Processing math: 6%

Friday, November 18, 2016

Loose ends of smooth manifolds

 Preliminary exam prep

Here we round up some theorems that have escaped previous roundings-up. Let X,Y be smooth manifolds and f:XY a smooth map.

Theorem: (Inverse function theorem) If dfp is invertible for some pM, then there exist Up and Vf(p) connected such that f|U:UV is a diffeomorphism.

Corollary: (Stack of records theorem) If dim(X)=dim(Y), then every regular value yY has a neighborhood Vy such that f1(Y)=U1Uk, where f|Ui:UiV is a diffeomorphism.

Proof: Since yY is a regular value, dfx is surjective for all xf1(y). Since dim(X)=dim(Y) and dfx is linear, dfx is an isomorphism, hence invertible. By the inverse function theorem, there exist Ux and Vy connected such that f|U:UV is a diffeomorphism. Before we actually apply this, we need to show that f1(y) is a finite set.

First we note that by the preimage theorem, since y is a regular value, f1(y) is a submanifold of X of dimension dim(X)dim(Y)=0. Next, if f1(y)={xi} were infinite, since X is compact, there would be some limit point pX of {xi}. But then by continuity,
y=limso p\in f^{-1}(y). But then either p cannot be separated from other elements of f^{-1}(y), meaning f^{-1}(y) is not a manifold, or the sequence \{x_i\} is finite in length. Hence f^{-1}(y) = \{x_1,\dots,x_k\}. Let U_i\owns x_i and V_i\owns y be the sets asserted to exist by the inverse function theorem (the U_i may be assumed to be disjoint without loss of generality). Let V = \bigcap_{i=1}^k V_i and U_i' = f^{-1}(V)\cap U_i, for which we still have f|_{U_i'}:U_i'\to V a diffeomorphism. \square

Theorem: (Classification of manifolds) Up to diffeomorphism,
  • the only 0-dimensional manifolds are collections of points,
  • the only 1-dimensional manifolds are S^1 and \R, and
  • the only 2-dimensional compact manifolds are S^2\# (T^2)^{\#n} or S^2\#(\R\P^2)^{\#n}, for any n\geqslant 0.
Compact 2-manifolds are homeomorphic iff they are both (non)-orientable and have the same Euler characteristic. Note that
\chi\left(S^2\# (T^2)^{\#n}\right) = 2-2n, \hspace{1cm} \chi\left(S^2\#(\R\P^2)^{\#n}\right) = 2-n.
These surfaces are called orientable (on the left) and non-orientable (on the right) surfaces of genus n.

Theorem: (Stokes' theorem) For X oriented and \omega\in \Omega^{n-1}_X, \int_X d\omega = \int_{\dy X} \omega.

Proposition: The tangent bundle TX is always orientable.

Proof: Let U,V\subset X with \vp:U\to \R^n and \psi:V\to \R^n trivializing maps, and \psi\circ \vp^{-1}:\R^n\to \R^n the transition function. To show that TX is always orientable, we need to show the Jacobian of the induced transition function (determinant of the derivative) on TX is always non-negative. On TU and TV, we have trivializing maps (\varphi,d\varphi) and (\psi,d\psi), giving a transition function
(\psi\circ \varphi^{-1}, d\psi \circ d\varphi^{-1}) = (\psi\circ \varphi^{-1}, d(\psi \circ \varphi^{-1})). 
The Jacobian of this is
\det(d(\psi\circ \varphi^{-1}, d(\psi \circ \varphi^{-1}))) = \det(d(\psi\circ \varphi^{-1}), d(\psi \circ \varphi^{-1}))) = \det(d(\psi\circ \varphi^{-1}))\cdot \det(d(\psi \circ \varphi^{-1})) \>0,
and since d(\psi\circ \varphi^{-1})\neq 0 (as \psi \circ \varphi^{-1} is a diffeomorphism, its derivative is an isomorphism), the result is always positive. \square

References: Lee (Introduction to smooth manifolds, Chapter 4), Guillemin and Pollack (Differential topology, Chapter 1

Sunday, November 13, 2016

Covering spaces

 Preliminary exam prep

Let X,Y be topological spaces.

Definition: A space \widetilde X and a map p:\widetilde X\to X are called a covering space of X if either of two equivalent conditions hold:
  • There is a cover \{U_\alpha\}_{\alpha\in A} of X such that p^{-1}(U_\alpha)\cong \bigsqcup_{\beta\in B_\alpha} U_\beta.
  • Every point x\in X has a neighborhood U\subset X such that p^{-1}(U) \cong \bigsqcup_{\beta\in B}U_\beta.
We also demand that every U_\beta is carried homeomorphically onto U_\alpha (or U) by p, and the U_\alpha (or U) are called evenly covered.

Some definitions require that p be surjective. A universal cover of X is a covering space that is universal with respect to this property, in that it covers all other covering spaces. Moreover, a cover that is simply connected is immediately a universal cover.

Remark:
Every path connected (pc), locally path connected (lpc), and semi locally simply connected (slsc) space has a universal cover.

Theorem: (Lifting criterion) Let Y be pc and lpc, and \widetilde X a covering space for X. A map f:Y\to X lifts to a map \widetilde f:Y\to \widetilde X iff f_*(\pi_1(Y))\subset p_*(\pi_1(\widetilde X)).

Further, if the initial map f_0 in a homotopy f_t:Y\to X lifts to \widetilde f_0:Y\to \widetilde X, then f_t lifts uniquely to \widetilde X. This is called the homotopy lifting property. Next, we will see that path connected covers of X may be classified via a correspondence through the fundamental group.

Theorem:
Let X be pc, lpc, and slsc. There is a bijection (up to isomorphism) between pc covers p:\widetilde X\to X and subgroups of \pi_1(X), described by p_*(\pi_1(\widetilde X)).

Example:
Let X=T^2, the torus, with fundamental group \Z\oplus \Z. Below are some covering spaces of p:\widetilde X\to X with the corresponding subgroups p_*(\pi_1(\Z\oplus\Z)).
Definition: Given a covering space p:\widetilde X\to X, an isomorphism g of \widetilde X for which \id_X\circ p = p\circ g, is called a deck transformation, the collection of which form a group G(\widetilde X) under composition. Further, \widetilde X is called normal (or regular) if for every x\in X and every \widetilde x_1,\widetilde x_2\in p^{-1}(x), there exists g\in G(\widetilde X) such that g(\widetilde x_1)=\widetilde x_2.

For path connected covering spaces over path connected and locally path connected bases, being normal is equivalent to p_*(\pi_1(\widetilde X))\leqslant  \pi_1(X) being normal. In this case, G(\widetilde X)\cong \pi_1(X)/p_*(\pi_1(\widetilde X)). This simplifies even more for \widetilde X a universal cover, as \pi_1(\widetilde X)=0 then.

Theorem: Let G be a group, and suppose that every x\in X has a neighborhood U\owns x such that g(U)\cap h(U) = \emptyset whenever g\neq h\in G. Then:
  • The quotient map q:X\to X/G describes a normal cover of X/G.
  • If X is pc, then G = G(X).
A group action satisfying the hypothesis of the previous theorem is called a covering space action.

Proposition: For any n-sheeted covering space \widetilde X\to X of a finite CW complex, \chi(\widetilde X) = n\chi(X).

References: Hatcher (Algebraic Topology, Chapter 1)

Thursday, November 10, 2016

Differential 1-forms are closed if and only if they are exact

 Preliminary exam prep

The title refers to 1-forms in Euclidean n-space \R^n, for n\geqslant 2. This theorem is instructive to do in the case n=2, but we present it in general. We will use several facts, most importantly that the integral of a function f:X\to Y over a curve \gamma:[a,b]\to X is given by
\int_\gamma f\ dx_1\wedge \cdots \wedge dx_k = \int_a^b (f\circ \gamma)\ d(x_1\circ \gamma)\wedge \cdots \wedge d(x_n\circ \gamma),
where x_1,\dots,x_n is some local frame on X. We will also use the fundamental theorem of calculus and one of its consequences, namely
\int_a^b \frac{\dy f}{\dy t}(t)\ dt = f(b)-f(a).

Theorem:
A 1-form on \R^n is closed if and only if it is exact, for n\geqslant 2.

Proof: Let \omega = a_1dx_1+\cdots a_ndx_n\in \Omega^1_{\R^n} be a 1-form on \R^n. If there exists \eta\in \Omega^0_{\R^n} such that d\eta = \omega, then d\omega = d^2\eta = 0, so the reverse direction is clear. For the forward direction, since \omega is closed, we have
0 = d\omega = \sum_{i=1}^n \frac{\dy a_1}{\dy x_i}dx_i \wedge dx_1 + \cdots + \sum_{i=1}^n \frac{\dy a_n}{\dy x_n} dx_i\wedge dx_n \ \ \ \implies\ \ \ \frac{\dy a_i}{\dy x_j} = \frac{\dy a_j}{\dy x_i}\ \forall\ i\neq j.
Now fix some (\textbf{x}_1,\dots,\textbf{x}_n)\in \R^n, and define f\in \Omega^0_{\R^n} by
f(\textbf x_1,\dots,\textbf x_n) = \int_{\gamma(\textbf x_1,\dots,\textbf x_n)}\omega,
for \gamma the composition of the paths
\begin{array}{r c l} \gamma_1\ :\ [0,\textbf x_1] & \to & \R^n, \\ t & \mapsto & (t,0,\dots,0), \end{array} \hspace{5pt} \begin{array}{r c l} \gamma_2\ :\ [0,\textbf x_2] & \to & \R^n, \\ t & \mapsto & (\textbf x_1,t,0,\dots,0), \end{array} \hspace{5pt}\cdots\hspace{5pt} \begin{array}{r c l} \gamma_n\ :\ [0,\textbf x_n] & \to & \R^n, \\ t & \mapsto & (\textbf x_1,\dots,\textbf x_{n-1},t). \end{array}
By applying the definition of a pullback and the change of variables formula (use s=\gamma_i(t) for every i),
\begin{align*} \int_{\gamma(\textbf x_1,\dots,\textbf x_n)}\omega  & = \sum_{i=1}^n \int_{\gamma_i} a_1 dx_1 + \cdots + \sum_{i=1}^n \int_{\gamma_i}a_n dx_n \\ & = \sum_{i=1}^n \int_{\gamma_i} a_1(x_1,\dots,x_n)\ dx_1 + \cdots + \sum_{i=1}^n \int_{\gamma_i}a_n(x_1,\dots,x_n)\ dx_n \\ & = \sum_{i=1}^n \int_0^{\textbf x_i} a_1(\gamma_i(t))\ d(x_1\circ \gamma_i)(t) + \cdots + \sum_{i=1}^n \int_0^{\textbf x_i}a_n(\gamma_i(t))\ d(x_n\circ \gamma_i)(t) \\ & = \int_0^{\textbf x_1} a_1(\gamma_1(t))\gamma'_1(t)\ dt + \cdots + \int_0^{\textbf x_n}a_n(\gamma_n(t))\gamma_n'(t)\ dt \\ & = \int_{(0,\dots,0)}^{(\textbf x_1,0,\dots,0)} a_1(s)\ ds + \cdots + \int_{(\textbf x_1,\dots,\textbf x_{n-1},0)}^{(\textbf x_1,\dots,\textbf x_n)}a_n(s)\ ds \\ & = \int_0^{\textbf x_1} a_1(s,0,\dots,0)\ ds + \cdots + \int_0^{\textbf x_n}a_n(\textbf x_1,\dots,\textbf x_{n-1}, s)\ ds. \end{align*}
To take the derivative of this, we consider the partial derivatives first. In the last variable, we have
\frac{\dy f}{\dy \textbf x_n} = \frac\dy{\dy \textbf x_n}\int_0^{\textbf x_n}a_n(\textbf x_1,\dots,\textbf x_{n-1}, s)\ ds = a_n(\textbf x_1,\dots,\textbf x_n) = a_n.
In the second-last variable, applying one of the identities from \omega being closed, we have
\begin{align*} \frac{\dy f}{\dy \textbf x_{n-1}}  & = \frac\dy{\dy \textbf x_{n-1}}\int_0^{\textbf x_{n-1}}a_{n-1}(\textbf x_1,\dots,\textbf x_{n-2}, s,0)\ ds  +  \frac\dy{\dy \textbf  x_{n-1}}\int_0^{\textbf x_n}a_n(\textbf x_1,\dots,\textbf x_{n-1}, s)\ ds \\ & = a_{n-1}(\textbf x_1,\dots,\textbf x_{n-1},0) + \int_0^{\textbf x_n}\frac{\dy a_n}{\dy \textbf x_{n-1}}(\textbf x_1,\dots,\textbf x_{n-1}, s)\ ds \\ & = a_{n-1}(\textbf x_1,\dots,\textbf x_{n-1},0) + \int_0^{\textbf x_n}\frac{\dy a_{n-1}}{\dy s}(\textbf x_1,\dots,\textbf x_{n-1}, s)\ ds \\ & = a_{n-1}(\textbf x_1,\dots,\textbf x_{n-1},0) + a_{n-1}(\textbf x_1,\dots,\textbf x_n) - a_{n-1}(\textbf x_1,\dots,\textbf x_{n-1}, 0) \\ & = a_{n-1}(\textbf x_1,\dots, \textbf x_n) \\ & = a_{n-1}. \end{align*}
This pattern continues. For the other variables we have telescoping sums, and we compute the partial derivative in the first variable as an example:
\begin{align*} \frac{\dy f}{\dy \textbf x_1}  & = \frac{\dy}{\dy \textbf x_1}\int_0^{\textbf x_1}a_1(s,0,\dots,0)\ ds + \sum_{i=2}^n\frac\dy{\dy \textbf x_1}\int_0^{\textbf x_i}a_i(\textbf x_1,\dots,\textbf x_{i-1}, s,0,\dots,0)\ ds \\ & = a_1(\textbf x_1,0,\dots,0) + \sum_{i=2}^n \int_0^{\textbf x_i} \frac {\dy a_i}{\dy \textbf x_1} (\textbf x_1,\dots,\textbf x_{i-1}, s,0,\dots,0)\ ds \\ & = a_1(\textbf x_1,0,\dots,0) + \sum_{i=2}^n \int_0^{\textbf x_i} \frac {\dy a_1}{\dy s} (\textbf x_1,\dots,\textbf x_{i-1}, s,0,\dots,0)\ ds \\ & = a_1(\textbf x_1,0,\dots,0) + \sum_{i=2}^n \left(a_1(\textbf x_1,\dots,\textbf x_i, 0,\dots,0) - a_1(\textbf x_1,\dots,\textbf x_{i-1},0,\dots,0)\right) \\ & = a_1(\textbf x_1,\dots,\textbf x_n) \\ & = a_1. \end{align*}
Hence we get that
df = \frac{\dy f}{\dy x_1} dx_1 + \cdots + \frac{\dy f}{\dy x_n}dx_n = a_1dx_1 + \cdots + a_ndx_n = \omega,
so \omega is exact. \square

References: Lee (Introduction to smooth manifolds, Chapter 11)

Tuesday, November 8, 2016

More (co)homological constructions

 Preliminary exam prep

Recall a previous post (2016-09-16, "Complexes and their homology") that focused on constructing topological spaces in different ways and recovering the homology. Here we complete that task, introducing cellular homology. Recall a cell complex (or CW complex) X was a sequence of skeleta X_k for k=0,\dots,\dim(X) consisting of k-cells e^k_i and their attaching maps to the (k-1)-skeleton.

Cellular homology


Definition: The long exact sequence in relative homology for the pair X_k,X_{k-1} shares terms with the long exact sequence for the pair X_{k+1},X_k, as well as X_{k-1},X_{k-2}. By letting d_k be the composition of maps in different long exact sequences, for k>1, that make the diagram
commute, we get a complex of equivalence classes of chains
\cdots \to H_{k+1}(X_{k+1},X_k) \tov{d_{k+1}} H_k(X_k,X_{k-1})\tov{d_k} H_{k-1}(X_{k-1},X_{k-2})\to \cdots \to H_1(X_1,X_0)\tov{d_1} H_0(X_0) \tov{d_0} 0,
whose homology H_k^{CW}(X) = \ker(d_k)/\text{im}(d_{k-1}) is called the cellular homology of X. The map d_1 is the connecting map in the long exact sequence of the pair X_1,X_0, and d_0=0.

This seems quite a roundabout way of defining homology groups, but it turns out to be very useful. Note that for k=1, the map d_1 is the same as for a simplicial complex, hence

Theorem:
In the context above,
  1. for k\>0, H^{CW}_k(X)\cong H_k(X);
  2. for k\>1, H_k(X_k,X_{k-1})=\Z^\ell, where \ell is the number of k-cells in X; and
  3. for k\>2, d_k(e^k_i) = \displaystyle\sum_j\deg(\underbrace{\dy e^k_i}_{S^{k-1}}\tov{f_{k,i}} X_{k-1}\tov{\pi} \underbrace{X_{k-1}/X_{k-1}-e^{k-1}_j}_{S^{k-1}})e^{k-1}_j.
Example: Real projective space \R\P^n has a cell decomposition with one cell in each dimension, and 2-to-1 attaching maps \dy(e_k) =2X_{k-1} for k>1. This gives us a construction
X_0 = e_0, \hspace{1cm} X_1 = e_1 \bigsqcup_{\dy(e_1)=e_0} X_0, \hspace{1cm} X_2 = e_2 \bigsqcup_{\dy(e_2)=2e_1} X_1, \hspace{1cm} X_3 = e_3 \bigsqcup_{\dy(e_3)=2e_2} X_2, \dots It is immediate that d_0=d_1=0, and for higher degrees, we have
d_k(e^k) = \deg(S^{k-1}\to \R\P^{k-1}\to S^{k-1})e^{k-1}.
Since this is a map between spheres, we may apply local degree calculations. The first part is the 2-to-1 cover, where every point in \R\P^{k-1} is covered by two points from S^{k-1}, one in each hemisphere. One covers it via the identity, the other via the antipodal map. As long as we choose a point not in \R\P^{k-2}\subset \R\P^{k-1}, the second step doesn't affect these degree calculations. The antipodal map S^{k-1}\to S^{k-1} has degree (-1)^k, hence for a the antipodal map, the composition has degree
\deg(S^{k-1}\to \R\P^{k-1}\to S^{k-1}) = \deg(\id_{S^{k-1}}) + \deg(a_{S^{k-1}}) = 1+(-1)^k = \begin{cases} 2 & k\text{ even}, \\ 0 & k \text{ odd.}\end{cases}

Products in (co)homology


Recall that an n-chain on X is a map \sigma:\Delta^n\to X, where \Delta^n=[v_0,\dots,v_n] is an n-simplex. These form the group C_n of n-chains. An n-cochain is an element of C^n = \Hom(C_n,\Z), though the coefficient group does not need to be \Z necessarily.

Definition: The diagonal map X\to X\times X induces a map on cohomology H^*(X\times X)\to H^*(X), and by Kunneth, this gives a map H^*(X)\otimes H^*(X)\to H^*(X), and is called the cup product.

For a\in H^p(X) and b\in H^q(X), representatives of the class a are in \Hom(C_p,\Z) and representatives of the class b are in \Hom(C_q,\Z), though we will conflate the notation for the class with that of a representative. Hence for a (p+q)-chain \sigma the cup product of a and b acts as
(a\smile b)\sigma = a\left(\sigma|_{[v_0,\dots,v_p]}\right)\cdot b\left(\sigma|_{[v_p,\dots,v_{p+q}]}\right).
Definition: The cap product combines p-cochains with q-chains to give (q-p)-chains, by
\begin{array}{r c l} \frown\ :\ H^p(X) \times H_q(X) & \to & H_{q-p}(X), \\\ (a, \sigma) & \mapsto & a\left(\sigma|_{[v_0,\dots,v_p]}\right)\cdot \sigma|_{[v_p,\dots,v_q]}. \end{array}
The cap product with the orientation form of an orientable manifold X gives the isomorphism of Poincare duality.

Remark: Given a map f:X\to Y, the cup and cap products satisfy certain identities via the induced map on cohomology groups. Let a,b\in H^*(Y) and c\in H_*(X) be cochain and chain classes, for which
f^*(a\smile b) = f^*(a)\smile f^*(b), \hspace{1cm} a\frown f_*c = f_*(f^*a\frown c).
The first identity asserts that f^* is a ring homomorphism and the second describes the commutativity of an appropriate diagram. The cup and cap products are related by the equation
a(b\frown \sigma) = (a\smile b)\sigma,for a\in H^p, b\in H^q and \sigma\in C_{p+q}.

References: Hatcher (Algebraic topology, Chapter 2.2), Prasolov (Elements of homology theory, Chapter 2)

Monday, November 7, 2016

Images of manifolds and transversality

 Preliminary exam prep

Let X,Y be manifolds embedded in \R^n, and f:X\to Y a map, with df_x:T_xX\to T_{f(x)}Y the induced map on tangent spaces.

Definition: The map f is a
  • homeomorphism if it is continuous and has a continuous inverse, 
  • diffeomorphism if it is smooth and has a smooth inverse,
  • injection if f(a)=f(b) implies a=b,
  • immersion if df_x is injective for all x\in X,
  • embedding if it is an immersion and df_x is a homeomorphism onto its image,
  • submersion if df_x is surjective for all x\in X.
Transversality is a mathematical relic whose only practical use is, perhaps, in classical algebraic geometry.

Definition: The manifolds X and Y are transverse if T_pX\oplus T_pY \cong \R^n for every p\in X\cap Y. The map f and Y are transverse if \text{im}(f) and Y are transverse.

Note that being transverse (or transversal) is a symmetric, but not a reflexive, nor a transitive relation. Recall that a regular value of f is y\in Y such that df_x:T_xX \to T_{f(x)}Y is surjective for all x\in f^{-1}(y). If y is not in the image of f, then f^{-1}(y) is empty, so y is trivially a regular value. Every value that is not a regular value is a critical value.

Theorem: (Preimage theorem) For every regular value y of f, the subset f^{-1}(y)\subset X is a submanifold of X of dimension \dim(X)-\dim(Y).

Now let M be a submanifold of Y.

Corollary: If f is transverse to M, then f^{-1}(M) is a manifold, with \codim_Y(M)=\codim_X(f^{-1}(M)).

Theorem: (Transversality theorem) Let \{g_s:X\to Y\ |\ s\in S\} be a smooth family of maps. If g:X\times S\to Y is transverse to M, then for almost every s\in S the map g_s is transverse to M.

If we replace f with df, and ask that it be transverse to M, then df|_s is also transverse to M.

Example: Consider the map g_s:X\to \R^n given by g_s(X)=i(X)+s=X+s, where i is the embedding of X into \R^n. Since g(X\times \R^n)=\R^n and g varies smoothly in both variables, we have that g is transverse to X. Hence by the transversality theorem, X is transverse to its translates X+s for almost all s\in \R^n.

Theorem: (Sard) For f smooth and \dy Y=\emptyset, almost every y\in Y is a regular value of f and f|_{\dy X}. Equivalently, the set of critical values of f has measure zero.

Resources: Guillemin and Pollack (Differential topology, Chapters 1, 2), Lee (Introduction to smooth manifolds, Chapter 6)

Friday, November 4, 2016

Tools of homotopy

 Preliminary exam prep

Let X,Y be topological spaces and A a subspace of X. Recall that a path in X is a continuous map \gamma:I\to X, and it is closed (or a loop), if \gamma(0)=\gamma(1). When X is pointed at x_0, we often require \gamma(0)=x_0, and call such paths (and similarly loops) based.

Definitions


Definition:
  • X is connected if it is not the union of two disjoint nonempty open sets.
  • X is path connected if any two points in X have a path connecting them, or equivalently, if \pi_0(X)=0.
  • X is simply connected if every loop is contractible, or equivalently, if \pi_1(X)=0.
  • X is semi-locally simply connected if every point has a neighborhood whose inclusion into X is \pi_1-trivial.
Path connectedness and simply connectedness have local variants. That is, for P either of those properties, a space is locally P if for every point x and every neighborhood U\owns x, there is a subset V\subset U on which P is satisfied.

Remark: In general, X is n-connected whenever \pi_r(X)=0 for all r\leqslant n. Note that 0-connected is path connected and 1-connected is simply connected and connected. Also observe that the suspension of path connected space is simply connected.

Definition:
  • A retraction (or retract) from X to A is a map r:X\to A such that r|_A = \id_A.
  • A deformation retraction (or deformation retract) from X to A is a family of maps f_t:X\to X continuous in t,X such that f_0 = \id_X, f_1(X) = A, and f_t|_A = \id_A for all t.
  • A homotopy from X to Y is a family of maps f_t:X\to Y continuous in t,X.
  • A homotopy equivalence from X to Y is a map f:X\to Y and a map g:Y\to X such that g\circ f \simeq \id_X and f\circ g \simeq \id_Y.
Definition: A pair (X,A), where A\subset X is a closed subspace, is a good pair, or has the homotopy extension property (HEP), if any of the following equivalent properties hold:
  • there exists a neighborhood U\subset X of A such that U deformation retracts onto A,
  • X\times \{0\}\cup A\times I is a retract of X\times I, or
  • the inclusion i:A\hookrightarrow X is a cofibration.
In some texts such a pair (X,A) is called a neighborhood deformation retract pair, and HEP is reserved for any map A\to X, not necessarily the inclusion, that is a cofibration. For more on cofibrations, see a previous blog post (2016-07-31, "(Co)fibrations, suspensions, and loop spaces").

Definition: There is a functor \pi_1:\text{Top}_*\to \text{Grp} called the fundamental group, that takes a pointed topological space X to the space of all pointed loops on X, modulo path homotopy.

This may be generalized to \pi_n, which takes X to the space of all pointed embeddings of S^n.

Definition: Let G,H be groups. The free product of G and H is the group
G*H = \{a_1\cdots a_n\ :\ n\in \Z_{\>0}, a_i\in G\text{ or }H, a_i\in G(H)\implies a_{i+1}\in H(G)\},
with group operation concatenation, and identity element the empty string \emptyset. We also assume e_Ge_H=e_He_G=e_G=e_H=\emptyset, for e_G (e_H) the identity element of G (H).

The above construction may be generalized to a collection of groups G_1*\cdots*G_m, where the index may be uncountable. If every G_\alpha=\Z (equivalently, has one generator), then *_{\alpha\in A} G_\alpha is called the free group on |A| generators.

Theorems


Theorem: (Borsuk-Ulam) Every continuous map S^n\to \R^n takes a pair of antipodal points to the same value.

Theorem: (Ham Sandwich theorem) Let U_1,\dots,U_n be bounded open sets in \R^n. There exists a hyperplane in \R^n that divides each of the open sets U_i into two sets of equal volume.

Volume is taken to be Lebsegue measure. The Ham sandwich theorem is an application of Borsuk-Ulam (see Terry Tao's blog post for more).

Theorem: If X and Y are path-connected, then \pi_1(X\times Y)\cong \pi_1(X)\times \pi_1(Y).

Now suppose that X = \bigcup_\alpha A_\alpha is based at x_0 with x_0\in A_\alpha for all \alpha. There are natural inclusions i_\alpha:A_\alpha\to X as well as j_\alpha:A_\alpha\cap A_\beta \to A_\alpha and j_\beta:A_\alpha\cap A_\beta \to A_\beta.
Both i_\alpha and j_\alpha induce maps on the fundamental group, each (and all) of the i_{\alpha*}:\pi_1(A_\alpha)\to \pi_1(X) extending to a map \Phi:*_\alpha \pi_1(A_\alpha)\to \pi_1(X).

Theorem: (van Kampen)
  • If A_\alpha\cap A_\beta is path-connected, then \Phi is a surjection. 
  • If A_\alpha\cap A_\beta\cap A_\gamma is path connected, then \ker(\Phi) = \langle j_{\alpha*}(g)(j_{\beta*}(g))^{-1}\ |\ g\in \pi_1(A_\alpha\cap A_\beta,x_0)\rangle.
As a consequence, if triple intersections are path connected, then \pi_1(X) \cong *_\alpha A_\alpha /\ker(\Phi). Moreover, if all double intersections are contractible, then \ker(\Phi)=0 and \pi_1(X)\cong *_\alpha A_\alpha.

Proposition: If \pi_1(X)=0 and \widetilde H_n(X)=0 for all n, then X is contractible.

References: Hatcher (Algebraic topology, Chapter 1), Tao (blog post "The Kakeya conjecture and the Ham Sandwich theorem")

Tuesday, November 1, 2016

Explicit pushforwards and pullbacks

 Preliminary exam prep

Here we consider a map f:M\to N between manifolds of dimension m and n, respectively, and the maps that it induces. Let p\in M with x_1,\dots,x_m a local chart for U\owns p and y_1,\dots,y_n a local chart for V\owns f(p). Induced from f are the differential (or pushforward) df and the pullback df^*, which are duals of each other:
\begin{array}{r c l} df_p\ :\ T_p M & \to & T_{f(p)}N \\[10pt] df\ :\ TM & \to & TN \\ \alpha & \mapsto & (\beta\mapsto \alpha(\beta\circ f)) \\[10pt]\\\\ \end{array} \hspace{1cm} \begin{array}{r c l} df^*_p\ :\ T_{f(p)}^* N & \to & T_p^*M\\[10pt] df^*\ :\ T^*N & \to & T^*M \\ \omega & \mapsto & \omega\circ f\\[10pt] \bigwedge ^k T^*N & \to & \bigwedge^k T^*M\\ \omega\ dy_1\wedge\cdots \wedge dy_k & \mapsto & (\omega \circ f)\ d(y_1\circ f)\wedge \cdots \wedge d(y_k\circ f) \end{array}
These maps may be described by the diagram below.
Example: For example, consider the map f:\R^3\to \R^3 given by f(x,y,z) = (x-y,3z^2,xz+yz), with the image having coordinates (u,v,w). With elements
2x\frac\dy{\dy x} - 5z\frac\dy{\dy y}\in TM, \hspace{2cm} 2uv+\sqrt w-5\in C^\infty(N), \hspace{2cm} \cos(uv)\in T^*N,
we have
\begin{align*} df_p\left(2x\frac\dy{\dy x} - 5z\frac\dy{\dy y}\right)(2uv+\sqrt w-5) & = \left(2x\frac\dy{\dy x} - 5z\frac\dy{\dy y}\right)\left(6(x-y)z^2+\sqrt{xz+yz}-5\right)(p),\\ df_p^*\left( \cos(uv)\right) & = \cos((x-y)3z^2), \\ \left(\textstyle\bigwedge^2 df_p^*\right)(\cos(uv)du\wedge dw) & = \cos((x-y)3z^2)d(3z^2)\wedge d(xz+yz) \\ & = \cos((x-y)3z^2)\left(-6z^2\ dx\wedge dz -6z^2\ dy \wedge dz\right). \end{align*}