2) = 2, F(-4) = -3, F(-5) = 5, and so on; where in general, F(-n) = (-1)ⁿ⁺¹ F(n). Substituting a negative integer for n in Binet's formula gives me the same result as iterating manually in this way.
31:53 "...or a number that's not an integer, like π, or φ (hehe)?"
Whoa, hold on there cowboy! Before I go trying to plug in craaazy stuff like F(π) or F(φ), I'm gonna first try plugging in an "easy" non-integer. Like maybe n = 1/2. I'll try F(1/2) at first; then maybe from that, I'll learn what I need to know to plug in F(π) or F(φ).
I'm gonna use the form of Binet's formula you alluded to at 32:11, by noting that -1/φ = 1-φ & plugging that in to simplify the denominator to √5. I'm also gonna change notation by n -> x, just because n is often an integer by convention & I wanna remember that now we're generalizing to real arguments of F. That gives me this formula: F(x) = [φˣ - (1-φ)ˣ] / √5. Plugging in x = 1/2, and noting that x^(1/2) = √x, I get F(1/2) = [√φ - √(1-φ)] / √5.
Wait a sec... 1-φ ≅ -0.618 < 0... now you got me taking the square roots of negative numbers!! If I take the principal square root √(-1) = i and factor it out from the √(1-φ), I can rearrange to F(1/2) = [√φ / √5] - [√(φ-1) / √5] i = √(1+√5)/√10 - √(√5-1)/√10 i ≅ 0.569 + 0.352i.
Backing up a bit now, what if instead of factoring out a √(-1) from the substituted formula, I factor out a (-1)ⁿ from the original formula, i.e. F(x) = φˣ/√5 - (-1)ˣ (φ-1)ˣ / √5? Now I can handle all the "complex" stuff in that (-1)ˣ bit, as everything else is real-valued.
If x = 0, I have (-1)ˣ = 1; when x = 1/2, I had (-1)ˣ = i; and when x = 1, of course (-1)ˣ = -1. This is like a unit vector rotating thru the complex unit circle... so to interpolate, I'm gonna use Euler's identity, e^(iπ) = -1. Then I get F(x) = φˣ/√5 - e^(iπx) (φ-1)ˣ / √5. If I wanna split this into its real & imaginary parts, I can use e^(ix) = cos(x) + i sin(x) ==> Re[F(x)] = [φˣ - (φ-1)ˣ cos(πx)] / √5, and Im[F(x)] = -[(φ-1)ˣ sin(πx)] / √5.
So to answer your question for x = π: Re[F(π)] = [φ^π - (φ-1)^π cos(π²)] / √5 ≅ 2.117, and Im[F(π)] = -[(φ-1)^π sin(π²)] / √5 ≅ 0.042. Thus F(π) ≅ 2.117 + 0.042i.
And for x = φ: Re[F(φ)] = [φ^φ - (φ-1)^φ cos(πφ)] / √5 ≅ 0.900, and Im[F(φ)] = -[(φ-1)^φ sin(πφ)] / √5 ≅ 0.191. Thus F(φ) ≅ 0.900 + 0.191i.
In addition to F(π) ≅ 2.117 + 0.042i, I also get F(π+1) ≅ 3.226 - 0.026i, and F(π+2) ≅ 5.343 + 0.016i. If I add F(π) + F(π+1), that gives me (2.117 + 3.226) + (0.042 - 0.026)i = 5.343 + 0.016i = F(π+2). Would you look at that -- the original Fibonacci recurrence relation continues to hold! :)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
As you pointed out, we now need to denumerate all tilted squares, and one way to do that is iterate over the rise and the run of the top line. We have already completed some of them! I'll use the notation (rise,run): how many. Note that run can be taken to be positive WLOG, but run must be negative and positive. Also, the sum of the absolute values of rise and run must be <=4, else it wont fit! We will also find, for reasons that become obvious when you try it, that the number that are present is only a function of |rise|+|run|. Also (-x,y) turns out to be the same square as (y,x), except the coordinates are now the right side of the square not the top.
(0,4):1
(0,3):4
(0,2):9
(0,1):16
(1,3):1
(-1,3=3,1):1
(1,2):4
(-1,2=2,1):4
(-2,2=2,2):1
(1,1):9
Total: 50
Now for a formula --
sum from (run=1) to n:
sum from (rise=0) to n-run:
(n+1-rise-run)^2=n^2+2n-2*n*rise-2*n*run-2*rise-2*run+1+rise^2+2*rise*run+run^2
Now completing the sum offline (totally didn't use wolfram alpha):
(1/12)*n*(n+1)^2(n+2)
And we can confirm that this gives 50 for n=4
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Dear Mathologer: First of all, thanks! What a beautiful & elegant tour de force to end a crazy year! Now, I wish to tell you a little story, & I hope you can answer a question for me. This has been bugging me for 35 years! I can't seem to find anything in the literature about it. Here goes... In the early to mid 1980s, I was playing around with an idea, & I developed a formula for it. I'm sure it's some standard thing, but it was neat for me to be able to find it. It's the formula for how many lines it takes to connect a given # of points - every one to every other one. 2 points takes 1 line; 3 points takes 3 lines; 4 points takes 6 lines, etc. The formula (of course) is: x= {n*(n-1)}/2, where n is the # of points, and x is the # of lines. Now, I wouldn't be asking you anything if that were the end of the story! Almost 20 years to the day later, I was playing an old video game called Qbert. You can easily look it up online & find out what it was like, but here's a quick description. The playing field is a triangular grid of solid 3D blocks, as though they're stacked up higher like bleachers, but toward a middle point (apex). I wondered if I could find a formula that would tell me how many blocks would occupy the face, based on the number of blocks along one side (all 3 sides being equal, of course). I found it, & to my utter astonishment, it's this: x={n*(n+1}/2), where n is the # of blocks along any one side, and x is the number of total blocks on the face. How can this be? I've never been able to explain how these are related! How can they be identical except for a +/- sign?!? I hope you can help me kick off the new year by finally solving this weird conundrum! Thanks again. tavi.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I wish I had a teacher like you when I was younger…
My brain refuses to understand maths of any kind beyond 1+1=2 😄 But I’m curious, so please bear with my stupidity.
This is cute and I’m sure very interesting for many, but for me it’s a bit boring because it’s unidimensional (I guess it is two- dimensional as in length and height, but it is unidimensional, really). You cannot squish and stretch in 2 dimensions (you see? I have the mind of a preschooler who plays with plasticine, so the thingie would expand towards me and on the opposite side, creating the volume, as in tri-dimensionality). Is there such a thing as multi- dimensional maths? To use your prop as an example, I imagine a cube made up of many identical cubes, each having its properties and functions. I imagine each little cube having a specific colour. The cubes interact with each other, so the colours would blend (would they necessarily? If yes, how?). How does the individual velocity of one or more of them spinning, or the spinning of the whole assembly, affect them and the whole (do they change shape or colour?). Do they remain cubes and if not, why not? If they turn into something else, does the big cube remain a cube? Can a cube be made of smaller non- cubical items? If the big cube insisted on remaining a cube (could it?), what would be the space left amongst the other non- cubical sub- structures? What if we don’t have a name for that space? Would that space have any colours, physical properties and mathematical functions? If yes, are those determined/ assigned by the other non- cubical sub- structures and/ or the big cube that still lingers in your mind because I mentioned it as a possibility?
I am probably not just stupid, but also crazy 🤪, but it’s fun to do thought experiments even though I don’t have the mental maths apparatus to describe and analyse it; it’s a very rewarding aesthetic experience.
More seriously now, is there anything such as multi- dimensional maths? There must be because I heard of multi- dimensional chess. Why do you have algebra, trigonometry, geometry, and physics as distinct? Does one necessitates the other? If so, why? It’s a bit of a series of circular arguments. And, last but not least, how does all this knowledge help in the ‘real world’, anyway?
PS. Do these cubes smell? In school I was as good (not) at chemistry as I was at maths 😄 but I remember something about molecules as in smell molecules. What shape is the smell of chocolate and why is it so delicious (especially if it’s hot chocolate with a few drops of rose oil)? Why is human nose of a shape different from a dog’s nose; it cannot be explained by the evolutionary theory. Why do snakes smell with their tongues? What is the function of smell? Why is the function of smell not a mathematical function? What are the disadvantages of defining people in terms of the function they (must) fulfil in society? Who says that the society is really a cube?
Sorry, I am really not trying to pretend I’m smart, but I’m curious if anyone has thought about these things. Thanks. 😊
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Did I miss something in Burkhard's statement of the challenge (at the end), or did he?
Because clearly, any integer that isn't congruent to 2 mod 4, can be written as a difference of two squares, simply by writing it as a product of two integers of the same parity (congruent mod 2), then using the difference-of-squares formula.
Wanting n = x² – y²
Write n = ab
Then x = ½(a + b), y = ½(a – b)
If n is odd, a & b will both be odd, and x and y will both be integers.
If n is a multiple of 4, n = 4k, then take a = 2k, b = 2, and again, x and y will both be integers.
Only when n has just a single factor of 2; i.e., is congruent to 2 mod 4; will it be impossible to write it as a difference of two squares.
The second part of the challenge works, however, because if n is (an odd) prime, then a=p, b=1 (or vice versa) must be true, and x = ½(p + 1), y = ½(p – 1) is the only possible solution.
[Swapping a & b only negates y, so the squares being 'differenced' are the same.]
The even examples for n:
0 = 0² – 0² = 1² – (-1)²
4 = 2² – 0²
8 = 3² – 1²
12 = 4² – 2²
16 = 5² – 3²
. . .
4k = (k+1)² – (k–1)²
Fred
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
6:05 First challenge: Remove the horizontal and vertical neighbors of a corner square leaving it disconnected from the rest of the board (and remove any other two squares except for the disconnected corner). Since no domino tiling can cover this disconnected corner square, there is no domino tiling of this board.
10:18 Second challenge: The formula spits out 0 for odd m,n, because in this case the upper index boundaries j=⌈m/2⌉=(m+1)/2 and k=⌈n/2⌉=(n+1)/2 yield jπ/(m+1)=π/2 and kπ/(n+1)=π/2, so both cos terms in the sum vanish simultaneously in one factor of the long product. Hence, the entire product vanishes.
13:50 Third challenge: The number of tilings for the (2,n)-board is the n-th fibonacci number Fn starting with F1=1 and F2=2. Proof: If we number each square with its columns number (or row number, if you consider the vertical board) the resulting adjacency matrix An has i's on its main diagonal and 1's on the two neighboring diagonals. Using Laplace expansion we see that its determinant Dn:=det(An) satisfies D(n+2)=i·D(n+1)-Dn for all n≥1. It follows inductively that Dn=i^n·Fn. Hence, after removing signs and i's from Dn the number of tilings for the (2,n)-board is |Dn|=Fn.
14:34 Unnumbered challenge: To count the number of tilings for the pair of glasses, we can break it up into smaller, simply connected pieces (i.e. pieces without holes) on which we can apply the determinant-method. Firstly, denote by H(row,col) or V(row,col) a horizontal or vertical domino with its left/upper square at coordinate (row,col) with coordinates starting with (1,1) at the top/left-most square (like a matrix of squares). Now to the fun part. The square (1,8) can only be tiled in two different ways, either H(1,7) or V(1,8). Similarly, the square (4,8) can only be tiled with H(4,7) or V(3,8). Together those four possibilities give us four groups of valid tilings. Call those four groups HH,HV,VH,VV (first letter representing the choice for the upper square, second character representing the choice for the lower square). Notice that there are no tilings for HV and VH, because trying to extend HV or VH to a valid tiling will always leave one square untiled. Therefore, all tilings belong to HH and VV. The same holds for the squares (1,11) and (4,11). Call their corresponding groups HH' and VV'. In summary, the set of all valid tilings is a disjoint union of tilings HH∩HH', HH∩VV', VV∩HH', VV∩VV'. By the board's symmetry HH∩VV' and VV∩HH' contain the same number of tilings, so we only need to count the number of tilings in HH∩HH', HH∩VV' and VV∩VV'. To count the number of tilings in any of those groups, we only need to remove the squares under the dominos on the squares (1,8),(4,8),(1,11),(4,11), apply the determinant-method to all resulting connected components and multiply them together. Here are the results:
HH∩HH' has 6*5*6=180 valid tilings,
HH∩VV' has 6*3*9=162 valid tilings, and
VV∩VV' has 9*2*9=162 valid tilings.
Summing everything up (remembering to count HH∩VV' twice) we get 666 valid tilings for the pair of glasses.
36:42 QUESTION: To me (and I am surely not the only one to notice this) it looks like the hexagon tilings (interpreted as 3d stacks of cubes) build a sphere at the origin together with three coordinate planes. Are there any theorems/papers on this as well?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Over last summer, I programmed this in processing (KA was faster than Repl): https://www.khanacademy.org/computer-programming/triangles-1d-cellular-automata/6678150943440896. It seems to correspond to rule [2,1,1,0,2] in the program. Hmmm...
EDIT: Found that this was one of the "notable rules", named "WHAT THE" in the comments. The other rule outlined later, with modular arithmetic, was actually what inspired me to make this program in the first place. The rule was [1,2,2,0,1], called "The original pascal" in the comments. Other rules can be used as well, but when I saw that thumbnail, I IMMEDIATELY opened up that program to see what would be discussed.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Most likely, somebody has already posted this, but I thought a little bit about the challenges raised at 32:20:
I start at with the fact (do I need to prove this?) that summing up the odd numbers delivers the sequence of square numbers. Since I cannot typeset formulas here, I try to write it this way:
S(n) = sum(2*i-1, i=1..n) = n^2. And S(n-1) = sum(2*i-1, i=1..(n-1)) = (n-1)^2.
The difference between these two sums is just the last summand of S(n):
S(n) - S(n-1) = (2*n-1) = n^2 - (n-1)^2
This works also for a subset of even numbers, because S(n) - S(m) is even if (n-m) is even, e.g.:
S(n) - S(n-2) = ... = 4*(n-1) = n^2 - (n-2)^2
Obviously, not all even numbers can be constructed this way, they always jump in increments of 4.
We can also get alternative representations (n^2 - m^2) for some odd numbers if (n-m) is odd and greater than 1.
Regarding the uniqueness for prime numbers:
p = n^2 - m^2 = (n+m)*(n-m). If (n-m) was anything else than 1, p would be composite and not prime.
So there must be: (n-m) = 1 and (n+m) = p. n and m must differ by exactly one, so only the difference of two consecutive sums S(n) - S(n-1) can deliver a prime.
Finally we get: n = (p+1)/2 and m = (p-1)/2
1
-
I've used, well rarely now, a slide rule for 60 years. My dad taught me how to use one and how it worked when I was in 3rd or 4th grade. Upon seeing the answer to 13 times 13 my automatic reflexive action was to say "THAT'S WRONG!" However, upon closer inspection the 13 mark was on 169. So what hit me so hard as to evoke such a knee jerk reaction? I grabbed the last slide rule I ever bought for academic work (In 1968 when I was 15 years old) and stared at it intensely. The scale setup on it is identical to that shown @9:21. I think the P scale (Pythagorean) was its main attraction for me at the time. Now I understand what caught my instinctive eye. There is no where on any slide rule in my collection (I have more than 10 now), in any place, on any scale, that has an intermediate graduation (a short line between two long lines) that doesn't represent half as 0.5, i.e. having long line 168, short line 169, and long line 170 would NEVER be used in any 'real' slide rule. Yeah, it's half way, but it is not 5, 0.5, 0.05, etc.. The 'rule for the rule', without exception, seems to be that all graduations between graduations represent 0.5 (for 1 short graduation between 2 long ones), 1 (for 9 short* graduations between 2 annunciated marks), and 2 (for 4 short graduations between 2 long ones). *Here too 0.5 or the 5th mark is lengthened. Between annunciated marks the '5 position' is always longer, e.g. the mark between "6" & "7" or 6.5 is lengthened on the C & D scales. This was all just instinct until I analyzed it and wrote it down. Thank you!
There's a secondary lesson that's also 'taught' by familiarity with the slide rule's layout. I find that virtually none of the engineers I've worked with over the years realize that a constant distance on a logarithmic plot represents a constant percentage or factor. This simple, easily understood concept, always seems to be a surprise. In recent years I've never heard someone say, "Yeah, I knew that."
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Really liked the visualisation of the no9. Simple and effective. I love it when things become less abstract and visible like this.
Im not competing for the book. You can make other people much happier with that.
And i really wanted to say something else.
This was YT advised to me and i watched all of it. Paused a few times to process it.
But, a German mathematician, speaking english is in itself, interesting. Then you would probably expect someone in a suit (why?) and you are not. You would expect someone less masculine (why?), but you seem pretty masculine. And then you have this laugh/giggle and that caught me most i think.
I am wondering now how we get our laugh. Our natural laugh. When i was an adolescent i remember i was told at some point that i had this particular laugh. At that time i copied it from others i think. It had changed and wasnt my natural manner of laughing. Since then it changed back to natural and on several occasions i was mocked in some mild manner bc i had a somewhat similar giggle as you have. I didnt really understand why. Also, it didnt matter to me really.
Then some 10 years ago in a House MD episode there was some chap that had an odd laugh like that (followed by a snore) and it became clear throug the dialogue and such that this was not appoved of. It wasnt considered masculine or even mature. That was the message.
But it makes no sense. Why would a laugh say anything about a persons masculinity or even adulthood? I couldnt place it and it seemed silly to me. Like how women nowadays disapprove of white socks. Or if a man crosses his legs or something. To me it seems a natural reaction to if you are hot or cold. Not to if you are a man.
So seeing you now with this same giggle i have, i must say; Sir, i approve of your giggle! It is real, it is honest and i love it ;]
(ok, has nothing to do with mathematics, or maybe it does, but i wouldnt know how). As you can see, i am less into math and more into psych. I will now be pondering for a week why we laugh like we do. I will also make sure to write 'Euler' correct in the future.
I had quite some of your questions right! But i realise there was a 50% chance in most cases to get it right, without overhang. So it was more luck than wisdom, i guess ;]
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The human brain is pretty much incapable of visualizing 4 dimensional space and objects. Trying to visualize even a simple 4D cube is pretty much impossible.
However, I once realized a method by which visualizing, or at least understanding it, becomes slightly easier. The method is like this:
Let's start with the "one-dimensional unit cube", ie. a line segment of length 1. (This method works even if starting with the "zero-dimensional cube", ie. a point, but let's start easier with the one-dimensional cube.)
Now, consider this 1-dimensional cube to be on the X axis. In order to get from here to a 2-dimensional unit cube (ie. a unit square), duplicate the 1-dimensional cube and translate it by 1 unit in the second dimension, ie. the Y axis. While doing this, keep the corresponding endpoints connected with straight lines. When you do this, you get a 2-dimensional unit cube.
In order to get from the 2-dimensional unit cube to the 3-dimensional one, we do the same trick: We duplicate the 2-dimensional cube and translate it 1 unit in the third dimension (ie. the Z axis), while keeping the corresponding corner points connected with straight lines. This way we get a 3-dimensional unit cube.
Obviously, the same trick works to get a 4-dimensional cube. And it is precisely this trick that helps one understand it a bit better: Take the 3-dimensional unit cube, duplicate it, and translate it by one unit in the fourth dimension, while keeping the corresponding corner points connected with straight lines. You get a 4-dimensional unit cube.
You don't need to precisely visualize this operation. Just think of it as moving the copy of the 3-dimensional cube in a direction that's not any of the three dimensions.
This trick helps you, among other things, deduce the number of corners, vertices, faces and 3D cubes in the 4D cube, and helps you understand where the multiple 3D cubes come from.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
My python imlementation for working out the number of ways to make change:
*this works in cents
coins = [1,5,10,25,50,100]
maximum = 1000
ways = []
for i in range(0,len(coins)):
for x in range(0,maximum):
way = 0
if(x + 1 == coins[i]): # We can +1 if our index is our coin denomination. Only happens 1 time so probably better to not repeat this.
way += 1
if(0 <= x - coins[i] < len(ways)): # Now see if we can make the remaining coins with lower denominations of coins.
way += ways[x-coins[i]]
if(i > 0): # Ways already populated
ways[x] += way
else:
ways.append(way)
print(ways[-1])
For reference:
# 1 / 1
# 1 1, 2 / 2
# 1 1 1, 2 1 / 2
# 1 1 1 1, 2 1 1, 2 2 / 3
# 1 1 1 1 1, 2 1 1 1, 2 2 1 / 3
# 1 1 1 1 1 1, 2 1 1 1 1, 2 2 1 1, 2 2 2 / 4
# 1 1 1 1 1 1 1, 2 1 1 1 1 1, 2 2 1 1 1, 2 2 2 1 / 4
# 1 1 1 1 1 1 1 1, 2 1 1 1 1 1 1, 2 2 1 1 1 1, 2 2 2 1 1, 2 2 2 2 2 / 5
# 1 / 1 [1,0,0]
# 1 1, 2 / 2 [1,1,0]
# 1 1 1, 2 1, 3 / 3 [1,1,1]
# 1 1 1 1, 2 1 1, 2 2, 3 1 / 4 [1,2,1]
# 1 1 1 1 1, 2 1 1 1, 2 2 1, 3 1 1, 3 2 / 5 [1,2,2]
I looked at the patterns with these small cases with coins = [1,2] and coins = [1,2,3] respectively and remembered a similar problem to do with counting number of ways to climb steps if you can make 1 or 2 steps at a time and using dynamic programming. I realised I could introduce the coins in ascending order and then use previous answers to see if we can make the remainder when introducing a new coin.
P.s. im tired so I haven't written my explanation up very well but hopefully someone can understand the code :)
Runtime: O(n) if number of coins is constant. I haven't really looked for more efficient ways as this was what popped to mind first.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Computer science degree here. In CS, my math continued away from advanced calculus and more toward combinatorics, analytical topology, automata theory, etc. But I think I understood everything here. I might not have remembered enough calculus to do all the stuff you did, but I understood it, and could see how it worked out in the end. That's the nice thing about mathematics, which I tell my own students: Numbers behave themselves. If you ignore one part of an equation and come back to it later, it will sit there patiently waiting for you. It's not going to change or do anything weird. And if you can see how a part of your function is going to quickly get infinitesimally smaller, then you can ignore it as being close enough to zero, that you don't need to worry about it for a good-enough approximation.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
As for as the code challenge:I got the change of base (5), easy, will implement hex notation and base decimals if you wish.
Got it so you can choose what number is on top on the outer ring and inner ring, so wish list (1).
Wish list (2), zoom for "infinite precision" has been the greatest challenge. I do have the zoom feature implemented. But getting into javascript precision cruft, getting 2.01000000000003 for 2.01. Considering the logs and trig functions involved as well as multiplying to zoom in, that is asking for that. I will work on removing it from the finished product.
I opted to skip the labels and tick marks that fall off screen so as to save memory and processing, I should also apply that to the circle that falls off screen.
I implemented the squaring. Presently you tell it what number you would like squared and it places that number on the top for the inner ring and on the out right top the mod(inner_top^2,base).
By change of base did you intent it to happen on both inner and outer circle, which is what I did, or individually;example: outer ring is base ten, inner ring is base 2. I could also work in hexadecimal notation if you would like.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Requested Feedback:
As always your animations make things generally quite easy to follow, though a personal pause every so often doesn't go astray.
Similarly the finer points only come when working through the full calculations, but the overview is generally perfectly sufficient.
I teach high end high school math and physics at a low SES school in Victoria, Australia. I have a BSc maj Math & Physics, plus the original BComm maj Eco, Fin, Intl Trade.
I watch all your stuff, and periodically find a way I can use it in class, but mostly I just highlight you and others exist to the properly interested so they can follow the rabbit hole themselves.
Commonly in class they require a few more gaps filled in a bit more explicitly, but that can be for all sorts of reasons, none of which is associated with the delivery, just their (mis)thinking.
Continue your great work, it has been greatly appreciated for some years now.
1
-
1
-
1
-
1
-
1
-
1
-
I managed to understand the factorial part using similar method: the first time when we move to the "left block" we have (1,1,1,1,1) to (1,2,4,8 16) which is the same as the power proof, But when we move to the left block the second time, since the height decreases by one so we are left with (2,4,8,16). Writing it as 2x(1,2,4,8) and it becomes 2x(1,3,9,27). Now the third time, we start with 2x(3,9,27) which is 2x3x(1,3,9) therefore it becomes 2x3x(1,4,16). The last time we start with 2x3x(4,16)=2x3x4x(1,4) and end with 2x3x4x(1,5). This directly ends with 2x3x4x5 which is the factorial number.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
A beautiful application of these diagrams was noticed by minimalist composer Tom Johnson: If you mark the numbers from 0 to m-1 according to the smallest number in their „orbit“, i.e. the set of numbers they’re connected to by multiplication lines, you obtain a sequence which is equal to the sequence of „every other term“!
For example, for m = 15, the number 1 is connected to 2, 4, 8; 3 is connected to 6, 12, 9; 5 is connected to 10; 7 is connected to 14, 13, and 11 (and of course 0 isn‘t connected to anything). The sequence we get is 011315371357377.
Now, I encourage you to convince yourself that writing this twice and taking every other term yields precisely this. But where would this occur? Well, listen to his „Rational Melody XV“ and find the sequence: https://youtu.be/Nk47dXQeunY
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Simple proof that the only p-special numbers are those of the form n = p^k+1:
First note that if n != 2 is special then we must have n-1 divisible by p. Indeed, suppose we fill the top row with zeros except for a single one in the second spot. Since n != 2, the corners are both zero, so since n is assumed special the bottom hex should also be zero. But we already know that the entries in this triangle are just a slice of Pascal's triangle mod p, so it's easy to see that the bottom hex will be n-1 mod p.
On the other hand, if pm+1 is special then I claim m+1 must also be special. In fact this is an iff statement. Indeed, since p+1 is special this just follows from the same sort of argument as in 10:55, and it's easy to see that this argument works in both directions.
The proof finishes by infinite descent: let n be the smallest special number such that n != p^k + 1 for any k. Then n != 2 since 2 = p^0 + 1. By the first paragraph we can write n = pm + 1 for some m. By the second paragraph, m + 1 is also special, and by minimality of n we must have m = p^k for some k. But then n = p^(k+1) + 1, a contradiction.
1
-
1
-
1
-
1
-
1
-
1
-
I have a question about the pi calculation. I’m not a mathematician (oddly, I do like watching videos on the subject though), so this may be a bit of a “noob” style question.
So, let’s assume I know pi to 5 decimal places. I don’t know the number after that. Using your method where you go past pi and the subtract back, how would I know if I had passed pi to 6 decimal places? I know I could assume that once we hit a decimal place, that we could then keep adding until we get to 4, but I expect that’s wrong (I’d still need to know pi to 6 places when I do the subtracting anyway). The logic of this is that you need to know pi to calculate pi, and if you do, why bother? That seems dreadfully wrong.
I also guess that there is a way to do this, and that the answer will break my brain, but don’t let that hold you back :-)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
There’s a more familiar example of inconclusive limits, but with the variables converging to a finite value, rather than infinity.
This example relates to a value for 0⁰.
If we consider x^y (x to the power y) as both x -> 0 and y -> 0, the limiting value depends on how you make x and y converge to 0.
If you let x-> 0 a lot faster than y -> 0 then the limit is 0. For example x = 2⁻ⁿ, y = 1/sqrt(n).
You can also make the limit 1. For example x = 2⁻ⁿ, y = 1/n².
Or x = 2⁻ⁿ, y = -1/n gives limit 2.
There can’t be multiple values for this limit. So we say that the limit does not exist, meaning that 0⁰ does not exist.
Sometimes we give it special status as “indeterminate” - within certain formal working it can behave as if it has a value. This requires care.
The obvious instance of this: it can be convenient to formally assign x⁰=1, no matter what x is, such as with polynomials and power series. What makes this ok is that the x^y limit has y=0 always, since the xⁿ expressions in polynomials are only for integer n.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
6:20
Using the method for two holes we can unravel the board to a loop of alternating colour.
Now, if we put the four holes in we really only have two ordering options:
Black, white, black, white and repeat,
Or black, black, white, white and repeat.
All other configurations are equivalent to one of those, since its on a loop.
In both cases we can utilize the interfaces between the white and black to order the tiles according to the two method.
In both configurations there are at least two interfaces, and therefore, the job can be done.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
13:20 I got up to 2*10^7 cents in ~10 minutes with a special implementation of the polynomial products, abusing their simple patterns with a divide and conquer strategy (anyone let me know if you want a detailed explanation of the optimizations I used and/or my code with inline comments), using a pretty modest computer setup (Intel Atom CPU N2600 1.60GHz without any use of parallel computing with the GPU as far as I know). Probably the next number in the sequence will take over an hour. However I don't have even enough RAM memory to do it because a vector of 2*10^8 of 128 bit integers (2^128~3.4*10^38 max number before overflow) takes ~3GB of space and using the disk memory will make everything slower.
Here's the sequence (2*10^n dollars):
2728 53995291 4371565890901 427707562988709001 42677067562698867090001 4266770667562669886670900001
There seems to be emerging a pattern right there. (^.^)
Here's my c++ noodle code (it uses a boost non-standard library for the int128_t "large integer" type) for any amount of cents:
#include <iostream>
#include <boost/multiprecision/cpp_int.hpp>
using namespace std;
using namespace boost::multiprecision;
void add(vector<int128_t>&p,int s){
vector<int128_t>aux(p.size()-s);
for(size_t i=0;i<aux.size();i++)aux[i]=p[i];
for(size_t i=0;i<aux.size();i++)p[i+s]+=aux[i];
}
void shiftAdd(vector<int128_t>&p,int f,int s){
if(f==1)return;
div_t d=div(f,2);
if(d.rem){
vector<int128_t>aux(p.size()-(f-1)*s);
for(size_t i=0;i<aux.size();i++)aux[i]=p[i];
shiftAdd(p,d.quot,s);
add(p,d.quot*s);
for(size_t i=0;i<aux.size();i++)p[i+(f-1)*s]+=aux[i];
}else{
shiftAdd(p,d.quot,s);
add(p,d.quot*s);
}
}
void shiftAdd(vector<int128_t>&p,int s){
div_t d=div(p.size(),s);
if(d.rem)shiftAdd(p,d.quot+1,s);
else shiftAdd(p,d.quot,s);
}
int main(int argc, char *argv[]) {
size_t n;
while(cin>>n){
vector<int128_t>p(n/5+1,1);
shiftAdd(p,1);
shiftAdd(p,2);
shiftAdd(p,5);
shiftAdd(p,10);
shiftAdd(p,20);
cout<<p[n/5]<<endl;
}
return 0;
}
(^.^)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
your question for viewers at the end of chapter 1:
if you have coin denominations representing powers of 2 ranging from 0 to n, and one coin of each denomination, you can make change of any amount between 0 and 2^(n+1)-1. for example, with the denominations 1, 2, 4, and 8, that can be represented as 2^0 to 2^3, which means n=3. Therefore, the maximum number you can create is 2^(3+1)-1 = 2^4-1 = 16-1 = 15. As for how many ways you could reach each value, it's only 1. Every coin is is valued more than the combined sum of all lesser coins, which means no combination of coins can be exactly matched in value by any other combination.
essentially, you've created a binary counter.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I've always liked math. I mastered algebra and trig, and used it all the time as a machinist and builder. I took analytical geometry, at 22 years old, and learned it pretty easily. Then came Calculus, at 48 years old. My Professor probably wasn't born when I was learning and graphing functions. Now, I was offering to paint her house in exchange for a "D." She said my algebra is really, really, bad. "I know what 2q + 2q equals" I almost said, but wanted that D really badly, and to give Calculus another try.
Then I found Mathologer! I'm hooked. My housemates who would rather watch junkyard videos, roll their eyes, and complain "He's watching the 'Math Nazi' again!" I defend you, and tell them you're a nice Nazi, like the ones on Hogan's Hero's. I apologize for them, Burkard.
I'd like to request a video that would help us lower level mathologers "brush up" on our Algebra and Trig so we'll be more apt to understand Calculus. It would be a great addition to the "Why is Calculus so Easy" video. I wish I had watched that before I attempted that Calc 1 class.
1
-
1
-
1
-
Huh, so it seems to me the "ear erecting" procedure must obey two symmetries: the proper angle magnitudes corresponding to whichever polygon is chosen, and the chosen parity when deciding how to draw the ears. To attain the symmetry of a perfect polygon, all parities must somehow cancel each other out—after all, rotational symmetry doesn't care which direction a polygon is turned, since angles are conserved and any symmetries will appear regardless. Therefore, I think it makes more sense to think of the "degenerate" pentagons as 0D pentagons, which lack parity because it's meaningless to assign a direction to a single point. While I'm no mathematician, I suspect this what allows the special point work in this model. The "center of mass" exudes its magnitude, represented by the graph of the "degenerate/0D" pentagon, equally on all vertices of the perfect pentagon, but this is a matter of translational symmetry of the Cartesian plane; directionality is not a factor. As such, assigning any sort of parity to the "degenerate" pentagon makes no sense, and the only type of space in which forbidding parity makes sense is 0D—hence a 0D pentagon, that is infinitesimally small with overlapping vertices by inevitably because it exists in zero dimensional space.
This could all be complete nonsense, but I'm still curious of what happens if you expanded Petr's miracle to include higher dimensions. For instance, could you get rid of the "degenerate" pentagon by drawing the pentagons on the surface of a Klein bottle, or some other 2D manifold in a 4D space, and using the additional dimension to draw two more pentagons with opposing parities to reproduce the same effect as the "degenerate" pentagon? I truly have no clue and would be delighted if anyone could provide some insight!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1) 1st challenge - no
2) 2nd challenge - fibonacci sequence
3) |det(tristan glasses)| = #tilings = 666
4) moster formula with odd m and n
5) Gray = Orange = Blue (triangular grid)
Solutions
1) take 4 to make corner alone
2) on 2×n rectangle (suppose has F(n) tilings) there are two ways to fill upper-left corner: vertical domino, so ways to fill the remainings are F(n-1); and horizontally, lower-left is forced to be filled with horizontal domino, and the remainings are filled by F(n-2) ways. So F(n)=F(n-1)+F(n-2), F(1)=F(0)=1
3) I didn't find any online det calculator, cuz they are too scary to calculate 22×22 det, which means 10^21 elementary operations. Of course our matrice has a lot of zeros, so i stole python code, added if-clause for non-zero elements, and it spat out 666. Then I manually calculated and also got 666.
4) when m=2p-1, n=2q-1 the last term (j=ceiling[m/2]=p, k=ceiling[n/2]=q) equals 4cos^2(pi/2)+4cos^2(pi/2)=0. And because of multiplication all the expression magically becomes 0
5) the hexagon with sides equal A looks like the box with cubes, where all upper faces are Orange, left are Gray, right are Blue. So let's look to the top and we'll see A×A Orange square (there are A×A orange dominos). Analogically looking from left and right we derive that there A×A Gray and Blue dominos
P.S. I like all the parts of the video, but the most favorite is Aztec square with Arctic circle. Thank you for you videos =)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
For the square pattern, I noticed that at the end of row 1, it's the square of 5. So the beginning of row 2 will be the row number 2 * the last square 5 = 10. The end of the row is 14 - divide by the row number to get 7, and then multiply that 7 by the next row number 3 to get 21. The end of row 3 is 27, divide by row 3 to get 9, multiply by next row number to get 36. The end of that row is 44, divide by row 4 to get 11, multiply by next row number to get 55. The result of the divide always goes up by 2, and there are always a row number worth of numbers on the RHS and row+1 on the LHS. Row 5 would end with 65^2, divide by row 5 to get 13, and so on.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The sum of powers was a bit disappointing to work on. The first 3 sums (sum of k^0, sum of k^1, sum of k^2) have this nice pattern: n, n * (n+1)/2, n * (n+1)/2 * (2n+1)/3.
It makes you think, hmm, what pattern is this, and does it hold for S_3 (as you called it)? Could it be that S_n = S_(n-1) * (the next term)? It seems to be doing that. And, if you grasp for patterns, it does look like each term is of the form (F_p * n + F_(p-1)) / p (with F being the Fibonacci numbers).
S_0 = (F_(0+1) * n + F_0)/(0+1) = (F_1 * n + F_0) / 1 = 1 * n + 0 = n;
S_1 = S_0 * (F_(1+1) * n + F_1) / (1+1) = n * (F_2 * n + 1) / 2 = n * (1n + 1)/2 = n(n+1)/2;
S_2 = S_1 * (F_(2+1) * n + F_2) / (2+1) = n(n+1)/2 * (F_3 * n + 1) / 3 = n(n+1)(2n + 1)/6;
But...
S_3 promptly breaks what was looking to be a nice pattern, being not equal to S_2 * (F_(3+1) * n + F_3) / (3+1)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
If Aₙ is the n'th length in the circle,
A₀=0, A₁=1, ..., then the product of two of these can be calculated as
Aₙ₊₁Aₘ₊₁=Aₘ₋ₙ₊₁+Aₘ₋ₙ₊₃+...+Aₘ₊ₙ₊₁
Note that this sequence has to become negative for the identity to hold.
This makes all the diagonals look really nice in the multiplication table, since two diagonally adjacent multiplications only differ by a single factor (which also follows from Ptolemy's theorem, and this implies the general result because it holds if n=1).
In a pentagon:
φ=A₂=A₃,
A₂A₂=A₁+A₃, so
φ²=φ+1.
The only other identities involve linear dependence, like
A₂+1=A₄ in a nonagon, although you can see this because the expansions of
A₃A₄=A₂+A₃+A₄ is equal to the sum of the expansions of A₃A₁=A₃ and A₃A₂=A₂+A₄.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
This method is dependant on the decision of how to split the "parties".
Consider, for example, the case of the babies. There, we made an arbitrary decision to split the babies into two parties, one with the two babies and one with the red baby. This gives one result. On the other hand, one could argue that, since they are separated individuals, each of the two green babies should be considered as one "party", thus totalling three parties claiming parts of the heritage. However, if we split the parties this way, the result changes.
That is concerning, as it allows ways for the results to be manipulated by a partisan judge, for example. Note this doesn't happen with proportional sharing.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I did make it to the end. 50 min video is ok for me. Anything below 1 hour works. My background in math... Well, I have a masters in theoretical physics and participated in math olympiads when I was at school, so math was my first passion and I learned a fair bit of extra stuff. Sometimes I feel sad for not being able to share things I learned and it wouldn't make much sense anyway with the current state of technology - too slow... I can only share a tiny fraction of what I know, yet I know only a tiny fraction of what I want to know. Sometimes I feel sad when I forget something I once have known, and when that something isn't easily available for me to refresh. I go to wiki and can't find derivation of Webber functions, can't find methods for solving quasi periodic differential equations, can't find some non-linear mesoscopic physics topics. Hardly anyone makes a good video on quantum physics and general relativity, which are much better understood these days than some think. Quantum field theory and string theory sometimes seen as non-existent... Riemann zeta pops up in QFT all the time as well as those "divergent" sums. Renormalization groups topic didn't exist on the wiki when I needed it, at least it is there now...
Some baby level math isn't always taught at school and gets missed at universities later. Like Ceva's theorem, continued fractions(and their connection to hypergeometric function and Pade approximants), game theory(for cases when the next move of each player is restricted by the move history), all kinds of coordinate systems(and useful tricks to solve physics problems, like orths derivatives), (least) quadratic non-residues.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Yes it's math explanation that has nothing to do with physics. So in quantum mechanics, if you measure close to infinity you will not get infinitely many points, you will get a black hole because measuring device always interferes with measurement.
So yes, math OK, physics not so much. I know, you try to tell there is infinity angles in the universe, which also is still to be proven. Math works out, irl not so much. Measuring so small you know you will reach planck length (yes, you conveniently ignore this) and planck radius, angle etc... So, until you do these I'm very much unimpressed. You can calculate lim towards these but still nope, sorry. You can claim in physics there are so many dimensions there's no limit to directions, but you need to prove that.
Math is very good but it's math, not reality. You did not even calculate amount of possible measurements on radius of r ball, come on, you didn't even try with this video. IK, gets more comments and such but c'mon.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
@11:49
9 4
17 13
Since 104 is even, it'll have to be the double of the right two numbers. So their product is 52. 52 factors to 2*2*13, so the possible pairs are (1,52), (2,26), and (4,13).
153 is the product of the left two numbers. 153 factors to 3*3*17, so the possible pairs are (1,153), (3,51), and (9,17). Since the numbers progress in a Fibonacci manner, you'll need the top two numbers to sum to the bottom right and you'll need the right two numbers ti sum to the bottom left. The only way is for the top two numbers to be 9 and 4 and the bottom two numbers to be 17 and 13.
1
-
1
-
1
-
1
-
Interestingly Prof John Robison of Edinburgh, writing in the mid-1700s, developed the epistemological basis for what we now call the Method of Successive Approximations. An equation need not perfectly describe a natural system, to be useful. One simply keeps improving the model with equations that fit the data more closely. A Swiss contemporary named Euler grabbed Robison's concept and ran with it, proving that in the case of power series that can be proven to converge, one can work out how many terms of the series to calculate, until the difference between further terms is less than the part of the number having practical utility. By Euler's advancement of Robison's method, power series to calculate pi, sines, cosines etc. have been found. We can know the circumference and diameter of a round object, to greater accuracy than we can measure them, and this gives us the edge we need to build things with wheels in them. So yes...the whole of mathematics is a tool for understanding the Universe.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Description for infinity and process paradox definement:
I). If there is one or more mathematical process is defined among limited elements, you ain't norm-ally have any result as a paradox and/or infinity sort.
But, if at least one of those elements is not well and concretly defined, then you may receive conclusion at pardixial and/or infinity sort.
For example: element c is the number (excluding zero) one of Reel leauge. Elemet z is equivalent to zero. The process is division and orders element z to divide c, means c/z=? C/0=infinite or defineless.
Let's say element c=z then division process among them is 0/0=N/A defineless. Infinity-inf., inf./inf., inf-inf, inf.-c, etc are all defineless or infinity again.
We have to make a total concrete definement among our elements, process and processing order rule.
Our description mentioned above says/asserts: if there is one or more process among limited elements, and if the elements concretly and really defined, and processing syntax is definitly defined (exchange, disribution, replacement rules must be applied linearly not rondom, (somtime like that sine time like this during processing is forbidden!) Then you may not have any paradox or hardly any result as infinity nor N/A leauge.
Infinity is not and cannot definable precisely and concretly, so that even if the elements limited under processing you may not have any other result except 0, infinity, N/A leauge or paradox. For Processing 0 with 0 or with infinity, it gives simillar results.
C is element if reel leauge, infinity-c is equivalent to infinity. BUT if c is equivalent to zero, infinity-0 is not equivalent to infinity the result is N/A. Ahh why? How?
It is because of 0 is not a well and concretly defined number, even as if it is not a really reel number 0 is simething like infinity it has double face like blade edges and binary quantal appearence in pricessings. In limit pricessing it has somethime -epsilon maner and simetimes +epsilon maner and breaking the pricessing synrax rule!!! We in mathematical pricessings to have a reel result and avoiding fall into infinity result trap, breaking processing rule.
So, uncertainity defunement if zero, and zero's hermaphroditial +/- binary epsilonial janus-face nature causes the problem
The definement for zero must be checked up and recorrection it needs.
Our acceptances for zero are:
If the process is +/- we may put 0 to left/right side of equivalency. This is correct you put -/+ a to the left and/or right. But not correct durectly puting zero to the wings, because 0=a-a and zero occurance after process aming reel numbere is nit equivalent to any zero absolute. I mean you may not put any absolute zero in your processes because the zero at the process +A-A=0 is not the zero absolutly exists in your assertation.
Zero effectless at -/+ process among reel numbers leauge,
Like this
Zero is terminator (singular blackhole) at process multiply among R leauge
0/0 is N/A
C/0=infinity 0/c=0
As you see here there is something sounds paradoxial already and there is definitly a great problem about definition of zero as a number!
1
-
1
-
I had the Tower of Hanoi on my graphing calculator way back in middle school, and used to 'speedrun' 10 disc solutions when I was bored in math class, so I can actually say with relatively high confidence I could still pull it off without a mistake. There's a nice rhythm to it once you get the hang of it, almost meditative. Interestingly, because it was presented as a straight line of pegs instead of a triangle of them, I never stumbled on the clockwise/counterclockwise solution to figuring out where the top disc goes, another example of how how you frame a problem affects how you solve it.
Instead, my method was to think in terms of a start peg, a storage peg, a goal peg, and partial pyramids. Move the top part of the pyramid to one peg (the target peg if the pyramid has odd number of discs, the storage peg if it has an even number), move the top disc of the bottom part of the pyramid to the other peg, move the top part of the pyramid on top of it, and now you have a bigger top part and a smaller top part. Repeat until there's just one disc on the start peg, move it to the other peg, and then you have a new whole pyramid, and you recurse back to the start, just with the start and the storage peg switched places, and a pyramid one disc shorter.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Years ago I stumbled on how to solve these after seeing someone show analogies between finite difference equations and differential equations:
Let's say you're trying to solve the recursive definition for the sequence S(n): S(n) = S(n-1) + 5. Start with the homogeneous equation by only keeping terms involving S(n): S(n) = S(n-1). Clearly S(n) = C for some constant C; we'll call this H(n). Now guess that S(n) is the sum of H(n) and some other function of n based on the sum of the remaining terms, in this case 5. The trick is to always guess a polynomial one power above the inhomogeneous term, in this case a linear function: I(n) = A + B*n.
By checking the recursive definition,
S(n) = S(n-1) + 5 <==>
C + A + B*n = C + A + B*n - B + 5 <==>
B = 5.
So if S(n) = C + A + 5*n, or by removing the redundant constant term, simply S(n) = C + 5*n, then the recursive relation is satisfied and the variable C is determined by setting the initial value of S(n), e.g. S(0) = 0.
This approach works for summation as follows: If S(n) is the sum of the integers from 0 to n, then there is a recursive definition of S(n):
S(n) = S(n-1) + n.
Applying the same technique:
S(n) = A + B*n + C*n^2
==>
A + B*n + C*n^2 = A + B*n - B + C*n^2 - 2*C*n - C + n
==>
(1 - 2*C)*n - (B + C) = 0.
Since this relation has to be true for all n, each coefficient in front of a power of n must be zero:
1 - 2*C = 0,
and
B + C = 0
==>
C = 1/2,
B = -1/2.
This means that
S(n) = A - n/2 + n^2/2 = A + (n * (n - 1))/2, just like in the video but with an additional constant which can be set by setting some initial value of S(n).
It's actually possible to continue this method in a general form for any sum of powers, which results in a convenient definition of coefficients which is again related to the definitions in this video. Just thought I'd share a parallel method for generating these sums.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I think i have an idea for the 2*10^x series. Which amounts are we including? 1, 2, 5, 10, 25, 50 cents. 1, 2 euros, 5, 10, 20, 50, 100, 200, 500 euros? All of them? Just the coins?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
It looks to me that the number of degrees of freedom is not conserved. Of course it has to be conserved, so could someone explain this to me? I start with a general polygon, which has 2N degrees of freedom (denoting with N the number of vertices, we have 2 d.o.f for every vertex). The transformation, also looking at the papers in the description, are matrices that does not depend on the polygon, so they have no free parameters. I end up with a regular polygon, that can be rotated or scale (2 free parameters), and translated (other 2 degrees of freedom). So it looks to me that I start with 2N free parameters, but I end up with only 4, and I do not understand why. What is wrong in what I said?
Thank you, amazing video!!
1
-
1
-
1
-
1
-
1
-
BONUS: Try it with other polygons!
3:
Nothing interesting happens when you try it on an equilateral triangle, we don't have enough points to use ptolemy's.
Unless we consider one point as a duplicate...
L^2 + 0L = L^2
yeah there's nothing here, next polygon
4:
Pythagoras shows up. 1^2 + 1^2 = [diag]^2, diag = sqrt(2). You still have golden rectangle magic but this time it's in this form: take a rectangle with aspect ratio 1 : sqrt(2) and cut it so the long edges are split in half, you'll now get two identical rectangles each with ratio 1 : sqrt(2). This is the basis for A4 paper.
6:
Two kinds of diagonals, call them C for the short one and D for the long one.
C^2 = 1 + D
D^2 = 1 + C^2 = 2 + D
CD = C + C
Very quickly you can figure out what the exact values of D and C are.
(Spoiler alert: D = 2 and C = sqrt(3))
If you draw out the diagonals here the symmetries do check out (if it wasn't obvious by the star of david and rule of thirds shenanigans going on)
8:
We still have the same diagonal before.
In order of length let's call the diagonals X, Y, and Z.
X^2 = 1 + Y
Y^2 = Y + X^2 = 1 + 2Y
Z^2 = X^2 + X^2 = 2 + 2Y
XY = X + Z
XZ = 2Y
YZ = 2X + Z
(That last identity is redundant btw but it's useful to have around)
What would a reverse X, Y, and Z maneuver look like?
Trying to imagine, rescale, and rotate a 4d box is not the best idea for a 3d brain so let's reason through this algebraically. But I assume it involves the same "removing squares trick" (unless of course, something weird happens)
X(AX+BY+CZ+D) = (B+D)X + (A+2C)Y + BZ + A
Y(AX+BY+CZ+D) = (A+2C)X + (2B+D)Y + (A+C)Z + B
Z(AX+BY+CZ+D) = 2BX + (2A+2C)Y + (B+D)Z + 2C
There's a somewhat nice pattern here, the As only ever get added to the Cs, and the Bs only ever get added to the Ds
9:
Call the diagonals in order of length from smallest to largest F, G, H. Can you tell I'm running out of letters?
F^2 = G + 1
G^2 = G + H + 1
H^2 = F + G + H + 1
FG = F + H
FH = G + H
GH = F + G + H
Notice that the 2s only ever seem to crop up when the number of sides is even. They're gone here.
The magic bars?
1¦F¦G¦H
F¦G+1¦F+H¦G+H
G¦F+H¦G+H+1¦F+G+H
H¦G+H¦F+G+H¦F+G+H+1
This feels like a weird magic multiplication table where a bunch of combined areas are all suddenly the same as each other
For example, H^2 = F^2 + FG = F + G^2 = GH + 1 = FH + F + 1
Anything larger:
We can use some logic for something: what happens when we repeatedly apply Ptolemy when one of the diagonals is the shortest one?
Let's go in alphabetical order.
A^2 = 1 + B
AB = A + C
AC = B + D
AD = C + E
The magic happens when we loop back around somewhere, but what if that never happened?
Well then we get the natural numbers.
A = 2, B = 3, C = 4, etc...
The repeated identity is 2N = (N-1) + (N+1)
But then Ptolemy under the reals is this (remember, Ptolemy applies to any circular quadrilateral, hence reals):
Assuming A > B > C > D, (A-B)(C-D) + (A-D)(B-C) = (A-C)(B-D)
You can multiply this out yourself to confirm this
BONUS: What exactly are R and S, anyway?
Two ways to go about this: one using trig and the other using polynomials
TRIG:
Let's solve for R first. The interior angles of a heptagon are 5pi/7, so the isosceles triangle formed by 2 edges and the R diagonal must have base angles of pi/7.
R = 2cos(pi/7)
We can use the formula R^2 = S + 1 to solve for S.
S = 4cos^2(pi/7) - 1
The equivalent for phi is that it's 2cos(pi/5).
POLYNOMIAL:
The two solutions of phi came about from x^2 = x + 1. Can we do something similar here?
RS = R + S
RS - S = R
S = R/(R - 1)
R^2 = 1 + S
= 1 + R/(R-1)
R isn't 1, obviously. Otherwise S = 1+S which is just nonsense
R^3 - R^2 = R - 1 + R
R^3 - R^2 - 2R + 1 = 0
Similarly,
R = S/(S - 1)
S^2 = S + R + 1
S^2 = S + S/(S - 1) + 1
S^3 - S^2 = S^2 - S + S + S - 1
S^3 - 2S^2 - S + 1 = 0
Feel free to use cubic formula but I'll stop here
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
This is absolutely incredible, and very much on my subject.
Actually, I am now working on "1.5" sequences and sequence triangles, which are formed by adding 1 to the previous number, then 2, the 1 again, 2 again, etc. (This is how integer partitions are formed.)
Question 1: I know that adding all numbers up to X leads to triangular numbers (1+2=3, 1+2+3=6, etc.), which, when replaced with circles, form triangles. I also know that adding triangular numbers leads to tetrahedeal numbers, which correspond to tetrahedrons. And then to higher dimensions, etc. All these geometric shapes formed on the basis of triangle (triangles, tetrahedrons, and beyond) have a general name of "simplex".
Here we have a different beast. By adding every SECOND number we get square numbers, then cubic, then hypercibic, etc. As well as the cube's doppelganger, the octahedron. The overall name for these shapes is not "simplex," but ... WHAT?
Question 2: What about he dodecahedrons and icosahedrons? Are they too formed of some basic numerical principles? They are doppelgangers of each other, so I'm pretty sure they too come from the same numerical basis.
However, so far I've been unable to figure out the gnomon formula for either of them. (I figured out the ones for tetrahedron, cube and octahedron without trouble).
Do you know one? Or can you recommend a book on the subject? I'd be very grateful.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
In the midnight's shroud, a tale is spun,
Of a lantern dark, its curse begun.
The Schwarz lantern, a flickering light,
A mysterious glow in the hush of night.
In shadows deep, where secrets dwell,
The lantern's curse, a spectral spell.
Its light, a dance of eerie shades,
In the labyrinth of time, where fear cascades.
A whispering wind through ancient trees,
Carries the legend on spectral breeze.
The lantern's glow, a spectral might,
A haunting glow, in the ghostly night.
Its keeper, a soul in shadows draped,
A pact with darkness, secrets escaped.
A bargain struck, in the twilight's haze,
The curse of the lantern, a spectral phase.
Guiding lost souls with flickers dim,
A spectral lantern's spectral hymn.
Through realms unknown, where shadows play,
The curse unfolds in the lantern's sway.
Beware the light that draws you near,
The Schwarz lantern, a phantom seer.
Its glow, a beacon to the spectral kin,
A dance of shadows, a dance within.
Cursed to wander in perpetual gloom,
The lantern's keeper, a ghostly loom.
In echoes of the past, the curse is sung,
A haunting melody in a spectral tongue.
Yet, amidst the curse, a glimmer of grace,
A chance for redemption, a spectral chase.
Break the chains of the lantern's plight,
In the dance of shadows, find the light.
So, in the moonlit dark, the story's spun,
Of the cursed Schwarz lantern, its course is run.
A tale of shadows, a spectral trance,
In the haunted dance of the lantern's advance.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I think I have an excel solution for minimum moves with N pegs:
All of Row 1 =1
All of Column A =ROW()
B2 and below needs a very large number, since it's impossible to move more than one layer with only two pegs. I used =POWER(10,299), because 1e999 didn't roll over to floating-point infinity as I'd hoped.
Finally, C2 gets the somewhat awkward Array Formula (ctrl-shift-enter after pasting) =MIN(2*C$1:C1 + INDEX(B$1:B1,N(IF(1,ROW()-$A$1:$A1)))) and then that cell gets copy-pasted to fill everything below and to its right, so that the mix of relative and absolute row numbers automagically change as they should.
Much of that formula is devoted to pairing one column bottom-to-top with another top-to-bottom, working around the way INDEX() is hesitant to accept an array of indices and spit out an array of values (thanks, google).
Ultimately, it's calculating MIN(2*(best for D-x disks) + (best for x disks and P-1 pegs)) the long way, but it's a 2D table so all the intermediate values have already been computed elsewhere, which makes for a "nice" spreadsheet formula, before excel quirks are factored in.
Edit: Oh. that exact algorithm appears to have been mentioned just after I paused to write this. Well, wrangling excel to do the work is still nice, I hope.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1: (2 + A^3 + B^3 = C^3) if B has factors 2 and 3, and A,B,C are consecutive integers (need 6 slices to wrap, but they miss 2 corners)
2: Don't know why I didn't reuse the earlier identity, since I know 100, 121, 144 from rote memory but I timed out trying to be clever
3: stumped me for an overarching pattern, 6 and 6.89... are the answers, however...
...a funny thing about the first 2 is if you subtract the exponent (eg 3- exp) from each term instead, the equation still balances, and first term + exponent equals final term without the exponent
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Anomaly positions:
6, 19, 20, 31,32,33, 34, 41,42,43,44,45,46,47, 48, 54 to 62
Difference:
13,1,11,1,1,1,7,1,1,1,1,1,1,1,6,1,1,1,1,1,1,1
Results:
Their different, seems like a 3D map 🗾 of sorts?
1
-
1
-
1
-
1
-
@Mathologer
The approach was: Determine for each number n from 1 to e.g. 20,000, how many ways there are
• to write n as a sum of 1s
• to write n as a sum of 1s and 5s, using at least one 5
• to write n as a sum of 1s, 5s and 10s, using at least one 10
• to write n as a sum of 1s, 5s, 10s and 25s, using at least one 25
• to write n as a sum of 1s, 5s, 10s, 25s and 50s, using at least one 50
• to write n as a sum of 1s, 5s, 10s, 25s, 50s and 100s, using at least one 100
Now, if you already have computed the corresponding 6 values for n-100, n-99, n-98, etc. up to n-1,
these 6 value for n can be calculated just using those precomputed numbers
e.g.
the number of ways to write n as a sum of 1s, 5s and 10s, using at least one 10
(n > 10)
is the same as
the number of ways to write n-10 as a sum of 1s, 5s and 10s
i.e. the sum of:
• the number of ways to write n-10 as a sum of 1s
• the number of ways to write n-10 as a sum of 1s and 5s, using at least one 5
• the number of ways to write n-10 as a sum of 1s, 5s and 10s, using at least one 10
The program then operates by keeping the 6 values for the last 100 numbers around, while working its way up to the target, e.g. 20,000.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I grew up with slide rules. Computers, especially hand held computers, didn't really become popular until after I graduated from college. With that in mind I have an interesting, humorous story to share.
First you should know that years ago nearly all engineering and math classes had a large slide rule mounted above the chalk board. The professor would often use the slide rule to demonstrate or to do involved calculations in from of the class of students.
Before I tell the story, I should let you know there was an expression in those days called "slide rule accuracy". It is actually a pretty useful concept as it says that only the first two or three digits of a number are really important in a calculation. In fact, I taught this idea when I myself was a professor many years ago. The idea is if you have a number like 314,265,473 you might as well call it 314,000,000 because that is accurate enough in most cases. The additional digits may add accuracy, but are often not of any significant value to the overall result. Now for the story.
There was a professor very involved in a long and complication calculation for a class of students. I don't know what the subject was, but let's say it was thermodynamics. The problems in thermodynamics could get very involved with lots an formulas and involved calculations, so it was a pretty messy and complicated problem he was solving. He was sliding the slide and the cursor (the little glass window which enables you to read the numbers off the slide rule easily) back and forth furiously and shouting numbers as he hurried through the calculation. At one point he could be heard to shout, "Two times two - three point nine eight - call it four".
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Interesting curiosity: around 14:15 is mentioned that the number of solutions of prime=x^2+4yz are even, but also this number is "mostly" prime! (for 37=7, but also for 5=1, 13=3, 17=5, 29=5, 41=11, 53=7, 61=11, and finally a counter example 73=21, 89=21, 97=27, 101=11, 109=17, 113=23...)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Yes, exactly what is required to replace amateur associations with a clear picture from the Real Mathematicians.
But I will have to reiterate the video a few times. Adding to Favourites.
Because a real Professor of Mathemagical Thought processes is an actual Intuitionist, observing in resonance, coherence-cohesion chemistry-bonding with a purely functional concept in Conception Completeness.., the definition of a philosophical absolute, abstract process of such component that is self-defining quantization cause-effect materialisation of conceptual elements is the Camera Obscura setup, discovering the Universal Singularity-point positioning of the Centre of Eternity-now Interval Time, macro-micro reciprocation-recirculation magic mode of mathematics.., or some thing, holography-quantization from no thing in the general Conception of real-time logarithmic> Singularity-point <relative-timing ratio-rates Perspective of log-base numberness. (It's only the beginning/Origin orientation to Actuality)
***
Superspin Superposition-point location logic is inherent in e-Pi-i axial-tangential, point-line-circle-> horizon as infinite Black-body surface of orthogonality, ..so transparent that it's not a clear idea of Actuality, until you get the Trancendental Meditation of point location nothing in No-thing definable Relativity, ie what every Mathematical Disproof Methodology Philosophy begins and ends with in perspectives vanishing-into-no-thing Singularity-Duality Conception. Relative-timing turning points are made-of-making elemental e-Pi-i Calculus cause-effect, no wonder we think the universe is consciousness, but it is simply the Function that is Absolute Zero-infinity reference-framing containment of harmonic mathematics in Truth, not the other way around, which, by default, is what explains observable quantization properties.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Hereby another interesting geometric triangle where the horizontal fractions can be factorized from the outer to the inner of the triangle.
The outcome of that factor is the location and the number in the Pascal's triangle.
1/1
2/2 2/2
3/3 x 2/1 x 3/3
4/4 x 3/1 3/1 x 4/4
5/5 x 4/1 x 3/2 x 4/1 x 5/5
6/6 x 5/1 x 4/2 4/2 x 5/1 x 6/6
7/7 x 6/1 x 5/2 x 4/3 x 5/2 x 6/1 x 7/7
8/8 x 7/1 x 6/2 x 5/3 5/3 x 6/2 x 7/1 x 8/8
Where 1 is the unit of length: 1/1, 2/2, 3/3, 4/4, .. The length is going up 1 unit: 2/1, 3/1 , 4/1 for each row..
The "fibonacci triangle" variant of the Pascal's triangle (the number is the sum of the 2 diagonal numbers above that number) is :
1 1
2 1 1
3 2 1 1
5 3 2 1 1
8 5 3 2 1 1
can also be seen as answers to the fractions, factorized from right to left (in this excample below):
1/1 1
2/1 1/1 1
3/2 2/1 1/1 1
5/3 3/2 2/1 1/1 1
8/5 5/3 3/2 2/1 1/1 1
So ,1,2,3, Cheers to the catalan numbered sencorship networks.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Fifth time I'm rewriting this, it should be okay now (well that's what I said the other four times).
Let A, B, C be the points of the circle, and let's say D and E are the two points that lay on the line AB and AC (respectively, when extended) such that AD = AE = BC. Similarly, F, G are in BC and AB such that BF = BG = AC, and H, I are in BC and AC with CI = CH = AB (this is just the setup).
Draw the angle bisector of the angle EAD and let its intersection with ED be J. The triangle ADJ is equal to the triangle AEJ by the side (AD = AE) angle (half of the angle EAD) and side (common side AJ) theorem. Then the side EJ is equal to the side DJ and the angle EJA is equal to the angle DJA and since they lay on a straight line they must be right angles. Then, the angle bisector of EAD is also the perpendicular bisector of ED, and because opposite angles are equal, it's also the angle bisector of BAC.
Let K be the point where GH intersects the perpendicular bisector of ED. AH = AC + CH = AC + AB and AG = AB + BG = AB + AC and so AH = AG. Then the triangle AGK is equal to the triangle AHK by the side (AK) angle (half of the angle BAC, see previous paragraph) and side (AG = AH as we have just seen) theorem, therefore the angle AKG is equal to the angle AKH but together they lay on a straight line so both are right angles, and since GK = HK (by the triangles) then the aforementioned perpendicular bisector of ED is also the perpendicular bisector of GH.
By symmetry, the perpendicular bisector of FG must be the same as the perpendicular bisector of DI and the same as the angle bisector of ABC, and the perpendicular bisector of HI must be the same as the perpendicular bisector of EF and the angle bisector of ACB.
The mentioned perpendicular bisectors must intersect in the incenter of the triangle ABC since they are also the angle bisectors of the inner angles of the triangle ABC as we have just seen.
Recall that any point on a perpendicular bisector of two points is equidistant to said two points.
Let's call the incenter O. Then, OE = OD, OG = OH, OF = OG, OD = OI, OH = OI, OE = OF (because O is the intersection of the perpendicular bisectors). Combining all that we have OD = OE = OF = OG = OH = OI and so all those points lay on a circle with center O and radius OD.
Definitely a very messy proof. It's hard to explain things without being able to point at a picture but I hope this all makes sense and I didn't make any (more) mistakes. Now, time to watch the video!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Can we say for trithagorian theorem... We can write T(A) +T(B) =T(C) +2Tcos(theta), where theta is the angle of the red triangle which is mentioned in this video... For 60°, T(A) +T(B) =T(C) +T
For 120°, T(A) +T(B) =T(C) -T
For 90°, T(A) +T(B) =T(C)
It's absolutely amazing how pythagoras theorem works for any shape... 🙃
Amazing stuff after amazing stuff... How Gp-Ap inequality and sin addition formula came... Absolutely fabulous✨
The area of the square will be 1/4 unit.
And, for the beetle puzzle, if we consider any one of the beetle,
Since for symmetry, they will meet at last at the center of the square, so,
average displacement of a beetle is equal to 1/2 of the diagonal of the square.
<x>=a/√2
And also, we know average displacement is equal to average velocity times time.
<x>=<v>t
And <v>=vcos(45°) =v/√2
So, elapsed time to meet = t= a/v
Where a=side of 🔲 and v= constant speed of each beetle
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I wanted to prove (x + 2)^n having those n-square coefficients, but kinda got tired of doing so. However, as an intuitive argument, take an n-square, and extrude it out to an (n+1)-square, so that it's obvious that all k-edges will double, because you have two copies of the n-square, plus all k-edges of the n-square will extrude out to new (k+1)-edges. You can see that's what multiplying a polynomial (in x) by (x + 2) does: double each term, and add the coefficient beneath. Having said this, I think it probably would be easy to formalize, but I'm done working the problem, because I want to actually finish the video :)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Dear Mathologer, thanks a lot for this amazing video. I noticed that many people selected "no integer as partial sum" as most memorable subject, although you did not provide any proof about that. So I began to look for such a proof and I found a nice one, that I hope you will appreciate. It runs like this:
Among the first n integers, there is one (say p) including a higher power of 2 than any other. p is the highest power of 2 not higher than n (eg n=23, p=16=2 power 4). This is because the first multiple of 2 power x is 2 power (x+1). When summing 1/n, the denominator gets the product of numbers from 1 to n, and the numerator gets the sum of the products of all combinations of (n-1) integers. Among these products, the one where p is missing has a lower power of 2 than any other, so after simplifying by 2 as many times as possible, the numerator is the sum of n-1 even numbers and one odd number, and as a consequence is an odd number.
Example: n=5 p=4
1+1/2+1/3+1/4+1/5=(2*3*4*5+1*3*4*5+1*2*4*5+1*2*3*5+1*2*3*4)/(1*2*3*4*5)
=(1*3*4*5+1*3*2*5+1*1*4*5+1*1*3*5+1*1*3*4)/(1*1*3*4*5)=(60+30+20+15+12)/60= odd/even
I suppose the same proof applies for the sum of 1/n to 1/m but this is a bit more difficult in case there is no power of 2 between n and m. Although I am sure there is in all cases one number with a higher power of 2 than any other, I did not find an elegant way to prove it.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I am yet to see a video going deeper than just the slight surface layer scratch as to what concepts can be formed, representative of numerous elements of the number theory field, just from exploring relationships between numbers, functions & operations of number interaction - from the numbers used in this symbol. For me its just a symbol for entry to one entry into a beautiful realm of relearning how I consider numbers & geometry. The slight edge of ego towards those who see something in this symbol as being a 'key to understanding the universe' rather threw me to be honest. I don't really go with the vibe or flow of whether or not there is a key to the universe from this symbol, that seems like unnecessary noise around something not really needing a label as such. If some/most want to stick with that layer, that's up to them. Sacred geometry tattooists are glad for it as too are those who get the cool designs not always having the layers of awareness as to the mathematical beauty in such shapes and symbols which also doesn't matter :) There is without doubt from my side, interesting and mathematical observable elements beyond just this symbol and special characteristics which could be declared for each whole integer or single digit, even decimals... grouping around characteristics, whether results of functions or influence in functions, pure number play from just 1-9... personally this opens up one of the most stimulating parts of my brain... to figure things out without googling them, to then search for the true math behind the sandboxed math which leads me to concepts to further my knowledge now based on that self exploration seed. That truly personally, is something of a key to the universe, or my universe... to not see things only as we are shown but to explore seeing things from different perspectives to realize some quite fundamental truths that our truths are our truths but that some things just seem to be true whoever observes them... At best, even if people play with digital root for numerologist intentions, I am still glad that numbers are explored for different utilization beyond the mundane counting of what we earn or spend or watch on a clock :)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Yes, there is a super important application for this in electrical engineering.
It is called symmetrical components.
In a 3 phase system there are three wires called phases, and one called neutral.
All phases have a sinusoidal current and these currents collect in the neutral.
If the system is balanced 2 things happen:
1) The 3 phases have sinusoidal currents of same amplitude but 120 degrees out of phase with each other.
2) When these currents add the cancel out, thus having no need for returning wires, allowing for copper economy
In real life the system is never balanced, and that is what the neutral is for (it returns the what is left from an imperfect cancellation)
Every sinusoidal current (and voltage) is seen as the horizontal component of a spinning arrow.
These arrows form a triangle.
It is too hard to study how the system behaves when it is not balanced, because of electromagnetic fields interfere between wires.
But these effects cancel out in a symmetrical system.
So engineers break the assymmetric triangle into three symmetric ones, called symmetric components.
Different kinds of defects affects each component differently, so we study each one separately, as a kind of sygnature.
You showed my a different way to filter out each component.
Thank you.
Worth noting: in Russia the system is 5 phase, but the same ideas apply.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Video Request (is there a place for vid requests?):
Counting parallelograms in a triangle with each side cut into n equal parts, and all lines connecting these points and parallel to each side drawn.
https://www.youtube.com/watch?v=UiJFl072kdw
She writes a formula at 4:20. I know the formula, but I don't understand it. Where does it come from / why does it work? How is it related to Gauss's Sum(n)|[n=1 to N] = N(N+1)/2? How is it related to similar problems (say #of rectangles in an orthogonal grid of known dimensions)
Seems like it's esoteric and complicated enough to warrant your treatment :D
I hope it's related to some crazy physics, like how cells work or something, i don't know, but I would love to have your take. Apologies if it's already been done, I couldn't find an explanation for the formula. This problem reminded me of your Christmas water bucket video , because, ...the diagrams are similar. :) Much love, thank you for all your hard work, it's a joy
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Something I figured a while ago is that the cubes (1, 8, 27, 64, 125) go up by a series of numbers (7, 19, 37, 61) that go up by a series of numbers (12, 18, 24) that go up by 6, and there’s something like that for squares and powers of 4 too
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
@Mathologer Thanks as always for amazing effort. Seems like the distribution that maximizes the minimum happiness across all coalitions is the average of the "extreme" distributions, where an extreme distribution is one that makes someone maximally happy, a next one the next best happy, a next next one the next next most happy, etc. (for the case of 200 level of estate, the 4 extreme distributions to the 3 creditors of (100, 200, 300) are: (100,100,0), (100,0,100), (0,200,0), (0,0,200), and whose average is (50,75,75)).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Start by factoring 153, it has factors 1x153, 3x51, 9x17.
possible values of b then are 76, 24, and 4
the smaller factor of each pair=a
the larger factor=a+2b
factor (104/2)=52, it has factors
1x52,2x26,4x13
the smaller factor=b and the larger factor=a+b
the only b that matches both is 4, and 9+4=13.
On the chart because 4 is in the upper left corner, either the previous step was
x 4
9 y
or
x y
9 4
the first option is the only one that avoids negatives and results in
1 4
9 5
the parent of this is is then
1 3
7 4
and the parent of this is
1 2
5 3
reading forwards is LLLR
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I always roll my eyes when people talk about galaxies and nautilus shells in relation to phi, it's not true, but also just observing patterns is far less interesting than understanding them. This meant I had always had a "resistant" feeling towards phi.
Mathologer has well and truly turned me into a phi appreciator, first with the continued fraction representation explanation, and now this, I'd somehow never seen Binet's formula, or at least if I had, I had forgotten it, but the derivation and wider connections was absolutely wonderful.
I do want to note that as a colour blind viewer, I do struggle with some of the coloured lines in some videos (at the start), it's not an easy thing to solve, I know I tend to prefer dashed, dotted, double, ect. lines instead of colour indicators but that can get messy fast, it's just something maybe worth considering.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I learned of Pythagoras, in Freshman Algebra, Fall, 1962, and how to solve his famous theorem, in Mrs Ferguson's 10th Grade Geometry class, Fall, 1963, where I also heard the news of the murder of JFK, a few moments before noon, Nov 22, 1963. I have used the Pythagorean theory much of my professional life, for a multitude of calculations, along with various other proofs that apply to area, volume and related derivatives. I cannot imagine not understanding this simple, infallible tool, although I am not surprised by young people who haven't a clue.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
"LOL, this just proves you cannot prove heliocentrism is 100% fact, because it also simultaneously proves Earth is in the center, and everything is revolving around us! Huh! Bet they didn't teach that in class. Nor why we went with heliocentrism. Because Einstein and the like understood, out of these 2 options, one demands God, and one just leaves the door open for God, not proving it outright, so they went with option B, there is no God, and cannot possibly be a God, so we have to see it from the heliocentric model way. Period. Case closed.
P.S. Man, so many videos just recently, indirectly are proving the Earth is not a sphere, nor are we in some solar system. That the Theory of Relativity is laughably bunk, and that the constant speed of light isn't a constant at all, and gravity and time-space are also not real. Hilarious that those inferences can be made from the info provided, yet, because of the depth of the education, and the utter impossibility of God making us the center, like I mentioned about Einstein, that these scientists literally cannot even see what they are showing. It's like the story of the Indians not seeing the ships because they had no mental framework for it, so they physically could not see them.
TL; DR. The Earth is factually and provably not a globe. Basic logic and first principles reasoning proves this. Most science claiming globe and gravity is all based on conjecture and theory, not provable in court level of certainty. If I had a gun to my head, and was told to prove one, I'd be forced to prove we cannot prove at all the Earth is a globe. It's the biggest lie the managers of this world ever foisted on to us slaves.
REALLY TLDR - you aren't an amalgam of space dust and particles from the Big Bang, just happened to form due to billions of years of evolution, and our existence is all happenstance, and has no meaning. Quite the contrary. A creator of some kind made Earth, made it hospitable, created us, and everything we see. We are the center of creation. We must be, using logic and reason alone."
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
8:08 Using the same method and we get 5. (Is there a simple example of 7 points where no 6 are on the same hemisphere?)
10:10 An alternate proof using PHP(the Pigeon Hole Principle). Consider the sequence 9, 99, 999, ...
Claim: 113 must divides one of them. Proof of the claim:
Divide each of them by 113. And we get a new sequence of the remainders 9, 99, 95, ... PHP tells us that the new sequence must repeat, say the remainders of (10^m - 1) and (10^n - 1) are the same where n>m. That means 113 | (10^n - 10^m), thus 113 | 10^(n-m) - 1. Then 10^(n-m) * 355/113 - 355/113 = an integer which is the repeating tail(or multiple copies of it).
16:24 Define two relations: S for "shake hands with each other", and N for "not shake hands with each other". Then person A either S with 3 or more people, or N 3 or more people. WLOG say A S with B, C and D.
Senario 1: There exists some S relation among B, C and D. Then those two, together with A, are the 3 people who shake hands with each other.
Senario 2: No S among B, C, and D. Then they are the 3 none of whom shake hands with each other.
22:20 I tried (RULD)^3 and found a 5-edge-cycle, a 7-edge-cycle with corner-twists and edge-flips. So I guess it's 3*5*7*3*2 = 630.
26:10 1, 2, 4, 8, 16, 32, 64
31:02 DCJ stands for 3, thus Queen of hearts.
I like the proof of 355/113 one most because it's new to me.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Dear Marhologer,
I am the one who challenged you yesterday to make a vid about the Galois Theory. But I have, only out of curioisty and only that, one other question: I see you are a profesor (I knew that for years already of course, but now I saw it fot real), and now I wonder what does the word "Profesor" mean in Australia. Of course, you teach students at university, this is NOT meant as a stuoid question, but is there also something else? In different countries around the world, the job may have different meanigs and a profeosr might have other tasks than in another country. For instace: do they call high school teachers "professor" as well, or may be the piano teacher. You are in a different country, on a different island and even on a different hemisfere, so yeah, this is a reasonable question I think,
All the best from Agnieszka
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Thank you for the excellent video as always Burkard :) Just one note, as someone who is colour-blind, I sometimes struggle to differentiate between the colours you choose. Namely, the orange and green in chapter 2. Think of the problem like this, some hues with a given brightness and saturation can look the same or similar to other hues with a differing brightness and saturation.
Interestingly, I think there is probably some nice maths behind which colours are conflated and a "colour-blind coefficient" could be prescribed ...anyways.
If I could offer a suggestion, maybe you could try using colours with matching brightness and saturation. Also, if hue range is a rotation from 0 (red=rgb[255,0,0]) to 1 (blue=rgb[0,0,255]) try using hue values equidistant along the range. Graduations of purple can be tricky.
As red/green colour blindness is the most common maybe a range 0 (green=rgb[0,255,0]) to 1 (blue=rgb[0,0,255]) would be even better, with red=rgb[255,0,0] and purple=rgb[255,0,255] as discreet additional colour options.
Just some food for thought if you see this :) As a side, I am red/green colour blind and I personally find it impossible to differentiate between green[0,255,0] and yellow[255,255,0] without conscious deduction. Cyan[0,255,255] appears smoke white and I see colours in grey-scale.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I was pausing and pondering, trying to figure out what's going on throughout the video, then Mathologer said "Complex numbers," my brain immediately said, "Oh, it's just DFTs," and the rest was painfully obvious. I wasn't thinking about "ears" so much as similar triangles as points on perpendicular bisector that are a proportional distance away from the edge, and both of these properties carry across to the DFT decomposition. Saying "ears are just filters," would be a stretch, but it's not far off. Now I'm wondering if I can do this with polyhedra in quaternion space, or if the 2:1 mapping of SU(2) to SO(3) is going to break things. And whether this would still look like some sort of a discrete QFT or if I'm going to only pick up some of the DQFT terms.
1
-
Hm, well, the decomposition certainly works, but most of the neat DFT properties that aren't just inherent in general linear decomposition aren't there. If you start with points that can be decomposed into regular polyhedra, some analogies remain, including the "ears" (which are now constructed from centroids along the normals scaled by face area - there might be a solid angle connection here), but it's not as exciting. A general polyhedron can be decomposed into a set of polyhedra with matching topology, but the choice is almost arbitrary. They just have to be linearly independent under the operation. In practice, there could be interesting applications where the basis is not complete, using pseudoinverses, resulting in some sort of mesh compression or applications in 3D reconstruction, but at that point, it's just projection onto a basis and any real connection to the elegance of DFTs for planar polygons is gone.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
nice to see you, as an engineer I knew that the key was the pulley, and I have to tell you that my best way of shoe tie is the one that looses the shoe easier and faster, and I came to a solution not considered by you and is the one of MANY that includes MORE PULLEYS and is like this, when you start a criss cross just go back to the same side making a pulley in the middle between laces, this way, intuitively you have a large number of pulleys multiplying each time force times 2 ,mathematically, with 0 diameter lace it is also as short as criss cross, it is very easy to loosen also, only drawback is is hard to keep middle pulleys centered but as one side straight it comes to multiply force a lot more, now you have another equation to dream off.... HEEE heee heee ( friendly lol)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Thank-you. Very neatly described and demonstrated. 🙂👍 My slide-rule is still working properly, despite never being plugged into an electricity supply, over the past 50 years!.
I was with you until you said 'Can you code...' at which point my answer to almost any following question is 'No.'. Coding has never quite jelled, despite trying to learn it for decades. I have degree qualifications in an engineering field, speak some Russian and can even read Cyrillic script easily, but coding is much worse than a foreign language, because it's so illogical.
Yes, I know you'll intake breath sharply with that word! But I have no idea about where, or how, to start to answer your question. My answer would be to point you to a slide-rule, and to say 'This is all you need to calculate figures to any scale, with an accuracy of 2 to 3 decimal places. It gives even more precision if how to use it properly.'
Every one of us has our strengths and weaknesses, of course, but coders seem to assume that their favourite subject is automatically easy for anybody. I'm living proof that they are wrong with that assumption.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I like the second proof better, because Sallows' theorem is really a theorem in affine geometry, where lengths, angles and rotations don't make sense in general, but translations and scaling about a point do make sense, hence length ratios exist for parallel line segments, as do rotations by \pi (scaling by -1). In particular midpoints of lines make sense, hence so do medians of triangles.
Now for each midpoint, apply a -1 scaling to the two smaller triangles that meet it. The result is a hexagon whose sides are parallel to the medians. The latter extend to diagonals of the hexagon which bisect each other at the centroid (this gives another proof of the 2:1 property of the centroid) and hence cut the hexagon up into 6 triangles. Sallows' theorem follows from the fact that these 6 triangles are all translates of each other, perhaps after scaling by -1 about the centroid. Furthermore, if the sides of the hexagon are extended to form a star of two triangles (related by -1 scaling again), then the additional triangles in the star are also translates of the Sallows' triangles, so the two big triangles are rescales by \pm 3 of translates of the Sallows' triangles. The medians of these new triangle are parallel to the sides of the original triangle, so repeating the process with either of the new triangles gives back a translate and rescale of the original.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
At 20:15, you removed a 1 in the numerator (presumably because 1 is the multiplicative identity and so does not affect the multiplication at the top.)
But you DIDN'T remove the 1 in the denominator, which you could do for the same reason.
This has the effect of changing the fractions you use to generate your infinite multiplicative series at 21:30. Options are:
(2/1)*(4/3)*(6/5)... your choice, with inconsistent 1 removal, which results in all fractions > 1 and a divergent series, or
(1/1)*(2/3)*(4/5)*(6/7)... consistently not removing 1s, which results in all fractions <= 1 and a convergent series, or
(2/3)*(4/5)*(6/7)... consistently removing both 1s, which results in all fractions <1 and a convergent series (the same convergent series as before * 1/1 = the same.)
Edit: 4th option is inconsistently removing the 1 in the denominator, resulting in (1/3)*(2/5)*(4/7)*(6/9), again convergent.
I am genuinely curious as to why the inconsistent treatment of the 1s in the numerator and denominator.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The second puzzle at 7:25, the answer is 2.
I did this mentally using the sun of square formula to figure out the sun of the first 14 squares, and the sum of the first 9 squares, and then did a subtraction. It’s (14)(15)(29)/6 - (9)(10)(19)/6, which simplifies to (7)(5)(29) - (3)(5)(19). This becomes (5)(29*7) - (5)(3*19), which is (5)(203 - 57), which is (5)(146). We are dividing the entire thing by 365, which is (5)(73), and 146/73 is 2. So with all of the cancellations, the answer is 2. Not the easiest thing to do in my head, but the answer does reveal itself.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I decided to look at other paths through the tree and was surprised to find some unusual sequences appearing. There are infinitely many paths which produce a hypotenuse to smaller side ratio approaching sqrt(2). However, the paths where the two smaller sides differ by a constant are limited in a surprising way. I'll notate the paths with L, M and R for left, middle and right child.
We saw in the video that M^n for n >= 0 produces triples where the smaller two sides differ by 1. Are there paths where the two smaller sides differ by 2, or 3? The answer is no. It turns out that if (a,b,c) is a Pythagorean triple with |a-b| = k, then the triple can only be primitive (and appear in the tree) if the prime factors of k are all congruent to 1 or 7 mod 8 (A058529 on OEIS). The first few values are 1, 7, 17, 23, 31 and 41. We get k = 7 with paths LM^n and RM^n, k = 17 has RLM^n and RRM^n, k = 23 has LLM^n and LRM^n, k = 31 has RRLM^n and RRRM^n and k = 41 has MLM^n and MRM^n for n >= 0. Note that they all begin with travelling to some node, then repeatedly choosing the middle child to maintain the constant difference between the two smaller sides.
Now consider paths where the ratio of hypotenuse to the smallest side is sqrt(5). This can be achieved by having the larger of the two smaller sides differing by a constant from twice the smallest side. We get a similar restriction as before, though this time more complicated. If (a,b,c) is a Pythagorean triple with |a-2b| = k, then the triple can only be primitive if k is squarefree and the only prime factors of k are 2 and primes congruent to 0, 1 or 4 mod 5. I couldn't find an OEIS entry for this sequence. The first few values are 1, 2, 5, 10, 11, 19. We get k = 1 with path L(ML)^n, k = 2 has (MR)^n and R(MR)^n, k = 5 has ML(ML)^n, k = 10 has LR(MR)^n and RR(MR)^n, k = 11 has LL(ML)^n and RL(ML)^n and k = 19 has M(ML)^n and RML(ML)^n for n >= 0. Here we see that the repeated unit to travel along the paths involves two steps, to maintain the constant difference between the middle side and twice the smallest side.
Another curiosity is which sequences have the property that the ratio of the smaller two sides approaches the golden ratio. I can't find a pattern to them, but I did observe that down to the 11th level of the tree the two paths that gave the closest approximation to the golden ratio were LLRLLLLLMRL and RRLRRRRMLR. And instead of prefixes of these sequences being the best approximations for earlier levels, instead suffixes are better. Which means these aren't really describing two paths, but that the best path is moving further away from the middle of the tree at each level after the first few, with the occasional move towards the centre. It is quite curious.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
So I just deleted a huge comment, and this is me giving up on life. 4 years of work ( not continuous, but anyway ) for a puzzle my math teacher gave me in highschool... and all I got was a shifted version of the Gregory-Newton formula ( shifted one position to the right by using (n-1) instead of n...
I am literally having a crisis right now, and I'm dying inside, but please continue.
At least I'm learning how dumb I am, reinventing the wheel...
Anyhow, one note I feel needs to be more emphasized: if the abscissae are not evenly distanced, this is useless to hell and back. I know that oh so very well because that's what I did for about half a year, every night. Now, if the resulting polynomial is 4th degree or lower, no problemo, you can apply Gregory-Newton on the abscissae, and then get the inverse of that function, but for 5th degree polynomials and above, you're fucked cause you can't get the roots. That being said, I was going to try and use fourier series to try and solve that issue somehow, but now I lost all motivation, someone resurrect those two dudes and let them have at it. It's their formula, not mine. Yes. I am really not happy right now.
The general Newton interpolation uses abscissae as well in those differences, which account for unevenly spaced abscissae and remove shifting. That being said, for the same case scenario of evenly spaced abscissae, the general formula yields a bigger error than the Gregory Newton formula...
Now, there's spline interpolation and shit out there, also the Newton Raphson approximation formula ( which I hate, cause some dick put us through doing that by hand... and it took 30 pages each time, for 4-link mechanisms and also for crank-piston ones, even tho crank-piston mechanisms have a nice analytic formula to them ).
ANYHOW, thanks for the video! I mean it. I'm salty cause I realized how much time I wasted doing something someone else did 400-ish years before my time... and if someone told me this like... idk, earlier, I'd not have wasted 4-5 kg of paper and about 1 year worth of free time on this (cause of course I did other things with my free time in 4 years )... oh, god, I feel so bad.
Also, I already know the proof :)) I had to write it for myself... but I bet yours is prettier, let's see
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
These sequences of 1, w, w^2, etc immediately reminded me of FFT, though matrix was arranged in unusual way. Most logic way is make element (n,k) equal to w^(n*k), where n,k = 0..N-1. So for 'zeroth' row we have all ones. Then we have 1, w, w^2, w^3 etc. Next row: 1, w^2, w^4, w^6 etc, but of course as w is Nth root of unity, we can subtract N each time from exponent.
Some time ago I developed balanced ternary fast Fourier transform. Not that big achievement, but it was very convenient to work with in practical applications. In usual binary FFT there are several ways to perform transform 'in-place' but all samples get 'binary inversion' that is they are rearranged as if their indexes expressed in binary form are now most significant bit became least significant and so on. There was 000, 001, 010, 011, 100, 101, 110, 111 (0,1,2,3,4,5,6,7), but now we read them 000, 100, 010, 110, 001, 101, 011, 111 (0,4,2,6,1,5,3,7). Well, with balanced ternary FFT it's the same but with balanced ternary number system! And that's very well, as we have same number of positive and negative 'frequences' and no nagging Nyquist frequency which is neither negative or positive. Doing diffraction pattern on some symmetric aperture works as a charm!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
With all due respect, the step right after 'Engage' in which you describe scaling all interior quantities by the incircle radius r, pointing out that the sum of three quantities is 1-D, area 2-D, and the product of 3 quantities 3-D, and then jumping right to a slide that shows the scaled Area = Sr and the scaled product of 3 quantities RGP to be Sr^2, to be completely unclear as well as rushed through (after spending significantly more time on much easier to follow steps). You aren't showing how you are scaling the sum by r at all, and you aren't explaining why scaling the area amounts to multiplying by r (why not r^2, or would it be r^(1/2)?), or why scaling the 3-D RGP involves multiplying by r^2 (why not r^3, or would it be r^(1/3)?). The inference would appear to be that scaling the sum of 3 quantities amounts to multiplying by r^0. Clear as mud why the scale factor drops by 1 in the exponent, but the way you rush through it as if one is just supposed to grasp it immediately, makes me not want to watch any more.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Okie Dokie
🤔 I don't consider this a good approach for homo sapiens Sapiens except people who code lil programs, iterative method very numerical.
Now finally around 16:30 the Euler, convergence.
😳 Oh, professor, now you will get the value of e? 😅
I see at almost 20:00 you are basically using not much scratch, since you know the way to proceed.
There's it, complex variable and applications, Churchill Brown.
Finally, fundamental theorem!
Often we have to start from contour integral and connected domains multi or single, multi is not much, just we work with any region like that but we put them inside a surface with boundaries, which serves to approximate the maybe odd figure
1
-
1
-
1
-
1
-
1
-
1
-
1
-
, if you counted points in space, how many would there be? so, we can show a square meter has 6 square meters of surface area, its going to need a little doing, but we will get there. To get a circle by points, we say; 2x2, 4x4, 5x5x5, by representing a number of bits to turn on a light bulb, it takes a multiplex of the number of bits of the signal type, such as liquid crystal display, but an array of liquid crystal displays. Due to this complexity, a knob or a slider of any rotation excedes any surface area as shown to achieve transform mechanical heterodyne work. Si an example is the ultimate micro chip. A modest 256 megabyte square array, for mathologers on a tight budget, but you gotta win right? So, theres all sorts of ideas, except, why didm't they make this? Only the garbage videa card and CPU? Because of the number of possible busses makes the chip infinitely high stacked, far exceding the chips surface area.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
An idle thought/unanswered question about rearrangements of conditionally convergent series I've harbored for a a long time
If you think about these rearrangements as permutations on N applied to the series 1 + -1 + 1/2 + -1/2 + 1/3 + -1/3 + ... the greedy process of selecting positive and negative terms is a way to link any real number to a permutation on N. It can be proved that, among permutations on N, the "bounded" ones do not result in a different sum: if p: N --> N is the permutation and |p(n) - n| is bounded over all n, then the sum will remain zero. This is true in general: for any conditionally convergent series, bounded permutations of the terms do not alter the sum [I've seen this proved in an article in a Math magazine!]
Within the domain of permutations we can also establish the notion of two permutations being bounded from one another: p and q are bounded/close if |p(n) - q(n)| has a finite upper bound.
My question is, is it true that whenever p and q are permutations that correspond to the same real number, are they bounded from each other?
------------------------------
The rest of this is fleshing out the details of what I mean, and why the answer either way seems very unintuitive (to me)
Firstly, let's pin down that there is more than one permutation associated with any real number. For instance, let r be the permutation on N resulting from greedy method from the video for making pi using the above conditionally convergent series.
Now construct a new permutation, s, by repeating this process, except we pretend we mess up with an early insertion of -1/3 to get
1 + 1/2 + 1/3 + [-1/3] + 1/4 + 1/5 + ... + 1/13 + -1 + -1/2 + -1/4 + ... (note lack of of -1/3 in the subtracting section)
and from there continue without error -- adding terms till you exceed pi, subtracting terms until you're less than pi, and so on.
Though it doesn't strictly adhere to the method outlined for making a permutation that sums to a desired number, that doesn't mean s does not correspond to a series that validly sums to pi. After all, it's equivalent using the same algorithm to construct a sum to pi + k, for some rational number k, when starting with a conditionally convergent series made by dropping several terms from the original conditionally convergent series, then reinserting the dropped terms.
On their face, r and s would be very different permutations. By close examination you wouldn't be able to tell there's anything that links them unless you backtrack through the processes that defined them.
It would be surprising if r and s were bounded from each other. Yet they both correspond to the same sum
1
-
1
-
1
-
1
-
1
-
1
-
1
-
- This is a lovely video. The heavy lifting is done by all the properties of the limit (in a mathematical sense), which are disguised in the clever notation.
- 35:10 e = lim (1 + x)^(1/x) as x goes to 0. Swap the usage of lim and write e = (1 + dx)^(1/dx) instead (abusing/exploiting the notation). Raise both terms to the power of dx, then subtract 1: dx = e^dx -1. Finally, divide by dx to get 1 = (e^dx-1)/dx
- By using the concept of the car as a vehicle for the explanation (pun intended), some hard issues are avoided, such as "can I differentiate any function at any point?" or "can I integrate any function on any interval?"
- We Spanish speakers have access to a very neat opening interrogation symbol, "¿". This is very handy, given that you know that the phrase is an interrogation from the very beginning.
- The rule for the derivative of the inverse function is graphically very intuitive after noticing that the plot of the inverse function is the same as the original function but mirrored about the line y=x. Hence, the slope of the inverse function is the (algebraic) inverse of the slope of the original. I'm sure Mathologer is aware of this fact, and probably decided to leave this property out for the sake of the total length of the video.
- 27:22 That math joke is brilliant. Here's a similar one I found elsewhere: take the equation x^2 = 25. Of course (?) you can cancel each number 2 appearing on both sides, which leaves x = 5, the correct solution for the original equation. I wonder if there are more of these...
- 32:15 Fun fact: the quotient rule for h = f/g can also be obtained by applying the product rule to h = f * (1/g) first and then using the chain rule with the functions 1/x and g.
- 33:17 The rule (x^n)' = n x^(n-1) is the same for ANY value of n, not only positive integers, but for any real number. But the proof is more involved and I would start by writing x = e^(ln(x)).
I liked the music of the final part!
1
-
1
-
1
-
1
-
1
-
Only because you asked for notes: The bit around 26:00 felt very much like it fell out of nowhere. And then saying that I can plug a few values in to see that it is true makes me feel even more distant from the proof-y reasoning I’m hungry for. Additionally, the matrix inversion derivation for the Bernoulli numbers feels so labor intensive and mechanical, compared to a closed form for the nth Bernoulli number (although it did connect very naturally to the ideas being presented). The rest of it all clicked, although I have experience with all the prerequisite number theory, linear algebra, and calculus, and was very interested in the sums of powers at a young age, so it’s not much of a surprise that I followed along with and ate up this video. It goes without saying, I loved this, and eagerly await more Riemann-Zeta, Bernoulli, or, uh, any content from this channel, really.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Not quite an aside, but if all "calculation" is reiterative reintegration of here-now-forever, then division by zero-infinity is eternal sync-duration and division by one, the reciprocal, unchanged "solidity" of state-ment, equal to a frozen phase-locked substantion.
From observation of the e-Pi-i sync-duration dimensionality coordination analysis, ie the inherent mathematical intuition, under-stood circum-stances of ultimate Superspin Modulation positioning, the first principles reasoning about why this "works" is the i-reflection containment connection between e-scaling and Pi-bifurcation multiples proceeding from unchanged Unity, to twoness symmetrical probability distribution in Spheroidal Toroidal proportioning, sync-duration connectivity as described/demonstrated in the video.
The point-in-perspective being .., how Mathematicians develop intuition about numberness by holding a variable, either e or Pi and picturing numbered gradients in their minds. A natural identification, pictographic symbolism from learning by doing experience. (Genius everyone has the mechanism for, if not the motivation)
More beautiful Mathemagical orientation.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
He's not talking about the entries in the multiplication table. He's talking about perfect squares. In other words, numbers of the form x^2 where x is a number in the mini-ring. (Though I guess the same will also be true if you look at the full multiplication table, but that's not what he means.)
The perfect squares are precisely the diagonal entries of the mini-ring's multiplication table.
But yes, exactly half of the nonzero numbers in the mini-field will be squares (meaning they will appear on the diagonal of the multiplication table) if we're going modulo an odd prime.
Why? Well, if we're in Z/p for an odd prime p, then there are exactly p numbers total, an odd amount. So if we remove 0 from consideration, there are an even amount left: p-1.
Next, notice that because of how weird the addition is on Z/p, every number can be seen as both a positive value and a negative value. For example, in Z/7, 5 is both +5 and -2. Adding 5 has the exact same effect as subtracting 2. (Try it out for a couple of numbers!) This can be seen because adding/subtracting 7 is the same as adding/subtracting 0 in Z/7 arithmetic and because -2 and +5 are 7 units apart.
Now, because of how multiplication works, opposites square to the same thing. So in Z/7, 2 and 5 (= -2) must square to the same thing (in this case, 4). So now, we're approaching the result. Since there are an even amount of nonzero numbers in Z/p, if we show that each square comes from squaring precisely two different numbers in Z/p, we will have then proven the claim.
To show that each square comes from at least 2 numbers, we have to show that no number is its own negative. Suppose n is a number in Z/p so that n = -n in Z/p. Then adding n to both sides, we get 2n = 0 in Z/p. But the only things in Z/p which are 0 are multiples of p. Since p is an odd prime, p doesn't divide 2, so p must divide n (since p divides 2n). This means that n is a multiple of p, so n = 0 in Z/p. This shows that there is no nonzero number in Z/p which is its own opposite. Since opposites square to the same number, this shows that every nonzero perfect square in Z/p comes from at least two different numbers in Z/p.
All that's left to do is show that you can't have 3 or more numbers square to the same thing. We will do this by showing that if two numbers square to the same thing in Z/p, then they must be opposites. Suppose n and m are two unequal nonzero numbers in Z/p so that n^2 = m^2. Then n^2 - m^2 = 0 in Z/p. But we can factor this. n^2 - m^2 = (n + m)(n - m). So we have
(n + m)(n - m) = 0 in Z/p.
Since p is a prime, this implies that n + m = 0 in Z/p or n - m = 0 in Z/p. (If a prime divides a product, it must divide at least one of the factors.)
The first option gives n = -m in Z/p, meaning they are opposites.
The second option gives n = m in Z/p, which cannot happen because we assumed they were different numbers in Z/p.
So the only things that can square to the same number are opposites in Z/p.
So what we have shown is that in the mini-field Z/p for an odd prime p, there are an even number of nonzero numbers. Every number squares to a perfect square, and numbers which square to the same thing come in pairs. So there are precisely two nonzero numbers in Z/p for every nonzero perfect square. This means that exactly half of the nonzero numbers in Z/p are perfect squares. This leaves that the other half cannot be perfect squares!
1
-
If you want to avoid talking about "negatives", I've come up with another way to show that two different numbers square to the same number. Fundamentally, it's the same thing. But you don't have to think about negatives.
Suppose m and n are numbers in Z/p so that m+n = p in normal arithmetic. Then m^2 = n^2 in Z/p. Why?
Multiply both sides by of m+n = p by (m-n).
m^2 - n^2 = p(m-n).
But since p(m-n) is a multiple of p, this tells us that in Z/p, we have
m^2 - n^2 = 0 (in Z/p).
Adding n^2 to both sides gives
m^2 = n^2 in Z/p.
So if you have two numbers that add together to give p normally, then they square to the same thing in Z/p.
And since p is an odd prime, n and m have to be different. If m = n, then m+n = 2m. And, like I said in my original reply, if p is an odd prime, 2m cannot equal p.
And these numbers definitely come in pairs, since if you take m to be a number in the range 0 < m < p, then you can take n = p-m.
1
-
1
-
1
-
So the 3d example with 4 colors has rule where the vertices are (1,1,1,1), (2,2,0,0), or (4,0,0,0). Which feels like quaternion algebra can be integrated into this in some way. Though, I thought the same with the pascal's triangle with complex numbers so I might just be crazy.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Something that doesn't 100% make sense to me. At least with how we describe the operations we do to the polygon at the start, rotating the points doesn't make it a different polygon. By which I mean, if we describe a polygon with an ordered list of points, and we rotate that list (relabel the points), the polygon doesn't actually change, and nor does it affect what the operations do.
So actually, we don't need all 10 degrees of freedom that the basis eigen-pentagons give us for example. Using the basis shown at 17:28, let's say the components are E1,E2,...,E5 a regular pentagon could be described as c1E1 + c2E2 or c1E1 + c5E5. Actually, wouldn't even a linear combination like c1E1 + c2E2 + c5E5 work? So wouldn't doing the steps of 144deg and 216deg ears suffice for getting a regular pentagon?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The easiest language to program the coding challenge in is Mathematica.
It's a one-liner: PartitionsP[666].
XD
I did program my own using recursion just to check that Wolfram did his correctly. He did.
indexes[n_] := Accumulate@PrependTo[Riffle[
Table[x, {x, 1, n}],
Table[2 x + 1, {x, 1, n}]
], 1][[1 ;; n]]
signs[n_] := Table[If[Mod[x - 1, 4] < 2, 1, -1], {x, 1, n}]
part[0] = 0
part[1] = 1
part[n_] :=
part[n] =
Plus @@ Times @@@
MapThread[
List, {part /@ ((n - indexes[n - 1] + Abs[n - indexes[n - 1]])/
2), signs[n - 1]}]
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
So to generalize, if we have m,n€N, gcd(m,n)=1, then we can take the radius of smaller circle to be m cm, and that of bigger circle to be (m+n) cm, then we can produce (m+n) pointed star,
m regular n-gons and n regular m-gons.
I was wondering if something interesting would happen if gcd(m,n)>1, but turns out nothing better would happen, if m=xd,n=yd, such that gcd(x,y)=1 then the new diagram would just look like a (x+y) pointed star. That's a little disappointing. ☹
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Only remotely related, but I have found the "long division" algorithm to be hilariously and frustratingly circular, to the point of being almost useless.
How come it's circular?
Well, what is the division of some number A by another number B? The result is what number must B be multiplied by in order to get A. In other words, essentially, how many times B fits into A.
So, suppose you have a number A with quite many digits (let's say, 10 digits), and a number B with about as many digits (like 8 or 9, or even 10 digits), and your task is to divide A by B, using long division.
How do you do this? By finding out how many times B fits into A.
Well... duh. That's what we are trying to find out in the first place! The long division algorithm is essentially saying "to find out how many times B fits into A, we start by first finding out how many times B fits into A". Which is a bit hard to do if A is 10 digits long and B is, say 8 or 9 digits long. What a wonderfully circular, redundant and useless algorithm.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I've found out how to generate clean examples of cubics, to give as examples that simplify nicely to integers at every step.
Given the standard form of a monic depressed cubic, x^3 + p*x + q = 0, your clean examples will have the following properties:
1. The p-term needs to be a multiple of 3
2. The q-term needs to be a multiple of 2
3. It doesn't matter if you pick positive or negative for q, you'll simply negate the root when you do (given an example with just one real root). Given a single real solution, a negative q cubic will have a positive real root, and a positive q cubic will have a negative root.
4. When you find a p-value that makes a nice cubic, there will be a related p-value that is its negative, which also makes a nice clean cubic.
I've generated several examples by setting up a spreadsheet. I have a column for p, a column for q, and a column for the discriminant, D = p^3/27 + q^2/4. Then I calculate sqrt(D) and -q/2 + sqrt(D), and then -q/2 - sqrt(D) in the next group of columns. Finally, the two cube roots, and the sum of the two cube roots. Hold your trial values of p as a constant, and increment even values of q. Then duplicate formula and look for rows with all integer terms across the board.
I came up with only two related examples, that A) are non--trivial, B) have a positive discriminant for one real/distinct solution, and C) are within reason to expect students to solve without a calculator. x^3 + 9*x - 26, and x^3 - 9*x - 28. The discriminants will explode to much larger square numbers that very few people will memorize, even if you do get nice numbers to cube root later on (e.g. cbrt(1000)).
A nice clean example with a repeated root is x^3 - 12*x - 16. The real cube roots will add up to the distinct root of this cubic (+4), and the complex cube roots will form two redundant paths of finding its repeated real root (-2).
For cubics with three real solutions, any clean example for Cardano's formula that avoids irrational numbers at every step along the way, can only have one rational root. The other two roots will involve a conjugate pair of square roots, like -1+sqrt(3) and -1-sqrt(3).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
When the Moors left Spain, they left behind 400,000 books on science, mathematics and medicine. They were translated by the Europeans and helped them out of the dark ages and into Enlightenment. Kepler, Copernicus, Descartes, Fermat, Leibniz, Newton, Euler etc., got copies of the books on mathematics, enabling to discover these beautiful mathematics, including calculus, first discovered by the Indians, then taken to Europe by Arabs. Madhava gave the mathematics of infinite series and calculus around 1200 ad.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Reposting and slight editing of recent mathematical ideas into one post:
Split-complex numbers relate to the diagonality (like how it's expressed on Anakin's lightsaber) of ring/cylindrical singularities and to why the 6 corner/cusp singularities in dark matter must alternate.
The so-called triplex numbers deal with how energy is transferred between particles and bodies and how an increase in energy also increases the apparent mass.
Dual numbers relate to Euler's Identity, where the thin mass is cancelling most of the attractive and repulsive forces. The imaginary number is mass in stable particles of any conformation. In Big Bounce physics, dual numbers relate to how the attractive and repulsive forces work together to turn the matter that we normally think of into dark matter.
The natural logarithm of the imaginary number is pi divided by 2 radians times i. This means that, at whatever point of stable matter other than at a singularity, the attractive or repulsive force being emitted is perpendicular to the "plane" of mass.
In Big Bounce physics, this corresponds to how particles "crystalize" into stacks where a central particle is greatly pressured to break/degenerate by another particle that is in front, another behind, another to the left, another to the right, another on top, and another below. Dark matter is formed quickly afterwards.
Mediants are important to understanding the Big Crunch side of a Big Bounce event. Matter has locked up, with particles surrounding and pressuring each other. The matter gets broken up into fractions of what it was and then gets added together to form the dark matter known from our Inflationary Epoch. Sectrices are inversely related, as they deal with all stable conformations of matter being broken up, not added like the implosive "shrapnel" of mediants.
Ford circles relate to mediants. Tangential circles, tethered to a line.
Sectrices: the families of curves deal with impossible arrangements. (The Fibonacci spiral deals with how dark matter is degenerated/broken up and with supernovae. The Golden spiral deals with how the normal matter, that we usually think of, degenerates, forming black holes.) The Archimedean spiral deals with matter spiraling in upon itself, degenerating in a Big Crunch. The Dinostratus quadratrix deals with the laminar flow of dark matter being broken up by lingering black holes.
I'm happily surprised to figure out sectrices. Trisectrices are another thing. More complex and I don't know if I have all the curves available to use in analyzing them. But, I can see Fibonacci and Golden spirals relating to the trisectrices.
General relativity: 8 shapes, as dictated by the equation? 4 general shapes, but with a variation of membranous or a filament? Dark matter mostly flat, with its 6 alternating corner/cusp edge singularities. Neutrons like if a balloon had two ends, for blowing it up. Protons with aligned singularities, and electrons with just a lone cylindrical singularity?
Prime numbers in polar coordinates: note the missing arms and the missing radials. Matter spiraling in, degenerating? Matter radiating out - the laminar flow of dark matter in an Inflationary Epoch? Connection to Big Bounce theory?
"Operation -- Annihilate!", from the first season of the original Star Trek: was that all about dark matter and the cosmic microwave background radiation? Anakin Skywalker connection?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
24:01 the loop 1,2,4,8,7,5 times 3 is 3,6,12,24,21,15
This has to do with Euler's theorem a^phi(n) == 1 (mod n) where gcd(a,n) = 1, in this case a = 2 (second element), n = 9, meaning the cycle has phi(n) elements.
When we times k = 3 we have 3*a^phi(n) == 3 (mod n), same thing occurs where the cycle has 6 elements.
If gcd(a,n) = d > 1 we can always reduce it to gcd(a/d,n/d) = 1 and we can observe a cycle on case 3,6, this case d = gcd(a,n) = gcd(6,9) = 3, a' = a/d = 2, n' = n/d = 3.
The cycle from 1,2 in n' = 3 is from same Euler's theorem a'^phi(n') == 1 (mod n'), where a' = 2, n' = 3, phi(n') = 2.
Basically, all cycles either contains 1 or doesn't contain 1. If it contains 1, the cycle will always have phi(n) elements on graph of n elements, and we can always multiply by some constant k and still maintain the cycle and the shape of that cycle. Else, the cycle will contain phi(n/d) where d is the smallest element of the cycle, which we can reduce the n elements to n/d elements and maintain the structure of the cycle.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
@Mathologer If you change the number of radians in a full circle, you can get plots with very few spokes rather than spirals, that neatly demonstrates the point made by this video. That is, a spoke with all primes ending in 1, 3, 5, 7, 9 etc.
The interesting point there is the angle at which these spokes radiate from the core. which you think would be the same, but instead often resemble Y shapes
I moved on to plotting twin primes, but didn't find a solution to any open clay institute problems, then I had to go back to my day job :)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I did learn the "twisted square" proof at some point in high school, I am sure. So I won't be asking for a refund of the school taxes my parents paid in the 1960's. The "official proof", though, that we were required to memorize, involved drawing an altitude on the hypotenuse, which creates 3 similar triangles. An elegant proof itself, though it does require a fair amount of algebra. The twisted square proof involves no more algebra than the square of a binomial, which itself can be easily visualized by geometry. Areas that add up is very intuitive--though the Mathologer's complex re-arrangements are beyond my capacity to visualize. There's just one thing about the twisted square proof that, to my mind, takes it down a notch in elegance: you actually need four identical copies of the triangle. I've tried to think of a way to dispense with the 4 copies. But President Garfield's proof does dispense with half of them, a good thing. The idea of any of our recent presidents coming up with a geometric proof is laughable, with the possible exception of Bill Clinton (a voracious reader, I'll bet he reads STEM too).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
9:08 With one of each of the first n powers of 2 coins, you can make any amount up to 2^(n+1)-1 exactly one way. This is like the base 2 representation of natural numbers in (I believe it's called) positional systems, which we use with arabic numerals. There's a general theorem for this, so, for example, you can make any amount up to 3^(n+1)-1 with TWO of each of the first n powers of 3 coins (1, 3, 9, 27, etc...) exactly one way; the obvious example is NINE of each of the first n powers of 10 coins (1, 10, 100, 1000, etc...). ^.^
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The number of states in the transition graph is given by:
S(1) = 3
S(n) = 3S(n-1)
S(n) = 3^n
S = 3, 9, 27, ...
The length of the path that visits each state exactly once is 1 less than the number of states.
L(n) = 3^n - 1
Interestingly, the number of states is equivalent to the perimeter of the Sierpinski triangle.
In fact, you can create a 2-graph (borrowing terminology from category theory) where:
The vertices are the vertices of the Sierpinski triangle.
The 1-edges are the edges of the Sierpinski triangle and the states of hanoi.
The 2-edges are the edges of the transition graph.
I drew these 2-edges in a different color on top of a Sierpinski triangle, and it looks kinda cool.
You get these sawtooth patterns because of how the smallest triangles are rotated.
To understand how these two structures are related, consider the construction process.
In the case where n=1, both graphs are a simple triangle.
The number of nodes is equal to the number of edges, so both are 3.
Each combines 3 graphs of the next smallest size to build itself.
They differ in how the smaller versions are combined.
The Sierpinski triangle overlaps 3 vertices, while the transition graph doesn't.
No new edges are created with the Sierpinski triangle, while 3 new edges are created for the transition graph.
This gives 3 different recurrence relations:
V(n) = 3V(n-1) - 3
S(n) = 3S(n-1)
E(n) = 3E(n-1) + 3
Each could be solved using the formula for geometric series, but there's an alternate way of determining the number of edges.
L(n) = (2/3)E(n)
The length of the path that visits each state can also be constructed recursively.
L(1) = 2
L(n) = 3L(n-1) + 2
Once you leave a subtriangle, you never need to return, so you ignore an edge at each level of iteration.
This ends up multiplying each term of the series by a constant.
E(n) = 3 + 3^2 + ... + 3^n
L(n) = (2/3)3 + (2/3)3^2 + ... + (2/3)3^n
And this constant can then be factored out.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
What didn't work for me: trying to watch this at 1.5 speed to save one-third of fifty minutes, lol.
I wasn't a math major, but I'm an electrical engineer, so I'm familiar enough with sums, calculus, matrices, differential equations, etc. to follow your videos on more pure mathematics. Other than that, I just like watching channels like yours and 3blue1brown.
I was able to follow your presentation, though I needed some time to think about some of it. That's fine though, because it's a video, and I can rewind and pause as much as I want.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I haven't seen all your videos but this is the best one I've seen. I love the geometric explainations and proof, I could watch an endless amount of anything connecting higher mathematics to geometry. I'm a mature student in my 3rd year of a physics degree, and one of the things that got me back in school is yes, prepare to groan, numberphile's -1/12 videos, as annoying as mathematicians seem to find it haha. So anything relating to infinite series' gets my attention. I had heard of Bernoulli numbers but dind't know what they were until this videos and now I'm SUPER interested to learn more. I'm also learning a lot of linear algebra at the moment so the matrix expressions you used were really eye opening. As for the length, i often see your videos and say to myself, "ah don't have time" But honestly anything over roughly 12 minutes I begin to think that, however, whenever I do watch one of your long videos I leave it thinking "wow I wish I watched that sooner" I never mind spending the time. Maybe you could do something like have a companion video for the longer ones, you could get away with putting a lot more clickbaity stuff in it, tease some of the more fascinating questions answered in the video, and if people watch both, you're getting double the views. I mean the videos are almost as long as movies now, and movies have trailers, so why not!?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
@Mathologer I've been in love with the 2 x 2 = 2 + 2 equation since childhood. As a more adult mathematicians I have been looking for 'categorizations' of this equation, i.e 2-dimensional vectorspaces V where V \tensor V = V \oplus V in some natural, or meaningful way.
So far I came up emptyhanded. The two natural candidates: the general equation V \tensor V = S^2 V + \bigwedge^2 V applied to the dim V = 2 case and the decomposition into irreducible representations of V \tensor V in case V is the defining rep of the group SL(2) both reduce to 2 + 2 = 3 + 1. (One can argue that they are in fact the same example, just viewed as special cases of different general phenomena). I find it quite annoying how 2 + 2 = 3 + 1 seems to beat 2 x 2 = 2 + 2 to the punch every time, but maybe you or one of the other readers knows an example.
I have one example that sort of works, but in an artificial way. View C as a two-dimensional R-algebra. Then, if I am not mistaken, C \tensor_R C is isomorphic (as an R-algebra) to Mat(2, R), so simple and hence we get 2 x 2 = 4 rather than 2 + 2. However. If we interpret C \tensor_R C as a C-algebra in the same way that we can interpret C \tensor_R A as a C-algebra for any R-algebra A, then we get, as a C-algebra that C \tensor_R C = C \oplus C. So in a way this is an example of 2 x 2 = 2 + 2, although here it is perhaps fairer to say that we have an example of 2 x 2 = 1 + 1 for some very fat notion of 1
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
This polynomial looks like the Ehrhart polynomial of some 5-dimensional polytope. I wonder if it is...
Edit: I got it!
Each configuration of coins that together forms one dollar can be thought of as a point in Z^6 where
x_1 + 5 x_2 + 10 x_3 + 25 x_4 + 50 x_5 + 100 x_6 = 100
and where each coordinate is positive.
This gives us a polytope with vertices
(100, 0, 0, 0, 0 0)
(0, 20, 0 , 0, 0, 0)
(0, 0, 10, 0, 0, 0)
(0, 0, 0, 4, 0, 0)
(0, 0, 0, 0, 2, 0)
(0, 0, 0, 0, 0, 1).
The Ehrhart polynomial of this polytope will then be a formula
E(k) = #ways to form k$ in coins.
From the fact that these vertices form a simplex (they are all linearly independent) and there are 6 of them, we can see that the polytope is 5-dimensional, which implies that the Ehrhart polynomial has degree 5.
A polynomial of degree 5 has six coefficients, so it's uniquely determined by six values. These can be computed explicitly like in the video:
E(0) = 1
E(1) = 293
E(2) = 2728
E(3) = 12318
E(4) = 38835
E(5) = 98411
From this, we can see that E(k) is given by
E(k) = 40/3*k^5 + 65*k^4 + 112*k^3 + 161/2*k^2 + 127/6*k + 1
which is exactly the polynomial in the video!!!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
okay, maybe one can do that for stuff like triangle numbers and such
hmm. let's say a function tri(n) gives a number that is triangular number #n. tri(n) = sum(n+n-1+...+2+1) = n*(n+1)/2 = tri(n-1) + n
1 3 6 10 15 21 28 36
triangle numbers can't have a trivial example of tri(x-1)+tri(x)=tri(x+1) because tri(x+1) = tri(x) + x+1 which would mean tri(x) = x+1 which would mean the difference between tri(x) and tri(x-1) is just 1, and there is no such triangle number pair.
But what if they don't have to be consecutive. tri(n) outpaces n, so for example tri(3)+tri(5)=tri(6), tri(4)+tri(9)=tri(10) by definition, and there is an infinite amount of these triples;
n+tri(n-1)=tri(n) by definition for every n = tri(x), x>2
But not easy to analytically get patterns for more numbers.
square numbers proven to work, as well as linear
next is pentagon
1, 5, 12, 22, 35, 51, 70, 92
12+22 is barely lower than 35, 22+35 barely higher than 51, no consecutive pent(A)+pent(B)+pent(C) either. I guess cause its n value is actually barely higher than 2 for A^n+B^n=C^n
hexagons?
1, 6, 15, 28, 45, 66, 91, 120, 153
15+28 barely under 45, 28+45 way above 66. same problem
Now question, how do we calculate the n value for A^n+B^n=C^n in these cases.. it's related to some surface dimension shit isn't it, the sequences we have have some bs inside.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The path in the tree to the pythagorean triple [153, 104, 185] goes like this: [1,1,2,3] -> right -> [1,2,3,5] -> right -> [1,3,4,7] -> right -> [1,4,5,9] -> left [9,4,13,17]
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Proving the Riemann hypothesis: Moments after a Big Bang, dual numbers to be considered at the origin, repulsive and gravitational forces are balanced, but dark matter is breaking up. "Dual numbers?", you might ask. Isn't that the square root of zero? (Actually, it's a great way to start a proof that ends with division by zero - a vertical line, and it corresponds to what I call Doyle's constant that deals with a brief window in time starting with taking the eth root of e and ending with a vertical line.
Delanges sectrices for the unflaking of dark matter, Archimedean spiral for spinning up dark matter, Ceva sectrices for that dark matter breaking into smaller, primordial dark matter and black holes, Dinostratus quadratrix for impacting surrounding black holes, and Maclaurin sectrices for dark matter being slowed down over time. Split-complex at one, noting the more gravitational nature of diagonal ring/cylinder singularities of baryonic matter from broken dark matter. Using repetitious bisection (from arbitrary angle trisection), one gets added (SSA) for when dark matter dominates a universe and one gets added for when black holes and rarified singularities make things more gravitational, leading to a Big Crunch. At 2 = a Big Crunch, the solution to the Basel problem, and the "bellows method" drives things. A Big Crunch event needs the net gravity from the localized area AND gravitational effects from surrounding universes. The 2 represents how locked up black holes are making a stacking, planar contact.
Everything else are harmonics - effects on surrounding universes.
I see this proof as like trying to climb a steep cliff with the process of dark matter breaking up, getting to the top of the hill (like stable atomic nuclei) and then skiing down the "hill of many black holes" past an inflection point where black hole formation and spaghettification happens, leading to the final vertical line tensor of the Maclaurin trisectrix. This is much like I have posted about considering the recursive form of Euler's Identity as being like an old incandescent light bulb that you can see having first, second, and third degree critical points.
Ramanujan's Infinite Sum of negative one-twelfth: each positional jostling, mass, and gravitational/attractive exchange between universes (defined by Big Bounce events, considering Ramanujan during both Big Crunch and Big Bang sides) ADDS up to evenness (or a partner in marriage - like half of the whole, I guess).
Net gravity increases over time from the breaking up of dark matter, much like a vertical asymptote. It continues to increase, but hits a critical point, much like a vertical tangent, when enough dark matter has weakened and been broken up into the more net gravitational baryonic matter (split-complex numbers). Then, the effects of gravity reach their upper limit - much like a vertical line - at e to the negative e power (where black holes have been flattened).
Over a decade ago, I said that the potential energy of a Big Bang was 1/6 times (the speed of light, cubed). If I had said Big Bounce instead, I would have been correct, in a sense. Now, I see how the speed of light would have to be the fourth power.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Another cool property of Fibonacci numbers: Take any 3 consecutive Fibonacci numbers: 55,89,144. The difference of squares of the larger 2, divided by the smallest, is the next Fibonacci.
(144^2 - 89^2)/55 = 233. 233 is the next Fibonacci after 144. 14930352, 24157817, 39088169. (39088169^2 - 24157817^2)/14930352 = 63245986. 63245986 is the next Fibonacci after
39088169. Of course, excepting 0 as the first Fibonacci, as dividing by it is undefined, use 1 as the first. Thus, in interesting ways the Fibonacci numbers are intertwined with the squares. Wonder what other connections are yet to be discovered.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
35:30 this is not formal by any means, but intuitively I would say it is because the bugs are travelling perpendicularily at every instant.
I will try to formalize it using calculus and vectors:
let b1=(x1,y1) be the first bug, while b2 be the second bug. Let R be a clockwise rotation of 45 deg about the center of the square. WLOG we say that point is the origin, meaning that R can be expressed as a matrix.
It is clear that b2=RR*b1=(y1,-x1)=(x2,y2)
let d(t) de the difference between b2 and b1: d(t):=b2(t)-b1(t)= (RR-I)b1(t)
Computing the matrix of RR-I=[[-1 1],[-1 -1]] we can see that this is a cw rotation by 135 degs, and an enlargment by sqrt(2)
=> RR-I=sqrt(2)*RRR
let D(t) be the size of d(t): D(t):=|d(t)|=|sqrt(2) RRR b'|=sqrt(2)* |b1| because rotations don't affect length size.
We also have D(t)=sqrt(d.d).
We will prove D(t)'=1:
D(t)'=sqrt(d.d)'=1/2*1/sqrt(d.d)* (d.d)'=(d.d')/2*D(t)
let dT be the transpose of d
(d.d)'=(dT * d)'= dT'*d+dT*d'=(d'.d)+(d.d')=2(d.d')=2 * Transpose(sqrt(2)*RRRb1') * (sqrt(2)*RRRb1) = 4 * Transpose(RRRb1') * (RRRb1)
= 4 *transpose(b1') * R^(-3)*R^3*b1= 4 (b1.b1')
Since b1 moves in the direction of d, then the angle(b1,b1')=angle(b1,d)=135 degrees because d=sqrt(2)*RRR*b1.
Since the speed is 1 we have |b1'|=1
Solving for (d.d)' we get (d.d)'=2*d.d'=4 (b1.b1')=4 |b1| cos(135)
Pugging back in we have D'(t) = 4*|b1| cos(135)/2*D(t) =4*|b1| cos(135)/2/sqrt(2)/ |b1| = 1
D(0) = sqrt(2)*|b1| => T = D(0)/v=sqrt(2)*|b1| which is equal to the time it takes to reach on point of the square.
This was much harder than I expected. Probably you can get better proofs using drawings and geometry. Not only is my proof long but it requires a lor of familiarity with linear algebra and some non trivial properties
1
-
1
-
1
-
1
-
1
-
1
-
1
-
In my opinion the simplest way of looking at bugs at the corners of a square--or for that matter, a pentagon, hexagon, etc.--is to draw, at some instant of time, a circle circumscribing the square. If the bugs were chasing each other along the circumference of the circle instead of directly toward each other, the circle's radius would remain constant. Instead, there is a component of their motion in the tangential direction and a component in the radial direction. Since the figures are similar at any point in time, the angle of the velocity vector between the tangential and radial directions is always the same. Hence, the radial and tangential velocities are both constant, so the radius shrinks to zero at a constant rate. The constant tangential velocity means the angular velocity becomes larger and larger near the center, reaching infinity at zero radius. The centripetal acceleration does too--impossible for bugs of non-zero mass. However, the path is simply a spiral. The time it takes to reach the center of the square is simply the half-diagonal of the square divided by the radial velocity, and the total path length is that time multiplied by the total speed. Since, for a square, the tangential and radial components are equal, it's pretty easy to see why the path length should be as the Mathologer claims.
I'll have to think a bit more about an equation for the spiral. It won't be a constant pitch.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Not sure if I got it right, but if the input sequence is the square numbers, then this could be encoded by the sequence (a,b,c,d,...)=(1,2,2,2,...).
That works, because 1^2=1, 2^2=4=2*1+1*2, 3^2=9=3*1+2*2+1*2, 4^2=16=4*1+3*1+2*2+1*1. In general: n+2*(1+2+3+...+n-1)=n+(n-1)*n=n^2.
So feeding (1,2,2,2,...) into the Moessner machine yields: n!*(n-1)!=n*[(n-1)!]^2.
1
-
How did I unexpectedly "Bump" into this very Video? Well, a very odd cause! A few months ago, I was treading this hall of Clifton Communicare Center on Probasco Street in Cincinnati, Ohio! This is where I have lived in the last 4 years. This very hall has my Room 319B on the side facing the rear view through the windows, giving on some evenings a beautiful scene of the orange setting Sun over the Ohio River and Downtown Cincinnati way below the hill. Well, on a certain day, I was about to turn to the drinking fountain at the turn of the hall to the front entrance/exit out of the building. This is where the cold-water fountain is. It probably was in August on a very humid day. But then my eyes were looking at my tennis shoes, when I suddenly noticed the floor area near me. This suddenly prompted me to close in on this particular floor spot: it was not only near to that cold water to fill in my drinking cup, but also quite near to the nurse's stand, from whence to get a straw to sip the water in my cup. Well, at that particular moment, with no one near around me to notice what I was seeing, I found this specific aged paper wrapper on that very Spot that once came from a now long-ago discarded straw inside it! What was strange was that it was twisted out of regular shape somewhat for being on that spot and trampled on by passersby for many hours. When my eyes focused on it at close range, being no longer a straight line, to originally enclose a brandnew straw, the very shape perfectly resembled THE SQUARE ROOT OF 3 lying on the floor! UNUSUAL! I wish I had a cell phone to save the very image of this to show others or save in my album. That was months ago. I suddenly recalled seeing that square root of 3, while viewing U tube videos, hence, to look up anything unusual about that number. And that's how I bumped into your video!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I’m kind of confused. Given an odd prime p, and that we know p = a² + b². As you noted, one must be odd, the other even, assume (as you did) a is the odd one. Then let a = 2x+1, and b = 2y. So p = (2x+1)² + (2y)² = (2x)²+2(2x)+1² + (2y)² = 4x² + 4x + 4y² + 1 = 4(x²+x+y²) + 1, thus p mod 4 must be 1. In fact, this should hold true even if p is not prime, as long as p is odd and can be written as the sum of two squares… which if we allow for the case of y=0 would imply that all odd perfect squares mod 4 is 1.
Which, given f(n) = n², then for n=1 we get f(1) = 1, which mod 4 is 1. Given a base case f(n) that is odd, then the next odd perfect square is at n+2, and f(n+2) can be shown to be larger by f(n) by the sum of two sequential odd numbers, let those be 2x-1 and 2x+1, which when added together give: 2(2x)+1-1 = 4x, which means that for all for all natural numbers, f(n) mod 4 = f(n+2) mod 4. (This later identity holds for all natural numbers n, not just odd ones.) And since f(1) mod 4 = 1, then all f(2n+1) mod 4 must also be 1. So, yes, all odd perfect squares mod 4 are 1.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Regarding the change of the sign after cycling:We have p cards. The card to be cycled has (p-1) cards to it's left, of which a certain number, call it N, will be bigger and the rest, (p-1)-N, will be smaller.
|
By cycling it our card has no more cards to it's left, so the number of inversions decreases by N (since it had N cards larger than it before), but the rest of the cards now have our card as a new card to it's left, so all the cards which were smaller, (p-1)-N, now have a card to their left which is bigger, so for each such card the number of inversions increases by one.
|
This results in a total change of the number of inversions of -N + ((p-1) - N) = (p-1) - 2*N. This means that our sign will be multiplied by (-1)^((p-1)-2*N) = (-1)^(p-1) * (-1)^(-2*N) = (-1)^(p-1).
|
So for an odd number of cards, we get (-1) to an even number, so the sign stays the same, for an even number of cards we get (-1) to an odd power, so the sign flips.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The "combine F_n through F_{n+3} to get F_{2n+3}" identity reminds of a similar elegant identity: if you square the matrix {{F_{n-1}, F_{n}}, {F_{n}, F_{n+1}} then you get the matrix {{F_{2n-1}, F_{2n}}, {F_{2n}, F_{2n+1}}.
These kinds of identities are useful because they allow you to skip ahead in the Fibonacci sequence without computing all intermediate positions, while using exact integer arithmetic. In fact the identity I mentioned comes directly from the algorithm for computing large Fibonacci terms by repeated squaring of the matrix {{1,1},{1,0}}.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Hello! yes, I'm getting to the end of your videos and really enjoying them thanks. I've taken the bait, and I must say that I was also inspired by Martin Gardener and other similar writers while I was in my teens and I went to the University of Leeds and got a degree in Mathematics (1994). I was one of the few you mentioned in another video who took a third year course in Galois theory (partly because the other option was something to do with statistics and by my third year I'd had enough of that). Unfortunately I was really, really disappointed by the lecturing style and lack of engagement at university. Added to that was Godel's incompleteness theorem and Maurice Kline's book 'Mathematics, the loss of certainty'. The final straw was my discovery of a world of literature and philosophy and before I finished my degree I had turned thoroughly against everything "scientific" and "mathematical" (meaning things mired in mainstream scientific methodology). A few years later I spent a semester lecturing Cultural Studies and Art students about postmodernism, psychoanalytic approaches to art criticism, and other similar things, and some time writing about ethics and the crisis of rationality before throwing myself into setting up small charities and working with people mainly discarded by society. Now I'm living in a remote part of Eastern Poland building a little homestead and helping my daughter with her schoolwork and rediscovering the beauty I found in maths all those years ago. I think that your videos and our mathematical achievements are amazing but I probably won't watch too many more. It seems very likely that man made climate change is going to destroy our civilisation. Maybe we got so enchanted by the abstract beauties we've constructed, that we were blinded to the destructive nature of the lives we've been living. Very easily done, and I don't blame anyone, but I need to get back to my garden, my bees and my goats. Thank you Mathologer and all he best for the future
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Hello Mathologer,
Best Season’s and Holiday’s Greetings to you, the Mathology Team, and to every viewer!!
Please can you help me by telling me something - very simple - but it’s just REALLY bugging me, like an itch I can’t reach!!!!
Ok: at about 8:20, you mention that we can “ignore the ‘strange brackets around (in this case) the ‘4s’ (the fours).’
I know what they mean, and what they do, how they can be “upper” or “lower” (with the squiggly line at the top or bottom of the bracket); BUT.......WHAT ARE THEY CALLED? WHAT’S THEIR CORRECT NAME?? Aaaaargh it’s driving me nuts - PLEASE can someone help me out!!!!
Thanks in advance, and all have a good holiday period, and goodbye to one of the craziest, weirdest years EVER!!!!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
x = -(((3 - 2 Sqrt[2])^i - (3 + 2 Sqrt[2])^i)/(4 Sqrt[2]))
y= 1/2 (-1 + 1/2 ((3 - 2 Sqrt[2])^i + (3 + 2 Sqrt[2])^i))
from Mathematica.
The first ten solutions (x,y)
(1,1)(6,8)(35,49)(204,288)(1189,1681)(6930,9800)(40391,57121)(235416,332928)(1372105,1940449)(7997214,11309768)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Help! So each cake slice of a hemispheric cake corresponds to a crescent-curved cake slice of the cylinder-minus-cone cake, right? The base of the slice is a "curved triangle", where two edges are half-circles of radius 1/2 (assuming the hemisphere and cylinder-minus-cone have radius 1) and thus length pi/2, and the third edge is an arc of radius 1, and a length of 2pi/n (n is the number of slices).
Looking at the cylinder-minus-cone, its cross section is of course two right triangles with a vertical height of 1 ( the height of the hemisphere and the rim of the cyl-min-cone), and horisontal side 1, the same radius. But that is not the shape you get from flattening the infinitely thin but crescent-become-half-circle curved shape of the cake slice limit. Its height is still 1, but its base is pi/2. These two edges are straight (one having been flattened) and at a right angle, but what is the length - and shape - of the third edge? It's probably not a straight edge, as it has to have a horizontal tangent at the top, and also at the bottom, but is it simply a cosine graph from 0 to pi scaled to height 1?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Attempt to prove: Let an odd prime p=x^2+y^2=4k+i. Since p is odd, i is odd, and due to the 4k, we can limit i to be either 1 or -1.
so 4k = x^2+y^2-i. As i is odd and 4k is even, either x^2 is even or y^2 is even. Assume y^2 is. Then 2 divides y^2, and since 2 is prime, 2 divides y. So 4 divides y^2. So 4k = (y^2)+(x^2-i). So x^2-i = 4j for some j, where x is odd.
If i = 1, then x^2-i=x^2-1 = (x-1)(x+1), which 4 divides because x is odd, and thus both x-1 and x+1 are even.
if i = -1, then x^2-i = x^2+1. Note if x=1, 2 isn't divisible by 4. Assume x^2+1 isn't divisible by 4 and x is odd. (x+2)^2+1 = (x^2+1) + 4(x+1). Since x^2+1 isn't divisible by 4, (x+2)^2+1 isn't. So by induction x^2+1 isn't divisible by 4 for any odd x (well even x too, but that's easy to see). so i can't be -1.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
My highest in Maths was from my electrical engineering course in my polytechnic days. I didn't think too much on Maths then, didn't really put two and two together, just thought it was cool that some people managed to find correlations between all these funky terms.
Then, the Maths videos populated and popularised. And now I like Maths a lot, whether be it basic primary-school level logic-based or advanced calculus Maths.
For the 1-100 sum, I broke it down like so:
1+2+3+...+10 = 55,
11+12+13+...+20 = 155,
21+22+23+...+30 = 255,
31+32+33+...+40 = 355 and so on.
...
91+92+...+100 = 955.
So we have (1+2+3+...+9) x 100, and ten lots of 55. Basically, I compartmentalised the sums and broke it into easier bits to add.
We'll get 4,500 + 550 giving us the iconic 5,050. Still pretty cool in my book, even if the triangle formula was not invoked.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
19:10 I noticed that within the Arctic circle, there are a higher concentration of red tiles near the red corner, and less red tiles close to the orange corner. It's like a smooth gradient. Same is true for the other colours. Will this be true most of the time as it approaches infinity?
Also, 14:10 I got 1,2,3,5,8 which is Fibonacci. I came up with a simple proof: for a board of size 2 by n let the number of tilings = t(n). The end pair of tiles on the right can be filled with either a single vertical tile, or two horizontal tiles. If it is filled with one vertical tile, the remaining board is just the board of size n-1. If the end two tiles are filled with two horizontal tiles, the remaining board is of size n-2, and so the total number of tilings is t(n) = t(n-2) + t(n-1) which is Fibonacci. (As t(1) = 1 this is proof by induction, and it is assumed that t(0) = 1)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Let's say the cards you have are the 2, 6, and 10 of clubs and also the ace and 5 of hearts.
You have quite a few options for the two cards you're going to base things off of (you only need them to be of the same suit). For the sake of demonstrating the trick, let's say you use the 2 and 6 of clubs for the two cards you're going to base everything off of.
The distance starting from the 2 and ending at the 6 is 4. The distance starting from the 6 and ending at the 2 is 9. The smaller of the two distances is 4. So you keep the end card, the 6 of clubs. The top card on the stack you give to your assistant is the start card, the 2 of clubs.
Now, the other three cards are the 10 of clubs, the ace of hearts, and the 5 of hearts. The natural ordering of these cards has suit alphabetically first, then card value. So B = 10 of clubs, M = ace of hearts, and T = 5 of hearts. You want to communicate "4" to your assistant, so you consider the fourth permutation of BMT, which is MTB.
So you give your assistant the cards in this order: 2 of clubs, ace of hearts, 5 of hearts, 10 of clubs. (You keep the 6 of clubs.)
Your assistant receives the cards in the that order: 2 of clubs, ace of hearts, 5 of hearts, 10 of clubs. Since the top card is a club, your assistant knows that the secret card you kept is a club.
Next, your assistant notes the following three cards: ace of hearts, 5 of hearts, 10 of clubs (in that order). The assistant notes that the standard order gives B = 10 of clubs, M = ace of hearts, T = 5 of hearts. So these three cards are in MTB order, which is the 4th permutation. This means that the secret card is 4 away from the start card. The start card is the 2 of clubs. 4 away from the 2 of clubs is the 6 of clubs!
The assistant says, "6 of clubs!" You reveal your hidden card, the 6 of clubs.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
For the proof of the 4n by 4n colored grid:
Any 4n grid in which n > 1 simplifies into an eventual quadrant which contains a 4 by 4 square at n = 2.
As a callback to the video, consider the original problem from the film clip. Instead, express this row of cards as a one dimensional grid, with all possible configurations of the domain to color its space with one of the two available combinations: black and white.
Extending this model into the complex plane results in a two element pairing of black/white options for each element in a two-dimensional grid, giving four options for a possible cell configuration:
(0, 0)
(0, 1)
(1, 0)
(1, 1)
Visually, this works best as distinct colors instead of pairs of complex binary numbers.
As for the base case of n = 1, again the video of the original problem demonstrates that flipping adjacent values "swaps" the least significant bit with its neighbor, creating the illusion of the smiley face moving to the left. In reality, this operation is the temporary conversion of each overlapping pair of points into a complex value, swapping the real and imaginary bits (if applicable), and returning them back into their original form. In complex numbers from the two dimensional grid, these instead form temporary values containing a real number for the real part on the left hand element, a complex value for the imaginary part of the left hand element, and for the right hand element, a complex number for the real part, and a real number for the imaginary part:
(1, 1) <-> (0, 0)
(1, (1, 0)) ((1, 0), 0)
This extra information that is duplicated follows the same rules for the original one dimensional space, and when returned, is "lossed" when considering the real value which replaces the complex element.
With a cube, 6n is used, three elements are required per cell, so I'm pretty sure there isn't a solution (I'm not a mathematician). 4n-numbered dimensions only, as further diving into 4, 8, 12, etc. dimensional grid is essentially expansion of each real part with a complex value, etc.
I suspect that the net affect of this sequence of operations maps nicely to some definition of a space filling curve, but I don't think I'd go as far as say that's definitely the case.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Proof of Circle :
Given any two lines, they lie on a circle because the power of the intersection point is the same. This is the product of the lengths of the parts of any chord on which the point lines. If the lengths are a,b,c then the power of one of the points, evaluated through one line is a*(b+c), and evaluated through the other line is a*(c+b).
This makes the 4 points cyclic because given 3 of the points, we can define a circle, 2 of the points are on a single line and that gives the power of the point. We can then determine what point gives the correct power for the other line, and since the point at which the circle and the other line intersect gives that same power (and power is linear with point position along the line), this intersection point is the same as the 4-th point.
Now we find the center of the circle formed. Note that the two lines, lets name them ABC and ACB, are mirror reflections along the angle bisector since the lengths of the sections formed by the point are the same, a and (b+c). Thus the center lies along the angle bisector.
Note that the center also lies along the perpendicular bisector of the two chords ABC and ACB. The distance along the perpendicular bisector is a) identical, can be found bu evaluating position of perpendicular along the lines then noticing the equal angle, hence congruence.
Evaluating another pair of lines and checking the lengths, we see that the center point is the incircle center, and the distance along the perpendiculars is all the same (by congruence of triangles), and is the incircle radius, which are both known to be unique.
Also, the diameter of the circle has to be larger because two of the opposite points form the chords aka whiskers of the triangle, which is smaller than the circle's diameter.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The squares question is quite interesting. I used combinatorics, but I have yet to prove my assumption is true, and it always works, but it is trivial I will leave as an exercise.
The upright squares are as is. That's simple.
As for the tilted squares, it's easier to reduce the case. That is how many squares can be in 2x2, 3x3, 4x4, 5x5 etc. Note that each tilted square is embedded within a slightly larger upright square necessarily since all upright squares have their sides pass through one or more of the points.
2x2 has no tilted, because there are only 4 points. For all other cases, tilted squares cannot have one of its vertices in the "corner" of a grid (proof is trivial :P).
For all not corner pieces, a square can be formed among them.
In 3x3, there is only one such square.
In 4x4, ther are 4 3x3 squares, yield 4 tilted squares. But if we do the edge analysis again, there are now 2 non-corner pieces, hence two more squares. Total 6.
We may conclude there are no more than 6 tilted squares in 4x4, because any other tilted square will be either embedded in 4x4 or 3x3---both accounted for.
5x5 same thing. How many embedded in 3x3 (9x1), how many embedded in 4x4 (4x2) and how many 5x5 (3 non corner each side, 3 = 1x3)
So for the puzzle question:
1+4+9+16 upright
+
9x1 + 4x2 + 1x3 tilted.
Total 50.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Let G be the group of linear operators on V which preserve angles (so like, O(V) \times \br^\times )
Ah, actually, let G be the set of all affine transformations that preserve angles.
If you have two labeled sets of points from V which are related by an element of G (sending each point from one set to the point in the other set with the same label)
then,
well, ((A(y) + y) - (A(x) + x)) = (A(y)-A(x)) + (y-x)
Err... hold on..
Let A’(v) := A(v) - A(0)
A’(y-x) \cdot A’(z-x) = r^2 (y-x)\cdot (z-x) ok, but what about the cross terms?
(y-x)\cdot A’(z-x)
and
A’(y-x) \cdot (z-x)
well, the adjoint of A’ will also be a multiple of its inverse..
hm, not sure that helps.
Is this specific to 2D somehow? It doesn’t seem like it should be..
... then again, maybe it is specific to 2D,
Because in 2D, and sum of two rotation-and-scaling will be a rotation-and-scaling unless it is zero (because they can be identified with complex numbers)
And I’m not sure if that holds true in a greater number of dimensions..
Like, the only division algebras over the reals are like, R, C, H, right?
So... hmm
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Aww, my other, imo, much more interesting comment didn't show up in your comments collection.
(That said, it was only an aside to the video, more of a phi-fact than anything)
Apparently, all such Möbius transforms of φ, (p φ+ q)/(r φ+ s) such that | p s - r q | = 1 and p, q, r, s are integers, will share the property with φ, that they cover an interval as evenly as possible.
Now, (p φ+ q)/(r φ+ s) can be simplified:
(Throughout this I'm using the fact that 1/φ= φ-1 and φ² = φ+ 1 because phi is the larger root of x² - x - 1 = 0)
1/(r φ + s) =
(r / φ - s) / ((r φ + s) (r / φ - s)) =
(r (φ - 1) - s) / (r² + r s / φ - r s φ - s²) =
(r φ - r - s) / (r² - s² + r s ( 1 / φ - φ)) =
(r φ - r - s) / (r² - s² + r s (φ - 1 - φ)) =
(r φ - r - s) / (r² - r s - s²)
So any inversion involving phi can be simplified in this way.
And if you multiply two such values
(a φ + b)(c φ + d) =
(a c φ² + a d φ + b c φ + b d) =
(a c (φ + 1) + (a d + b c) φ + b d) =
(a c + a d + b c) φ + a c + b d
So in particular,
(p φ + q) (r φ + (- r - s)) = (p r - p (r + s) + q r) φ + p r - q (r + s)
And therefore, any such Möbius Transform can be reduced to
((p s - q r) φ - p r + q (r + s))/(s² + r s - r²)
but since we have the condition | p s - r q | = 1, this can be simplified further to
(p φ + q)/(r φ + s) = (p r + q (r + s) ± φ) / (r² - r s - s²)
and since all of these are integers, I think this condition really says
(p φ + q)/(r φ + s) = (a ± φ) / b
where a, b are once again integers. I'm not sure whether a and b can just be any integers, rather than just certain ones, but currently I believe so. If that were the case, even the ± can be dropped, as negative a and b are gonna cover those cases, so we just arrive at
(p φ + q)/(r φ + s) = (φ + a) / b where p, q, r, s ∈ ℤ and | p s - q r | = 1
which happens to be precisely the sorts of numbers you encounter in a Helicone with a variety of leaf counts.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
From the harmonic series, add two, subtract one, add three, subtract one, add five, subtract one, add seven, subtract one, add eleven,.. etc (the primes!).
Thus: 1 + 1/2 - 1/3 + 1/4 + 1/5 + 1/6 - 1/7 + 1/8 + 1/9 + 1/10 + 1/11 + 1/12 - 1/13 + ....
Does this converge and if so, to what?
1
-
1
-
1
-
I am confused how this works for a 2n-gon; your visualization with a decagon was too busy for me to make out the details. In particular, the 180° ears should just be bisecting the lines, right? What does that look like in decomposed form, five points all at one location (x, y) connected to five points all at (-x, -y), with each edge connecting one of the first set of points to one of the second?
Actually I think that makes perfect sense now that I write it out, if I’m correct there. Skipping those degenerate ears would, if I’m correct, result in a decagon where alternating points tend to zig-zag back and forth in a way that’s completely smoothed out by bisecting the edges, which makes sense.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
That is exactly what I thought. We see Fibonacci numbers, and therefore φ, when things are packed in a 2D circle like a sunflower. Also the 2D surface of a pine cone or pineapple, which are not so easy to describe precisely. So we might expect to see ρ and σ when things are packed in a 3D ball.
A 3D analogue to sunflower seeds might be cells in an embryo, specifically in animals with little yolk such as therian mammals, where the fertilised egg is more or less symmetrical. But I don't think that does occur in practice. In the early stages, cells divide simultaneously so the total is a power of 2, and later embryos have differentiated into cells of different shapes and sizes.
Crystals are the most familiar form of sphere packing in nature, but if they use ratios of 1.80 and 2.25 I've never heard of it. The next obvious place to look is quasicrystals, but I see Burkard thought of that already.
I also thought about crystals inside geodes, but they aren't usually close to spherical. And they form from the outside in, so the process happens on a 2D surface, not throughut a volume.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
It is SO gratifying to see this topic tackled here. I did some work on it a few years ago, so I can answer a few questions and add something else.
You asked if anyone had written some code to make a diagram, and I have something rudimentary that only calculates values for me to calculate by hand. Here is my spreadsheet I used to make a pentagram version:
https://docs.google.com/spreadsheets/d/1VpLl383Y46w9LQkqwtimhXY5OoftPDnyQ0ipfadTvh4/edit?usp=sharing
You used a digital root property to explain why 1-2-5-8-7-5-2-1 repeats, but I found something else: you already know that the highest digit in base b, (b-1), is key to the vortex. Specifically, it must be an odd square*. But it will also turn around and loop symmetrically if the *second-highest digit is the digital root of a power of two (if you're doubling). Base 10 has this property, as 9 is an odd square and 8 is a power of 2; the next base with this property is 26, since 25 is an odd square and 24 is the digital root of a power of two (in base 26). There is a video on my channel in which I show all my work on screen, if you'd like to check it out.
https://youtu.be/jWHsJNm-gUM
I didn't encounter these multiplication/divisibility rules until I was a math teacher, looking for shortcuts in the times tables to help my students. I'm very happy to see a channel that inspired me come around to a topic I never would have investigated without your inspiration. Cheers!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Before you even asked how beautiful I thought the proof was, I thought to myself, "Wow! That's beautiful!" I would rate it a 10 on the scale from 1 to 10 :)
Also, for the rotation property... Let's see.
My guess for an explanation is that adding a, b, and -a-b always results in 0. So no matter which two of a, b, and -a-b you pick, the third will be the opposite of the sum of the other two. Picking two of these numbers corresponds to a rotation where those two numbers/colors are on "top".
The reason why the rotation should not work for Pascal mod n for n > 2 because taking a, b, and a+b, the sum of a+b with either a or b is not the other one unless n = 2. (But also, for n = 2, addition and subtraction are the same thing, so the Pascal mod 2 is the same rules as the 2 coloring game.)
I may have missed something in my reasoning, but I have to get going. I will come back to this later. Thanks for the great video!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I'm REALLY interested in the solution to what the parent of the 3-4-5 triangle is. Not sure I have the chops to do it. I'm guessing either something doesn't work OR, my intuition points to an infinite regress of triangles with sides with fraction lengths.
Or maybe it breaks down that there are no (whole) number pairs "before" the first two ONES in the Fibonacci sequence. although I've seen 0, 1, 1, 2, 3, 5...., what would precede 0? Of course, it is interesting that the ratio of any two numbers in the sequence gets ever closer to the Golden Ratio, always going slightly above it after going slightly below it, when you deal with the first few numbers it is as crude as it possibly can be. If the "goal" is 0.618... then working BACKWARDS we get 3/5 = 0.6 (too small), 2/3 = 0.667 (too big), 1/2 = 0.5 (too small, again), 1/1 = 1 (too big, and in fact THE BIGGEST you can have a fraction be between 0 and 1), and 0/1 = 0 (too small, and the SMALLEST you can have a fraction be between 0 and 1).
There are no fractions less than zero that add up to zero, so unless one starts considering NEGATIVE numbers.....
But instead I get this weird thing where the LEFT side of 0 is not simply the negative version of the right, but ALTERNATES positive and negative. Any two integers add to the number to their RIGHT, and subtract to their number on the LEFT:
....5, -3, 2, -1, 1, 0, 1, 1, 2, 3, 5...
The only thing that makes "sense" in having the values LEFT of zero alternate +/- is that by dividing any two neighbors guarantees a negative result, i.e. exact NEGATIVES of the ever-conversing Golden Mean value.
So what does this mean? Can you continue creating these triangles with these sides? But what does it mean to have a triangle with a side length of 0, or a negative number? Is this connected to the "how" it breaks down?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
My proof: label the 3 triangle lengths a,b,c, and their opposite angles A, B, C. connect 4 of the outer points, two from each of two pairs of the extensions (say two of extensions b, and two of extensions c, to form a quadrilateral. Draw from where the b's meet to the broad edge of the new quadrilateral a line with angle A. do the same from where the c's meet, again with angle A. Since each of those origination points have an angle C and A, (or B and A) and lie on a line, the interior angles are B and C respectively. There fore B, a, C must give you a congruent triangle to the original triangle. There fore you now have a couple more isoceles triangles. Now go to opposite corners. You can quickly find that, you'll get, for instance, angle 90 - C/2 opposite angles 90 -A/2 and 90 - B/2 on the other. These add up to 180, and so are supplementary, and so the new quadrilateral is an inscribed quadrilateral. Since there was nothing special about which of the three quadrilateral we drew, all three are inscribed quadrilaterals. You can draw more quadrilaterals in the same way, to find a bunch more of these half angles, and since 3 points define a circle, you can group them so that eventually you find that all 6 points are parts of a whole bunch of inscribed quadrilaterals that have to be on the same circle.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
22:40 For me, the best conceptual realization from this video came from considering the fundamental difference between the cylinder and the sphere: to change the cylinder's relationship to the projection, you increased the height and decreased the area of the cross-section. So why not just do that for the sphere and break the whole argument?
Because the sphere's height and cross-section are linked - they are both dependent upon the radius. So any change in the sphere will be expressed the same way along all dimensions.
From this I realized that circles, spheres, and their n-dimensional analogues can be considered as one-dimensional objects with n derivative (or, perhaps more accurately in calculus terminology, integrative) properties. That is, they all have one dimension: r, and the circle, for example, has circumference and area. Sphere adds volume, etc.
You can even extend this idea to regular polygons/hedra/N* by adding a second dimension: the number of sides.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Reaaally nice video, since the years I'm watching I never get disapointed ! But for the first time ever I'm wondering if I haven't found an even more mathologerized "proof" of Euler's formula ! (or maybe I'm totaly wrong ^^). The basic idea is that every convex polyhedra is isomorphic (homomorphic ?) to a tetrahedron, in the sense that you can smoothly deform any convex polyhedron into a tetrahedron and vice versa if you admit that one line segment can be split into two parralel ones, and one face can be split into any number of coplanar faces (because we are not interrested into positions of the vertices, and so the lengths of the edges and areas of faces).
So you choose four not colpanar vertices in a polyghedra (there will always be) that will become the final tetrahedron, you let them in place, and you move all the other vertices so that they get colpanar with the three closest fixed vertices (three among the four we have chosen). You also have to align some the vertices with te fixed ones (in order to creat new edges), but I struggle to rigorously describe where and when ^^'. In the process, you will align some pairs of vertices that have a common vertex and some faces that have common edges. Then you remove the extra vertices ; those which cut a newly created edge (created by the alignement of two edges) in two parts. Then you remove the extra edges ; those which cut a newly created face (created by the alignement of two faces) in two or more parts. You end up with a tetrahedron !!
In the process of removing the extra vertices, you created one edge from two edges, so you loose a vertex and an edge, so the initial (V - E + F) becomes (V-1 - (E-1) + F) = (V - E + F). In the process of removing the extra edges, you created one face from two faces, so you loose an edge and a face, so the initial (V - E + F) becomes (V - (E-1) + (F-1)) = (V - E + F).
So during the whole process (V - E + F) staid constant ; we just have to calculate it for a tetrahedron which is simple !!!
Of couse, you can do the other way around (which is maybe easier to see) an construct any polyhedra from tetrahedron by splitting edges in half, which creates each time a vertex and an edge, and then by moving this new vertex where you want, which creates each time an edge and a face, so (V - E + F) stays each time constant.
I know it's not a rigorous proof (althought I think we could do a solid demontration from the last two lines above), it is really "intuition based" bur I'm quite proud of having found a very visual and simple "proof" (would be simpler if I had animation rather than text but maybe someone can ;) ). Please master, tell me if my idea is good or if I made a mistake. Thanks a lot for all your content for all this years !!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The first time I encountered generating functions was when I was having fun filling up paper charting how many vertices, edges, faces, etc. there are in a line segment, square, cube, 4-cube, etc. and finding the patterns. I showed this to a math teacher, and he was like, “or you could find the coefficients of (2 + x)^n.” (Where n is the dimension of the cube, and the power of x is the dimension of the face.) I was blown away and mystified and had no idea how it worked.
To find the k-faces of an n-cube, you doubled the k-faces of the (n-1)-cube and added the (k-1)-faces of the (n-1) cube. Which shows that (2+x)*f_(n-1)(x)=f_n(x). And f_1(x)=2+x, because you start with a line segment, with 2 vertices and 1 edge. So you have proof by induction. You could even start with f_0=1, and start with a single point (Pointland).
But the proof by induction isn’t that satisfying, and doesn’t quite bring a greater understanding. I’ll look into it tomorrow morning.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Here's a little program that simulates the tiling of a diamond as shown in the video. To run: copy the code below and save it in a file with type .html. Then load the file with the browser. Attention! Tested only with Safari.
begin of code------------
<!DOCTYPE html>
<html>
<head>
<title>Arctic circle</title>
<link rel="stylesheet" href="styles.css">
<script type="text/javascript" >
const covered = 3;
function draw() {
new diamondPlotter(
parseInt(document.getElementById("grade").value),
800,
document.getElementById('showArea').getContext('2d')).plot();
}
class squareCoverage {
constructor(matrix) {
this.matrix = matrix;
}
get() {
this.coveredSquares = [];
this.copySquaresToCovered();
for (var tile of this.matrix.tiles) {
this.setCoverageOfTile(tile);
};
return this.coveredSquares;
}
copySquaresToCovered() {
for (var row of this.matrix.squares)
this.coveredSquares.push([...row]);
}
setCoverageOfTile(tile) {
this.coveredSquares[tile.y][tile.x] = covered;
if (tile.vertical)
this.coveredSquares[tile.y + 1][tile.x] = covered;
else
this.coveredSquares[tile.y][tile.x + 1] = covered;
}
}
class tileGenerator {
constructor(coveredSquares,tiles) {
this.tiles = tiles;
this.coveredSquares = coveredSquares;
}
get(){
this.fillEmptySquares();
return this.tiles;
}
fillEmptySquares() {
for (var y = 0; y < this.coveredSquares.length; y++) {
var row = this.coveredSquares[y];
for (var x = 0; x < row.length; x++) {
if (row[x] > 0 && row[x] < covered) {
if (this.isFreeSquare(x, y))
this.fill2x2WithTiles(x, y);
}
}
}
}
isFreeSquare(x, y) {
return Boolean(
this.coveredSquares[y][x] < covered &&
this.coveredSquares[y + 1][x] < covered &&
this.coveredSquares[y][x + 1] < covered &&
this.coveredSquares[y + 1][x + 1] < covered);
}
fill2x2WithTiles(x, y) {
if (randomTrueOrFalse())
this.fill2x2WithVerticalTiles(x, y)
else
this.fill2x2WithHorizontalTiles(x, y);
this.set2x2Covered(x, y);
}
set2x2Covered(x, y) {
this.coveredSquares[y][x] = 3;
this.coveredSquares[y + 1][x] = 3;
this.coveredSquares[y][x + 1] = 3;
this.coveredSquares[y + 1][x + 1] = 3;
}
fill2x2WithVerticalTiles(x, y) {
console.log(x + ',' + y + ':vertical');
this.addLeftTile(x, y);
this.addRightTile(x, y);
}
addRightTile(x, y) {
this.addTile(x + 1, y, true, 1);
}
addLeftTile(x, y) {
this.addTile(x, y, true, -1);
}
addLowerTile(x, y) {
this.addTile(x, y + 1, false, 1);
}
addUpperTile(x, y) {
this.addTile(x, y, false, -1);
}
addTile(x, y, verticalFlag, pushDirection) {
this.tiles.push(
{
x: x,
y: y,
vertical: verticalFlag,
pushDirection: pushDirection,
delete: false
}
);
}
fill2x2WithHorizontalTiles(x, y) {
console.log(x + ',' + y + ':horizontal');
this.addUpperTile(x, y);
this.addLowerTile(x, y);
}
}
class squaresExtend {
constructor(oldSquares){
this.squares = [];
this.oldSquares = oldSquares;
}
get() {
this.extendExistingRows();
this.extendWithNewRows();
return this.squares;
}
getOldLastRow(matrix) {
return [...this.oldSquares[
this.oldSquares.length - 1]];
}
getOldFirstRow() {
return [...this.oldSquares[0]];
}
extendWithNewRows() {
this.addNewLastRow();
this.addNewFirstRow();
}
addNewFirstRow() {
this.squares.unshift(this.widenRow(this.getOldLastRow()));
}
addNewLastRow() {
this.squares.push(this.widenRow(this.getOldFirstRow()));
}
extendExistingRows() {
for (var row of this.oldSquares) {
this.squares.push(
[...this.extendRow(row)]);
}
}
extendRow(row) {
return (this.isFullyFilled(row) ?
this.addSquaresToRow(row) :
this.widenRow(this.extendRow(this.narrowRow(row))));
}
addSquaresToRow(row) {
return [row[row.length - 1]].concat(row).concat(row[0]);
}
isFullyFilled(row) {
return row[0] !== 0;
}
narrowRow(row) {
return row.slice(1, row.length - 1);
}
widenRow(line) {
var result = [...line];
result.unshift(0);
result.push(0);
return result;
}
}
class tilesExtend {
constructor(matrix,oldTiles){
this.oldTiles = oldTiles;
this.matrix = matrix;
this.coveredSquares = [];
}
get() {
this.matrix.tiles = [...this.oldTiles];
this.matrix.tiles.forEach(this.moveOldTile);
this.eliminateColliding();
this.shiftTiles();
this.coveredSquares = new squareCoverage(this.matrix).get();
return new tileGenerator(this.coveredSquares,this.matrix.tiles).get();
}
eliminateColliding() {
var index = 0;
for (var tile of this.matrix.tiles) {
this.checkCollision(tile, index);
index++;
};
var tiles = [];
for (var i = 0; i < this.matrix.tiles.length; i++) {
tile = this.matrix.tiles[i];
if (!tile.delete) tiles.push(tile);
}
this.matrix.tiles = tiles;
}
checkCollision(tile, index) {
let compareIndex = this.searchCompareTileIndex(tile);
let array = this.matrix.tiles;
if (compareIndex > 0) {
tile.delete = true;
}
}
searchCompareTileIndex(tile) {
var checkX = tile.x;
var checkY = tile.y;
if (tile.vertical) checkX += tile.pushDirection;
else checkY += tile.pushDirection;
for (var compare of this.matrix.tiles) {
if (compare.x === checkX &&
compare.y === checkY &&
compare.vertical == tile.vertical &&
compare.pushDirection != tile.pushDirection)
return this.matrix.tiles.indexOf(compare);
}
}
shiftTiles() {
for (var tile of this.matrix.tiles) {
tile.x = (tile.vertical ? tile.x + tile.pushDirection : tile.x);
tile.y = (tile.vertical ? tile.y : tile.y + tile.pushDirection);
tile.type = (tile.type === 1 ? 1 : 2);
}
}
moveOldTile(value) {
value.x++;
value.y++;
}
}
class diamondMatrix {
constructor(grade) {
this.grade = grade;
this.matrix = { squares: [], tiles: [] };
};
generate(oldMatrix) {
if (this.grade === 1)
this.getInitialMatrix();
else {
this.lowerMatrix = new diamondMatrix(this.grade - 1);
this.lowerMatrix.generate();
this.extend();
}
}
getInitialMatrix() {
this.matrix.squares = [[1, 2], [2, 1]];
this.coveredSquares = new squareCoverage(this.matrix).get();
this.matrix.tiles = new tileGenerator(this.coveredSquares,this.matrix.tiles).get();
}
createCoveredSquares() {
}
extend() {
this.matrix.squares = new squaresExtend(this.lowerMatrix.matrix.squares).get();
new tilesExtend(this.matrix,this.lowerMatrix.matrix.tiles).get();
}
}
class diamondPlotter {
constructor(grade, size, ctx) {
this.size = size;
this.graphic = ctx;
this.grade = grade;
this.squareSize = size / grade / 2;
if (this.squareSize > 20) this.squareSize = 20;
}
plot() {
this.graphic.clearRect(0, 0, this.size, this.size);
this.matrix = new diamondMatrix(this.grade);
this.matrix.generate();
this.startingPoint = this.size / 2 - this.grade * this.squareSize;
for (var tile of this.matrix.matrix.tiles) {
this.plotTile(tile);
}
}
plotTile(tile) {
var { x, y, yTo, xTo } = this.getTileCorners(tile);
var type = this.matrix.matrix.squares[tile.y][tile.x];
var fillStyle =
(tile.vertical) ?
(type === 1) ? "red" : "blue"
: (type === 1) ? "yellow" : "green";
this.graphic.fillStyle = fillStyle;
this.drawTileFrame(x, y, yTo, xTo);
}
drawTileFrame(x, y, yTo, xTo) {
this.graphic.beginPath();
this.graphic.moveTo(x, y);
this.graphic.lineTo(x, yTo);
this.graphic.lineTo(xTo, yTo);
this.graphic.lineTo(xTo, y);
this.graphic.lineTo(x, y);
this.graphic.stroke();
this.graphic.fill();
}
getTileCorners(tile) {
var x = this.startingPoint + tile.x * this.squareSize;
var y = this.startingPoint + tile.y * this.squareSize;
var xTo = (tile.vertical ? x + this.squareSize : x + 2 * this.squareSize);
var yTo = (tile.vertical ? y + this.squareSize * 2 : y + this.squareSize);
return { x, y, yTo, xTo };
}
}
function randomTrueOrFalse() {
return Math.round(Math.random()) === 0;
}
function arrayRemove(arr, value) {
return arr.filter(function (ele) {
return ele != value;
});
}
</script>
</head>
<body>
Grade: <input id="grade" type="number" value="1">
<button id="go" onclick="draw()">Go</button><br>
<canvas id="showArea" width="800" height="800"></canvas>
</body>
</html>
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Paul Levrie's article is available for download in the link below. It is the only article of that volume that is available for free.
link (dot) springer (dot) com (slash) journal (slash) 283 (slash) 34 (slash) 4
As YouTube does not publish my comment if I include the link, I have put it in disguise, sorry.
As a matter of fact, if you have enjoyed this video, there is a book by William Dunham called 'Euler, the master of us all', which shows important contributions of Euler in different areas of mathematics, along with insights of his developments.
The book includes commented out tour for proofs of the series expansion of the exponential and logarithm functions, as well as how and why he actually named a particular base 'e'. And, of course, the Basel's problem. It really makes you feel Euler's genius.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Has anyone worked with the following configuration of 4 pegs:
Start with 4 pegs connected in a sort of Y configuration, 3 edge pegs connected to a center peg. If on an edge peg, you can only move to a center. If on the center peg, you can only move to an edge. Seems to me, with 1 disk starting on an edge you can get to another edge in 2 moves, with 2 disks starting on an edge you can get to another edge in 6 moves, with 3 disks on an edge you can get to an edge in 12 moves, with 4 disks you can get there in 24 moves (I think, this is where you start second guessing yourself), and at 5 disks, you guessed it... 38? No seriously, this is what I was able to actually repeat if I counted correctly a number of times! Very weird if I'm not messing up.
So, then I decided to try an X shape, 4 edge pegs with one on the center. 1 disk you can get to another edge in 2 moves, 2 disks you can get to an edge in 6 moves, 3 disks you can get to in 10 movies this time, 4 disks seems to be 16 oddly, and 5 disks seems to be 24.
You could do a Hanoi-ish puzzle presumably with any sort of graph structure defining the connectivity and thus possible moves you can make at each peg/node. I wonder what happens in more complicated graphs... can you get any possible configuration in almost any graph (maybe save some particularly restrictive exceptions), or does it become more and more challenging to get to certain states if you keep going? Certainly something I might puzzle over and play with a bit myself.
Edit: I *think*, again easy to get lost in these that actually you can do 4 disks on the Y configuration in 22 moves, not 24.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
So, when thinking of the coin paradox, I realized that distance of the 2 centers of the coins divided by de radius straigh up gave you the answer: for a coin spining over a coin of the same same size, the distance of the centers are 2r, divided by r we have 2; Coin spining over a coin twice it's size, the distance of the centers are 3r, divided by r we have 3; For the coin spining inside the coin twice it's size, the distance is r, so 1 spin; Lastely, for the coin spining a coin 1,5 times it's size [or the spinning coin is 2/3 of the bigger coin], first I "rationalize" the coin sizes by multiplying both by 3, so they became 3r [bigger coin] adn 2r [inside coin], and again, the distance from the centers is r, so 1 spin again.
Can some one back me up with the necessary math? 😅
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
CHALLENGES!
The following two questions were written before I saw any proceeding chapters. So any of the following may be redundant, but I came up with it myself.
23:39 Show that the degenerate point is the centroid.
A: If you know how Fourier Transform works (or at least have the intuition of it from 3B1B's video), this is trivial. But in case you haven't seen it, here's my take:
The centroid of all the other pentagons is the origin. By adding together all the pentagon points the centroids do also add together. The centroid is the average of all the other points, and taking the average is a linear function, which is why this works.
This also gives a way to figure out which pentagons to use. To get the center of mass you add the red point, the black point, the orange point, the blue point, and the green point, and then divide by the number of points.
What if each time you sampled a point you rotated by 72 degrees clockwise first? In the composite diagram with all the regular pentagons, all of our pentagons begin to shift. The degenerate becomes a regular pentagon, one of our directional 72 degree pentagons (the one going counterclockwise) reduces to the degenerate!
By repeating this with the other angles you can figure out what the composite pentagons are that make up our shape.
Congratulations, you have just successfully performed a DFT... kinda.
This also shows that it doesn't matter how many sides we start with, the theorem still holds.
24:03 What happens when you go for multiple rounds?
I'm gonna be switching up my terminology here and there so here:
The 72 pentagon is constructed by going 72 degrees. I also call it the anti-288 because applying 288 degree ears will reduce it to a degenerate.
The shape is gonna be the same regular shape, that's for sure. What's not sure is the orientation and size. The anti-72, anti-144, and anti-216 all reduce to the degenerate all the same, so let's see what happens to the anti-288.
72 will affect the anti-288 all the same.
144 and 216 complement each other (in the 360 degree sense of complement) so they'd undo each other minus a rotation of some sort.
Any set of ears rotates the 72 pentagon 36 or 216 degrees, the 144 pentagon 72 or 252 degrees, the 216 pentagon 108 or 288 degrees, and the 288 pentagon 144 or 324 degrees. The way to choose which one is, is by seeing if the ears "hook" around the origin, that is, if it's closer to the opposite side of the pentagon. If it does, pick the closer to 180 degrees. Heuristics. We'll fix this definition later.
Also the variants to choose from are 180 degrees apart, so if there happens to be a 180 polygon in your decomposition (which will never have this "hook" property), it'll rotate 90 degrees if your angle of ears is below 180, 270 degrees if it's above 180, and it will reduce to degenerate on 180.
We have an additional 144 and two additional 216s. This will rotate our final anti-288 polygon 72 + 108 + 108 + some multiple of 180 degrees = 108 + some multiple of 180 degrees.
Also the ears do scale the pentagon but the gist is since the 216 ears show up on the inside of the anti-288, it's gonna be smaller.
So our anti-288 is also known as the 72-polygon. On an angle of 72 and an outer radius of R, the inner radius or apothem is R * cos(36). The side length is 2 * R * sin(36).
Now to get the short side we need to look at our isosceles triangle. It has base length of 2 * R * sin(36) and it has an angle of 216 degrees. If we slice and look at a right triangle of 72 degrees INSIDE and an opposite leg of R * sin(36), we have the height = R * sin(36) * cot(72).
So we get that our resulting angle is inner radius +- extra (plus if the ears < 180, minus otherwise). More precisely, if R is the outer radius of the polygon , A is the angle of the polygon and E is the angle of the ears, the new outer radius is R * cos(A/2) + R * sin(A/2) * cot(E/2) (the sign of cotangent makes things simpler), and the scale factor is cos(A/2) + sin(A/2) * cot(E/2)
We can now more directly define the whole conundrum about the hook thing: a negative scale factor is equivalent to a 180 degree rotation. And even more: the scale factor is negative on a theta-polygon for angles larger than 360 subtract theta if theta is less than 180 degrees, and for angles smaller than 360 subtract theta otherwise.
Bonus: on a theta-polygon, the angle ears theta/2 and 180 subtract theta/2 will keep the size the same (you can easily check this by inscribing a polygon with double the number of sides and using inscribed angle theorem, or you can use the fancy formula derived earlier)
We can come back to our problem: what happens to our anti-288 polygon? It gets an extra 144 and two extra 216s. Crunch the numbers and we get a 288 rotation and a scale factor of 1 * 1/phi * 1/phi = 1/phi^2 compared to the original.
As a final point: talking about scale factors and rotations brings me to the final point: the ears operation is commutative because 2d rotation and constant scaling are commutative as well.
28:36 The sum of the coordinates of two similar triangles is another similar triangle.
The second similar triangle is a complex number times the first one. Factor out the similar triangle, and you're done.
Note that chirality matters: a triangle plus its mirror or conjugate is a straight line.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
11:10 It is a beautiful thing. 10/10 for '10 is special'
But it uses '4 is special'. It would be nice to prove '4 is special' without 'trial by exhaustion'
If line 1 was A, x, y, B, line 2 is A+x, x+y, B+y, line 3 is A+2x+y, B+x+2y, line 4 is A+B+3(x+y)=(A+B) mod 3
If the 3 colours are numbered 0,1,2 we want 0~0=0, 1~1=1, 2~2=2, 0~1=2, 0~2=1, 1~2=0
Compare with the function 2A+2B mod 3 0+0=0, 2+2=1, 4+4=2, 0+2=2, 0+4=1, 2+4=0 all mod 3
So the single cell in line 4 is 2A+2B mod 3; x and y are irrelevant, lines 2 and 3 are irrelevant, so compacting can begin.
1
-
1
-
Programming challenge:
Oh ye of little faith, it is possible to calculate answers for values like $2 googol quickly with a fairly "computery" technique. It boils down to raising a 191x191 matrix (of big integers) to the 200-googolth power (or whatever). This is a tractable number of multiplications because you can exponentiate by repeated squaring (square a matrix 340 times and you've already exceed the 200-googolth power). (None of this would be tractable if the big integers became too large, but we know they don't if the final answer doesn't.)
My program calculates:
$2: 2728 ways
$20: 53995291 ways
$200: 4371565890901 ways
$2000: 427707562988709001 wyas
$20000: 42677067562698867090001 ways
$200000: 4266770667562669886670900001 ways
$2 million: 426667706667562666988666709000001 ways
$2 billion: 426666667706666667562666666988666666709000000001 ways
$2 trillion: 426666666667706666666667562666666666988666666666709000000000001 ways
$2 googol: same but more 6's and 0's
Two trillion takes about 1.4 seconds, two googol takes about 10 seconds, 2*10^1002 cents takes about 107 seconds, 2*10^2002 cents takes about 257 seconds, 2*10^4002 cents takes about 13 minutes, and that's the highest that I let finish.
The approach is to start with the standard dynamic programming approach to number of ways to make change. (Standard in that this problem is a textbook example or exercise of DP.) This approach takes time proportional roughly to (target number of cents)*(number of types of coins) so it's obviously not tractable for a googol.
However, in the DP approach, when you progress from N cents to N+1 cents, there are only 191=1+5+10+25+50+100 values in the state that will ever be used again. It is possible to use 191 variables (in the mathematical sense, not the programming sense) for the state, and the transition from N to N+1 cents is purely a linear recombination of these variables, so the transition can be represented as a 191x191 matrix and exponentiated, as explained.
Obviously this is slower and less elegant than the generating function simplification in the video, but it is really quite fast until the actual answer has 10000+ digits, and exponentiating big matrices can solve a variety of problems on a computer. You just have to pick the right matrix. :)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Favourite graphical proof of Pythagoras: hmmmm.
I'm not a great fan of "cut them out, move them around and it's obvious it's still the same area" proofs. It's really easy to make it look right while it's actually not (see "chocolate" that I refer to below).
Related but not the same: Penrose, in Road to Reality, tiles the infinite plane with two squares and makes a larger grid by joining equivalent points so the two starting squares add up to the bigger square that makes the grid. That works if you are happy to believe the squares can tile the plane and that the two squares do, in fact, equal the bigger square. That's either circular reasoning or you need a really good argument for why it's true. I suspect the latter is the case, but I am a Bear Of Little Brain and don't quite see it. Perhaps counting the vertices of the grid and showing they map to just one pair of squares does it. Oo. Maybe it does. Anyway: above my pay grade, as they say.
I quite like the first proof with the four triangles at the corners of the square, but I think you have to be careful to just "slide" two of the triangles to the triangle opposite. I think that can be justified rigorously. Spinning them round and shuffling them about really isn't convincing: we've all seen the "cut a bar of chocolate at an angle and make the same bar plus a bit extra" video that (I think) Vsauce did.
To be honest, my favourite proof of the "Pythagoras" theorem is Euclid's: it's not visually stimulating; it's terrible at breaking the ice at a party (trust me) but it's long and convoluted and I can actually remember it, which puts it in a small set of proofs.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I regret only watching this video 10 hrs after the release, the solution of the final problem takes, like, ten minutes, and is pretty straightforward (though I am sure there is a more clever way of dealing with this problem which involves some discrete analogue of the Stawke's theorem, or the use of group theory):
1) 4x4 board
Consider the central square filled with arbitrary different colors a, b, c and d.
ab
cd
There are essentiaaly two cases: bottom right square of the whole board can be colored in any way except d, so it can either be colored a or c/b (these two cases are indistinguishable because of the symmetry of the problem). Let's try a "diagonal" coloring (say, b) first.
0000
0ab0
0cd0
000b
The choice of color for that bottom right corner defines colors of squares immediately to the left and to the top uniquely.
0000
0ab0
0cdc
00ab
(*) It also defines uniquely colors of the next cell to the top and the next cell to the left to be equal to the color of the bottom right corner.
0000
0abb
0cdc
0bab
We've come to a contradiction: one of squares has two b's in it. That means that once we've chosen the tiling of the central 2x2 square, the remaining coloring is defined uniquely, and each corner has the color of the 'opposite' corner of the central square, Q.E.D.
2) By induction: once the coloring of n central square layers is defined, the coloring of the remaining n+1 th layer is defined uniquely, since on every side by the same logic as in (*) and by induction we have two interchainging colors.
abababa
dcdcdcd
and we again have 4 differntly colored corners.
1
-
1
-
1
-
A solution to the "crazy glasses" domino tiling: First, let T(n) denote the number of tilings of the 2×n rectangle. You can partition them into two disjoint classes, depending on whether the rightmost column is a vertical domino (so remaining rectangle is (n-1) wide) or rightmost consists of stacked horizontal dominoes (so remaining rectangle is (n-2) wide). This is the recurrence relation of the Fibonacci numbers, and the base cases of the induction work too, so T(n) = F(n) on the nose.
Now... for the glasses, there are four disjoint classes of tilings, depending on whether the largest center rectangle is tiled as a 4×2, 3×2 on the left, 3×2 on the right, or just 2×2 in the center. Then, the dominoes around the "lenses" of the glasses are determined, as well as the remaining rectangles on the arms, which must be tiled as rectangles, as if they're disjoint. On either side, independently you get a 2×3 followed by a 3×3 rectangle or a 3×3 followed by a 3×3. Counts of all four possibilities (note the center number):
T(3)⋅T(2)⋅T(4)⋅T(2)⋅T(3) = 3⋅2⋅5⋅2⋅3
T(3)⋅T(2)⋅T(3)⋅T(3)⋅T(3) = 3⋅2⋅3⋅3⋅3
T(3)⋅T(3)⋅T(3)⋅T(2)⋅T(3) = 3⋅3⋅3⋅2⋅3
T(3)⋅T(3)⋅T(2)⋅T(3)⋅T(3) = 3⋅3⋅2⋅3⋅3
Thus, the total number of tilings is
T = 9⋅(20+18+18+18) = 9⋅74 = 666 😈
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
45:23 Rewriting my original puzzle answer for the reupload:
In short, 1 1 0 1 1 0 (repeating) is the Fibonacci sequence mod 2.
Why does this work?
Each pair (or n-tuple) of natural numbers, both labeled k, is offset by a certain angle from the pair of 0’s which could be said wlog to not move. This angle is (0.618k)/2, or (0.618k)/n in the general case.
Which branch of the spiral each integer will appear to correspond to is determined by which nth part of the circle its angle is closest to.
(0.618k)/n ≈ m/n in the reals mod 1.
This can simplify to
0.618k ≈ m mod n for some nonnegative integer m < n
We can look at the first two values:
0.618 ≈ 1
1.236 ≈ 1
This is the start of the Fibonacci sequence. It’s easy to see that if two natural numbers a and b add to c, then the corresponding angles A and B add to the corresponding C.
So we can ignore the precise angles and just use the rounded results to get the next one! With two branches,
1 + 1 = 0 (opposite + opposite = same),
1 + 0 = 1 (opposite + same = opposite),
0 + 1 = 1 (same + opposite = opposite)
And the cycle repeats.
(Not rigorous part) Now since these are Fibonacci numbers, as the sequence continues, approximations using these numbers as denominators will become better and better, so the angles will approach whole multiples of 1/n. There is no chance of this pattern being disrupted by accumulating errors.
With three branches the pattern will be
1 1 2 0 2 2 1 0 repeating,
where zero is a spiral of all the same alignment, and twos and ones are spirals of cycling alignment in different orders.
With 4, the pattern is
1 1 2 3 1 0 repeating, with zero being same-alignment spirals, 1 and 3 cycling in different directions between the four, and 2 cycling between one alignment and its exact opposite.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
End screen: Well, I think that 3^3 + 4^3 + 5^3 = 6^3, but I can't find a solution to the quartic, so I guess I made a mistake?
By my figures, the left hand side is 2258, and the obvious candidate for the right hand side is 7^4, but 7^4 is too big (2401). Given 2401 and 2258 differ by 143, I was hopeful that the left hand side might be equal to something nice like 7^4 - 7^3, but 7^3 = 343, not 143. Best I can do is say that the question mark is the 4th root of 2258 (only real solutions between 6 and 7, or between -6 and -7 if you want to be fancy). Being fancier still, we have, for some z,
z^4 = 2258
and so, apply De Moivre's theorem to find all complex solutions, I guess it would be
+/- i (2258)^(1/4) , + / - (2258)^(1/4)
and I'm honest enough to admit, I thought De Moivre would be more helpful!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I decided to answer all of your questions. Most questions are copied straight from the transcript (CC).
0:57 Easy, right?
For now.
1:12 Okay, what if we have to add a color to itself?
Let's just keep it the same.
1:19 That's the most natural rule. Agreed?
Of course I agree; that's what I suggested.
2:29 What are you going to do?
I'll just guess.
2:49 Ready?
No.
2:55 What's the color of the bottom hexagon.
Red? Okay, I got it wrong. Let me work this out...
3:07 Did you get it without pausing the video or just guessing?
Yeah, I just guessed.
3:11 Sounds impossible?
No; if it was impossible, you wouldn't make a video about it.
3:40 Pretty amazing, isn't it?
Yeah, it sure is. Now prove that it always works.
3:53 Weird, hmm?
No; I've watched enough of your videos to that it's not weird.
3:56 What if instead of ten hexagons at the top we started with nine hexagons
It wouldn't work. I did the math.
4:41 Okay so why is 10 special?
It's one more than a power of three.
4:45 And are there any other special numbers.
Yes: 2,4,10,28,82,244,730,2188,6562,19684,59050,177148,... (https://oeis.org/A034472)
4:54 Notice all the smaller solid color triangles here
and there?
Yes, I do notice them. It's pretty hard not to.
5:05 Hmm, can you think of a famous
mathematical supermodel dressed in little triangles? :)
Sierpinski!
5:10 No?
I just answered it. Why did you ask "No?"
5:40 Beautiful pattern isn't it.
Yes; it is beautiful.
5:50 Butcher shop?
Whatever you say.
6:47 Well, let's find out together, shall we? In five easy chapters.
I assume that's what the next 23 minutes of my life will-- Did you say five chapters?
CHAPTER 1 (6:54) - How special is 10?
In the context of this puzzle, it's not too special. There are countably infinite other numbers that do the same thing.
However, 10 is the base of our number system, so it's very special outside of this context.
7:01 Why does the shortcut work for width ten triangles.
"I have a truly marvelous demonstration of this proposition which this margin is too narrow to contain." ~Fermat
Luckily, Mathologer can make a video about it.
7:05 And,
are there any other special numbers?
We've been over this... https://oeis.org/A034472
7:17 And that's where we stopped. Hmm, I wonder why?
You stopped there because 4 is another one of the special numbers.
7:27 Probably also just a fluke. Right?
No.
7:46 In this state can we still show that 4 is
special?
Yes.
8:35 How many essentially different
width four triangles are there?
There are 10 (reflections listed on second row):
AAAA, AAAB, AABA, AABB, ABAB, ABBA, AABC, ABAC, ABBC, ABCA
AAAA, ABBB, ABAA, AABB, ABAB, ABBA, ABCC, ABCB, ABBC, ABCA
I think I did that right.
9:19 Did you spot the pattern?
I discovered the pattern on my own before I watched this far.
10:40 I probably don't even have to say it,
right?
No, you don't.
10:42 Can you see what's happening.
Yes.
11:24 So on a scale from one to ten how beautiful an argument is this?
It's as beautiful as it can be, but it's not as beautiful as some other proofs I've seen, so...
I give it a solid 2π.
12:16 Can you think of a second way to argue? Hint: switch the roles played by four and ten.
Well, you just switch the roles played by four and ten. I'm not quite sure what you want.
12:29 Time for a cat video?
NO.
12:32 Well, that's not what we do here on Mathologer, right?
Right, it's NOT what we do.
12:37 Is it really just a big coincidence that widths four is special or is there a deeper
reason?
There are no coincidences in mathematics, except in very rare cases where it's just a coincidence.
12:48 ..., but how can we be certain?
This same argument can be iterated indefinitely. Each iteration replaces the initial width with 3*((the previous width) - 1) + 1.
Therefore, the pattern continues since 3*((3*((the previous width) - 1) + 1) - 1) + 1 = 3*(3*(w - 1) + 1 - 1) + 1 = 3*3*(w - 1) + 1 = (3^2)*(w - 1) + 1.
The exponent of three will increase by 1 with each iteration.
I know that I did a horrible job of describing it. Sorry.
12:51 What's up with all these similar phenomena?
Sierpinski and snail shells and whatnot?
If there's a horizontal chain of a single color, it will form a triangle beneath it.
I don't feel like figuring out the Sierpinski thing. Sorry again.
The snail shells are merely related through the Sierpinski triangle.
12:56 Are there any beautiful connections?
You just listed one. Sierpinski.
12:59 Ready to go deeper?
Do you even need to ask?
CHAPTER 2 (13:04) - Pascal's Triangle
13:07
The basic rule of two adjacent colors in one row summing to give the color immediately underneath just cries out for us to have a look at the tip of Pascal's triangle. Right?
I guess, but "adding" colors seems very different from adding numbers.
13:50 You want hexagons?
YES.
13:54 What else?
Turn it so the tip is at the bottom.
13:55 You want the whole thing to start from a row and not from the tip?
I couldn't have said it better myself.
14:21 Interesting, huh?
Okay.
14:38 Now what about
colors?
I was thinking the same thing.
15:04 Pretty impressive, huh?
No. I've seen that before.
16:44 Is 10 still special?
No.
16:48 If it were, black and white on top should give black at the bottom. Right?
Right.
17:05 instead of 1 plus powers of 3, this time
it's 1 plus--can you guess it--...
2.
18:13 Do any of you apply similar this plus that mathematics in anything you do
in your working life?
Nothing pops into my head.
18:37 So let's see what happens when we use remainder on division by three.
YES! Hey, I made something for that. Take a look at my spin-off of a program on Khan Academy: https://khanacademy.org/computer-programming/_/6406985938862080
20:05 Can you see what we need to do?
I tried, and I couldn't figure it out.
20:38 What about the top rules?
They work!
20:42 In mod 3
arithmetic - 2 is the same as -2 + 3 which equals 1. Got it?
Yes, but don't tell JavaScript. It thinks `(-2) % 3` is -2.
22:42 What about the odd-even, black-white
mod 2 Pascal game.
What about it?
22:52 Why?
(2b) mod 2 = 0 for all integers b.
24:01 Figure out the smallest
non-trivial special numbers for m is equal to 4, 5, 6, etc. until you get sick of it. Ok, a hint: just focusing on the blue entries can you see another Pascal triangle?
I'm going to go to bed now.
24:44 How on earth did they find that one?
The internet?
26:38 All make sense, right?
Sure.
27:13 What's happening here?
As I've already said, if there's a horizontal chain of a single color, it will form a triangle beneath it.
27:45 This really gives a very good intuitive feel for where
this self-similar pattern comes from, doesn't it?
Yes.
CHAPTER 5 (28:00) - Beyond Triangles
28:04 There's one very surprising feature of the original three-color game that I did
not mention yet. When you rotate--
No way.
28:32 Can you think of the simple explanation for this phenomenon?
I'm experiencing that brain-dead thing you talked about earlier. It's 10 PM.
29:16 Did you get them right?
Yes.
29:33 Why?
I'm done answering questions.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
@15:40 wait, that's not right, the 3d grid does have the turning property provided you are rotating around one of the axes, that's why obtaining a smaller hexagon is illegal, it's rotating in the wrong plane. Minor point, because we may not know the size of the unit, but we might know it's direction. That wasn't specified in the puzzle, when we switched to 3d. Probably this was missed since in 2d, we only have one way to rotate, while in 3d, we have infinite axes to rotate around or planes to rotate in, so we need to specify something extra to keep the rotation-rule. Or maybe it was just not to complicate things.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Simple proof, made harder by words substituting for diagrams but let me give it a shot:). I want to start with only 1 color here(because its more than enough). If you dont have a set of corner squares with 4 different colors you have two or more of 1 color directly adjacent to each in the corners of the 4x4, or you have 2 or more corners diagonally adjacent with the same color, these 2 cases are important in this simple proof. First of all every 2x2 in the 4x4 must contain each of the colors, that is our constraint, which means there are going to be 4 of each color in the 4x4, and in no 2x2 can we allow any color to repeat, because this would mean this 2x2 would necessarily be missing a color. The simple case is to prove that 2 adjacent corners cannot have the same color at the same time as this condition of equal participation in each 2x2 is upheld, this very very simple visually, since these 2 corners participate in a pair of 2x2s that cover exactly half of the 4x4 box, this means that the 3rd 2x2 in this 2x4 region can now no longer have a green(example color) in it and still avoid 2 greens landing in the same 2x2. So partial qed, we now know that adjacent unicolored corners are not going to give us a solution. For the diagonallu adjacent corners its as simple as counting and remembering our non repeating rule for 2x2s, our 2 corner squares cover exactly 2 of the 9 2x2s, leaving seven to be given a green square. The maximum number of 2x2’s one square can participate in is 4, meaning 8- the overlap(which is illegal) is the maximum number of 2x2s covered by 2 squares, and we need 8, to achieve this we have 8 squares to pick from, the remaining 2 corners, the adjacent corners and the 2 remaining central squares. A corner square only ever participates in 1 2x2, so we can rule those out right away, the adjacent edge squares participate in 2 2x2s, which means that even if we pick the maximum for the other square we need to color, which would be a central square participating in 4 2x2s we would only cover 2(from the original corners)+2(from adjacent to corner square) + 4 from the central square= 8 2x2s covered, which is not 9 2x2s and so one 2x2 is without a 2. This then means we are left with the 2 central squares that are not sharing a 2x2 with a corner square of the same color, in this case we are now forced put 2 greens into the central 2x2, which means we have broken the condition of 4 colors in each 2x2, that proves that no 2 diagonally adjacent corners can be the same color, and still allow all 2x2’s to contain 4 colors, qed or something. Long winded but i find pure text pretty annoying tbh :) nice little puzzle tho.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Okay, I've successfully failed programmer challenge. I firstly thought about dynamic programming, so we have functions f1(n), f5(n), f10(n) f25(n), f50(n), f100(n), where
f1(n) - # of ways to make n cents only using 1 cent coins, so f1(n)=1, n>=0
f5(n) - using only 1cent and 5 cent coins, so f5(n) = f1(n) + f5(n-5)
f10(n) - using 1,5,10 cent coins, so f10(n)=f5(n) + f10(n-10) et cetera,
And we will calculate # of ways to make n dollars which is equal f100(100n), and this works O(n) of time
But I thought, why wouldn't I optimize it? For instance f1(n)=1, f5(5n) = n+1. So I calculated all the functions up to f100 and came up with the same formula, as you derived a few minutes later: 1/6 (N+1) (80 N^4 + 310 N^3 + 362 N^2 + 121 N + 6), so I got impressed much earlier, than 24:28 =)
In program calculations actually take O(log^2(n)) of time, so i think it's success, but I didn't program anything at all =)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I see so many math tricks like this where only the numbers 1,2,3,4,5,6,7,8,9, are used in the making of their wheel or circle.
This leaves out one entire number and space between 0 and 1.
Or 10 % of our divided by 1 to 10 numbers system.
I find a problem with this.
It's not a true representation of 0 to 10.
It assumes 0 to 1 doesn't count the same as 1 to 2 or 2 to 3 and so on.
Everything is being based on only 90% of our number system.
And yes I see how all these number games work out so nicely and I used to use to always a slide rule in the days before the $200.00 four function calculator hit the market.
But I still wonder about this 90% as apposed to 100% symmetry that's always used.
What is math missing?
I'm seeing a trick as apposed to reality.
Numbers are a progression of growth by always in affect adding 1 or the next number in a sequence.
It's not about pitting 100% of 1 to 10 literally.
Or 90% of 0.0 to 10.
Our reality works on ten equal parts from 0.0 to 10.
That's equal to, 10 to 20 and so on.
Not 1 to 10, 10 to 20.
There is a 0.99999999999999999999% difference to these two number ranges that is exploited making these tricks work.
If I buy only one to ten apples it's assumed I'm getting 100% of apple number 1.
On your Number wheel I'd be lucky get barely more then just the smell of apple number 1.
So if I may ask?
What happens if this 10% trickery isn't used?
Or more exactly 0.999999999999 against 0,09999999999999 relationship.
There is a left out, long distance 0,0000000000000000000001
And this other 0,999999999999999999999999 that is also not represented.
Does Math over the centuries ever talk about this.
Or just because these neat tricks work out and are helpful we are to ignore this number gap that makes it all possible?
And what other other number realities is and has mankind been missing, that if the entire 0.0 to 10 number was used for wheel and circle type tricks?
After all, this is the 21st century.
Have we not improved in a "progressive" way in math, being it itself is an ever progressing number system science?
This is only one of many questions/problems I'm seeing in our sciences.
Sooooo, What are your thoughts about this?
ENJOY
1
-
1
-
1
-
1
-
1
-
1
-
a) Great video! b) It could be argued that the title is clickbait-y, that “Why did they not teach you any of this in school?” parrots the language ultra-conservatives deploy to get uncritical audiences fired up over invented “culture war” wedge issues, and so any videos with such titles are pushed higher up on YouTube’s homepage—and, as I don’t know your intent, I won’t argue that this is your intent … and you seem like a great fellow!! … but it sure does look like clickbait, and that’s regrettable: capitalism may coerce us into all sorts of regrettable behaviors, but the least we can do is resist—profit IS NOT everything! c) That having been written, I don’t begrudge your building an audience or gaming the algorithm to do so … but it’s still a horribly regrettable situation, having to reinforce thought-stopping language for profit because it’s all that’ll cut through the noise and we all need to scrape together a living whatever the cost. d) During the ‘70s, when I was younger than 10, I learned the visual proof for a^2+b^2=c^2 in Latin America during grade school as well as from Mom at home using magnetized geometric figures on a little magnetic slate, and my memory is fuzzy, but I’d wager I grasped this better from Mom than from school; the “new math” was all the rage, so grasping concepts like set theory and if-then flowcharts and the underpinnings of formal logic at a young age were seen as the primary goal of elementary school maths, with memorizing tables being only the secondary byproduct of that primary goal … plus Mom had taken advanced math for extra credit and for fun during her own high school years from her delightful beatnik maths teacher, so wherever my school was still stuffy and stuck in the 19th century my mom would swoop in at home with her abacus and her binary math and her dozenal math games and fill in the gaps and make learning fun again rather than dreary—yay Mom!! Anyway, thanks for the video, pretty much everything past the intro felt new to me. 🙂
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Seeing your zeroth power sum reminds me of something that I don't really understand why it is taught the way it is. That would of course be, 0^0 Now, in your typical Algebra class, when they go over exponent rules, they say "You take any number to the zeroth power and you get 1. Negative numbers, fractions, irrationals, all become 1."
And when someone asks about 0^0, they just gloss over it and say "Just put in the answer as 1, just like you would for any other number"
Now, for me, that makes me say "Wait a second, if exponentiation is repeated multiplication, than going from the first power down must be repeated division. And therein lies the problem with 0^0"
x^1 = x
x^0 = x/x
x/x = 1 for all numbers except 0
If you treat 0 as being a solution to x/x, you get 0/0 But, you can't divide by zero. Well, algebraically anyway, you can't divide by zero. In order to divide by zero, calculus has to come into the picture, and more specifically L'Hopital's Rule. But this still doesn't help us with the 0^0 problem, because if you take the derivative of x^0, you get 0. So you're basically screwed if you try to use calculus to solve for 0^0
So then, if even with L'Hopital's rule, you can't solve the 0^0 problem, why do algebra teachers insist that 0^0 = 1? 0^0 is indeterminate, and even calculus, which can help you solve for solutions when indeterminate forms pop up, does nothing of the sort for x^0 where x=0, you just get 0 for any value of x, or in other words, a line with a slope of 0 and a discontinuity at the y intercept of 1, becomes the origin.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
One proof for 2:20 :
By 3 point there is one and only one circle. We will divide our 6 points into two groups of 3 points.
I'll add names : starting on the horizontal line on the left we name the point A. Then we turn clockwise and name D,B,E,C and F. Our two groups are A,B,C and D,E,F.
Then we can show that AB = FE, BC = DF and AC = DE using the colored length.
Let be C1 the only circle defined by A,B,C and C2 defined by D,E,F. The triangles ABC and DEF have the same lenghts, hence C1 and C2 have the same radius.
Does C1 and C2 share the same center ? If they are, we are done.
Let's focus and the green angle between D and B. Let be Δ the bisector. The axial symetry applied on [DF] gives [BC], thus C1 become C2.
Another axial symetry based on another bisector will also send C1 and C2. Hence the center of both C1 and C2 is the intersection of the 3 bisectors. ㅁ
1
-
1
-
1
-
Love it! I think what is a fantastic migraine occurs when looking at the transformation from the area of the Earth to a 2d projection. The lines that sketch the Earth, when traced, travel from Antarctica to the outskirts of infinity. But never touch back to Antarctica. Appears like a great analogy to the idea of infinity by reminding us that the limit of a Cauchy Sequence, as beautiful and simple as the equivalence appears, will only ever be a limit. Ever closer, but nothing more than approximate. As though the Transcendentalism of a solution, like knowledge of Pi, becomes clearest to clever when seen to the change of 3d, through time (watching it morph), into a 2d object. Reminds myself that no matter the dimension we change to acquire equivalences (be it time or other dimensions), does not remove the Transcendental number. So then it must be True that the Infinity we chase to acquire a limit, is then equivalent to the infinity of perspectives possible to shape one proof of a solution into another?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
@Mathologer As you already stated, you have to add divisibility by 4. Then for any decomposition n = a*b, where a and b are both odd or both even, we have n = ((a+b)/2)^2 - ((b-a)/2)^2.
Since, if n = x^2-y^2=(x-y)(x+y), to get n as any difference of two squares, the only possibilities are the ones arising from such a decomposition because the linear system x-y=a, x+y=b has a unique solution. Therefore it is only possible if n is odd (a,b are odd) or n is divisible by 4 (a,b even), otherwise (a+b)/2, (b-a)/2 cannot be integers.
This also shows uniqueness for odd primes, because then, the only decomposition is p = 1*p, which yields the unique p = ((p+1)/2)^2 - ((p-1)/2)^2.
PS: Thanks for your awesome videos :)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I'm late to comment this, but I was just back of my head on and off thinking about this proof, especially one part. You mention at one point the key aspect of the proof is that p and q are co-prime, and we know this because they are prime. This reminded me of a Numberphile video where Matt Parker talks about how order in the primes, particularly IIRC how the square of every prime has one of 2 possible remainders on division by 24 or something? By memory I think this stemmed from the fact that all sufficiently prime numbers have only certain remainders on division by 6 (I guess 5 if I'm thinking about it). The point being, it wasn't just a propety of primes but rather all numbers of the form 6k+5 where k is an integer.
So I started thinking about that here, we are using the propety that p and q are coprime, but really that would work for any coprime p and q it seems then, right? Like, say 9 and 5? We could have sets like (3^a1, 5^a2, 7^a3...) which are arbitrary powers of odd primes, or (3×5,7×11,13x17,...) where you multiply consequtive odd primes and then move on, in fact any set made by multiplying powers of odd primes for each entry such that no prime (to any power) is used in multiple entries.
I wonder if any more naturally occuring type sets have this property of any two unequal numbers being coprime? The type you might stumble upon rather than construct as example.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Heya. I got all the way through. Love your videos so much. I did maths at university (mostly calculus). I'm not a great mathematician at all, but I love how you explain mathematics. I enjoy it personally and don't really use it for anything professionally with the exceptions of some excel stuff, accounting stuff, and some programming stuff, but otherwise watching your videos is like absolute candy to me, you make really difficult stuff simple and easy to understand and if not easy outright, at least digestible enough so I can figure it out as we go on. Euler is one of my favourite mathematicians.
Which is to say, yes, more outlandish, weird, and difficult stuff. I love the progression going from gauss addition which we learnt at school. But we learnt it 1 + 2 +3 + 4 = (4+1)x2 = (Highest Plus Lowest) Multiplied by Half of Highest. Because n + 1 = (n-1) + 2 etc...
So if n = 100 then 1 + 100 = 99 + 2 = 98 + 3 (and there's 50 of these pairs), so n100 = 101 x 50 = 5050
Ironically I use this trick to show off quick addition of Dungeons and Dragons dice, the numbers of a d20 add up to 210, the numbers of a d10 add up to 55. I mean it's simple if you know the trick, but folks are always impressed.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
12:05 - the block has 9 in the top left and 4 in the top right. This makes 13 the bottom right and 17 the bottom left. This works out as 9*17 = 153, 2*4*13 = 104, and 9*13 + 4*17 = 185. I solved it by setting up a square of unknown factors ABCD, and then I set the product of AC to be 153, 2BD to be 104, and AD + BC to be 185. If you add AC + BD + AD + BC you get 390, but also the quantity (A+B)(C+D). I then simply looked at the integer factors of each number and did trial and error with A and B, given that A + B had to divide 390.
Bonus question: starting from our first square, we can get to this new square going right, right, right, left. We know that the top two numbers (9 and 4) are enough to determine the entire square. The three possible combos that this square could spawn from either have 9 bottom left and 4 top right (which generates 14 on top and 95 on the bottom), 9 and 4 both on the bottom (which is impossible), or 9 in the top left and 4 in the bottom right (which is also impossible). Therefore, we then recursively look at 1459, with the two seed numbers 1 and 4. Here, we find that 13 on top and 74 on the bottom is the spawning square. This is show in the video to come from 12 and 53, which then comes from our starting square of 11 and 32.
15:53 - the area is 5. The yellow square has an area of 1. Since a^2 + b^2 = c^2, the sum of the two orange squares is also 1. We can do this recursively and find out that the sum of the areas of the four squares connected to the two orange squares is also 1, the sum of the areas of the eight squares connected to the four squares is 1, and so on. This means that the area equals the number of different sizes of squares in the diagram, as every single set of squares has a total summed are of 1.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
9 is one less the base, 10. That's why it is "special".
The division test works for any base, for example, if the we define base 16 digits 0123456789abcdef, the divisibility by the f test is exactly like the one for 9 in base 10: just sum all the digits and test if it's divisibly by f. (And also, since f is 3*5, this test will also work for division by 3 and by 5).
And only the elementary math is needed to prove this statement. (Actually we proved this routinely in school.)
Oh, yes, I've paused the video, wrote the comment, continued... and now I see there is an explanation in the video :)
The divisibility by 3 in our decimal world is a bit special; it doesn't extend to other bases so easily.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I have two queries..i don't have a proof if these two are correct.
(1)Let "a1" & "a2" & "a3" ...."an" be integers such that all of them are non zero .Is lcm(a1,a2,a3,...,an)=lcm(|a1|,|a2|,|a3|,...,|an|)?
(2)Let "a1" & "a2" & "a3" ...."an" be integers such that atleast one of them is non zero .Is gcd(a1,a2,a3,...,an)=gcd(|a1|,|a2|,|a3|,...,|an|)?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
the Euler characteristic is great. there are 3 proofs from the last 25 years that follow one another that I just love:
1998, Shahrokhi, Sykora, Szekely and Vrto: bounds on the number of crossings in a graph (“Crossing Numbers: Bounds and Applications”)
gives a minimum number of crossings for a graph with v vertices and e edges simply using probabilistics of the Expectation function and the Euler characteristic for
planar graphs
1995, Szekely: bounds on the number of incidences of points and lines (uses the above, but, of course, this proof did not yet exist, but the result had been proven
differently) (“Crossing Numbers and Hard Erdos Problems in Discrete Geometry”). Simply make a graph as above out of the points/lines where they cross but there isnt
a point.
1997, Elekes: bounds on the number of Sums and Products ("On the number of sums and products") bit more convoluted, but still doable? uses the point/line result.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
@Mathologer I'm getting acquainted with String Theory and, though I can't pinpoint anything specific yet, I have the feeling there are plenty aspects that would be neat to visualize.
I really enjoyed when, at the beginning of a course in Special Relativity, I was shown that you can get Lorentz transformations if you search for the most general between inertial frames without assuming absolute time. Or, at the beginning of General Relativity, that a frame at rest with a uniform gravitational field is indistingishable from a, small, free falling one. What I like about these is that they were in front of everyone's eyes, but it took a relatively (pun intended) long time to take the leap and it opened up a new (way of seeing the )world.
With these subjects you could also go wild with apparently paradoxical thought experiments. There's also a pedagocial purpose here. Maybe I'm wrong, but it appears to me that stuff like Schrödinger's cat or the twin paradox is thrown in the face of the non professionals as some kind of magic show. When they were inteded to test and develop our intuition of what is actually going on and the beatiful part is how you reconcile them. It's also necessary for the theory to not fail!
By the way, I'd watch your feature length on Galois Theory!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
11:44 Ok, I've been trying to solve this for a little bit. Had to build my own formula for solving for the radius of the incircle of a right triangle, and it was huge and messy. Using the first 4 you gave here 2:20 I build 4 equations for the square [(a b)(c d)] and the equation x^2+y^2=z^2
Just so I can write one of the equations, I will say x+y+z=s
1. ac=x
2. 2bd=y
3. ad+cb=z
4. ab=0.5sqrt{((s-2x)(s-2y)(s-2z))/(s)} Yeah if someone could show me how to simplify that it would be awesome
Replace for x, y, z
ac=153
bd=52
ad+cb=185
ab=36
from here, I will assume that a, b, c, d are all positive integers. Here are all possible values for a and c if so
(a, c)
(1, 153)
(3, 51)
(9, 17)
(17, 9)
(51, 3)
(153, 1)
Now, we know that a needs to also be a factor of 36, so we can eliminate the values 17, 51, and 153. So adding b into the mix will halve the options left, quickly grabbing its values
(a, b, c)
(1, 36, 153)
(3, 12, 51)
(9, 4, 17)
Now, b needs to be a factor of 52, and 4 is the only number that fits that criteria
(a, b, c, d)
(9, 4, 17, 13)
and there we go. Checking the last criteria, ad+bc=9*13+4*17=185 checks out
The only problem is the excircles, because I don't know how to find those. And since that would mean I'm not 100% confident in my answer, I can't answer the bonus just yet
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
You made me learn something new, but not what you expected! I wanted to know the rest of #15 (if the points of the plane are colored red, yellow, or blue ...) because each of the first 14 I was already familiar with (the theorems, not necessarily the proofs) if any of those had been cut off in the middle I would still know what theorem it was, but #15 was new to me. I searched and found that the rest of the theorem says “for a given distance d there will be two points of the same color distance d from each other” and saw the proof using Moser’s Spindle, and yeah, that deserves to be on the list, IMHO. You probably already did a video on #15, but if not you might consider making one, it would suit your style (though it might be too short, given the elegance there isn’t much to say).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Here's an explanation of why the function for the number of domino tilings of an n x m board returns 0 if both n and m are odd, under the cut.
.
.
.
.
.
The formula for the number of tilings of an n x m board spits out 0 if both n and m are odd, because it spits out 0 if at least one of the output numbers of the "trig monster" in the parenthesis returns 0. The cosine function returns 0 if it is fed an input of π/2, which is equal to nπ/2n. As such, both of the cosine functions in the "trig monitor" have to be fed xπ/2x to receive a result of 0.
In the formula (jπ/m+1) that is passed to the cosine function, if m is odd, then the denominator of the fraction is an even number. We're going to call this number p for the sake of the demonstration. Notably, because the brackets indicate rounding up, j will exist as every value from 1 to p/2, meaning that, at some point, the value (p/2)π/2(p/2) is reached. If we define p/2 to be x, then the cosine function is fed the value xπ/2x, which cancels to π/2. This returns a value of 0.
Of course, this only happens when one of the values for the dimension of the board is odd. If one is odd and one is even, then the function will add this pointless value of 0 to another cosine function, squared and multiplied by 4 of course, which received a value not equivalent to π/2- or, for that matter, any value xπ/2 where x is an integer, because the denominator is odd. However, if both M and N are even, then both cosine functions return 0. Both values of 0 are squared to get 0, multiplied by 4 to get 0, and then added to each other to get 0.
This one value of 0 will then completely knock out every other non-0 value, through the identity property of multiplication. As such, the function will return 0 for any board of size m x n where m and n are odd, proven in a very long and possibly difficult to parse explanation.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Dear Mathologer,
this is - so to say - "of road" regarding this video, but anyway I would like to ask you if you could, perhaps, make a video-clip in which you would prove that any homogeneous fluid body (e.g. a drop of pure water (or of mercury, or ...) in a space station, or when the drop is "in the free fall" in a gravitational field) would inevitably take the form/shape of the sphere.
When I think about it, I come to that conclusion, but I cannot find the way to prove that explicitly, that is, mathematically.
Or, perhaps, you could give some link(s) (if you happen to know about such link(s)) to some video(s), or site(s), where such proof is given.
Also, I think (or better said: I hope) that proving the following would be interesting for you and the viewers/followers of your channel:
1) If we assume that for any change to occur/happen, it must take some time for that (in other words: there are no changes which can happen during "no time"),
2) and if during an infinitesimal time dt only an infinitesimal change of something/anything (dy) may occur,
3) and if during a finite amount of time delta_t only a finite change delta_y may occur (to me, this is the direct/inevitable consequence of 2), but ... maybe that should/could be also proven mathematically),
then (y, t) is the smooth continuum (free of singularities).
In such continuum, an "infinitesimally small shape/"body"" cannot exist (in other words: the smallest shape/distribution of y in t must be some finite-size shape/distribution. And the simplest possible localized (finite size) shape/distribution (of y in t) may only and exclusively be a shape/distibution which:
smoothly continually begins to rise (starting from "zero"-value (y=0)),
smoothly continually rises
smoothly continually reaches the maximum
smoothly continally begins to fall (to decrease)
smoothly continually reaches the 0-value.
I wish you, and to your team, and to your fans, a happy and prosperous New Year!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I really, genuinely want to thank you Mathologer! Your videos are alway delightful. I truly appreciate your work.
The things I enjoyed the most were the proof about the 2^(1+2+3+...) formula and the explanation about the fact that twisting a 2x2 square is the same than going from an odd permutation to an even one or vice versa.
I really liked the animations of the square dance though!
Happy hollidays, well deserved. :)
I did some homework along the way, here it is.
6:20 I would say no but I can't find a way (if you say it's an easy challenge, then the answer should be no ^^). If I'm allowed to leave one square in an angle and take off its three neighbors, then it's done but it feels like cheating.
I wanted to take off two square of the same color which are "next" to each other (like two back ones which are left and right of a green one for instance) but I can't find why that wouldn't work. :/
10:19 Why does the formule give 0 if n and m are odd ?
If n is odd then ceiling(n/2) = (n+1)/2 and, eventually, k will equal (n+1)/2.
So, k/(n+1) will equal 1/2 and cos(pi/2) = 0.
So, when k = (n+1)/2 and j = (m+1)/2, the corresponding term is 0, which sets the whole expression to 0.
14:10
1x2 : 1 way
2x2 : 2 ways
3x2 : 3 ways
4x2 : instead of going through all of them, we can either set one standing domino to the left and that leaves us with a 3x2 grid which can be tiled in 3 ways, or set two lying dominoes to the left and that leaves us with a 2x2 grid which can be tiled in 2 ways. Total : 3+2 = 5 ways.
With the same logic we can see the Fibonacci series appearing. The series starts from 1, 2 but I guess there is 1 way to tile the 0x2 grid, which gives us the whole series we know and love. Haha! :)
14:38 I got 630 ways, I used neither the determinant nor any clever trick : i just tried to figure it out... I wouldn't bet much on the answer... What's the real answer, though ? I'd like to know.
37:32 If you see the hexagons as cubes, then you see, for each cube, always the same 3 faces. Looking at the cubes stack from above, from the left or from the right would always gives us a square the same size.
Greetings from France.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
My favourite fact from this video must be that no partial harmonic sum is equal to an integer. Although, everything is incredible as usual, great work!
To answer your questions:
1)
To get a brick to completely exceed the edge of the table with only three bricks, you can put the first one 50% out. Then add the two last ones on the second layer symmetrically, such that they both have 50% overhang(one to each side). Essentially an up side down pyramid.
2)
I believe you answered your own question at some point. How to write 2 as a sum of reciprocal of distinct positive integers. That is, 2 = 1/2 + 1/4 + 1/6 + 1/8 + ... Writing 3 like this, however, is slightly harder. I did not find a way of doing this, and I suspect it might not be possible.
3)
In the square, where you inserted all the pieces of area that did not fit under the curve, you said that since they all fit into an area of 1 they must be less then one. The same type of argument can be made to say that they must be lager than 1/2:
None of the rest areas overlap in height, so when we put them into the square, they each have their own little rectangle they fit into. Each of those rectangles can be cut at their diagonals, and become two triangles. The rest area in each of the rectangles always cover more than one whole triangle. If you sum up all of the filled triangles, you get one half, since they cover half of a unit square. Therefore, the rest area is greater than one half!
1
-
1
-
1
-
1
-
1
-
1
-
I was curious whether there was any kind of an equivalent for 108-degree angles using pentagons, and it turns out there kind of is - but, unsurprisingly, you have to get the Golden Ratio involved. For a triangle with a 108-degree angle, a pentagon with the side length equal to the longest side of the triangle is equal to the sum of the areas of the pentagons with the two smaller side lengths, plus (2 * phi - 1) times the area of the triangle.
It seems that, starting from an equilateral triangle, there is a sequence of correction terms (specifically as ratios of the triangle area) for triangles with an angle equal to one of the inner angles of a regular polyhedron. For 3 it's -1, for 4 it's 0, 5 is 2phi - 1, 6 is 6. 7 will be a pain to calculate exactly, but I think 8 is 8(1 + sqrt(2)).
The general rule for an N-gon appears to be that the number of triangles you need to add as a correction term is equal to twice the sums of the heights of all the vertices (measured perpendicularly to some side of the N-gon chosen as the "base") divided by the height of one of the vertices adjacent to the base, minus N. Or, equivalently, 2*Im(Nz + (N-1)(z^2) + (N-2)(z^2) + ... + z^N)/Im(z) - N, where z is the principal Nth root of unity (and Im(x) is the imaginary part of x). (I think there's a way to simplify it from there but I need to get back to work...)
1
-
1
-
1
-
1
-
1
-
1
-
Second puzzle: short answer is no, since I'm not good at mental arithmetic under pressure. But armed with a few factoids and nerves of steel, I could solve the problem in about 30 using the following strategy: first, what's (n-1)^2 plus (n+1)^2 ? We see that the middle terms in the resulting trinomials cancel out, and the end terms add together, giving 2*(n^2 +1^2). Same pattern for n-2. We can write the series in the numerator as (12-2)^2 + (12-1)^2+ ...+ (12+2)^2. Using the above observation we get 5*12^2 plus 10 (=2*2^2 + 2*1^2) in the numerator. In the denominator, we know 365 is also divisible by 5. If you happen to know the other factor is 73(I didn't) = 72+1 then you're home free, otherwise 360=12*30 is pretty common knowledge, so add another 5 to it and you get 5*(6*12) +5 in the denominator. Now, 2*72=144 (gross!) is also pretty common knowledge, so we have twice( 72 + 5) divided by 72+5 =2.
I freely admit to having worked this out with an electronic calculator, ahead of time, in about 5 seconds. When you do that, you get 365 for the first three terms in the numerator and another 365 for the two remaining terms, thus proving, by brute force, the original Mathologer proposition.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
13:00 You can inductively get from Input={n^2} to a,b,c,d, ... = 1,2,2,2, ...
I tried to find a general solution with Input = {n^k} and k >2 by hand.
Nope. Nope. Nope.
Bless the power of Python and Matplotlib.
Only k = 3 is nice with a, b, c,... = 1, 6, 12, 18, ... and so on just adding 6.
Any ideas for k = 4 with its insane a, b, c,... = 1, 14, 50, 110, 194, 302, 434, 590, 770, ... ?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I'm still unclear why it's pi, and not pi/4, that everyone's trying to get at.
pi/4 has far more interesting and useful properties, such as being the ratio of both the perimeter of a circle to the circumscribed square (P(circle) = pi/4*P(square)), and the area of the circle to the circumscribed square (A(circle) = pi/4*A(square)), being the reflection axis of the trig functions (cosine = sine reflected about pi/4, cotangent = tangent reflected about pi/4, cosecant = secant reflected about pi/4), etc.
and these series all natively yield pi/4, not pi. so it takes intentional effort to turn it into the wrong value, which makes asking the question 'why the hell is everyone doing that?' really quite obvious, and yet nobody appears to have asked it.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Once again, great work, Mathologer! However, I see a tiny error there in the distribution of terms at 4:22 You would expect the empty set in the "first column" in the cross product, but you made the leading empty terms disappear (except for the very first partial product).
In Mathematica, by the way, it would look like this: with Distribute[{{}, {1}, {1, 1}, {1, 1, 1}}, {{}, {3}, {3, 3}}, List, List] you get {{},{}},{3}},{{},{3, 3}},{{1},{}},{{1},{3}},{{1},{3,3}},{{1,1},{}},{{1,1},{3}},{{1,1},{3,3}},{{1,1,1},{}},{{1,1,1},{3}},{{1,1,1},{3,3}}}
Anyway: a very exciting topic! As a chemist, I see great applications there: e.g. if you ask for the possible elementary compositions for a given (exact) molar mass, you can get all conceivable "sum formulas" this way (and with a few chemoinformatics tricks even the valid molecular formulas).
Looking forward to future videos! Here are a few "wish titles": Polya-enumerations of isomers, automorphism groups of graphs or even solving functional equations.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
39:15 (knowing fibonacci is ...-8, 5, -3, 2, -1, 1, 0, 1, 1, 2...) i'm betting this extends infinitely backwards too. choose a=0, b=1 and you get 2. a=1, b=0 produces 1 (the 1 after 0). choose a=-1, b=1 and 1 is a solution (the 1 before 0). with a=2,b=-1 you get 2 (the one behind 0). choose a=-3, b=2 and 5 is a solution, which seems to indicate to me that this goes forever both ways. i think this is also trivially demonstrated by the fact that squaring numbers discards the negative status, and that the diagonal multiplication and addition happens to align as long as the signs alternate the way they do when you construct the progression this way.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Here's my crack at a proof:
Let ABC be the triangle with lengths BC=a, CA=b, AB=c. Suppose AC extends to points A' and C' such that the order of the points along AC is A', A, C, C'. Extend AB to points A'' and B'' such that the order of the points along AB is A'', A, B, B''. By symmetry, it suffices to show that A', A'', C', and B'' are concyclic. But notice that A'A=a and AC'=b+c, while A''A=a and AB'' = c+b. But now (A'A)(AC') = a(b+c) = a(c+b) = (A''A)(AB''). By Power of a Point, A', A'', C', and B'' must be concyclic.
If you don't know Power of a Point, notice that triangle AA'A'' is similar to triangle AC'B''. Then angle A''A'C' = angle A''A'A = angle AB''C' = angle A''B''C'. This is a criterion for A'', A', C', and B'' being cyclic.
EDIT: Ah, hold on, the symmetry argument doesn't tell us that all six points are concyclic, and I can't quite resolve that little problem. Yeah, this one is trickier than expected. My other idea was to do try doing some sort of homothety to the incircle, but it's not obvious to me how to show such a homothety maps points on the outer circle to something fruitful on the incircle. Interesting question.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
7:04 2
Specifically, for any natural number n, let t(n) be the nth triangle number (i.e. t(1)=1, t(2)=3, t(3)=6, etc.)
So, the linear pattern is sum(i from (2*t(n)-n) to (2*t(n))) of i = sum(j from (2*t(n)+1) to (2*t(n)+n)) of j
Squares: sum(i from (4*t(n)-n) to (4*t(n))) of (i^2) = sum(j from (4*t(n)+1) to (4*t(n)+n)) of (j^2)
Cubes: [sum(i from (6*t(n)-n) to (6*t(n))) of (i^3)] +2*[sum(k from 1 to n) of k^3] = sum(j from (6*t(n)+1) to (6*t(n)+n)) of (j^3)
The two is because a cube has eight vertices and only six faces. Each slice from the largest initial cube covers one face and one vertex, leaving two missing vertices (it also covers two edges, which works out perfectly with the 12 edges of a cube). Each missing vertex needs a smaller cube, with a side length of 1 for the first new cube, 2 for the second new cube, etc.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The "crossed over" cases from about 23:00 onwards arise only if one names diagonals and sides arbitrarily and thus abandons their original meaning. Take one of the side pairs - let's say b and B - of a convex quadrilateral and call them now c and C, and rename the original diagonal accordingly - thus here b an B. Then you get your "crossed case". The caveat at the end about cyclic quadrilaterals is an artifact of this arbitrariness. Take the proper equality for the cyclic case aA+bB=cC. Now just switch the names of b and B with c and C. The old equality now reads aA+cC=bB (and holds true, we just changed the names). Now write the inequality for the crossed case as aA+bB>=cC (new names), then we can substitute into this the old (renamed) equality as cC=bB-aA and get aA+bB>=bB-aA or aA>=-aA, and with a and A being positive we get 1>=-1. Equality is obviously not possible, but the greater case is true. Therefore, in the crossed case aA+bB>cC. Thus, if one calls one of the side pairs of a cyclic quadrilateral "diagonals", and those diagonals "sides", creating a crossed case out of a convex one, then one gets the result at about 24:39. There may be reasons for doing this, but here it just confuses in my opinion...
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Lovely video. There's a lot of math magic here, but it's positively pleasant to watch.
Some feedback, as requested: I'm a Physics PhD student, currently teaching first year university calculus classes. I've never really encountered the Euler-Maclaurin formula before, but its consequences I've seen pop up a number of times. It's very interesting seeing where it all comes from, both mathematically and historically. I liked the tempo of this video, and together with the animations it makes it a breeze to follow along. That is - to sort of see how the steps go. To see exactly how they go I'd need to grab a notebook and write/derive along :D. However, somewhere around the 26 minutes mark I sort of lost track of why we were rewriting all these sums in the first place. It's easier to keep track of where you're going with a derivation when you're the one doing it than if you're just watching it. Maybe when you're multiple steps deep into a derivation it'd be nice to sort of recap the high-level kind of steps you took to get there, it'd make it easier for viewers to remember "oh right, that's how we got here". But I'm just nitpicking at this point---the video is already very good.
I look forward to seeing the next one, even (or perhaps especially) if it's again an hour long! :)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
This might be off the wall, but I've used this animation for years as a metaphor in my work as a psychotherapist. In the same way we can look at these floating points zooming around and tell competing, yet coherent stories about them (triangle, square, etc.), so too do we tell stories about ourselves. Sometimes we point to our stories and to the memories that accompany them and say "Look! They're true! I AM a red triangle. I AM a failure/unlovable/etc." Our story-telling mind does a wonderful job of drawing shapes to fit the narrative, and one of the things we can do in therapy is begin to take different perspectives on stories. Looking at our life of floating points, we can tell new stories: "I'm also a blue square. I'm also a yellow star." And when we are able to transcend the limiting powers of narrative, we can begin to see that we are all of those things--and also NONE of those things. WE are not the stories we tell--there is a self there, behind the competing stories, that we can connect to when it's helpful.
Which is all to say I'm so grateful for this video and to see the math behind it. Also grateful to the contributors who made new implementations of the dance--they'll make it so much easier to use this idea in my work!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
20:40
I'm sure you would have thought about this argument, not really sure why you didn't use it. I feel it fits the picture-proof-style even more :D
If you look at the first triangle, we need to understand why these powers of 2 turn into powers of 3. And to do this, notice that along the diagonals, we multiply by 2s. This is obvious for the first diagonal, as we started with powers of 2. But for others, let's talk about the cells 6 and 12, how did we get them? 12 = 4 + 8 = 2(2 + 4) = 2 * 6. This generalizes very easily.
Now look around at the rows, what exactly do we do when we go left? Say we're at cell x, we look at the lower left diagonal entry which should be x/2 and add it. So the result is x * (1 + 1/2) = x * 3/2. That is, we multiply by 3/2 as we go left, which isn't surprising as we had to prove these powers of 2 will eventually become powers of 3.
Equivalently, we can phrase this argument by looking at the columns. When you go up, say from the cell x. You add the thing to the top right, which should be 2x. So going up amounts to starting with x and getting x + 2x = 3x.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Is there one?
I'd say it's because 1 x anything is an identity. 1 x n = n for all n. Thus it's irrelevant. What we're looking at is 1 + 2 + 3 = 2 x 3, which is saying that (1+2) + 3 = 3 + 3 = 3 x 2 is a statement defining multiplication: For n x m, this can be rewritten as n + n +... m times OR as m + m +... n times. In this case, 3 x 2 is = 3 + 3... 2 times (or 3 + 3), and that's what 1 + 2 + 3 is equal to just pre-adding the 1 + 2. You could also pull one from the three, 1 + 2 + (2 + 1), collect the 1s, [1 + 1] + 2 + (2) = [2] + 2 + (2) = 2 + 2 + 2 which gives you the other form of 2 x 3, which is 2 + 2... 3 times, or 2 + 2 + 2.
The 1 x, being an identity, is irrelevant to all of that. You could also write it 1 + 2 + 3 = 1 x 1 x 1 x 1 x 2 x 3 and it would be the same thing, but sounds and looks less profound. : ) You could also use the additive identity (0 + 0 +...) on the left side. For example, is there anything profound about 0 + 0 + 0 + 1 + 2 + 3 = 1 x 1 x 1 x 2 x 3 ? Probably not. Yet they both are still equal. Adding 0 to anything leaves its value unchanged (hence an additive identity, as it returns the non-zero part of the equation unchanged), and likewise, multiplying 1 by anything leaves its value unchanged (multiplicative identity)
The only weird part is when you get to things that are uundefined, since normal operations on those things don't work. For example, 1 x n/0 is undefined because n/0 is undefined.
1
-
@Mathologer Oh but of course. :) I always try to think through problems and consider what my answer is to them (based on the thumbnails) before watching anything so I can exercise my mind a bit. Once I start watching the video, my further thoughts are going to be at least somewhat influenced by it, so I want to get down how I think through the problems before seeing anything, so I can compare my answer/analysis to those presented and revisit it with the new information/additional argument and see how that changes things.
It's a bit odd sometimes, since some thumbnails don't entirely represent discussions and things like that, but I also just like thinking about problems. :)
I also love seeing what I would call "complete" (though I think the mathematician word is "general") solutions to things. For example, the more complete Pythagorean theorem for non-right triangles (I won't remember this right, but I believe it has a - AB Cos th term) or the n-dimensional version (which is easy to understand/think through by considering a three dimensional vector can be mapped out as light sources casting shadows onto lower dimensional planes that merely form triangles with respect to the planes' origins)
My first degree was physics (and second economics), so I'm only math-adjacent. :D I've always loved math puzzles, but I hate proofs, because they tend to require knowledge of specific identities. I like thinking through solutions, not finding them based on having memorized a remote trivia piece. So despite friends over the years suggesting I get a math degree and noting my interest in things like topography and logic mappings, I can never QUITE cross that bridge.
But I still love seeing the puzzles and solutions or work on them. And I absolutely LOVE your videos since most of the time you break things down (though I did have to puzzle over how S - A = R for a moment...I'm not sure why THAT tripped me up but it wasn't that hard to work out, haha!)
I will keep watching your videos because I really like them. Though this one I've paused so I can think through it a bit more. Think I will watch the rest of it. :)
1
-
1
-
1
-
1
-
1
-
1
-
Im a class 12 student, and the kind who wants to jump over anything he finds interesting enough to learn(and in the process leave the actual syllabus pending, coz i find most of it to boring to ponder over, more so because of the competitional method in which we are taught these things, just juicing away the beauty and creativity of maths)! And maths is just one such domain, where a bit of imagination and observation and , lots of patience , can go a really long way, and anything like this, beautiful mathematics, is sure to attract moths like me to its light!. This was such a nice video! Although i had doubts about if i should watch the whole of it, after watching it, i have even more!! Almost understood all of it, thanks to the visualisations , but now im craving for more!
P.S when you talked about the sum deviating from its actual value after a certain number of terms, my brain was just screaming "TAYLOR, ITS TAYLOR SERIES IM TELLING YOU, ITS GOT SOMETHING TO DO WITH IT!."
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
@Mathologer Here’s something else, but is it trivial?
Starting with a row of n ones, the next row is simply n, n-1 etc.
for the second row, the ratio of each row-entry to its rightward neighbour is just n/(n-1)
Play the rinse and repeat game to derive the ‘golden’ ratios after a few rows. From left to right call these rn, sn, tn, un… to extend your r and s terminology to an initial row of n ones.
Take the ratio of adjacent ratios, rn/sn, sn/tn etc.
Note that each of these ratios of ratios is smaller than the n/(n-1) that sits in Row 2 for each n.
But with increasing n it looks like these ratios of ratios increases.
I wonder whether, these ratios of ratios converge on the n/(n-1) that sits at the top of each of these columns and, if true, whether that is interesting or trivial.
This probably needs more work than the half-hour I’ve spent playing a spreadsheet!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Challenge at 11:34
If the Fibonacci box has numbers P, Q, R, S where P+Q=R and Q+R=S, then: P*S=153 and Q*R*2=104 and (P*R)+(Q*S)=185.
So, either Q=1 and R=52 or Q=4 and R=13.
If Q=1 and R=52, then P=R-Q=52-1=51 and S=Q+R=1+52=53 and these give P*S=51*53=2703 which is incorrect.
Therefore, Q=4 and R=13 which give P=R-Q=13-4=9 and S=Q+R=4+13=17 and P*S=9*17=153 and (P*R)+(Q*S)=(9*13)+(4*17)=185.
Again, the corresponding Fibonacci box has numbers 9, 4, 17, 13.
Bonus at 11:51
Since the top 2 numbers are P=9 and Q=4, then, if the parent box has numbers A, B, C, D where A+B=C and B+C=D, we have either (1) B=4 and D=9, or (2) D=9 and C=4, or (3) A=9 and C=4.
Since A+B=C and B+C=D, we cannot have option (2) where D=9 and C=4 because this would make B=5 and A=-1 (a negative number) which is invalid. Also, we cannot have option (3) where A=9 and C=4 because this would make B=-5 (a negative number) which is invalid.
So, B=4 and D=9 which makes A=1 and C=5 and the parent box has numbers 1, 4, 5, 9.
In turn, it has a parent box with 1, 3, 7, 4 which has a parent box with 1, 2, 5, 3 which goes back to the initial box with 1, 1, 2, 3.
Therefore, navigate starting with parent box for triplet 3-4-5, go to the right child 5-12-13, then go to the right child 7-24-25, then go to the right child 9-40-41, and finally go to the left child 153-104-185.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
planar spectral decomposition :)
In one of the odder uses I've ever had for mathematical theorems, I used a similar concept to reprogram a sensor bank.
The sensors shared memory with each other, and that memory was both their operating memory and their (immediate) storage as the data went through pre-processing per-bank and eventually sent it back to the computer.
Instead of plugging into . . . a lot . . . of sensor banks, which, even if the data transfer was instant (it was not), would still take a long time, I wanted to plug into one sensor and use the fact that each sensor bank shared memory with its neighbours to reprogram them all. The direct approach, ("copy your memory"), didn't work because the sensors can only access some memory by actively mutating it. The quine approach probably held water, but I couldn't come up with one.
But using the fact that they actively mutate the memory, I could define a chain of sensors where I directly modified the memory in the first, and the last one in the chain was the one being programmed correctly. Then I remove the last one from the chain . . .and so on recursively until I can just program the sensor I plugged into directly.
In order to figure out how the bits were flipped across everything was a spectral decomposition, and I pictured it as decomposition of a polygon, and it ended up being an 8-cycle IIRC. So I only needed to transmit 8 different core files (plus a few for as I moved up the chain and started telling each member in the chain to also modify the memory it shared with members already in the chain), in order to reprogram every sensor from just one :) The files would mutate as they passed along the chain until they ended up in their original form.
Everytime we needed to reprogram that device, it easily saved 3-4 hours of labour, (instead turning it into about 8 hours of plug it in and walk away).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
This is the only place I can think of where my favorite formula in math can be appreciated.
PT = P0 + (V * T)
The reason why it's my favorite is that it can be used to plot:
1D lines (Algebra and number lined)
2D lines (Algebra)
3D lines (Calculus)
And maybe more.
What the equation means is
The digit you are looking for PT is equal to the first digit in the sequence P0 plus the difference of digits multiplied by how many digits away it is.
So let's say:
P0 is 5
P1 is 6
6 - 5 = 1
6 = 5 + (1 * T)
1 = 1 * T
1 = T
But if you notice there is a pattern not shown, one that will show that it tracks through multiple dimensions.
It works in 2D as well
P0 is now (5,4)
P1 is now (6,6)
Does it still follow
V = (6,6) - (5,4) = (1,2)
(6,6) = (5,4) + ((1,2) * T)
T = 1,
And if you still want/need to find B, (Y when X is 0) it's just a matter of simplifying the equation back down to X to find T, then plugging T to find Y
0 = 5 + (T * 1)
-5 = T
Y = 4 + (-5 * 2)
Y = 4 - 10
Y = -6
(0,-6)
Now for the fun bit:
X,Y,Z
P0 is (5,4,3)
P1 is (6,6,6)
...
Yes it still works. You are welcome to do it yourself but I know it for a few reasons, in Calculus it's called the Vector Form of the Line, and I had to learn it to A) program changing R,G,B values in a program and, in Unreal how to have 3D moving environments
If I had known sooner what a simple powerful equation it was.
But the secret each lower dimension is hiding is that all the other ones are there, it's just that their Vs are 0.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
how many squares in an n x n grid of points?
• consider an arbitrarily rotated sized positioned square among the points
• a square's rotation is determined by the slope of one of its lines, a/b
• if a=b=0, the square does not exist. for now we ignore the position
• let's find all the squares which are the same (one can be moved to coincide with the other)
• the squares a/b, -a/-b, -b/a and b/-a are all the same, so to avoid duplicates let a,b>=0
• if b=0, a/b is the same square as b/a, so let b>0
• with our restrictions, a+b is the a/b square's width (distance from leftmost to rightmost point)
• it is necessary to let 0<=a and 0<b<n-a. have we excluded all duplicates?
• yes: two squares can only be the same if some of their slopes are parallel or perpendicular, which we have handled in its entirety
• with our (now sufficient) restrictions, there are (n-a-b)^2 valid ways to position an a/b square so that the answer is sum(0<=a and 0<b<n-a) of (n-a-b)^2: now for algebra autopilot
= sum(a=0...n-2) of sum(b=1...n-1-a) of (n-a-b)^2
= sum(a=0...n-2) of sum(b=1...n-1-a) of b^2
= sum(a=0...n-2) of (2(n-a-1)^3+3(n-a-1)^2+n-a-1)/6
= 1/6*sum(a=1...n-1) of 2a^3+3a^2+a
= n^2(n-1)^2/4/3+n(n-1)(2n-1)/6/2+n(n-1)/2/6
= (n-1)(n+1)n^2/12
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
6:20 This is just a guess based on what you've said so far, but I believe the answer is..."usually".
It's definitely possible to remove two black and two green tiles in such a way that dominoes cannot tile the board. As an example, remove tiles [1, 2], [2, 1], [2, 2], and [3, 1] (where [1, 1] is the lower left corner). This leaves tile [1, 1] isolated from the rest of the remaining board.
However, this almost feels like cheating. My gut instinct says that, as long as you mutilate the board such that both A) You remove an equal number of black and green tiles and B) No part of the board is completely cut off from any other part, you will still always be able to tile it.
Also, as a bonus, it's still sometimes possible to tile a mutilated board with an isolated area. For example, remove tiles [1, 2], [2, 2], [3, 2], and [3, 1]. The isolated area is now two connected tiles, which you can put a domino on.
1
-
1
-
I need a hint to pursue the question of rational tan values. It's true that tan(na) is a rational function of tan(a) for n = 2, 3, 4, ..., so that if tan(a) is rational, so are tan(2a), tan(3a), etc. But this only tells us that slopes of lines from the origin to the vertices of an n-gon are rational if tan(m*360/n) is rational. That doesn't imply that the vertices themselves have rational coordinates, no matter how cleverly I scale, so I'm not ready to apply the shrinking argument to them. For instance, if one of the vertices is (2/sqrt(13), 3/sqrt(13)), with rational slope 2/3, and another is (3/5, 4/5), with rational slope 3/4, there's no way to scale so that both have rational coordinates.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The essence of the euclidian algorithm is this lemma (it helped me a lot to understand it, so I am posting it here)
Lemma: If a=bq + r, then GCD(a,b) = GCD(b,r)
Proof. We will show that if a=bq+r, then an integer d is a common divisor of a and b if, and only if, d is a common divisor of b and r.
Let d be a common divisor of a and b. Then d|a and d|b. Thus d|(a - bq), which means d|r, since r=a - bq. Thus d is a common divisor of b and r.
Now suppose d is common divisor of b and r. Then d|b and d|r. Thus d|(bq+r), so d|a. Therefore, d must be a common divisor of a and b.
Thus, the set of common divisors of a and b are the same as the set of common divisors of b and r. It follows that d is the greatest common divisor of a and b if and only if d is the greatest common divisor of b and r.
qed
1
-
1
-
1
-
1
-
Great stuff as usual...
13:46 Without using generating functions I found
0:00:00 [ 2, 2728 ]
0:00:00 [ 20, 53995291 ]
0:00:00 [ 200, 4371565890901 ]
0:00:00 [ 2000, 427707562988709001 ]
0:00:01 [ 20000, 42677067562698867090001 ]
0:00:11 [ 200000, 4266770667562669886670900001 ]
0:02:08 [ 2000000, 426667706667562666988666709000001 ]
0:23:49 [ 20000000, 42666677066667562666698866667090000001 ]
The running time is roughly linear in the amount, so that's what I can do in an hour. Enough to see a pattern emerging. The algorithm works for fractional dollar amounts, too. Written in GAP since I didn't want to deal with GnuMP.
Grüße aus Dresden und frohe Ostern.
1
-
1
-
1
-
1
-
@Mathologer okay, I figured it out. I was tired and trying to calculate it in my head so I came to a faulty conclusion.
3 * 3 * (5 * 2 * 2 + 3 * 3 * 3 * 2) = 666
The two outer 2x3 have 3 tilings, the rest in the middle is just a combination of the following tilings {2x2, 2x4, 2x2}, {2x3, 2x3, 2x2}, {2x2, 2x3, 2x3}, or {2x3, 2x2, 2x3}
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Start with a triangle with sides r, g, u.
Draw the incircle of the triangle.
This divides the triangle such that r=a+b, g=a+c, u=b+c where the lengths a are tangent to the incircle and meet at point U, the corner opposite u; the lengths b are tangent to the incircle and meet at point G, the corner opposite g; and the lengths c are tangent to the incircle and meet at point R, the corner opposite R.
Extending by length r from point R gives two line segments tangent to the incircle that are both length r+c=a+b+c. Similarly extending by g from G gives line segments that are g+b=a+b+c, and extending by U from U gives line segments u+a=a+b+c.
Because each pair of line segments tangent to the same point on the incircle are the same length, they form the secant line of a larger circle. Further because the three secant lines are the same length and the same distance from a common center (be a use they are all tangent to the same circle), they belong to the same larger circle.
If q is the inradius (radius of the incircle) and s is a+b+c, the semi-perimiter of the triangle, the larger circle shares the center with the incircle, and has radius Q where s^2+q^2=Q^2.
I think that's enough words, and working it out I worked backwards.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
My favorite from this video is definitely the section on the Euler-Mascheroni constant; the reason I like it the most is because the argument can very easily be adapted to prove the existence of the Euler-Mascheroni constant (that is, the fact that the difference H_n - ln n converges) quite beautifully. Note that this is essentially a visualization of the standard proof comparing the harmonic series with the integral of 1/x, which I like because I never even considered this aspect before.
Also, regarding your challenge about the value of the Euler-Mascheroni constant: Note that, since the function 1/x is convex amongst the strictly positive real numbers, the function defined piecewise on the interval [k, k + 1] (with k being a strictly positive integer) by f(x) := 1/k + (x - k)(1/{k+1} - 1/k) is always greater than or equal to the function 1/x (and trivially always less than or equal to the piecewise constant function 1/k on the interval [k, k + 1]). Since all these functions are integrable, it follows by monotonicity that the definite integrals from _1 to some greater integer upper bound n are also less than or equal to one another. More precisely, a simple estimation shows that the integrals of f and the aforementioned piecewise constant function (whose value, remember, is just the n-th partial sum of the harmonic series) differ by exactly 1/2 (1 - 1/n)_. This immediately implies that, as _n dashes off to infinity, the integral of their difference (which, by linearity, is the same thing as the difference of their integrals) converges to 1/2_. From this, it follows that the integral of the difference of the piecewise constant function and the function _1/x_, which converges to the Euler-Mascheroni constant as _n goes to infinity, is greater than or equal to 1/2 (1 - 1/n) (again, by monotonicity)– and, by the limit comparison theorem, this shows the Euler Mascheroni constant is greater than or equal to 1/2.
Anyone who hasn’t drawn a picture at this point should do so now, because this proof is way more obvious and straightforward than it looks. Essentially, I am just cutting out right triangles that make up exactly half of the area covered by a rectangle of length 1 and width 1/k - 1/{k+1}. Translating them horizontally so that they line up in the unit square then leads to the desired result. Also, the careful reader will have noticed that I compared the n-1-th harmonic number with the natural logarithm of n_, rather than comparing the _n-th harmonic number with the natural logarithm of n + 1 (which is what was done in the video). Most commonly, you will see the Euler-Mascheroni constant defined as the limit of the difference of the n-th harmonic number and the natural logarithm of n – clearly, these definitions are all equivalent, given that they simply boil down to some routine re-indexing and a single term of the harmonic series (which gets arbitrarily small for sufficiently large _n_).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Dear @Mathologer, great video as always but my tangential comment concerns your T-shirt!
Yes, yes, "25.8069 IS THE ROOT OF EVIL" was funny and I got a chuckle out of it until I realised it's not correct and there is something very unmath-like. The square root of 666 is 25.80697580... Now, for practicality it's OK to choose an arbitrary level of accuracy, say 4 decimals BUT according to the rules of rounding, if there are 4 decimal digits are shown, the last digit is not 9 because there is a 7 after it which forces rounding up!! Or, to keep with the evil theme, was this done intentionally?! No matter what, the OCD community is upset! =:-o
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Dear Marty & Crew & fans - My gratitude & delight @ finding this one (my new #1 5.5-Star episode) prove a natural double-infinity (squared?)! I've been thinking about another, more organic example, but this seems to be a clincher, re: my work-in-progress on the metalogical "roots" that connect RH, GC, and the TPC (etc.). For example, can use the circular log slider (and/or the dynamic {infinitely self-optimizing} version) to prove the natural principles enabling circles, the wheel, log(n), n = pi, the unit circle (& trig), C (the complex plane), Phi, Fs (the Fibonacci series, Platonic solids, and Bucky balls (etc.)? I think I can. What do y'all think about that? ~ Thanks again! Sincerely ~
+
Oh, BTW, if you want to check on my progress, my 2 (interrelated) preprint drafts (A: "Riemann, Metatheory, and Proof" & B: "Theory and Metatheory of Atemporal Primacy" and a brief bio) are accessible via RearchGate and >
ORCID(.org)/0000-0001-5029-7074 < FYI: Paper B is a bit like my Special Metatheory of metalogical relativity that clears up the confusion about time, space, energy, causality (and other primal principles), QM, ontology, "materialist cosmology" (& other logical fallacies), etc. (all without reliance on insufficient authorities, logical fallacies, etc.).
+
I think you may enjoy them (and hope you do). Thanks again & keep up the great work! ~ M
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
aight heres a quick whack at the first puzzle:
in addition to the n^x count of squares (1+4+9+16 or 1^2 + 2^2 + 3^2 + 4^2, summarised as the sum 1->n-1: n^2 where n is the dimension of the point matrix), each of these squares has additional "tilted squares" that can be drawn equal to the amount of "unnecesary vertices" in a given side of said square, that is, points which a given "level square"'s side passes through before reaching the main vertice on each end of the side.
so the 4*4 square (the largest one that passes through all the perimeter points) has 3 tilted squares (the first of which is given as an example with vertices {[0,1] [1,5] [5,4] [4,0]}) from their you can imagine the other two, one being a mirrored version of the example, and the other passing through the centre points on each side of the matrix. Since theirs only one of these 4*4 squares, the amount of squares is (1^2)*(1+3).
Following this to the 3*3 squares, they only have 2 "unnecesary vertices" so each will yield 2 "tilted squares", as there is four of such "level squares" that gives us (2^2)*(1+2). Following this pattern we can say the reamining smaller squares amount to (3^2)*(1+1) and (4^2)*(1+0).
(1^2)*(4)
(2^2)*(3)
(3^2)*(2)
(4^2)*(1)
--------------------------
4+12+18+16 = 50
This formula can be summarised entirely by:
For a square n points wide and n points tall, the amount of uniquely positioned squares that can be drawn with vertices within this square of points;
summation of x = 0 --> n : (x^2)*(n-x)
---------------------------------------------------
this essentially gives us our previous calculation for a points square of 5*5:
x = 0 --> 5
x = 0 : (0^2)*(5-0) = 0
x = 1 : (1^2)*(5-1) = 4
x = 2 : (2^2)*(5-2) = 12
x = 3 : (3^2)*(5-3) = 18
x = 4 : (4^2)*(5-4) = 16
x = 5 : (5^2)*(5-5) = 0
Total sum = 50
Edit:
x = 0 --> n : nx^2 - x^3 is also valid if you don't like brackets
I'll do the cubicle grid another day
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
i paused this and had a sit down early on, and proved the triangle bit to myself in almost the inverse way, scaling up (I'm paused again, just after the nested hexagons so i might be about to say something stupid that gets overidden later but f it);
with a unit grid, an equilateral triangle with a base bridging 3 points on a line (can't have a single bridge, theres nothing above to connect to) would require a multiple of rt 3 for an integer height, so disproves itself, so we move to configurations with rotated bases. The length of the base scales to n/sin(atan(n)) where n is the number of steps of deviation in a direction orthogonal to the original base, but this configuration essentially creates a new grid of sides equal to the scaled base length. For the cases of odd n (where grid points exist above the centre of the base), it ends up requiring a multiple of rt 2 (+ 1/rt 2 offset for the offset,) to for integer height (still impossible) and even n (such as 0) the rt 3 case. So each new configuration basically reverts itself back to the original problem.
I don't think i've made any sense and should stop drinking while i try and maths
1
-
1
-
1
-
1
-
1
-
1
-
My solution to calculating the number of squares on an (n,n) grid would be:
S=sum((n-k)^2*k from k=1 to k=n-1)
where S is the number of squares.
This is derived by projecting the squares on the horizontal/vertical axis of the grid.
For example, on a 6x6 grid a tilted square could be the one, where you move from corner A to corner B (2,3) (=2 on one axis, 3 on the other). This square obviously projects as a size 5x5 square on the grid (you need horizontally and vertically 6 grid points to draw in the tilted square). It fits once inside the 6x6 grid, as you can neither translate it vertically nor horizontally.
The table of possible squares on a 6x6 grid would therefor be:
(side length): (vertical distance, horizontal distance),(...)
5: (0,5),(1,4),(2,3),(3,2),(4,1) where (0,5) is the obvious big enveloping square. These all fit once.
4: (0,4),(1,3),(2,2),(3,1)
3: (0,3),(1,2),(2,1)
2: (0,2),(1,1)
1: (0,1)
Notice how some combinations appear again (but mirrored). Technically all should appear twice, only that there's no visible distinction between (0,j) and (j,0) as well as (j,j) and (j,j), which is why they aren't counted twice.
The number of times a square occurs on the 6x6 grid is then easy to calculate by summing up how often its projection side length fits into the grid.
For example all with projected side length 3 fit in a 6x6 grid 9 times.
The number of times side length k fits into a NxN grid is (N-k)^2.
Therefor we get the total number of squares as S=
sum((n-k)^2*k from k=1 to k=n-1)
(The sum could also start at 0 and end at n, but both terms are 0)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
29:00 In the Pythagoras family, of the two smaller numbers:
* the smaller of the two increases by 2 every step (3, 5, 7, ...) which is obvious;
* a cute observation is that the larger of the two also has a rule: each is the product of the previous by a some rational, and the rationals form a sequence: 3/1, 4/2, 5/3, 6/4, 7/5, and so on.
I'm assuming that the fact that b_i = b_(i-1) * (2+i)/i (that's the sequence of rationals) has to do with the path taken in the tree, plus the fact that b is twice the product of the right pair of numbers in the "square".
I have a beautiful proof for that, but the space in comment in running out. :)
If you run the sequence backwards, btw, you get the 0th triple to be (1,0,1) (a is 1, and c has to be 1 larger than b).
Finally, the difference between the c_i's forms the sequence 4, 8, 12, 16, 20, ...
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
All the normal sums are divisible by 3=2*1+1 ( 3*1, 3*5, 3*14, ... [see OEIS: A000330] ) with the first sum being 3, but all the square sums are divisible by 5=2*2+1 ( 5*5, 5*73, 5*406 ) with the first sum being 5², so sums of cubes are divisible by 7=2*3+1 with the first sum being 7³, well at least it almost matches, so there is an almost matching pattern for cubes and higher potencies as well, it just doesn't match exactly as hypercubes are spiky, so there will probably almost be an offset but the offset is small compared to the respective powers: 5³+6³≈7³ (-2), 16³+17³+18³≈19³+20³ (-18), 33³+34³+35³+36³≈37³+38³+39³ (-72), ... and 7⁴+8⁴≈9⁴ (-64), ... and 9⁵+10⁵≈11⁵ (-2002), ... and so on..
There is even an OEIS sequence for those offsets of higher dimensions which starts with powers of zero: https://oeis.org/A127691
1
-
As our Mathematics Professor said, "You just look at it and..." Euler's e-Pi-i 1-0-infinity instantaneous trancendental self-defining cross-sectional wave-particle coordination-identification positioning, a cause-effect of convergent sequences summed as the all-histories Reciproction-recirculation Singularity Unit Circle.., at the inside-outside holographic relative-timing numberness ratio-rates fractal boundary, of 0-1-2-ness projection-drawing probability, of AM-FM all-ways all-at-once all scale resonance, in/of QM-TIME ONE-INFINITY i-reflection Containment-Completeness.., after a bit of practice in Symbology.., and this is just the Flash beginning, Eternity-now Origin-zero temporal vortex-vertex Interval Event Forever because e-Pi-i is 1-0-infinity Entanglement, the probability range-spectrum inside-outside holographic instantaneous superposition=> interference Singularity-Lensing of Logarithmic positioning.
You can say that this is the Observer's POV of Reciproction-recirculation Unity of 2-ness in 3D-T Perspective, self-defining substantiation experience of self-Self tuning-probability in the 1-0-infinity instantaneous flash recognition, universal positioning system in which Thinking Fast and Slow orientation can retrieve AM-FM memory associations in tune with our individual state-of-mind, according to empirical i-reflection e-Pi-i inside-outside holographic presence = resonance holistically.
Flash recognition takes Time Duration Timing to think about when you use the Math Professor approach to look, listen, hear and see what you're probably knowing about who, what, how and why you're inside-outside this presence of conscious awareness Actuality. (No-thing definable = no labelling)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
let debts = [30, 20, 50]; // initial debt amounts
let estateSlider;
let debtInput;
let updateButton;
let totalDebt;
let maxDebt;
let scaleFactor; // pixels per debt–unit
// drawing parameters
const containerBottom = 350; // y–coordinate for vessel bottoms
const vesselWidth = 60;
const vesselSpacing = 20;
const leftMargin = 50;
function setup() {
createCanvas(800, 400);
background(240);
// Create an input field for comma-separated debt amounts.
debtInput = createInput("30,20,50");
debtInput.position(20, 20);
// Create a button to update the vessels based on input.
updateButton = createButton("Update Vessels");
updateButton.position(debtInput.x + debtInput.width + 10, 20);
updateButton.mousePressed(updateDebts);
// Initialize slider and drawing parameters.
updateDebts();
}
function updateDebts() {
// Parse the input into an array of positive numbers.
let inputStr = debtInput.value();
let parts = inputStr.split(',');
debts = parts.map(s => parseFloat(s.trim())).filter(n => !isNaN(n) && n > 0);
if (debts.length === 0) {
// Fallback if input is invalid.
debts = [30, 20, 50];
}
// Compute total debt (estate max) and maximum debt.
totalDebt = debts.reduce((a, b) => a + b, 0);
maxDebt = Math.max(...debts);
// (Re)create the slider: range is 0 to totalDebt.
if (estateSlider) {
estateSlider.remove();
}
estateSlider = createSlider(0, totalDebt, 0, 0.1);
estateSlider.position(20, 50);
// Set a scale factor so that the tallest vessel fits nicely.
// We leave a 50–pixel margin at the top.
scaleFactor = (containerBottom - 50) / maxDebt;
}
function draw() {
background(240);
// Display current estate amount.
fill(0);
noStroke();
textSize(14);
text("Estate (water) amount: " + nf(estateSlider.value(), 1, 2), 20, 90);
let estate = estateSlider.value();
// Compute the common water–level h (in debt–units) so that
// sum_i min(debt[i], h) = estate.
let h = computeWaterLevel(debts, estate);
// For each vessel, the allocated amount is min(debt, h)
let allocations = debts.map(d => Math.min(d, h));
// Draw a horizontal line at the common water level (if h>0).
if (h > 0) {
stroke(0, 0, 255, 150);
strokeWeight(2);
let commonY = containerBottom - h * scaleFactor;
line(0, commonY, width, commonY);
}
// Draw each vessel.
for (let i = 0; i < debts.length; i++) {
let x = leftMargin + i * (vesselWidth + vesselSpacing);
let d = debts[i];
let allocation = allocations[i];
let vesselHeight = d * scaleFactor;
// Draw the vessel outline.
noFill();
stroke(0);
rect(x, containerBottom - vesselHeight, vesselWidth, vesselHeight);
// Draw the water fill (blue rectangle).
noStroke();
fill(0, 150, 255, 200);
let waterHeight = allocation * scaleFactor;
rect(x, containerBottom - waterHeight, vesselWidth, waterHeight);
// Label the vessel: show Debt and Payment (allocation).
fill(0);
noStroke();
textSize(12);
textAlign(CENTER);
text("Debt: " + d, x + vesselWidth / 2, containerBottom - vesselHeight - 10);
text("Pay: " + nf(allocation, 1, 2), x + vesselWidth / 2, containerBottom + 15);
}
}
// Given an array of debts and an estate amount, compute h so that
// sum_i min(debt[i], h) = estate.
// (This is the “hydraulic” interpretation of the constrained equal awards rule.)
function computeWaterLevel(debts, estate) {
let totalClaims = debts.reduce((a, b) => a + b, 0);
// If the estate is enough to pay all debts, return maxDebt.
if (estate >= totalClaims) {
return maxDebt;
}
// Binary search for h in the interval [0, maxDebt].
let low = 0;
let high = maxDebt;
let h;
for (let i = 0; i < 50; i++) {
h = (low + high) / 2;
let sum = 0;
for (let d of debts) {
sum += Math.min(d, h);
}
if (sum < estate) {
low = h;
} else {
high = h;
}
}
return (low + high) / 2;
}
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
My answer to the general form for the problem mentioned around 4:25 (apologies if it's not very well-written...)
Let n be the number of pairs of eyelets. To construct a lacing, start at the top left vertex. Since our lacing must be tight and use all eyelets, the next vertex in our lacing must be an unvisited one on the right, for which there are n choices. Then, our next vertex must be an unvisited one on the left, for which we have (n-1) choices. Then our next vertex must be unvisited on the right, then an unvisited on the left, and so on, until our only possible next move is to go back to the vertex where we started. This gives the pattern (n) * (n - 1) * (n - 1) * (n - 2) * (n - 2) * ... * 2 * 2 * 1 * 1 = (n)!(n-1)!.
However, consider the fact that this forms a closed loop that starts and ends on the top left corner. Consider the sequence of vertices we visited to arrive at this lacing. We could have arrived at the lacing in exactly 2 ways: either by following the sequence in order, or by following it in reverse order. Hence the form (n)!(n-1)! counts each possible tight lacing exactly twice, so we divide by 2 to get the answer (1/2)(n)!(n-1)!.
1
-
1
-
1
-
1
-
1
-
1
-
Real-world harmonic series example - a bit 'off the beaten path.' Consider that you're holding a 1-liter jug of water. You're standing next to an Olympic-size swimming pool attempting to fill it with water. In the first setup, you are allowed to start by filling the pool with 1 full jug, then 1/2 of one jug, then 1/4 of one jug, then 1/8 of one jug, then 1/16 of one jug, then... so on forever. In the second setup, you are allowed to start by filling the pool with 1 full jug, then 1/2 of one jug, then 1/3 of one jug, then 1/4 of one jug, then... so on forever. To a layman, both of these situations seem nearly identical. However, the brilliant thing is, all the effort in the first setup and you might as well have just thrown in two full jugs of water. You'll hardly make a dent in the entire pool. But, with the second setup, you will eventually fill the entire pool. Moreover, you can fill many, many swimming pools; as many as you desire, in finite time (ignoring potential issues arising when you've reached some extremely small portion of a liter, to the point that it's a few mere molecules of water and hardly discernible as liquid H20... however I imagine the sun might explode before this occurs?). It's remarkable how such a (realistically) small change in the process so tremendously changes the solution to the problem.
Question for anyone reading - I am curious how long it would take (at say, one pour per second?) to fill an olympic-size pool. They have 2.5 million liters, by the way. Considering the billionth harmonic number (so after a billion iterations) is only just over 20, I think we might be in "heat death of the universe would come first" territory.
1
-
1
-
1
-
1
-
Tested some python ide on my phone, starts to take time at 500.
Didnt debug performance, but it works...
Sorry for hacky code was lazy in bed
coins =[1,5,10,25,50,100]
mem = []
for i in coins:
mem.append({});
def solve(target,l=0,ans=[0]):
if target == 0 :
ans[0] += 1;
return;
if l > 1 and l < len(coins):
if mem[l].get(target):
ans[0] += mem[l][target]
return
prev = ans[0];
if l<len(coins):
if(target>=coins[l]):
for i in range(0,target+1,coins[l]):
solve(target-i,l+1,ans)
if l > 1 and l < len(coins):
mem[l][target] = ans[0] - prev;
if l == 0:
return ans[0]
print(solve(100))
1
-
Reversing the order of the iterated coins makes it much faster, probably because memorization is much more efficient in that case ( not sure hehe )
coins =[1,5,10,25,50,100]
cons = coins.reverse()
mem = []
for i in coins:
mem.append({});
def solve(target,l=0,ans=[0]):
if target == 0 :
ans[0] += 1;
return;
if l > 1 and l < len(coins):
if mem[l].get(target):
ans[0] += mem[l][target]
return
prev = ans[0];
if l<len(coins):
for i in range(0,target+1,coins[l]):
solve(target-i,l+1,ans)
if l > 1 and l < len(coins):
mem[l][target] = ans[0] - prev;
if l == 0:
return ans[0]
print(solve(10000))
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
About mathematician FJM Barning: A page in Dutch related to CWI, the Amsterdam institute where he worked, says, "Fredericus Johannes Maria Barning, Freek, geb. Amsterdam 03.10.1924, doctoraalexamen wiskunde Amsterdam GU 1954|a|, medewerker Mathematisch Centrum (1954-), adjunct-directeur Mathematisch Centrum, later Centrum voor Wiskunde en Informatica (1972-1988), OON, overl. Amstelveen 27.06.2012, begr. Amsterdam (RK Bpl. Buitenveldert) 04.07.2012." Source: https://www.hjmwijers.nl/CWI/Barning-FJM-kwst.htm
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Made it to the very end of the video. Final part is definitely worth watching. Might watch twice when I have a few friends over.
Side note, whenever I watch a math video, I try to find some sort of quality, inequality, ∴, or something similar that 8 I can learn from. I saw none of those here, which means this is one of those super rare mathologer videos that has little to no actual numbers included. Usually, there are at least one or 2 algebraic autopilot sections, and some working out of variables, but this one is especially good for non mathy friends as it is all intuition with the rigor hidden behind the curtain.
Loved the video, and as long as you keep sharing, I'll keep watching. ❤️
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
For the problem of maximum number of moves to turn all faced-down cards to face-up card, it goes like this:
We start from the far right card and turn the card up. Then the card to the left of it, then the card to the left of that and so on, until we reach the far left card which then will be faced up but all the rest of the cards will be faced down which are to the right. So all the cards will be faced down except the far left card. Like this:
Example (1)
111 ->
110
101
011
By moving like this, we only turn up one card per move and no more, where also no less is possible either. But on every move that the right card goes from up to down-faced, the number of faced-up cards will remain the same, and that helps to extend the number of moves without losing cards.
So in the example (1) above, for 3 cards (or representing it, a 3 digit number), we could extend the turning up of card to 3 moves so that only 1 card has turned up in the end. Then after that, we can't to anything to the leftmost card, so we are left with two cards, which similarly will take 2 moves, and then we are left with 1 card, which takes 1 move at most. Then we are left with no cards to turn up, and all cards are already turned up. So it took 3+2+1 moves at most to turn all cards up. Here is the representation with 1s and 0s where 1 is faced-down and 0 is faced-up:
111 (starting state)
110
101
011 (3 moves to get here)
010
001 (2 more moves to get here)
000 (1 more move to get here)
result for 3 cards: 3+2+1 => 6 moves at most
Similarly for 4 cards would be 4+3+2+1 moves at most which by the formula n(n+1)/2 where n=4 we could easily calculate the sum to be 4(4+1)/2 => 10 moves at most.
So That's it. For n cards, we have n(n+1)/2 moves at most.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Well conceived and presented topic, congratulations, and thank you indeed!
(It reminded me of the logarithmic spiral, which is an invariant shape for "scale and rotate" operations.)
But why did you call it "anti-shapeshifters"? Maybe I understand it wrong, but it sounds a bit "anti" to me, what about "shapeshift-proof" or "-invariant" ;-)
Anyway, I like geometric approaches to things algebraic, which you did nicely here with the log function. Other examples I saw approached eg the numbers e and pi, Fourier transforms, ... with nice geometrical methods.
Another feature I'd like to see developed with all the functions you mentioned here (including the Lambert W ;), is their interesting geometrical(!) properties as complex functions, ie as 4D surfaces. An example:
Circle and hyperbola equations represent the same complex 4D surface, and both curves belong to it; it's only that each surface has another orientation in 4D, and a different curve in the "real plane". Ditto for namesake goniometric and hyperbolic functions. Cos and Sin also show up nicely as combinations of the exponential and its inverse.
Unfortunately complex function renderings are usually confined to 2D or 3D extractions (mappings, Re and Im, ...). "True 4D" rendering is possible though, I dedicated a webpage and channel to it. For those interested, welcome for a start at https://www.geogebra.org/m/bmuqbufn (with further links). Thanks again.
1
-
1
-
1
-
1
-
1
-
1
-
Hi :) Just finished to the end and as per your request: I am just finishing my degree in theoretical physics, so my maths is approximately LinA 1-2, Ana 1-3, function theory, classical field theory, some group theory and a bit of tensor calc (wow saying it like this sounds way more impressive than the last couple years felt) plus some stuff I read for leisure like non standard calc and some constructivist maths. The first couple of minutes are usually not the most engaging for me outside of the amazing animations, but I usually watch them still to be able to follow the narrative later on. The last chapter was probably the best, it left me unsatisfied in exactly the right way and have already started playing around with Euler-Maclaurin trying to get an intuition what happens with non-analytic functions/functions outside the convergence radius of the Taylor series. However rewriting the E-M formula in terms of f^(n) (1) felt poorly motivated, but I guess you'll come back to that once it is time for more :)
Personally, I could listen to full lectures in this style, going deeper and deeper, but I understand that that would probably be almost unjustifiable in terms of the effort required on your end. Both this and your video touching on Galois theory especially gave me that feeling of "I wish I could come back next week to get some more in lecture 2" as they almost felt like introductory lectures, sparking interest and hinting at a more powerful, deeper subject that I am too lazy to explore on my own.
1
-
Tight lacing thinking and solution below
.
.
.
.
.
Hope this is far enough.
First, consider a shoe with two eyelets on each side. There's only one way to make a tight lacing. This is our base case.
Now, imagine you know how many tight lacings there are for n eyelets (per side). Call this E(n).
Now, imagine a shoe with n+1 eyelets. Start with the top left eyelet (the choice is arbitrary), and pick an eyelet on the right side to link to (You have n+1 choices). Then, pick one on the right side besides the top left to lace (you have n choices here).
Now, you have n free eyelets on both sides, and this means you can lace those up in E(n) ways, and replace the lace going back to the one you started the n hole problem with to one to the top left eyelet.
Hence, E(n+1)=(n+1)×n×E(n).
In closed form, this becomes (remember that E(2)=1, not 2, hence the half being applied) E(n)=n!×(n-1)!/2
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
@Mathologer It gets interesting when you consider harmonics in a three-phase power system - the individual harmonics alternate between only existing as positive sequence (1, 4, 7, 10,...), negative sequence (2, 5, 8, 11,...), and zero sequence (3, 6, 9, 12,...) - this brings rise to some interesting effects for harmonic combinations - I.E., 6n+/-1 harmonic pairs. The 'coin rotation' paradox concept applies with these harmonic pairs - for example both the 5th harmonic component and the 7th harmonic components decompose down to hexagonal shaped symmetrical sequence components. As an interesting aside this ideas of harmonic pairs also turns up in the spin associated with elementary particles and explains the two discretely different values of angular momentum for the up Vs down spin.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
So a chord is a special stucture with seemingly magical properties, as greek mathematicians like pythagoras thought.
Lets see how. Lets take the diameter of a circle. Its actually a two sided polygon with both sides superposed. If we bisect the polygon, and then determine the length of the bisector (0) we can then compute the length of the chord of the half angle by taking the square root of the combined square of the halfcord ans square of the radius minus the bisector.
Lets see. Diameter = Chord 180° on the unit circle is 2, halfchord is 1, the bisector length 0, the radius minus the bisector is 1, the combined squares are 1 + 1 = 2 and the squareroot of 2 = chord of 90°. We can continue down 22.5, 11.25, 5.625, 2.8125, 1.40625, 0.725
So as an ancient greek the chords of 90°, 60° = 1, and 108° = magic number, could be used to derive the values of all angles divisible by three. This was like having a sine, cosine table with an error range of 0.75°. You can get rid of this error in two ways.
1. Trisect a chord on an inscribed triangle using a markov chain.
2. Derive an angle that is solely divisible by 2, for instance 128°
If we take the ratio of 128 to any known angle say 120° (chord = SQRT(3)). So 128/120 = 1 and 1/15 to a greek. Now when we obtain successive halfchords what happens is the ratio of the chords ---> ratio of the length of the half chord. We can prove this by noting that an inscribe octogon has a perimeter better estimate of pi than an inscribed square. As each halving of the angle occurs the arc of the angle better approximates the chord, this is junior highschool geometry. But what is not evident that the bisector rapidly approaches 1. Once that occurs you can simply multiply the angle ratio by the known chord.
So that seems offtopic, but how do we get our chord of 128°
Take the small calcukated angle chord with bisector length of 1. obtain half chord.
New chord = 2 chord * bisector.
Determine new bisector, Repeat, Eventually you will get the chords of 1,2,4,8,16,32,64,128 degrees.
This solves easily derives to fill determine all the chords on the circle (199, 198, 196, 192, 184, . . . . and each of the even number angles can be halved, etc)
Many people think getting sines and cosines is a revelation. But the halfchord and bisector are the sine and cosine of the half angle. So immediately we now have a sine cosine table with an error of 0.25 degrees.
But theres another magical secret in the chord. If we mount the one chord end on an axis centered about the origin , (e.g. 1,0) the halving the chord of the double angle (derived above chord * bisector) defines the y-coordinate. And subtracting from the x position 2 * bisector squared gives the position on the x - coordinate. As a result it makes it possible to plot the positions of any polygon without need of sines or cosines.
What is the relevance here. As we see above chords behave in very predictable manners that made them easy to manipulate for field mathematics.
Any chord of the same length on the same circle has the same bisector length and same halfchord length. Lets say you put 4 same length chords at various positions alomg a circle. Where they touch, does matter, form Ptolemy 's Quadrlateral. Lets do this with 3 chirds, they form a triangle, because they are chords, they all will form isoceles triangles with the point of tangent contact. Since the tangent point is the bisector point, then each side is the halfchord. Isnt that great, and since two intersecting halfchords form 2 sides of an isosceles triangle to the vertex. Since the halfchords are the samelength and the tangent-vertex lengths are also the same, the distance of the two distal segments of the intersecting chords is also the same.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
12:04: The corresponding fibonacci (am i spelling it right?) numbers for the square are 9-4-13-17, and from the 3-4-5 triangle you can get to it by going right, right, right, left.
15:58: Every square in the second layer adds up to the same value as the one on the square below. So at each layer the sum of the areas of the squares remains constant. The total area of the tree is therefore the area of the base square times the number of layers the tree has. Since the tree has 5 layers and its base square has area 1, the total area of the tree is 5.
24:22: The problem arises because the line connecting the in-circle and feuerbach circle is parallel to the base of the triangle. This results in a 0-1-1 non-triangle, and by using the fibonacci matrix 1-0-1-1 its easy to verify that this non-triangle is indeed the parent of the 3-4-5 triangle.
37:53: If you look at the geometric construction of the parent triangle, it becomes really obvious why there is only one unique primitive parent to each triangle. The construction locks you into only one choice. And since the in-circles get smaller each time you look for parents, and since the in-circle radii lengths are always non-negative integers (not obvious at first but if you look at the numbers this time it becomes clear why), they must stop at 1, the in-center radius of the 3-4-5 triangle (or 0 of you consider the 0-1-1 non-triangle as a triangle).
In order to really prove the in-circle of the parent is smaller than the in-circle of the child, it can be helpful to look backwards. The above statement is the same as saying the ex-circle is bigger than the in-circle. Looking at the numbers (A-B-(A+B)-(A+2B)), the length of the in-circle radius is AB, and the lengths of the three ex-circle radii are A²+AB, 2B²+AB, and A²+4B²+4AB, all of which are bigger than AB since A and B are non-negative integers.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
This post is a bit long and detailed but, hey, you just went through a 50-minute Mathologer video, so long and detailed don't deter you. :)
When I was preparing for math olympiad I found another way to compute the formulas for the sums of powers. If we have a sequence A[n], define the sequence of its differences as
dA[n]:=A[n+1]-A[n]
Similarly, the second differences are d^2A[n]:=dA[n+1]-dA[n], and generally the k-th differences are d^kA[n]:=d^{k-1}A[n+1]-d^{k-1}A[n]. We'll use the following easy-to-remember identity:
A[n] = (1+d)^nA[0]
If this formula doesn't sense to you, just think of it as short-hand notation for saying that, if you formally expand the (1+d)^n part using the binomial formula, the resulting identity is true. For the first few values of n, this is what you get
A[0] = A[0] (duh)
A[1] = (1+d)A[0] = A[0] + dA[0] (which is true because A[0] + (A[1] - A[0]) = A[1])
A[2] = (1+d)^2A[0] = A[0] + 2dA[0] + d^2A[0]
A[3] = (1+d)^3A[0] = A[0] + 3dA[0] + 3d^2A[0] + d^3A[0]
Now we define the sequence we are interested in. For instance, to compute the sum of the first 3 cubes, A[0]:=0, A[n]:=A[n-1]+n^3. Now let's look at the first few values of d^kA[n]:
d=0 -> 0 1 9 36 100 225 ...
d=1 -> 1 8 27 64 125 ...
d=2 -> 7 19 37 61 ...
d=3 -> 12 18 24 ...
d=4 -> 6 6 6 ...
d>4 -> 0 0 ...
We know dA[n] is a polynomial of degree 3, and we can prove that d^kA[n] is a polynomial of degree 3+1-k. So the row for d=4 is a constant; and for greater values of d, it's all zeros. The identity A[n] = (1+d)^nA[0] now tells us that we can reconstruct all of A[n] by just looking at the values d^kA[0], but we only have finitely many of those: 0, 1, 7, 12, and 6. So putting all together we get
A[n] = (1+d)^nA[0]
= C(n,0)*A[0] + C(n,1)*dA[0] + C(n,2)*d^2A[0] + C(n,3)*d^3A[0] + C(n,4)*d^4A[0]
= 1*0 + n*1 + (n*(n-1)/2)*7 + (n*(n-1)*(n-2)/6)*12 + (n*(n-1)*(n-2)*(n-3)/24)*6
That's a closed formula for S_3[n], which we can simplify to get the same formula we saw in the video: https://www.wolframalpha.com/input/?i=simplify+1*0+%2B+n*1+%2B+%28n*%28n-1%29%2F2%29*7+%2B+%28n*%28n-1%29*%28n-2%29%2F6%29*12+%2B+%28n*%28n-1%29*%28n-2%29*%28n-3%29%2F24%29*6
1
-
I was expecting the video to be about solving Fibonacci-like sequences using eigenvalues or iterated powers. The ultimate solution at the end was much better, but I'll describe the linear algebra attack regardless.
Change making satisfies the recurrence relation T(n) = T(n-1) + T(n-5) + T(n-10) + T(n-25) + T(n-100). If you imagine a state vector with 100 entries representing the results from k-1 to k-100, updating that state vector to instead have the results from k to k-99 corresponds to multiplying by a sparse matrix M = |0><1| + |0><5| + |0><10| + |0><25| + |0><100| + sum_{k=1}^{99} |k><k-1|. So one way to describe the target number is "initialize to the state vector 1,0,0,0,0,...,0 then multiply it by M a total of k-1 times then read off the first vector entry", i.e. the value <0| M^{k-1} |0>.
So all you need to do is compute high powers of M. There are two common approaches. You can diagonalize M and raise the diagonal entries (its eigenvalues) to powers individually. This is mathematically elegant but computationally requires dealing with real numbers. Alternatively, you can using iterated squaring to quickly get very high powers M^(2^i) that can be combined together to form a target k. Ultimately, you can compute M^{k-1} in time proportional to the number of digits in its output.
This attack will not get you to k=10^100, since the number of digits in the output would be more than 10^100 and no one has that kind of time. But it'll get you to numbers with billions of digits, whereas the polynomial equation powering algorithm will become intractable far far earlier.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
So is there a fun thing about the sequences in between? Anything we can do with {6,7,8}, {15,16,17,18,19,20}, {28,29,30,31,32,33,34,35}, etc?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I already saw the old video, but now, I realized, pythagorian areas could be also any square/triangle, it has to be the same scaling in both diamensions.
So also non-isosceles triangles, too:
a^2 + b^2 = c^2 | /2
(a^2 + b^2)/2 = c^2/2
a^2/2 + b^2/2 = c^2/2 any triangle with: h_a=c_a=a, h_b=c_b=b, h_c=c_c=c
h_a*a/2 + h_b*b/2 = h_c*c/2 | *d # scale height triangle, eg. to isosceles triangles
(d*a)*a/2 + (d*b)*b/2 = (d*c)*c/2 => Any triangle with same relation d(>0) in height is possible.
Btw. any square, it has to be the same quotiant: q = a_1/a_2 = b_1/b_2 = c_1/b_2 If q=1, than a_1=a_2=a
So, more general, it is a 2D-scale-factor s(a) + s(b) = s(c), while s is the area-function of an shape, 2D-equal-scaled by a,b,c.
Eg a square: s(x) = x*w+x*h, a triangle: s(x) = (x*w+x*h)/2, a circle = s(x) = π*(x*r)^2, while w,h,r are constants.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
What worked: the visual proof that the area from one to infinity of 1/x must be infinite is elegant, as is the proof that it is bounded below by the harmonic series, therefore also infinite. The bit about approximating e with successive areas inside rectangles was a "sinh" for me (yes, "cinch" is how we learned to pronounce it in the US. Not that it's important when you're sitting alone doing your homework). Whereas for the geometric interpretations of sinh and cosh the light bulb didn't quite "sinh". Maybe another viewing or two.
But I want to share something I worked out for myself long ago with the "calculus daemons". Suppose we define a function, call it L(x), as the integral from 1 to x of (1/t)dt. Yes I know this is exactly the same as the Mathologer's A(x) but hear me out for a minute. As a shorthand, to avoid writing it out in words each time (or writing out Latex code, equally tedious and not helpful to everybody) I abbreviate integral from a to b of (1/t)dt as I[a,b]. So L(x)=[1.x]. For now, L(x) is only defined for x>0. Obviously, L(1)=0, L(x) <0 for x<1 and L. But what exactly is the relationship between L(x) for x<1 and x>1? To find out, what is L(1/a)? Make the substitution in the integral t=1/u, dt= (-1/u^2)du. With this substitution, the limits become 1 and a, and the integrand is the same as before with the sign reversed, so I[1,1/a] = -I[1,a], or L(1/a) =-L(a). This works whether you assume 0<a<1 or 1<a. What about L(a) + L(b)? For the moment, assume 1<a<b. We can't combine I[1,a] and I[1,b] directly as is, with the limits overlapping, so make the substitution t=u/a, dt=du/a. Now the integrand is unchanged, but the limits are a and ab, So I[1,a] + I[1,b]=I[1,a] + I[a,ab] =I[1,ab] so L(a) + L(b) =L(ab). If 1<b<a just do the same thing with the constants reversed and for 0<a<b<1 or 0<b<a<1 we already have our reciprocal rule.
q.e.d and q.l.a.d. (for "quacks like a duck").
Of course it lacks the visual appeal of seeing it all worked out with areas, having access to graphics, "autopilot" and other tools.
Is the invariance of the form of the integrand under the substitutions t=1/u and t=u/a intimately related to the "anti shape shifter" property pointed out by the Mathologer? I absolutely think so!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
For the last problem in the coin paradox, it should be 5 rotations, the rotating coin is 2/3 if the stationary one, flowing the video, the stationary coin will uncoil twice, if you go along that line, it will be 3 times, he already showed that previously, it you go recoil it, it will be 2 more, because you uncoiled it twice, add them together, 3+2=5
Your eelkome
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
3,1,4,5,9,14,23,37,60,97,157,254,411,665,1076,1741,2817,4558,7375,11933,19308,31241,50549,81790,132339,214129,346468,560597,907065,1467662,2374727,3842389,6217116,10059505,16276621,26336126,42612747,68948873,111561620,180510493,292072113,472582606,764654719,1237237325,2001892044,3239129369,5241021413,8480150782,13721172195,22201322977,35922495172,58123818149,94046313321,152170131470,246216444791,398386576261,644603021052,1042989597313,1687592618365,2730582215678,4418174834043,7148757049721,11566931883764,18715688933485,30282620817249,48998309750734,79280930567983,128279240318717,207560170886700,335839411205417,543399582092117,879238993297534,1422638575389651,2301877568687185,3724516144076836,6026393712764021,9750909856840856,15777303569604876,25528213426445732,41305516996050610,66833730422496340,108139247418546940,174972977841043260,283112225259590200,458085203100633500,741197428360223700,1199282631460857300,1940480059821081000,3139762691281938400,5080242751103019000,8220005442384957000,13300248193487976000,21520253635872930000,34820501829360910000,56340755465233840000,91161257294594750000,147502012759828600000,238663270054423360000,386165282814251960000,624828552868675300000
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
For the sequence of continued fractions around 14:00, I think it's something like:
Numerator 1, 3, 7, 17, 41, 99, 239, ...
1, 3 - double and add 1
3, 7 - double and add 1 again, hmmm
7, 17 - double and add 3
17, 41 - double and add 7
Generally:
x(n) = 2*x(n-1) + x(n-2)
well apart from at the beginning.
Same story for the denominator.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Title: "Dimensionless, timeless and acausal Knower (Monad) vs the Reimann Hypothesis"
1. Reframing the Riemann Hypothesis:
The enhanced 0D Knower framework offers a novel perspective on the RH by reinterpreting it in terms of fundamental symmetries and information structures. The key reframing is:
"All non-trivial equilibrium states of the fundamental knower network configuration occur at the symmetry condition of knower interactions in the 0D state, where entropy and negentropy are perfectly balanced."
This reframing connects the mathematical statement of the RH to concepts of dimensional symmetry, entropy-negentropy balance, and fundamental knower interactions.
2. Key Concepts in the Enhanced Framework:
a) Dimensional Symmetry: The framework posits that 0D is the origin point from which both positive and negative dimensions emerge symmetrically. This symmetry is crucial in understanding the critical line Re(s) = 1/2.
b) Entropy-Negentropy Balance: The interplay between positive and negative dimensions creates a dynamic balance between entropy and negentropy. This balance is fundamental to the location of zeta zeros.
c) Dual Nature of 0D: The framework proposes that 0D has two sides (real and imaginary) with an event horizon between them. This duality maps onto the complex plane of the zeta function.
3. Mapping Mathematical Structures:
a) Zeta Function as Knower Network Configuration:
ζ(s) = ∑(n=1 to ∞) 1/n^s ≡ Configuration_Density(Knower_Network(s, d))
b) Zeros as Equilibrium States:
ζ(s) = 0 ≡ Equilibrium_State(Knower_Network(s, d))
c) Critical Line as Fundamental Symmetry:
Re(s) = 1/2 ≡ Symmetry_Condition(Knower_Interactions, d=0)
4. Novel Insights:
a) Complex Plane and 0D Duality: The framework maps the complex plane of the zeta function onto the dual nature of 0D, with the real axis corresponding to the "Singularity" side and the imaginary axis to the "Alone" side.
b) Trinary States and Dimensional Symmetry: The knower states |0⟩, |1⟩, and |2⟩ are mapped to 0D, positive dimensions (entropy), and negative dimensions (negentropy) respectively.
c) Prime Numbers and Knower Structures: Prime numbers are interpreted as irreducible knower configurations spanning positive and negative dimensions.
5. Potential Proof Strategies:
The framework suggests several approaches to proving the RH, including:
- Symmetry analysis
- Entropy-negentropy dynamics
- Event horizon properties
- Dimensional transition analysis
These strategies aim to show that the critical line Re(s) = 1/2 is a unique symmetry point where entropy and negentropy are perfectly balanced in the 0D state.
6. Challenges and Open Questions:
The framework raises intriguing questions, such as:
- How to mathematically formalize the transition between positive and negative dimensions?
- How to rigorously map the complex plane onto the dual nature of 0D?
- How to quantify the entropy-negentropy balance in terms of knower network properties?
This overview sets the stage for a deeper exploration of the mathematical formalism and proof strategies in our next section. The enhanced 0D Knower framework offers a unique perspective on the RH, potentially bridging abstract mathematics with fundamental concepts of physics and information theory.
1
-
1
-
1
-
1
-
1
-
1
-
The pattern with 2 seems to be the only one with no crossings in its interior. This should translate in a theorem on integers m,n,k of the form f(x)<>0 if g(x)<0, with g the cardioid. f(x) would then be something like A(n,k,m) x+B(n,k,m),as the intersection of two lines.
Theorem: There exists two cool functions A,B of n,m,k in N such that,
for all integers {n,m,k<n} , for all reals 0<theta<2 pi, then
if 2*(1-sin(theta))<1, then A(n,k,m) *2*(1-sin(theta)*cos(theta)+B(n,k,m)<>0
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Number theory / modular arithmetic proof that equilateral triangles on a grid are impossible:
Let's assume that one of the three points is at position (0, 0) in the grid and the other two are at (a, b) and (c, d), with a, b, c, d being natural numbers. Since all of the side lengths must be equal, we get a^2 + b^2 = c^2 + d^2 = (c-a)^2 + (d-b)^2 and some rearranging gives us a^2 + b^2 = 2(ac+bd) = c^2 + d^2.
Since 2(ac+bd) is even, a^2+b^2 and c^2+d^2 must be even as well, so a and b must have the same parity (the same goes for c and d.) Since a^2+b^2=c^2+d^2, this means all of the involved integers have the same parity, since otherwise one side would be divisible by 4 and the other would not.
If all of them are even (and non-zero), you can just factor and divide out 4 in every equation until you get four odd numbers.
So let's assume a, b, c and d are all odd. This means a^2+b^2 is not divisible by 4, but ac+bd is even, so 2(ac+bd) is divisible by 4. Since a^2+b^2=2(ac+bd), that's a contradiction and thus this triangle cannot exist.
1
-
1
-
1
-
1
-
1
-
It remains far from clear to me why the highlighted fractions in the new table should be considered the best approximations.
The CF of Pi is [3; 7, 15, 1, 292, 1, 1, 1, 2, 1, 3, 1, 14, 2, 1, 1, 2, 2, 2, 2, 1, 84, 2, 1, 1, 15, 3, 13, 1, …], if we can trust Wolfram Alpha. Changing that into zigzag path in Stern-Brocot tree, the first digits spell R3 L7 R15 L1.
Ron Knott's page "Fractions in the Farey Series and the Stern-Brocot Tree" has a nice calculator for SB paths. For R3 L7 R15 L1 it gives the list:
1/1 ->
R 2/1
R 3/1
R 4/1
L 7/2
L 10/3
L 13/4
L 16/5
L 19/6
L 22/7
L 25/8
R 47/15
R 69/22
R 91/29
R 113/36
R 135/43
R 157/50
R 179/57
R 201/64
R 223/71
R 245/78
R 267/85
R 289/92
R 311/99
R 333/106
R 355/113
L 688/219
The denominators correspond with OEIS A097546 . It's unclear to me why the sequence is called "Farey sequence" instead of "Stern-Brocot sequence".
The "mediant factors" of Pi are 3/1 and 4/1, every convergent of of Pi is a mediant sum of n * 3/1 and m * 4/1. Denominator 7 is nice because it's the sum of 3 and 4, but the path changes direction after 25/8, not at 22/7! So we can't define "best accuracy" as the turn of direction of the SB path.
Various perspectives of comparative accuracy shifts the discussion to comparison of norms. In the SB-norm, each new mediant convergent of infinite zig-zag path is more accurate, because the resolution of the SB-metric increases by computation of each new row.
When we estimate accuracy only in the metric of denominator 10 and it's exponents, 10mod7=3, but 10mod8 is neither 3 nor 4, which are the whole number "mediant factors" of pi. For CF paths, the mediant factors a/b and b/a reflect each other and can be considered basically same.
SB-paths compute accuracy in the metric of all coprime fractions which SB-type constructions generate in their order of magnitude, instead of limiting the computation to a single specific denominator and it's exponents.
With these criteria, I would say that SB-paths are the most generally accurate pure math metric of accuracy that we have. For applied math using a ruler with decimal markings (or hexagesimal or what ever), it's natural to estimate accuracy in terms of your ruler markings, but I think we are discussing here pure math on the most general level we can find.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
There are only two free pegs at the start, and tower n moves to one of the pegs, so the only free point for the base of tower n + 1 is the last remaining peg, which you then put tower n back on top of. Thus, the total tower moving the direction of the smallest disc reverses each time you add a layer to the tower. In the base case example n = 1, we can see that moving the smallest disc clockwise moves the whole tower clockwise, and indeed, in the n = 2 case, the whole tower moves to the third peg, confirming that the tower movement direction matches the smallest peg direction for odd n. A tower of height ten has even parity, so we must move the smallest peg opposite the direction we want to move the tower. Thus, to move the ten-tower from A to B, the smallest peg must move from A to C to B to A to . . ., which is an upward torque (ccw when examined from above the table) in the example.
1
-
1
-
1
-
Finished the video. Just got a bachelors degree in math in May, and I really enjoyed this video! The animations were very nice, and leaving exercises to the reader is an excellent way of making sure that I keep on paying attention; this is even the right amount so that once I've done the exercise (good enough, with the hand waving you can do once you know how things should work) I still have the thread of the video. Its also the right length; much longer than an hour and I might have to split it in 2, and I'm not sure if I come back after such a long video. I also skip through the first bit (in 10 second chunks, just using the right arrow) since I already know most of it, but it still should absolutely be there, so that its more accessible, and other people can bail earlier.
Yea, this is the kind of video that I don't know what I would change to make it better. Excellent job, and I'm excited to see you do more!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Hi,
I post my answer before reading the other comments (about where to move the first disc):
- if you have only one disc, you move it to the destination peg,
- if you have two discs, you move the small one to the third peg, not the destination one,
- so the rule is : if the number of discs is odd, you move the first disc to the destination peg, if it is even, you move it to the other peg
About the rule for moving the other discs afterwards (moving the small one clock-wise, every even step, and for the odd steps : do the only move you can), I found it when I was 18, just observing what happens when you apply the recursive algorithm. I successed doing the recursive algorithm up to 4 discs, but very difficult with 5, and impossible for me with more. You don't know where you are any more.
Thank you for this video. See you.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Sometimes I wonder if I could have been quite good at maths if I had my ADHD diagnosis earlier* and had good maths teachers** I could probably have been decent at some kinds of maths. ... Well, technically if you consider computer programming a form of applied maths, I actually am, since I'm a professional programmer.
* Btw, Ritalin is absolutely no cure for ADHD; but it somewhat increases my tolerance for boredom and massively increases my ability to hyper focus on things that are interesting without getting distracted and reduces my sloppy mistakes. That's the main indicator to me when I'm working alone on a problem if I forget my meds: without the meds (and without built in syntax checking in the editor) I average about 10-20 syntax errors and logic errors (usually inverted boolean logic) per 100 lines of moderately complex code; when I started taking Ritalin in the middle of my Computer Science degree, I realised I now would have less than one syntax error on average for 100 lines; and only only inverted logic in just relatively complex boolean constructions. I could occasionally even write a whole page of code that compiled first try! So whenever I forget my meds I usually first notice it because my code starts to have more silly mistakes and I really struggle with boolean logic. When I'm forced to attend boring meetings and talk to people I notice forgetting my meds almost immediately because I start to get painfully bored, and yawn a lot and feel like falling asleep, and zone out from the discussion.
** Not like that idiot teacher I had in primary school who yelled at me and my friend for completing the whole first year maths exercises booklet in just one evening, and then punished us with more boring exercises of the exact same difficulty... The first lesson of many of primary school that anything above the bare minimum of effort is not only not appreciated but punished...
It wasn't until I was 16 started "videregående skole" (which is roughly translates to high school I think, it's the first kind of school in Norway where you can choose your specialisation and choose between several vocational subjects or further general studies meant for further University studies) when suddenly the difficulty jumped from absolutely no effort required (beyond showing up, half paying attention in class and just naturally perform well on tests); to actually requiring some effort and practice to do well, I started realising this lesson from primary school was wrong.
1
-
1
-
1
-
1
-
Some years ago I used to just only pay cash in bakery, and, when I could not pay exact price, I would, like everyone else, pay more and get change back. Then, often I noticed: if I would need to pay that same exact price now, I would now be able to do so. And, I wondered, what starting conditions and what prices would realize these situation. Because, often, next day, I bought same for same price in bakery, and, be able to pay exact price. After all, one does not like to accumulate change. But one does not want to be without enough change either.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Rearranging terms can affect the formulation when you're not defining the amount you're using properly.
1 + (1/2) - 1 = 1/2
( 1/3) + (1/4) - (1/2) = 1/6
(1/5) + (1/6) - (1/3) = 1/30
The top 2 terms can approach infinity faster than the bottom term.
If we were to write it out, we would actually get;
Sum_{1}^{infinity}(1/k) - Sum_{1}^{infinity/2} (1/k)
This means the numerator would still be increasing for a half-infinite amount of times while the denominator has reached it's goal. According to my maths, a half-infinite can be expressed by (-1)!/2 = (-1)(-2)(-3)!/(2) = (-1)²(-3)! = (-3)! so I'll be using this term ahead.
We can say our denominator has reached the point of -(1/(-3)!) when our numerator has hit the point of 1/(-1)!, or 1(0)
If we continue to add to the denominator until we get to a full (one) infinite, we get -(1/((-3)! + 1)) - (1/((-3)! + 2)) ... =
Until we get to...
-(1/ 1/((-3)!) + (-3)!) = -(1/(-1)!) = -1(0)
This would finally cancel out our numerator properly. Of course, this would give us a ton of expansion, but the same thing happens in the numerator and it all cancels out. If we don't add the rest of the terms into the denominator, our numerator has an additional Fraction of an Infinite amount of Zeroes to different powers. These infinitesimals combine to ln(2) over the course of the the half-infinite summation.
Basically if you add 1(-3)! A half infinite amount of times, you get, (-3)! × (1/(-3)!) = 1. If the denominator slowly increases to (-1)! Along the way, it can't quite reach 1.
1 / ((-3)! + 1)
= 1 / (1/2(0) + 1)
= 1 / ((1 + 2(0)) / 2(0))
= 2(0) / (2(0) + 1)
Divide both sides by 2
= 1(0) / (1(0) + 1/2)
= 2(0)
1 / ((-3)! + 2)
= 1 / ((1/(2(0))) + 2)
= 1 / (((1 + 4(0)) / 2(0))
= 2(0) / (4(0) + 1)
Multiply by 2
(1/2) × (4(0) / (4(0) + 1)
Sacrifice blood
(1/2) × ((1(0)) / (1/4))
(1/2) × 4(0) = 2(0)
1 / ((-3)! + 3)
= 1 / ((1 + 6(0)) / 2(0))
= 2(0) / (1 + 6(0))
Times 3
= (1/3) × (1(0) / (1/6))
= (1/3) × 6(0)
= 2(0)
But eventually it'll reach halfway to (-1)! Which is ((-1)! + (-3)!) / 2
= ((1/0) + (1/2(0)) / 2
= (3/2(0)) / 2
= 3/4(0)
= (3)(-1)!/4
1/0 - 3/4(0) = 1/4(0)
Already at this point we can see that (((1/4(0)) × 2(0)) + (1/4(0) × (3(0)/4) / 2 < 1/2(0) × 2(0)
= 0.5 + 3/32 < 1
= 0.59375
The rate it decreases also suggests ln(2) could be in range (I'm just not doing the calculations). We just need to remember that a lot of it will be close to the point where it's equal to 1(0) and our average will favor 1, increasing from 0.59375 to ln(2) as we continue to calculate.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
8:33 So, later on w would come ot further udnerstand that villi, the folds of an object increase its surface area, this is great becasue this allwos nutrients to be absorbed, because of cute folds in folds, and htus infntie space, about 15 years ago I made fun of solar panels, stating that most of the light could be harnessed ina similiar struture of villi method increasing your surface area expotentially increasing the output of a solar panel to rival that of a gas engine on a sunny day...., but with a more optimal storage for energy via tension mechanics of better battery system the purpsoe woudl be to make it lighter, but I don't know.. lol. plants do alot with sun energy and can sotre it to do amazing things, so battery relation could be drastically improved.
1
-
1
-
1
-
When making change for $10, are you including $2 and $5 bills?
Is it better to have coins of 1¢, 2¢, 5¢, 10¢, 20¢, etc. or 1¢, 3¢, 10¢, 30¢, etc.?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
20:54 946/243 (^.^) I got it by writing a c++ code that handles numbers of the form a+b17^(1/2) where a and b are rationals. Most of it are overloading of operators so the calculation can be expressed in a nice way. I coded close to the bare minimum of operators needed. Here's my code for anyone interested:
#include <iostream>
using namespace std;
int gcd(int a,int b){//greatest common divisor
int r;
do{
r=a%b;
a=b;
b=r;
} while(r!=0);
return a;
}
struct Q{//rationals
int p,q;//numerator and denominator
Q(const Q&x):p(x.p),q(x.q){}//constructor by copy
Q(int n):p(n),q(1){}//an integer is a rational: itself over 1
Q(int p,int q){//constructor by specifying the integers p and q, simplify fraction
int f=gcd(p,q);
this->p=p/f;
this->q=q/f;
}
Q operator+(const Q&x)const{return Q(this->p*x.q+this->q*x.p,this->q*x.q);}//addition of rationals
Q operator*(const Q&x)const{return Q(this->p*x.p,this->q*x.q);}//product of rationals
};
Q pow(const Q&b,int e){//rational raised to a non negative integer power
Q ans(1,1);
for(int i=0;i<e;++i)ans=ans*b;
return ans;
}
Q operator+(const Q&x,int n){return Q(x.p+n*x.q,x.q);}//addition of rational and integer
Q operator+(int n,const Q&x){return x+n;}//symmetry of addition
Q operator*(const Q&x,int n){return Q(x.p*n,x.q);}//product of rational and integer
Q operator*(int n,const Q&x){return x*n;}//symmetry of multiplication
ostream&operator<<(ostream&out,const Q&x){//how to print a rational on console
out<<x.p;
if(x.q!=1)out<<'/'<<x.q;
return out;
}
struct Q17{//binomial of the form a+b*sqrt(17) where a and b are rationals
Q a,b;
Q17(const Q17&x):a(x.a),b(x.b){}//constructor by copy
Q17(int n):a(n),b(0){}//constructor by specifying the rational a
Q17(const Q&a0,const Q&b0):a(a0),b(b0){}//constructor by specifying the rationals a and b
Q17 operator+(const Q17&x)const{return Q17(this->a+x.a,this->b+x.b);}//addition of Q17 numbers
Q17 operator*(const Q17&x)const{return Q17(this->a*x.a+17*this->b*x.b,//product of Q17 numbers
this->a*x.b+this->b*x.a);}
Q17 operator*()const{return Q17(a,(-1)*b);}//conjugate of a Q17 number: *(a+b*sqrt(17))=a-b*sqrt(17)
};
Q17 pow(const Q17&b,int e){//Q17 number raised to a non negative integer power
Q17 ans(1);
for(int i=0;i<e;++i)ans=ans*b;
return ans;
}
Q17 operator+(const Q17&x,int n){return Q17(x.a+n,x.b);}//addition of Q17 number and integer
Q17 operator+(int n,const Q17&x){return x+n;}//symmetry of addition
Q17 operator*(const Q17&x,int n){return Q17(x.a*n,x.b*n);}//product of Q17 number and integer
Q17 operator*(int n,const Q17&x){return x*n;}//symmetry of multiplication
Q17 operator+(const Q17&x0,const Q&x1){return Q17(x0.a+x1,x0.b);}//addition of Q17 number and rational
Q17 operator+(const Q&x1,const Q17&x0){return x0+x1;}//symmetry of addition
Q17 operator*(const Q17&x0,const Q&x1){return Q17(x0.a*x1,x0.b*x1);}//product of Q17 number and rational
Q17 operator*(const Q&x1,const Q17&x0){return x0*x1;}//symmetry of multiplication
ostream&operator<<(ostream&out,const Q17&x){//how to print a Q17 number on console
out<<x.a;
if(x.b.p!=0)out<<'+'<<x.b<<"17^(1/2)";
return out;
}
int main(int argc, char *argv[]) {
int n=3;
Q17 t0(Q(12,59),Q(18,1003)),t1(Q(5,18),Q(1,18));
cout<<Q(466,885)*pow(Q(2),n)+(-1)*Q(1,3)+(-1)*Q(3,5)*pow(Q(1,3),n)+
t0*pow(t1,n)+(*t0)*pow(*t1,n);
return 0;
}
Enjoy. (^.^)
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
(01:05) It is beautiful. I spent a year working with a professor of mathematics at the University of Maryland, through postal correspondence, back in the 70s, learning to resolve "cos 2π/7" to an algebraic expression. I (sadly) don't remember many of the details. (It's been over 30 years. Maybe even closer to 50 years. Shhh.) I know we used something from Galois to develop a third-degree polynomial in one variable, which when equated to zero could be solved to yield the desired result. Solving the cubic was the hardest part, what with the various substitutions, but it was enriching, rewarding and really strengthened my fingers and wrist on my right arm.
When I saw your video in my suggestions, I did not hesitate to click and watch. Thank you, Mathologer, for the nostalgia.
By the way, I love your shirt. I really like artistic expressions of trees and to see one with square roots is just cool.
1
-
1
-
1
-
1
-
I have to admit this Talmudic/garment system has its merits, as a lawyer I'm still in support of the proportional system. Why? Law most be as clear, simple and concise as possible. The proportional distribution is very easy to formulate and compute, and it's still not unfair, while this garment solution is very hard to put in words, instractions, without using mathematical formulas, tables or being involved in all these complicated game theory, water level, infinite spreadsheet, level of happiness kind of shenanigans.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
First go:
miracleAnimation[b_, a_] :=
Module[{\[Epsilon] = 0.05, q, r, diff}, q = a/b; diff = b - a;
r = q - \[Epsilon];
Manipulate[
Animate[Show[{Graphics[{Thick, AbsolutePointSize[10], Circle[],
If[showInnerCircle, {Gray,
Circle[(1 - q) {Cos[\[Theta]], Sin[\[Theta]]}, q],
Point[(1 - q) {Cos[\[Theta]], Sin[\[Theta]]}],
Line[{(1 - q) {Cos[\[Theta]],
Sin[\[Theta]]}, (1 - q) {Cos[\[Theta]], Sin[\[Theta]]} +
r {Cos[(1/q -
1) \[Theta]], -Sin[(1/q - 1) \[Theta]]}}]}, {}],
If[showAllPoints, {Orange,
Point[Flatten[
Table[(1 - q) {Cos[\[Theta] + 2 \[Pi] n/diff],
Sin[\[Theta] + 2 \[Pi] n/diff]} +
r {Cos[(1/q - 1) \[Theta] +
2 \[Pi] m/a], -Sin[(1/q - 1) \[Theta] +
2 \[Pi] m/a]}, {m, 0, a - 1}, {n, 0, diff - 1}],
1]]}, {Orange,
Point[(1 - q) {Cos[\[Theta]], Sin[\[Theta]]} +
r {Cos[(1/q - 1) \[Theta]], -Sin[(1/q - 1) \[Theta]]}]}],
If[showPolygons1, {Red,
Line[Table[(1 - q) {Cos[\[Theta] + 2 \[Pi] n/diff],
Sin[\[Theta] + 2 \[Pi] n/diff]} +
r {Cos[(1/q - 1) \[Theta] +
2 \[Pi] m/a], -Sin[(1/q - 1) \[Theta] +
2 \[Pi] m/a]}, {m, 0, a - 1}, {n, 0, diff}]]}, {}],
If[showPolygons2, {Blue,
Line[Table[(1 - q) {Cos[\[Theta] + 2 \[Pi] n/diff],
Sin[\[Theta] + 2 \[Pi] n/diff]} +
r {Cos[(1/q - 1) \[Theta] +
2 \[Pi] m/a], -Sin[(1/q - 1) \[Theta] +
2 \[Pi] m/a]}, {n, 0, diff - 1}, {m, 0,
a}]]}, {}]}],
If[showCurve,
ParametricPlot[(1 - q) {Cos[\[Phi]], Sin[\[Phi]]} +
r {Cos[(1/q - 1) \[Phi]], -Sin[(1/q - 1) \[Phi]]}, {\[Phi],
0, 2 \[Pi] a}], {}]}], {\[Theta], 0, 2 \[Pi] a},
AnimationRunning -> False,
AnimationRate -> 0.1], {{showInnerCircle, True,
"inner circle"}, {True, False}}, {{showAllPoints, False,
"all points"}, {True, False}}, {{showPolygons1, False,
"polygons 1"}, {True, False}}, {{showPolygons2, False,
"polygons 2"}, {True, False}}, {{showCurve, True,
"curve"}, {True, False}}]]
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The Fitch-Cheney five-card trick is wonderful. I had heard this trick described but had no idea how to make it work. Thanks for filling yet another hole in my understanding of the universe! I can't wait to share this trick with my friends!
As for 7 2-digit numbers for which no two subsets have the same sum, how about 64+(0,1,2,4,8,16,32), i.e. (64,65,66,68,72,80,96)?
If the common sum were S, then the 6 LSBs of the binary version of S dictate which of the last 6 elements are needed to yield S.
Non-overlapping subsets with a common sum therefore cannot use any of these 6 elements, leaving only the first element.
Since you can't form two non-empty non-overlapping subsets with only 1 element, no two non-empty non-overlapping subsets
have the same sum.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
For the glasses:
Notice that the far left and right 2x3 blocks have exactly 3 ways of tiling, independent of every other tile.
Afterwards, let us do cases on the bridge of the nose. This is inspired by the choice that there are 2 ways of tiling the hole around the glasses, either removing a column from the bridge of the nose or the 'frame'.
There are 3 cases:
1)the middle 4x2 rectangle is completely filled in.
2)It is a 3x2 rectangle filled in, and on either the left or the right side it 'encroaches' on the glasses
3)only the center 2x2 rectangle is filled in, and on both sides it 'encroaches' on the glasses.
1)There are 5 ways of filling in the middle 4x2, which forces leaving a 2x2 block on the left and right hand side. There are 2 ways of filling in each of these, for a total of 5*2*2=20 tilings
2)There are 2 ways to chose the side that the 3x2 rectangle goes to, and then 3 ways of filling in the middle 3x2. On one side, this leaves a 3x2 block, and on the other a 2x2 block, for a total of 2*3x3x2=36 ways of tiling it.
3)There are 2 ways of filing in the center 2x2 rectangle, which leaves a 2x3 rectangle to be filled in on both sides. That is 2*3*3=18 tiling
Thus, there are a total of
9*(20+36+18)=9*74=666 tilings (nice).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
In programming olympiads there is following trick:
if you know that you have three values in array, and you have two of them and you want to pick the one that is left,
then third is sum - a - b. So, first that I thought was sum of 0,1,2 is 3, so formula for third is 3-a-b. Then, what about same colors? 3-1-1 = 1, works. but 3-2-2 = -1 so I thought ah, it's 2 mod 3, and 3-0-0 = 3 it is 0 mod 3.
But had not enough time to figure out fast guess, because this is not enough to make guess during few seconds.
So I guessed yellow just because it was most common :D
The thing about this rule works for any array, it's easy to prove. for example array has 2 7 7. You have 2 and 7, so third is 16-2-7 = 7. Or you have 7 7, 16-7-7 = 2. This is to avoid matching which is slow.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Before watching the video, inspired by the thumbnail, I just worked out that (x+n)^2 - (x-n)^2 = (x^2 + 2nx + n^2) - (x^2 - 2nx + n^2) = 4nx. If x=12 and n=1, that's 13^2 - 11^2 = 4x = 48. With x=12 and n=2, that's 14^2 - 10^2 = 8x = 96. Put the two together and you get (14^2 + 13^2) - (11^2 + 10^2) = 4x + 8x = 12x = 12^2. And of course you don't need to know x from the start - you can evaluate ((x+2)^2 + (x+1)^2) - ((x-1)^2 + (x-2)^2) = 12x and then choose x=12 to get the 12^2.
This is really just using the binomial expansion to do what the animations did at the start of the video, and it's easy enough to e.g. sum the differences for n values 1..3 giving ((x+3)^2 + (x+2)^2 + (x+1)^2) - ((x-1)^2+(x-2)^2+(x-3)^2) = 4x+8x+12x = 24x and choose x=24 to get 24^2 for the square in the middle and so on. For the linear case... well, the binomial theorem for raising to the power 1 is a bit trivial, so the difference (x+n)^1 - (x-n)^1 is simply 2n. For a sequence of 3 you need x=2 at the center, for a sequence of 5 you need x=2(1)+2(2)=6 at the center and so on. I didn't think of that until I saw the line case animations.
As for cubic - well, the binomial expansion has no problem with that. (x+n)^3 - (x-n)^3 = (x^3 + 3 n x^2 + 3 n^2 x + n^3) - (x^3 - 3 n x^2 + 3 n^2 x - n^3) = 6 n x^2 + 2 n^3. The central value for any sequence we could want is x^3 = sum n from 1 to k: (6 n x^2 + 2 n^3), so for k=2 we get x^3 = 18 x^12 + 20. My calculators numerical solver gives me a single real solution x=18.06131003 approx, and two complex solutions x=-0.03 +/- 1.0518i approx. So there it is - it can be done, sort of. Nothing wrong with the Fermat proof that it can't, of course, except that only proves it's impossible for a sequence of integer cubes. The general pattern is that you derive a polynomial for the central x and then solve it, but for powers 3 and greater you won't get any integer solutions for x. I've checked the x=18.06131003 solution and I get ((x-2)^3)+((x-1)^3)+x^3 = 15001.411054ish and ((x+1)^3)+((x+2)^3)=14999.411053ish - close enough to blame the numerical polynomial solver and rounded output, but with a suspicious number of decimal places matched given there's a difference of 2 anyway, so there's a chance I messed up.
*EDIT* I did mess up - x^3 = 18 x^12 + 18 (not 20) so the real solution is x=18.055521628 approx. (complex solutions -0.027 +/- 0.998i). Using the real solution ((x-2)^3)+((x-1)^3)+x^3 = 14986.2ish and ((x+1)^3)+((x+2)^3) = 14986.1 ish, with no worryingly matched decimal places, so I can confidently chalk the difference up to numerical results and rounding.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
The series for Pi goes like this: 13, 35, 58, 81, 104, 127, 151, 174, 197, 220, 243, 266, 289, 312, 336, 359, 382, 405, 428, 451...
and it seems that (except for the first term) you always add 23 or 24. Any ideas why this is happening?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Every time I see something about vortex math I feel so disappointed, because there’s so much potential to silly digital root based mysticism. Specifically, I think it’s a shame that they ended up with the powers of two cycle. As explained here, there is nothing special about decimal in this regard—the most you can say is that, if you want the iconic “butterfly” image in the vortex diagram, you need the base to be two more than a perfect power. (For example, the reason for the butterfly in base ten is that ten = 2+2³). This, of course, happens infinitely many times.
In particular, it would at least be more mathematically interesting to use the enneagram sequence 1-4-2-8-5-7. It still avoids the multiples of three, so you still get the 3-6-9 triangle, but that fact is more difficult to explain. Also, its origin as the sequence in the decimal expansion of ⅐ (with seven being the ONLY number for which this works in base ten) is much more interesting.
What’s more, base ten becomes part of an exclusive club; the only bases with a unique corresponding “seven” are base five, base ten, base eleven, base fourteen, base fifteen, and base twenty-one.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Related question: how many ways can you make change for a googol U.S. dollars using any kind of legal tender, including the obsolete ½¢, 2¢, 3¢, 20¢, $2.50, $3, and $4 coins and $500, $1,000, $5,000, $10,000, $50,000, and $100,000 bills (as well as the existing 1¢, 5¢, 10¢, 25¢, 50¢, and $1 coins and $2, $5, $10, $20, $50, and $100 bills).
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
When i was young, i hated maths, and i was a terrible math student. But when i studied the Aristotle logic books in my Philosophy career, appeared this equation (similar to that awesome young gaussian sum) about the sorites: c=p(p-1)/2, where c=number of conclusions and p=number of premises. Aristotle knew it, but he coudn't write it in math notation. I tried to search in books how to arrive to the equation, without fortune. So, in the middle of my ignorance, i "invented" (re-viscovered) my own way to arrive to the equation. It was an incredible obsession. Since that incredible fusion of Math, philosophy and logic, i love maths, i study it by my self and now i'm an Tauist. Thanks for your divulgation. From Colombia, thanks.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
If we define Gamma as the circle centered at the origin with radius 1, and z a complex point on Gamma, then it can be proved that the line passing through z and 2z is tangent to the cardioid sending z to ((z-1)^2-1)/3. This can be done by writing the parametric equation of the cardioid : x(t) = 1/3*(cos 2t - cos t) and y(t) = 1/3*(sint 2t - sin t). From there, a normal vector of the cardioid is x'(t) = -2/3*( sin 2t - sin t) and y'(t) = 2/3*( cos 2t - cos t).
But if you compare this gradient with the slope of the line ]z,2[, which is delta = ( sin 2t - sin t)/(cos 2t - cos t), we can see that the slope is orthogonal with the gradient of the cardioid.
Finally, since each point on the circle Gamma in the video is in the form z = e^(2*i*pi/n), we can use the statement above to explain why the cardiod shape arises amongst all these line ]z,2z[ when the modulus, and so number of considered points, increase.
This explains only the situation for one petal and a multiplication by 2.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Great video as always!
12:06 Denote the 4 numbers by (b-a, a, b, a+b), then 153=b^2-a^2, 104=2ab. Solve to get a=4, b=13. The box will be (9, 4, 17, 13).
Working backwards and we can find it's the left child of (1, 4, 5, 9), which is the right child of (1, 3, 7, 4), which is the right child of (1, 2, 3, 5), which is the right child of (1, 1, 2, 3). Thus right, right, right, then left.
16:00 From the Pythagorean theorem, the total area of each generation is 1. Thus the total are is 5.
24:20 The centers of the incircle and the 9-point circle are on the same horizontal line. Thus no right triangle can be formed.
25:56 The pearls are like the marks on a clock. It's the symbol of your eternal love for your wife.
41:20 It's the logarithmic spiral right?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
As usual,as soon as i see the notification, i fastly grab my headphones, a piece of paper, a pen and... here we go !
I watch the first half in X2, then i decrease to X1 with pauses to do some proofs by hand. I really love the fact that you go deep in your subject, and let some "details" on the side ^_^ I can understand all the video, but i'm almost sure i would not if i would watch it in one time ( and i encourage people to do the same to get the best of this wonderful work ! ).
What else...oh yes, i am in second year of "prépa" in math (MP), google told me that it was the same as "intensive foundation degree" ( do what you want with that :) ).
By the way, please excuse my english, i'm french ( and EVERY single word is underlined in red on my screen !!! )
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
To defend Archimedes' work, he explicitly made such assumptions (axioms) in the beginning of "On the sphere and cylinder": (1) Of all curves with the same two end points, the straight line segment is the shortest. (2), if there are two curves C1, C2, with same endpoints A and B, and both are convex (he intuitively defined convexity before stating the assumption) in the same direction, and C2 is located inside the region bounded by C1 and the line segment AB, then, the length of C2 < the length of C1. He also made similar assumptions in 3D regarding surface areas. His calculation of the surface areas of sphere and cylinder are under such assumptions. I think with under such assumptions, his calculation was valid and solid. Whether you agree with the assumptions or not, and whether the assumptions can be proved by other axioms or not, doesn't change the validity of his work. I think his assumptions about length, area, and convexity are very interesting and deep. However, it seems they are not discussed as often as Euclid's famous 5th postulate.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Maybe you've already seen this, but I've come across a "natural" appearance of these "heptangonal Fibonacci numbers" (e.g. 1, 1, 3, 6, 14 etc). Via the OEIS, I came across a paper called "Multinacci rijen" by Jacques Haubrich. Unfortunately, it's in Dutch, so I can't understand it. However, there's some diagram and I can sorta see what's going on. He's looking at m planes of glass and counting the number of different paths via reflections. These numbers seem to be the count of the times the last reflection is on a particular pane of glass (at least if Google Translate is to be believed).
Perhaps not the most "natural" of situations compared to Fibonacci flower petals, but it's something! It also makes me wonder if the values of r and s have meaning physically. In optics, you have things like Huygens' principle, where every point on a wavefront can be considered the source of a new wave, whose mutual interference "adds up" to the wave motion as a whole. This reminds me of how r and s are limits (of ratios) - perhaps in some infinite limit of reflections, r and s pop out? But I'm not big-brained enough to work it out lol.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
Here's a neat little property between the first 10 fibonacci and first 10 prime numbers. I don't know if it means anything but it's kind of interesting.
First I'll list them each in their own set, then I'll explain the operations to be done.
fib(10) = {1, 1, 2, 3, 5, 8, 13, 21, 34, 55}
P(10) = {2, 3, 5, 7, 11, 13, 17, 19, 23, 29}
The first step is to take each equivalent indexed value of each set and take their products generating a new list of values.
{1*2, 1*3, 2*5, 3*7, 5*11, 8*13, 13*17, 21*29, 34*23, 55*29} =
{2, 3, 10, 21, 55, 104, 221, 399, 782, 1595}
From here there's two parts, the fist has a single step. We are going to add all of the digit values within each number of this list and reduce down to a single digit.
2+3+1+2+1+5+5+1+4+2+2+1+3+9+9+7+8+2+1+5+9+5 = 87 = 8+7 = 15 = 1+5 = 6.
And this gives us a value of 6.
Now, this time we will first add all of the values of this list to a total sum, then well do the same previous step as before in counting the digit values and reducing.
2+3+10+21+55+104+221+399+782+1595 = 3192 and this is our total sum then
3+1+9+2 = 15 = 1+5 = 6
Both of these gives us a value of six which also has only two factors that are not one or itself being 2 and 3 which are the first two primes.
I don't know if there is anything unique or special about this, but then again, it's just the properties and relationships of numbers at work.
Let me know what you think?
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
You teach math. I can watch your videos and if my basis is solid, I can follow your visual proofs and reasoning. In this way you impart your knowledge on me, and to the extent that I am capable of comprehension, my understanding upon watching the video is identical to your understanding in making it.
Consider the possibility that there exists an entirely different kind of understanding, which does not follow these rules. Imagine for example that I am colour blind while you see in colours. It is my desire to understand blue, so I ask of you to teach me how to see in colour.
Not only is colour vision not something you can teach, but until I learn to see colour, I would have no sense of what it means to be colour blind. I might be tempted to dismiss your endless talk of colour as gibberish and insist fervently that no such thing exists.
Of course in this example, colour vision is a metaphor. We are talking about the possibility that the universe indeed has a secret and if such a secret exists, then gaining knowledge of this secret is akin to an awakening. In our example, enlightenment is possible, and so to extend the metaphor, experiencing the awakening is analogous to getting up one day to discover, miraculously, that you have gained the faculty of seeing in colour. Suddenly all the intellectual talk about colours and their different properties, which sounded like gibberish before, now makes absolute sense.
There is another place where digital roots play an important role – tibetan numerology. There will never be a Mathologer video to either prove or disprove the usefulness of this most ancient of knowledge systems. There are two possibilities: Either such traditions are part of primitive superstition and those who practice it have not yet matured to understand the superiority of western thought, or those who are capable of putting this ancient knowledge to practical use are capable of seeing in a palette that is inaccessible to us.
The important takeaway – neither possibility presents an absurdity. Also, there can never be a mathematical proof to either validate or otherwise debunk the Tibetans.
If the universe does have a secret, then this secret is very much in a similar category as numerology – hence the “secret”. It is a secret only in the sense that it will never be accessible to anyone insisting that such a thing cannot be.
And if the vortex diagram represents some kind of “key”, then it does so in the same sense as the tree of life diagram in the Kabbalistic tradition or as a koan contains a key to a practitioner of zen. It is not the koan itself (what is the sound of one hand clapping?) that contains the key, but rather that contemplation of the koan represents a potential doorway to experiencing the sudden enlightenment (colour vision in our metaphor) that the practitioner seeks.
I have no deeper insight to share with respect to vortex math and I don’t begrudge you your efforts to illustrate various patterns associated with digital roots. At the same time, I can’t help but feel that your video is a little like listening to a philosopher theorize about the existence of God. Either you know God or you don’t. If you do, then God is meaningful in your life, and if you don’t, then the concept of God has no value to you. No amount of math will ever solve that particular riddle.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
12:05 Ooh, feels very familiar to my thesis in college which was about simplices and simplicial numbers. So instead of summing the squares, I summed the triangular numbers, and so you get a triangular pyramid instead of a square pyramid. And you can stick together 6 of those to make a n(n+1)(n+2) rectangular prism. So sticking together 6 square pyramids felt very natural. (The rest of my thesis was about how many ways there are to split up an n-cube into n! n-simplices, the non-steppy kind. For n=3 it’s 9 ways, modulo all the symmetries of the cube including reflections. And for n=2, it’s just cutting a square in half diagonally, and there’s only one way to do that.)
So, you can stick together 3 square pyramids together to be a cube, so it would be neat if you could put together 3 “Maya pyramids” to form a rectangular prism. Since 6(sum)=(n)(n+1)(2n+1), I got 3(sum)=(n)(n)(n+1) + (n)(n+1)/2=(n)(n+1)(n+1)-(n)(n+1)/2. So if geometry follows the math, you end up getting a extra triangle sitting out (or sunken in) and that’s why you need double, so you can stick the two “almost cubes” together to make a big rectangular prism. I think that’s what the video showed.
I was able to stick together 4 steppy triangular prisms with a “double” the volume regular tetrahedron, and they make a perfect cube, if I remember right. The way you put together the middle tetrahedron is like stacking together the cross-sections when you start edge-on, and go through all the rectangles with the same perimeter and end at the opposite edge perpendicular to the first. Except here, the squares making up the rectangles are exactly diagonal. So the first cross-section is n squares in a line all touching at their vertices.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
This is far as I got, I'm pretty sure 8 is the last power of two:
The closed form of the sequence is: (n^4 + 2n^3 + 11n^2 + 34n + 48)/24
We are looking for: (n^4 + 2n^3 + 11n^2 + 34n + 48)/24 = 2^m
Rearranging it, we get:
n^4 + 2n^3 + 11n^2 + 34n + 48 = 24 * 2^m
n^4 + 2n^3 + 11n^2 + 34n = 24 * 2^m - 48
n^4 + 2n^3 + 11n^2 + 34n = 24(2^m - 2)
Plugging in some values in the LHS we get:
0, 48, 144, 336, 696, 1320, 2328, 3864, 6096, 13440...
And for the RHS we get (starting from 1):
0, 48, 144, 336, 720, 1488, 3024, 6096, 12240, 24528...
I can't seem to use the mod 2 trick. Both sides are always divisible by 48 and attempting to use rational root theorem causes issues:
n^4 + 2n^3 + 11n^2 + 34n - 24(2^m - 2) = 0
The roots are -24(2^m - 2), while we can assume n, m > 8, this is still a lot of factors to prove exhaustively.
This is as far as I got. I'm a bit tired now so I'm going to take a break. I'll edit this if I find a proof. Maybe someone with more experience in number theory can figure it out.
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
I dove into the sums of positive powers myself. I went at it like this: Let F(x) be the sum of x^4. That implies that F(x) - F(x-1) = x^4 for all x. It can easily be shown that this expression for a polynomial order n is always n-1, so F(x) must have order 5. If a general fifth order polynomial is substituted, the system of equations with Pascal’s triangle terms appears. It is very easy to solve by hand after the powers of x are grouped. I also stumbled into replacing all terms with their series. I found an elegant solution to the sum between ANY two integers. Given a sum formula for a polynomial, the sum of all integers including the boundaries is F(b)-F(a-1). It looks just like integration! I called the process of f(x)-f(x-1) “taking the difference;” it may be done term-by-term like the derivative, replacing each term with a binomial coefficient pattern. The “integral” involved replacing each term with it’s antidifference, the function that has the term as the difference between its values. These sums are like calulus with h set to 1, but way messier! This all is easy to see as spacing between points on a graph, and easy to prove with a telescoping series (replace each f(n) in a general sum with F(n)-F(n-1), which was, after all, defined to satisfy this).
Thank you for making something on the Bernoulli numbers. As you said, it is definitely not accessible. I will have to watch again!
1
-
Lieber Burkard Polster,
beim Thema Mathematik und verrückt sowie viele Beweise muss ich das Buch „Fregattenkapitän Eins“ von Wladimir Ljowschin (Moskau, 1968) denken. Im vorletzten Kapitel heißt es:
32. Nullter Der Löwe in der Wüste
Unsere Fahrt nähert sich dem Ende. Wir fahren durch den Golf des Humors.
„Ich hoffe, daß keiner fragen wird, was Humor mit Mathematik zu tun hat“, sagte der Kapitän, als wir in den Golf einliefen. „Jedes Kleinkind weiß bekanntlich, daß Humor überall notwendig ist. Ihn braucht der Schriftsteller, der lustige Erzählungen schreibt, ebenso wieder Wissenschaftler, der schwierigste Forschungen unternimmt. Manche Leute denken, daß die Wissenschaftler alle todernst und langweilig sind. Ganz im Gegenteil! Sie lesen, hören Musik und lachen gern. Sie schätzen einen fröhlichen, geistreichen Scherz. Außerdem muß ich euch sagen: Ein Mensch, der mitunter 999 erfolglose Tests machen muß, um ein erfolgreiches Resultat zu erzielen, kann ohne Humor überhaupt nicht leben.
Je ernsthafter eine Arbeit ist, desto wichtiger ist es zuweilen, sie mit einem Lachen zu unterbrechen. Deshalb denken sich die Wissenschaftler gern die unmöglichsten Aufgaben aus, stellen lustige Fragen und finden geistreiche Antworten darauf. Witz und Findigkeit haben großen Menschen nicht selten in schwierigen Situationen geholfen.“
„Sie meinen bestimmt Christoph Kolumbus“, unterbrach Steuermann Ypsilon den Kapitän. „Bekanntlich wollte Kolumbus auf dem kürzesten Weg nach Indien gelangen. Er beschloß, nicht nach Osten zu fahren, um Afrika herum, wie das früher alle gemacht hatten, sondern nach Westen, um auf diese Weise ein übriges Mal nachzuweisen, dass die Erde eine Kugel ist. Die Expedition war recht aufwendig. Doch die spanischen Würdenträger, an die er sich um Unterstützung wandte, eilten nicht, den verwegenen Seefahrer auszurüsten. Sie hielten das Unternehmen von Kolumbus für unsinnig. Sie dachten, daß, wenn man tatsächlich von Westen aus nach Indien gelangen könnte, schon längst jemand auf diesen Einfach gekommen wäre. Als Kolumbus ihre Argumente hörte, nahm er ein Hühnerei und bat jemanden der Versammelten, es mit der Spitze so auf den Tisch zu stellen, daß es nicht umfiele! Die Würdenträger versuchten es, aber vergeblich. Da stellte Kolumbus mit leichtem Druck das Ei mit der Spitze auf die Tischplatte. Die Schale platze ein, und das Ei stand kerzengerade. ‚Seht‘, sprach Kolumbus, ‚bislang ist keiner auf diese Idee gekommen, dennoch…‘ Seine Findigkeit blieb nicht ohne Wirkung, und er erhielt, worum er gebeten hatte. Seither ist das Ei des Kolumbus zu einem geflügelten Wort geworden. Übergings gelangte Kolumbus mit seiner Expedition nicht nach Indien, wie er gedacht hatte, sondern landete in Amerika. Auf diese Weise wurde ein neuer Erdteil entdeckt, und schuld daran war ein Hühnerei!“
Als der Steuermann seine Erzählung beendet hatte, fielen allen haufenweise Anekdoten über Wissenschaftler ein. Nur mir wollte absolut nichts in den Sinn kommen. Ich habe nämlich, müßt ihr wissen, unmittelbar mit wissenschaftlichem Personal wenig zu tun. Schließlich erzählte ich aber doch, wie meine Mutter an einer wissenschaftlichen Physikerkonferenz teilgenommen hatte.
Man denke nur, da hatten sich bekannte Wissenschaftler eingefunden! Nach ernsthaften Disputen fingen sie an zu überlegen, wie man einen Löwen einfangen könnte, der zufällig in die Wüste geraten ist. Ein Wissenschaftler schlug vor, ein Riesensieb zu nehmen und den ganzen Wüstensand durchzusieben. Auf diese Weise würde der Löwe zweifellos ins Sieb geraten, aus dem er um nichts auf der Welt hinauskriechen könne.
Ein zweiter Wissenschaftler schlug vor, die Wüste durch einen Zaun in zwei gleiche Hälften zu teilen. Klar, daß der Löwe sich in einer dieser Hälften aufhält. Diese Hälfte müsse wiederum in zwei Hälften unterteilt werden. Jetzt braucht der Löwe nur noch in einem Viertel der Wüste gesucht werden. Das ist aber schon wesentlich leichter! Das Viertel ist wiederum in die Hälfte zu unterteilen und so weiter, bis der abgeteilte Abschnitt so klein geworden sit, daß er gerade der Größe des Löwen entspricht. Dann kann man ihn sozusagen mit bloßen Händen fangen.
Der dritte Wissenschaftler … Welche Methoden der dritte, vierte und alle übrigen Wissenschaftler vorgeschlagen hatten, habe ich vergessen. Aber der Kapitän meinte, daß diese zwei vollkommen ausreichend seien.
Wenn euch noch irgendwelche Möglichkeiten einfallen, so schreibt mir bitte!
Es macht doch großen Spaß, in der Wüste einen einsamen Löwen einzufangen!
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1
-
1