Piper Harron discusses her artistic and wonderful math Ph.D. thesis

This is the most “unique” PhD thesis I have ever seen. Very special, and humorous to read, and coming from the most elite institution Princeton, under the guidance of Fields Medalist Manjul Bhargava.

Cathy O'Neil, mathbabe's avatarmathbabe

Piper Harron is a mathematician who is very happy to be here, and yes, is having a great time, despite the fact that she is standing alone awkwardly by the food table hoping nobody will talk to her.

Piper, would you care to write a mathbabe post describing your thesis, and yourself, and anything else you’d care to mention?

When Cathy (Cathy?  mathbabe?) asked if I would like to write a mathbabe post describing my thesis, and myself, and anything else I’d care to mention, I said “sure!” because that is objectively the right answer. I then immediately plunged into despair.

Describe my thesis? My thesis is this thing that was initially going to be a grenade launched at my ex-prison, for better or for worse, and instead turned into some kind of positive seed bomb where flowers have sprouted beside the foundations I thought I wanted to crumble…

View original post 649 more words

2015 in review

The WordPress.com stats helper monkeys prepared a 2015 annual report for this blog.

Here’s an excerpt:

The Louvre Museum has 8.5 million visitors per year. This blog was viewed about 170,000 times in 2015. If it were an exhibit at the Louvre Museum, it would take about 7 days for that many people to see it.

Click here to see the complete report.

Rouche’s Theorem and Applications

This blog post is on Rouche’s Theorem and some applications, namely counting the number of zeroes in an annulus, and the fundamental theorem of algebra.

Rouche’s Theorem: Let f(z), g(z) be holomorphic inside and on a simple closed contour K, such that |g(z)|<|f(z)| on K. Then f and f+g have the same number of zeroes (counting multiplicities) inside K.

Rouche’s Theorem is useful for scenarios like this: Determine the number of zeroes, counting multiplicities, of the polynomial f(z)=2z^5-6z^2-z+1=0 in the annulus 1\leq |z|\leq 2.

Solution:

Let K_1 be the unit circle |z|=1. We have

\begin{aligned}|2z^5-z+1|&\leq |2z^5|+|z|+|1|\\    &=2+1+1\\    &=4\\    &<6\\    &=|-6z^2|    \end{aligned}

on K_1.

Since -6z^2 has 2 zeroes in K_1, therefore f has 2 zeroes inside K_1, by Rouche’s Theorem.

Let K_2 be the circle |z|=2

\begin{aligned}    |-6z^2-z+1|&\leq |-6z^2|+|-z|+|1|\\    &=6(2^2)+2+1\\    &=27\\    &<64\\    &=|2z^5|    \end{aligned}

on K_2. Therefore f has 5 zeroes inside K_2.

Therefore f has 5-2=3 zeroes inside the annulus.

We do a computer check using Wolfram Alpha (http://www.wolframalpha.com/input/?i=2z%5E5-6z%5E2-z%2B1%3D0). The moduli of the five roots are (to 3 significant figures): 0.489, 0.335, 1.46, 1.45, 1.45. This confirms that 3 of the zeroes are in the given annulus.

Fundamental Theorem of Algebra Using Rouche’s Theorem

Rouche’s Theorem provides a rather short proof of the Fundamental Theorem of Algebra: Every degree n polynomial with complex coefficients has exactly n roots, counting multiplicities.

Proof: Let f(z)=a_0+a_1z+a_2z^2+\dots+a_nz^n. Chose R\gg 1 sufficiently large so that on the circle |z|=R,

\begin{aligned} |a_0+a_1z+a_2z^2+\dots+a_{n-1}z^{n-1}|&\leq|a_0|+|a_1|R+|a_2|R^2+\dots+|a_{n-1}|R^{n-1}\\    &<(\sum_{i=0}^{n-1}|a_i|)R^{n-1}\\    &<|a_n|R^n\\    &=|a_nz^n|    \end{aligned}

Since a_nz^n has n roots inside the circle, f also has n roots in the circle, by Rouche’s Theorem. Since R can be arbitrarily large, this proves the Fundamental Theorem of Algebra.

Star Wars Ball Droid: How is the head attached to the body?

Just watched Star Wars: The Force Awakens, here is my review on it. Overall a good movie, enjoyed watching it. The storyline and lightsaber duels are a bit weak in my opinion. How Rey, an untrained person holding a lightsaber for the first time, managed to defeat Kylo Ren with his crossguard lightsaber remains a mystery to me. My favorite episode remains Episode 1: The Phantom Menace.

Many mysteries remain unanswered, like the identities of Rey and Snoke. Looking forward to the next episode.

Something I find very interesting is the Ball Droid BB-8. Something even more interesting about the droid is that it is not CGI effects, it is a real prop. How the head of BB-8 is being attached to the body seems to be via strong magnets.

The toy-version of BB-8 is being sold on Amazon, a possible gift idea for those who are Star Wars fans. Sphero BB-8 App-Enabled Droid

Galois Group (Example)

This post is about the Galois group of K over \mathbb{Q}, where K is the splitting field of f(x)=x^p-2, where p is an odd prime.

First we show that the polynomial f(x)=x^p-2 is irreducible over \mathbb{Q}. This follows immediately by Eisenstein’s Criterion, since 2\mid (-2), 2\nmid 1 and 2^2\nmid (-2).

Next, we show that the splitting field K of f(x) in \mathbb{C} is Q(\sqrt[p]{2},\omega), where \omega=e^{2\pi i/p} is a primitive p-th root of unity. The roots of f are \sqrt[p]2, \sqrt[p]2\omega, \sqrt[p]2\omega^2, \dots, \sqrt[p]2\omega^{p-1}.

The splitting field K contains \sqrt[p]2 and \omega=\frac{\sqrt[p]2\omega^2}{\sqrt[p]2\omega}. Thus \mathbb{Q}(\sqrt[p]2,\omega)\subseteq K.

On the other hand, \mathbb{Q}(\sqrt[p]2,\omega) contains all the roots of f, hence f splits in \mathbb{Q}(\sqrt[p]2, \omega). Thus K\subseteq\mathbb{Q}(\sqrt[p]2,\omega), since K is the smallest field that contains \mathbb{Q} and all the roots of f. All in all, we have that the splitting field K=\mathbb{Q}(\sqrt[p]2, \omega).

The next part involves determining the Galois group of K over \mathbb{Q}. We have |Gal(K/\mathbb{Q})|=[K:\mathbb{Q}]. Since [\mathbb{Q}(\sqrt[p]2):\mathbb{Q}]=p (minimal polynomial x^p-2), and [\mathbb{Q}(\omega):\mathbb{Q}]=p-1 (minimal polynomial the cyclotomic polynomial 1+x+x^2+\dots+x^{p-1}), thus |Gal(K/\mathbb{Q})|=p(p-1). Here we have used the lemma that suppose [F(\alpha):F]=m and [F(\beta):F]=n with \gcd(m,n)=1, then [F(\alpha,\beta):F]=mn.

What the Galois group does is it permutes the roots of f. Let \sigma be an element of the Galois group. \sigma(\sqrt[p]2) can possibly be \sqrt[p]2, \sqrt[p]\omega, \dots, \sqrt[p]2\omega^{p-1}, a total of p choices. Similarly, \sigma(\omega)=\omega, \omega^2, \dots,\omega^{p-1}, a total of p-1 choices. All these total up to p(p-1) elements, which is exactly the size of the Galois group.

The above Galois group Gal(K/\mathbb{Q}) is described by how its elements act on the generators. For a more concrete representation, we can actually prove that the Galois group above is isomorphic to the group of matrices \begin{pmatrix}a&b\\0&1\end{pmatrix}, where a,b\in\mathbb{F}_p, a\neq 0. We denote the group of matrices as M.

To show the isomorphism, we define a map \phi: Gal(K/\mathbb{Q})\to M, mapping \sigma_{a,b} to \begin{pmatrix}a&b\\0&1\end{pmatrix}.

Notation: \sigma_{a,b} is defined on the generators as follows, \sigma_{a,b}(\sqrt[p]2)=\sqrt[p]2\omega^b, \sigma_{a,b}(\omega)=\omega^a.

We can clearly see that the map \phi is bijective. To see it is a homomorphism, we compute \phi(\sigma_{a,b}\circ\sigma_{c,d})=\begin{pmatrix}ac&a+bd\\0&1\end{pmatrix}=\phi(\sigma_{a,b})\phi(\sigma_{c,d}).

Cellular Approximation Theorem and Homotopy Groups of Spheres

First we will state another theorem, Whitehead’s Theorem: If a map f:X\to Y between connected CW complexes induces isomorphisms f_*:\pi_n(X)\to\pi_n(Y) for all n, then f is a homotopy equivalence. If f is the inclusion of a subcomplex X\to Y, we have an even stronger conclusion: X is a deformation retract of Y.

The main theorem discussed in this post is the Cellular Approximation Theorem: Every map f:X\to Y of CW complexes is homotopic to a cellular map. If f is already cellular on a subcomplex A\subset X, the homotopy may be taken to be stationary on A. This theorem can be viewed as the CW complex analogue of the Simplicial Approximation Theorem.

Corollary: If n<k, then \pi_n(S^k)=0.

Proof: Consider S^n and S^k with their canonical CW-structure, with one 0-cell each, and with one n-cell for S^n and one k-cell for S^k. Let [f]\in\pi_n(S^k), where f:S^n\to S^k is a base-point preserving map. By the Cellular Approximation Theorem, f is homotopic to a cellular map g, where cells map to cells of same or lower dimension.

Since n<k, the n-cell S^n can only map to the 0-cell in S^k. The 0-cell in S^n (the basepoint) is also mapped to the 0-cell in S^k. Thus g is the constant map, hence \pi_n(S^k)=0.

Arzela-Ascoli Theorem and Applications

The Arzela-Ascoli Theorem is a rather formidable-sounding theorem that gives a necessary and sufficient condition for a sequence of real-valued continuous functions on a closed and bounded interval to have a uniformly convergent subsequence.

Statement: Let (f_n) be a uniformly bounded and equicontinuous sequence of real-valued continuous functions defined on a closed and bounded interval [a,b]. Then there exists a subsequence (f_{n_k}) that converges uniformly.

The converse of the Arzela-Ascoli Theorem is also true, in the sense that if every subsequence of (f_n) has a uniformly convergent subsequence, then (f_n) is uniformly bounded and equicontinuous.

Explanation of terms used: A sequence (f_n) of functions on [a,b] is uniformly bounded if there is a number M such that |f_n(x)|\leq M for all f_n and all x\in [a,b]. The sequence is equicontinous if, for all \epsilon>0, there exists \delta>0 such that |f_n(x)-f_n(y)|<\epsilon whenever |x-y|<\delta for all functions f_n in the sequence. The key point here is that a single \delta (depending solely on \epsilon) works for the entire family of functions.

Application

Let g:[0,1]\times [0,1]\to [0,1] be a continuous function and let \{f_n\} be a sequence of functions such that f_n(x)=\begin{cases}0,&0\leq x\leq 1/n\\    \int_0^{x-\frac{1}{n}}g(t,f_n(t))\ dt,&1/n\leq x\leq 1\end{cases}

Prove that there exists a continuous function f:[0,1]\to\mathbb{R} such that f(x)=\int_0^x g(t,f(t))\ dt for all x\in [0,1].

The idea is to use Arzela-Ascoli Theorem. Hence, we need to show that (f_n) is uniformly bounded and equicontinuous.

We have

\begin{aligned}|f_n(x)|&\leq |\int_0^{x-\frac{1}{n}} 1\ dt|\\    &=|x-\frac{1}{n}|\\    &\leq |x|+|\frac{1}{n}|\\    &\leq 1+1\\    &=2    \end{aligned}

This shows that the sequence is uniformly bounded.

If 0\leq x\leq 1/n,

\begin{aligned}|f_n(x)-f_n(y)|&=|0-f_n(y)|\\    &=|\int_0^{y-\frac{1}{n}} g(t,f_n(t))\ dt|\\    &\leq |\int_0^{y-\frac{1}{n}} 1\ dt|\\    &=|y-\frac{1}{n}|\\    &\leq |y-x|    \end{aligned}

Similarly if 0\leq y\leq 1/n, |f_n(x)-f_n(y)|\leq |x-y|.

If 1/n\leq x\leq 1 and 1/n\leq y\leq 1,

\begin{aligned}|f_n(x)-f_n(y)|&=|\int_0^{x-1/n} g(t,f_n(t))\ dt-\int_0^{y-1/n}g(t,f_n(t))\ dt|\\    &=|\int_{y-1/n}^{x-1/n}g(t,f_n(t))\ dt|\\    &\leq |\int_{y-1/n}^{x-1/n} 1\ dt|\\    &=|(x-1/n)-(y-1/n)|\\    &=|x-y|    \end{aligned}

Therefore we may choose \delta=\epsilon, then whenever |x-y|<\delta, |f_n(x)-f_n(y)|\leq |x-y|<\epsilon. Thus the sequence is indeed equicontinuous.

By Arzela-Ascoli Theorem, there exists a subsequence (f_{n_k}) that is uniformly convergent.

f_{n_k}(x)\to f(x)=\int_0^x g(t,f(t))\ dt.

By the Uniform Limit Theorem, f:[0,1]\to\mathbb{R} is continuous since each f_n is continuous.

Merry Christmas

Wishing all readers a Merry Christmas and Happy New Year!

For parents looking for an ideal Christmas gift for their child, do consider buying an enrichment book from Recommend Math Books. As a quote goes, “A book is a gift you can open again and again.” – Garrison Keillor

Some other excellent educational books for Christmas gifts are:

Another popular gift idea is the
All-New Kindle Paperwhite, 6″ High-Resolution Display (300 ppi) with Built-in Light, Wi-Fi – Includes Special Offers

2nd Isomorphism Theorem (Lattice Diagram)

tomcircle's avatarMath Online Tom Circle

I found this “lattice diagram”  only in an old Chinese Abstract Algebra Textbook, never seen before in any American/UK  or in French textbooks . Share here with the students who would find difficulty remembering the 3 useful Isomorphism Theorems.

Reference: 2nd Isomorphism Theorem (“Diamond Theorem”)

Let G be a group. Let H be a subgroup of G, and let N be a normal subgroup of G. Then:

1. The product HN is a subgroup of G,
The intersection H ∩ N is a normal subgroup of H, and

2. The 2 quotient groups
(HN) / N and
H / (H∩ N)
are isomorphic.

It is easy to remember using the green diagram below: (similarly can be drawn for 1st & 3rd Isomorphism)
image

image

This 2nd isomorphism theorem has been called the “diamond theorem” due to the shape of the resulting subgroup lattice with HN at the top, H∩ N…

View original post 123 more words

Quotient Ring of the Gaussian Integers is Finite

The Gaussian Integers \mathbb{Z}[i] are the set of complex numbers of the form a+bi, with a,b integers. Originally discovered and studied by Gauss, the Gaussian Integers are useful in number theory, for instance they can be used to prove that a prime is expressible as a sum of two squares iff it is congruent to 1 modulo 4.

This blog post will prove that every (proper) quotient ring of the Gaussian Integers is finite. I.e. if I is any nonzero ideal in \mathbb{Z}[i], then \mathbb{Z}[i]/I is finite.

We will need to use the fact that \mathbb{Z}[i] is an Euclidean domain, and thus also a Principal Ideal Domain (PID).

Thus I=(\alpha) for some nonzero \alpha\in\mathbb{Z}[i]. Let \beta\in\mathbb{Z}[i].

By the division algorithm, \beta=\alpha q+r with r=0 or N(r)<N(\alpha). We also note that \beta+I=r+I.

Thus,

\begin{aligned}\mathbb{Z}[i]/I&=\{\beta+I\mid\beta\in\mathbb{Z}[i]\}\\    &=\{r+I\mid r\in\mathbb{Z}[i],N(r)<N(\alpha)\}    \end{aligned}.

Since there are only finitely many elements r\in\mathbb{Z}[i] with N(r)<N(\alpha), thus \mathbb{Z}[i]/I is finite.

Behavior of Homotopy Groups with respect to Products

This blog post is on the behavior of homotopy groups with respect to products. Proposition 4.2 of Hatcher:

For a product \prod_\alpha X_\alpha of an arbitrary collection of path-connected spaces X_\alpha there are isomorphisms \pi_n(\prod_\alpha X_\alpha)\cong\prod_\alpha \pi_n(X_\alpha) for all n.

The proof given in Hatcher is a short one: A map f:Y\to \prod_\alpha X_\alpha is the same thing as a collection of maps f_\alpha: Y\to X_\alpha. Taking Y to be S^n and S^n\times I gives the result.

A possible alternative proof is to first prove that \pi_n(X_1\times X_2)\cong\pi_n(X_1)\times\pi_n(X_2), which is the result for a product of two spaces. The general result then follows by induction.

We construct a map \psi:\pi_n(X_1\times X_2)\to\pi_n(X_1)\times\pi_n(X_2), \psi([f])=([f_1],[f_2]).

Notation: f:S^n\to X_1\times X_2, f_1=p_1\circ f:S^n\to X_1, f_2=p_2\circ f:S^n\to X_2 where p_i:X_1\times X_2\to X_i are the projection maps.

We can show that \psi ([f]+[g])=\psi([f])+\psi([g]), thus \psi is a homomorphism.

We can also show that \psi is bijective by constructing an explicit inverse, namely \phi:\pi_n(X_1)\times\pi_n(X_2)\to\pi_n(X_1\times X_2), \phi([g_1],[g_2])=[g] where g:S^n\to X_1\times X_2, g(x)=(g_1(x),g_2(x)).

Thus \psi is an isomorphism.

Graph of measurable function is measurable (and has measure zero)

Let f be a finite real valued measurable function on a measurable set E\subseteq\mathbb{R}. Show that the set \{(x,f(x)):x\in E\} is measurable.

We define \Gamma(f,E):=\{(x,f(x)):x\in E\}. This is popularly known as the graph of a function. Without loss of generality, we may assume that f is nonnegative. This is because we can write f=f^+ - f^-, where we split the function into two nonnegative parts.

The proof here can also be found in Wheedon’s Analysis book, Chapter 5.

The strategy for proving this question is to approximate the graph of the function with arbitrarily thin rectangular strips. Let \epsilon>0. Define E_k=\{x\in E\mid \epsilon k\leq f(x)<\epsilon (k+1)\}, k=0,1,2,\dots.

We have |\Gamma (f,E_k)|_e\leq\epsilon |E_k|, where |\cdot|_e indicates outer measure.

Also, \Gamma(f,E)=\cup\Gamma(f,E_k), where \Gamma(f,E_k) are disjoint.

\begin{aligned}|\Gamma(f,E)|_e&\leq\sum_{k=1}^\infty|\Gamma(f,E_k)|_e\\    &\leq\epsilon(\sum_{k=1}^\infty|E_k|)\\    &=\epsilon|E|    \end{aligned}

If |E|<\infty, we can conclude |\Gamma(f,E)|_e=0 and thus \Gamma(f,E) is measurable (and has measure zero).

If |E|=\infty, we partition E into countable union of sets F_k each with finite measure. By the same analysis, each \Gamma(f,F_k) is measurable (and has measure zero). Thus \Gamma(f,E)=\bigcup_{k=1}^\infty\Gamma(f,F_k) is a countable union of measurable sets and thus is measurable (has measure zero).

Advanced Method for Proving Normal Subgroup

For beginners in Group Theory, the basic method to prove that a subgroup H is normal in a group G is to show that “left coset = right coset”, i.e. gH=Hg for all g\in G. Variations of this method include showing that ghg^{-1}\in H, gHg^{-1}=H, and so on.

This basic method is good for proving basic questions, for example a subgroup of index two is always normal. However, for more advanced questions, the basic method unfortunately seldom works.

A more sophisticated advanced approach to showing that a group is normal, is to show that it is a kernel of a homomorphism, and thus normal. Thus one often has to construct a certain homomorphism and show that the kernel is the desired subgroup.

Example: Let H be a subgroup of a finite group G and [G:H]=p, where p is the smallest prime divisor of |G|. Show that H is normal in G.

The result above is sometimes called “Strong Cayley Theorem”.

Proof: Let G act on G/H by left translation.

G\times G/H\to G/H, (g,xH)\to gxH.

This is a group action since 1\cdot xH=xH, and g_1(g_2\cdot xH)=g_1g_2xH=(g_1g_2)\cdot xH.

This action induces a homomorphism \sigma:G\to S_{G/H}\cong S_p. Let g\in\ker\sigma. \sigma(g)(xH)=xH for all xH\in G/H, i.e. gxH=xH for all x\in G. In particular when x=1, gH=H. This means that g\in H. So we have \ker\sigma\subseteq H.

Suppose to the contrary \ker\sigma\neq H, i.e. [H:\ker\sigma]>1. Let q be a prime divisor of [H:\ker\sigma].

We also have

[G:\ker\sigma]=[G:H][H:\ker\sigma]=p[H:\ker\sigma]

By the First Isomorphism Theorem, G/ker\sigma\cong\text{Im}\ \sigma\leq S_p. By Lagrange’s Theorem, [G:\ker\sigma]\mid p!, i.e. p[H:\ker\sigma]\mid p!. This implies [H:\ker\sigma]\mid(p-1)!. Finally, q\mid(p-1)! implies q\leq p-1<p.

However, q\mid [H:\ker\sigma] implies q\mid[G:\ker\sigma]=\frac{|G|}{|\ker\sigma|} which implies q\mid|G|.

This is a contradiction that p is the smallest prime divisor of |G|. Thus, H=\ker\sigma and therefore H is a normal subgroup.

This proof is pretty amazing, and hard to think of without any hints.

LaTeX to WordPress Converter

Just created a LaTeX to WordPress Converter: http://mathtuition88.blogspot.sg/2015/12/latex-to-wordpress-converter.html

Currently it is a very basic converter, just changes “$abc$” to “$ latex abc$”. To change back from WordPress to LaTeX, a simple text editor will do the job, with replace “$ latex ” with “$”.

Test code:

LaTeX: From the above inequality $|z^n|>|a_1z^{n-1}+\ldots+a_n|$ we can conclude that the polynomial $p_t(z)=z^n+t(a_1z^{n-1}+\ldots+a_n)$ has no roots on the circle $|z|=r$ when $0\leq t\leq 1$.

WordPress: From the above inequality |z^n|>|a_1z^{n-1}+\ldots+a_n| we can conclude that the polynomial p_t(z)=z^n+t(a_1z^{n-1}+\ldots+a_n) has no roots on the circle |z|=r when 0\leq t\leq 1.

Covering space projection induces isomorphisms

Proposition 4.1 (from Hatcher): A covering space projection p:(\tilde{X},\tilde{x}_0)\to (X,x_0) induces isomorphisms p_*:\pi_n(\tilde{X},\tilde{x}_0)\to\pi_n(X,x_0) for all n\geq 2.

We will elaborate more on this proposition in this blog post. Basically, we will need to show that p_* is a homomorphism and also bijective (surjective and injective).

Homomorphism

p_*([f]):=[pf]

p_*([f]+[g])=[p(f+g)]

p(f+g)(s_1,s_2,\dots,s_n)=\begin{cases}pf(2s_1,s_2,\dots,s_n)&s_1\in[0,\frac 12]\\    pg(2s_1-1,s_2,\dots,s_n)&s_1\in[\frac 12,1]    \end{cases}

p_*[f]+p_*[g]=[pf]+[pg]

(pf+pg)(s_1,s_2,\dots,s_n)=    \begin{cases}pf(2s_1,s_2,\dots,s_n)&s_1\in[0,\frac 12]\\    pg(2s_1-1,s_2,\dots,s_n)&s_1\in[\frac 12,1]    \end{cases}, which we can see is the same.

Thus, p_* is a homomorphism.

Surjective

For surjectivity, we need to use a certain Proposition 1.33: Suppose given a covering space p:(\tilde{X},\tilde{x}_0)\to (X,x_0) and a map f:(Y,y_0)\to (X,x_0) with Y path-connected and locally path-connected. Then a lift \tilde{f}:(Y,y_0)\to (\tilde{X},\tilde{x}_0) of f exists iff f_*(\pi_1(Y,y_0))\subset p_*(\pi_1(\tilde{X},\tilde{x}_0)).

Let [f]\in\pi_n(X,x_0), where f:(S_n,s_0)\to(X,x_0), n\geq 2. Since S^n is simply connected for n\geq 2, \pi_1(S_n,s_0)=0. Thus f_*(\pi_1(S_n,s_0))=0\subset p_*(\pi_1(\tilde{X},\tilde{x}_0)). By Proposition 1.33, a lift \tilde{f}:(S_n,s_0)\to (\tilde{X},\tilde{x}_0) of f exists, where p\tilde{f}=f.

i.e. we have \boxed{p_*[\tilde{f}]=[p\tilde{f}]=[f]}. Hence p_* is surjective.

Injective

Let [\tilde{f}_0]\in\ker p_*, where \tilde{f}_0:I^n\to \tilde{X} with a homotopy f_t:I^n\to X of f_0=p\tilde{f}_0 to the trivial loop f_1.

By the covering homotopy property (homotopy lifting property), there exists a unique homotopy \tilde{f}_t:I^n\to \tilde{X} of \tilde{f}_0 that lifts f_t, i.e. p\tilde{f}_t=f_t. There is a lifted homotopy of loops \tilde{f}_t starting with \tilde{f}_0 and ending with a constant loop. Hence [\tilde{f}_0]=0 in \pi_n(\tilde{X},\tilde{x}_0) and thus p_* is injective.

Rock-Paper-Scissors 石头 – 剪刀 – 布

tomcircle's avatarMath Online Tom Circle

A Chinese Mathematician Figured Out How To Always Win At Rock-Paper-Scissors – (Business Insider)

This is “Game Theory” demonstrating the Nash Equilibrium.
Very good to understand the “Kia-Soo” (Singlish means: 惊(怕)输 “afraid to lose”) syndrome of Singaporeans.

To win this game and beat the “kia-soo” mentality — 反其道而行 Adopt the reverse way of the opposition’s anticipated kia-soo way 🙂

Key points:
(1). Sequence : “R- P -S” or (中文习惯) “石头 – 剪刀 – 布”;
(2). Winner tends to stay same way in next move;
(3). Loser likely to switch to the next step in the Sequence (1).

Reflection:
In business,
(2) is where big conglomerates like IBM , HP, Sony, Microsoft etc lose because they stay put with the same strategy (Corporate Data Center, Sell thru Channel distributors with mark-up, CD/DVD music… ), and products (Mainframes, Servers, PC, CRT-TV, Packaged software…) which brought them to success but never…

View original post 78 more words

How to calculate Homology Groups (Klein Bottle)

This post will be a guide on how to calculate Homology Groups, focusing on the example of the Klein Bottle. Homology groups can be quite difficult to grasp (it took me quite a while to understand it). Hope this post will help readers to get the idea of Homology. Our reference book will be Hatcher’s Algebraic Topology (Chapter 2: Homology). I will elaborate further on the Hatcher’s excellent exposition on Homology.

This is also Exercise 5 in Chapter 2, Section 2.1 of Hatcher.

The first step to compute Homology Groups is to construct a \Delta-complex of the Klein Bottle.

klein bottle

One thing to note for \Delta-complexes, is that the vertices cannot be ordered cyclically, as that would violate one of the requirements which is to preserve the order of the vertices.

The key formula for Homology is: \boxed{H_n=\ker\partial_n/\text{Im}\ \partial_{n+1}}.

We have \ker\partial_0=\langle v\rangle, the free group generated by the vertex v, because there is only one vertex!

Next, we have \partial_1(a)=\partial_1(b)=\partial_1(c)=v-v=0. Thus \text{Im}\ \partial_1=0.

Therefore H_0=\ker\partial_0/\text{Im}\ \partial_1=\langle v\rangle /0\cong\mathbb{Z}.

Next, we have \ker\partial_1=\langle a,b,c\rangle. \partial_2U=a+b-c, \partial_2L=c+a-b. To learn more about calculating \partial_2, check out the diagram on page 105 of Hatcher.

We then have \text{Im}\ \partial_2=\langle a+b-c, c+a-b\rangle=\langle a+b-c, 2a\rangle, where we got 2a from adding the two previous generators (a+b-c)+(c+a-b).

Thus H_1=\ker\partial_1/\text{Im}\ \partial_2=\langle a,b,c\rangle/\langle a+b-c, 2a\rangle=\mathbb{Z}/2\mathbb{Z}\oplus\mathbb{Z}.

To intuitively understand the above working, we need to use the idea that elements in the quotient are “zero”. Hence a+b-c=0, implies that c=a+b, thus c can be expressed as a linear combination of a, b, thus is not a generator of H_1. 2a=0 implies that a+a=0, which gives us the \mathbb{Z}/2\mathbb{Z} part.

Finally we note that \ker\partial_2=0, and also for n\geq 3, \ker\partial_n=0 since there are no simplices of dimension greater than or equal to 3. Thus, the second homology group onwards are all zero.

In conclusion, we have H_k(K)=\begin{cases}\mathbb{Z}&k=0\\    \mathbb{Z}/2\mathbb{Z}\oplus\mathbb{Z}&k=1\\    0&\text{otherwise}    \end{cases}

Mean Value Theorem for Higher Dimensions

Let f be differentiable on a connected set E\subseteq \mathbb{R}^n, then for any x,y\in E, there exists z\in E such that f(x)-f(y)=\nabla f(z)\cdot (x-y).

Proof: The trick is to use the Mean Value Theorem for 1 dimension via the following construction:

Define g:[0,1]\to\mathbb{R}, g(t)=f(tx+(1-t)y). By the Mean Value Theorem for one variable, there exists c\in (0,1) such that g'(c)=\frac{g(1)-g(0)}{1-0}, i.e.

\nabla f(cx+(1-c)y)\cdot (x-y)=f(x)-f(y). Here we are using the chain rule for multivariable calculus to get: g'(c)=\nabla f(cx+(1-c)y)\cdot (x-y).

Let z=cx+(1-c)y, then \nabla f(z)\cdot (x-y)=f(x)-f(y) as required.

Function of Bounded Variation that is not continuous

This is a basic example of a function of bounded variation on [0,1] but not continuous on [0,1].

The key Theorem regarding functions of bounded variation is Jordan’s Theorem: A function is of bounded variation on the closed bounded interval [a,b] iff it is the difference of two increasing functions on [a,b].

Consider g(x)=\begin{cases}0&\text{if}\ 0\leq x<1\\  1&\text{if}\ x=1  \end{cases}

h(x)\equiv 0

Both g and h are increasing functions on [0,1]. Thus by Jordan’s Theorem, f(x)=g(x)-h(x)=g(x) is a function of bounded variation, but it is certainly not continuous on [0,1]!

Please help to do Survey

URL: https://career-test.com/s/sgamb?reid=210

The results of this Personality Test is quite surprisingly accurate, do give it a try to see if you are a Careerist, Entrepreneur, Harmonizer, Idealist, Hunter, Internationalist or Leader?

Do try out this Free Career Guidance Personality Test at https://career-test.com/s/sgamb?reid=210 while it is still available!

Benefits of doing the (Free) Career Test:

  • Get familiar with the top companies in Singapore
  • There are seven distinct career types based on career preferences, goals and personality. Get to know yours! (My result: Harmonizer)
  • Take part in the annual Universum survey and win prizes!

List of Fundamental Group, Homology Group (integral), and Covering Spaces

Just to compile a list of Fundamental groups, Homology Groups, and Covering Spaces for common spaces like the Circle, n-sphere (S^n), torus (T), real projective plane (\mathbb{R}P^2), and the Klein bottle (K).

Fundamental Group

Circle: \pi_1(S^1)=\mathbb{Z}

n-Sphere: \pi_1(S^n)=0, for n>1

n-Torus: \pi_1(T^n)=\mathbb{Z}^n (Here n-Torus refers to the n-dimensional torus, not the Torus with n holes)

\pi_1(T^2)=\mathbb{Z}^2 (usual torus with one hole in 2 dimensions)

Real projective plane: \pi_1(\mathbb{R}P^2)=\mathbb{Z}_2

Klein bottle K: \pi_1(K)=(\mathbb{Z}\amalg\mathbb{Z})/\langle aba^{-1}b\rangle

Homology Group (Integral)

H_0(S^1)=H_1(S^1)=\mathbb{Z}. Higher homology groups are zero.

H_k(S^n)=\begin{cases}\mathbb{Z}&k=0,n\\    0&\text{otherwise}    \end{cases}

H_k(T)=\begin{cases}\mathbb{Z}\ \ \ &k=0,2\\    \mathbb{Z}\times\mathbb{Z}\ \ \ &k=1\\    0\ \ \ &\text{otherwise}    \end{cases}

H_k(\mathbb{R}P^2)=\begin{cases}\mathbb{Z}\ \ \ &k=0\\    \mathbb{Z}_2\ \ \ &k=1\\    0\ \ \ &\text{otherwise}    \end{cases}

Klein bottle, K: H_k(K)=\begin{cases}\mathbb{Z}&k=0\\    \mathbb{Z}\oplus(\mathbb{Z}/2\mathbb{Z})&k=1\\    0&\text{otherwise}    \end{cases}

Covering Spaces

A universal cover of a connected topological space X is a simply connected space Y with a map f:Y\to X that is a covering map. Since there are many covering spaces, we will list the universal cover instead.

\mathbb{R} is the universal cover of the unit circle S^1

S^n is its own universal cover for n>1. (General result: If X is simply connected, i.e. has a trivial fundamental group, then it is its own universal cover.)

\mathbb{R}^2 is the universal cover of T.

S^2 is universal cover of real projective plane RP^2.

\mathbb{R}^2 is universal cover of Klein bottle K.