User Tools

Site Tools


math105-s22:s:jianzhi:start

Jianzhi Wang

Hi! I am Jianzhi from Singapore, currently an undergraduate studying CS and Mathematics at UC Berkeley.

Notes from Own Readings

1. Axioms of Measurable Sets

  1. Every open set in Rn\mathbb{R}^n is measurable. Every closed set in Rn\mathbb{R}^n is measurable.
  2. If Ω\Omega is measurable, then Rn\Ω\mathbb{R}^n \RM \Omega is also measurable.
  3. [Boolean algebra property] If (Ωj)(\Omega_j) is any finite collection of measurable sets, then the union and intersection are measurable.
  4. [σ\sigma-algebra property] If (Ωj)(\Omega_j) is any countable collection of measurable sets, then the union and intersection are measurable.
  5. [Empty set] The empty set ϕ\phi has measure 00
  6. [Positivity] For every measurable set Ω\Omega, 0m(Ω)+0 \leq m (\Omega) \leq +\infty
  7. [Monotonicity] If A,BA, B measurable and ABA \subset B, then m(A)m(B)m(A) \leq m(B)
  8. [Finite sub-additivity]
  9. [Finite additivity]
  10. [Countable sub-additivity]
  11. [Countable additivity]
  12. [Normalization] The unit cube [0,1]n[0, 1]^n has measure m([0,1]n)=1m ([0, 1]^n) = 1
  13. [Translational invariance] If Ω\Omega measurable set, m(x+Ω)=m(Ω)m (x + \Omega) = m (\Omega).

2. Properties of Outer Measure

  1. [Empty set] The empty set ϕ\phi has measure 00
  2. [Positivity] For every measurable set Ω\Omega, 0m(Ω)+0 \leq m^* (\Omega) \leq +\infty
  3. [Monotonicity]
  4. [Finite sub-additivity]
  5. [Countable sub-additivity]
  6. [Translational invariance]

Definitions (Lebesgue)

0. [Open box] An open box in Rn\mathbb{R}^n is any set of the form B=Π(ai,bi)B = \Pi (a_i, b_i). Define volume of the box to be vol(B)=Π(biai)vol(B) = \Pi (b_i - a_i)

1, [Outer measure] Let ΩRn\Omega \subset \mathbb{R}^n, then m(Ω)=inf{iBi,ΩiBi,BiRn open boxes} m^*(\Omega) = \inf \{\sum_i |B_i|, \Omega \subset \bigcup_i B_i, B_i \subset \mathbb{R}^n \text{ open boxes} \} The nice thing about outer measure is that it is defined for any set, not just measurable ones.

2. [σ\sigma-algebra] A σ\sigma-algebra is a collection of sets that includes the empty set, is closed under complement and closed under countable union.

3. [Zero set] Say ZRnZ \subset \mathbb{R}^n is a zero set if it has outer measure 00.

4. [Measurability] Let EE be a subset of Rn\mathbb{R}^n. Say EE is asurable if and only if ARn\forall A \subset \mathbb{R}^n, we have the identity m(A)=m(AE)+m(A\E)m^*(A) = m^*(A \cap E) + m^*(A \RM E). Denote its asure as mE=mEmE = m^*E.

5. [Abstract outer measure] An abstract outer measure on MM is a function ω:2M[0,]\omega : 2^M \rightarrow [0, \infty] that satisfies ω(ϕ)=0\omega(\phi) = 0, ω\omega is monotone, and ω\omega is countably additive. Say a set EME \subset M is measurable wrt ω\omega if it satisfies (4) but with MM instead of Rn\mathbb{R}^n.

6. [Measurable function] Let ΩRn\Omega \subset \mathbb{R}^n be a measurable set. A function f:ΩRmf:\Omega → \mathbb{R}^m is measurable if and only if f1(V)f^{-1}(V) is measurable for every open set VRmV \subset \mathbb{R}^m.

7. The collection MM of measurable sets with respect to any outer measure on any set MM is a σ\sigma-algebra and the outer measure restricted to this σ\sigma-algebra is countably additive. All zero sets are measurable and have no effect on measurability. In particular, asure has these properties.

Proof (written in my own words)

TODO

7.5 [Measure space] A measure space is a triple (M,M,μ)(M, \mathcal{M}, \mu) where MM is a set, M\mathcal{M} is a σ\sigma-algebra of subsets of MM and mumu is a measure on M\mathcal{M} (i.e. μ:M[0,]\mu : \mathcal{M} → [0, \infty]).

7.6 [GδG_\delta set] A GδG_\delta set is a countable intersection of open sets.

7.7 [FσF_\sigma set] A FσF_\sigma set is a countable union of closed sets.

Example: (just so that I can remember) In probability, MM is the set of outcomes. M\mathcal{M} is the set of events (defined as the set of subsets of outcomes). P\mathbb{P} (the probability measure) is a measure.

7.8 [Inner measure] Inner measure can be thought of as a dual to outer measure. m(A)=sup{m(F):F closed,FA}m_*(A) = \text{sup}\{m(F) : F \text{ closed}, F \subset A\}

7.9 [Hull] Denote the hull of AA as HAH_A: the GδG_\delta set that achieves the infimum of the measure of open sets that contain AA. The GδG_\delta set is unique up to a measure zero set. (note that HAH_A achieves the infimum)

7.10 [Kernel] Denote the kernel of AA as KAK_A: the FσF_\sigma set that achieves the supremum of the measure of closed sets that is contained in AA. The FσF_\sigma set is unique up to a measure zero set. (note that KAK_A achieves the supremum)

7.11 [Measure Theoretic Boundary] The measure theoretic boundary of AA is δm(A)=HA\KA\delta_m(A) = H_A \RM K_A. A bounded subset of Rn\mathbb{R}^n is measurable if and only if m(A)=m(A)m^*(A) = m_*(A) (i.e. m(δm(A))=0m (\delta_m(A)) = 0).

7.12 [Slice] Define the slice of ERn×RkE \subset \mathbb{R}^n \times \mathbb{R}^k at xRnx \in \mathbb{R}^n to be Ex={yRk:(x,y)E}E_x = \{y \in \mathbb{R}^k: (x, y) \in E\}

7.13 [Undergraph] The undergraph of f:R[0,)f:\mathbb{R} → [0,\infty) is defined to be Uf={(x,y)R×[0,):0y<f(x)}\mathcal{U}f = \{(x,y) \in \mathbb{R} \times [0, \infty): 0 \leq y < f(x)\}. Intuitively, the positive region bounded between f(x)f(x) and the xx-axis. Then ff is asurable if Uf\mathcal{U}f is measurable w.r.t. the planar asure, and denote the gral to be f=m(Uf)\int f = m (\mathcal{U}f) (permit \infty).

7.135 [Completed Undergraph] The completed undergraph is just the undergraph plus the graph.

7.14 [grable] A function ff is Lebesgue integrable if its integral is finite. Note by definition ff must also be measurable.

7.15 [Lower and Upper Envelope Sequence] Let f:X[0,)f:X → [0, \infty) be a sequence of functions. Then define the lower and upper envelope sequences to be fn(x)=inf{fk(x):kn}\underline{f}_n(x) = \text{inf}\{f_k(x) : k \geq n\} and fˉn(x)=sup{fk(x):kn}\bar{f}_n(x) = \text{sup}\{f_k(x) : k \geq n\}. Permit fˉn(x)=\bar{f}_n(x) = \infty.

8. [Simple function] Let Ω\Omega be a measurable subset of Rn\mathbb{R}^n and f:ΩRf:\Omega → \mathbb{R}. Say ff is a simple function if the image f(Ω)f(\Omega) is finite (i.e. \exists a finite set {c1,,cN}\{c_1, …, c_N\} s.t. xΩ\forall x \in \Omega, f(x)=cif(x) = c_i for some 1iN1 \leq i \leq N).

9. [Lebesgue integral (of simple function)] Let ΩRn\Omega \subset \mathbb{R}^n be a measurable set and f:ΩRf:\Omega → \mathbb{R} be a simple nonnegative function. Define the Lebesgue integral of ff to be Ωf=Σλf(Ω);λ>0λm({xΩf(x)=λ})\int_{\Omega} f = \Sigma_{\lambda \in f(\Omega);\lambda > 0} \lambda m (\{ x \in \Omega | f(x) = \lambda \}).

10. [Majorization] Let f:ΩRf:\Omega → \mathbb{R} and g:ΩRg:\Omega → \mathbb{R}. Say ff majorizes gg if and only if f(x)g(x)xΩf(x) \geq g(x) \forall x \in \Omega.

11. [Lebesgue integral (of nonnegative function)] Let ΩRn\Omega \subset \mathbb{R}^n be a measurable set and f:Ω[0,]f:\Omega → [0,\infty] be measurable and nonnegative. Define the Lebesgue integral of ff to be Ωf=sup{Ωss simple, nonnegative and s.t. f majorizes s}\int_{\Omega} f = \text{sup}\{\int_{\Omega} s | s \text{ simple, nonnegative and s.t. } f \text{ majorizes } s\}.

Let T:MMT:M → M' be a mapping between measure spaces (M,M,μ)(M, \mathcal{M}, \mu) and (M,M,μ)(M', \mathcal{M}', \mu').

12. [Mesemorphism] Say TT is a mesemorphism if EM\forall E \in \mathcal{M} to TEMTE \in \mathcal{M'}. Intuitively, TT sends a subset of MM to a subset of MM'.

13. [Meseomorphism] Say TT is a meseomorphism if both TT and T1T^{-1} are mesemorphisms. Intuitively, TT is an invertible mapping between the measure spaces.

14. [Mesisometry] Say TT is a mesisometry if TT is a meseomorphism and μ(TE)=μ(E)EM\mu'(TE) = \mu(E) \forall E \in \mathcal{M}. Intuitively, besides preserving measure structures, TT also preserves the measure. TT can be thought of isomorphisms between measure spaces, similar to orthogonal transformation of vector spaces in Linear Algebra. Note: mesisometry may disrespect topology (Pugh Ex 6.19)

15. [Diffeomorphism] A mapping that preserves smooth structure.

16. [Isomorphism] A mapping that preserves algebraic structure.

17. [Homeomorphism] A mapping that preserves topological structure.

18. [Absolutely Integrable Functions] Let Ω\Omega be a measurable subset of Rn\mathbb{R}^n. A measurable function f:ΩRf:\Omega → \mathbb{R} is absolutely integrable if the integral Ωf<\int_{\Omega} |f| < \infty. (i.e. is finite)

19. [Lebesgue integral]

Definitions (Multivariable)

  1. [Norm of Operator] For AL(Rn,Rm)A \in L(\mathbb{R}^n, \mathbb{R}^m), define its norm as A:=supx:x1Ax\|A\| := \text{sup}_{x:\|x\| \leq 1} \|Ax\|. Hence AxAx\|Ax\| \leq \|A\|\|x\|
  2. [kk-surface in EE] A kk-surface in EE is a continuously differentiable mapping Φ:DE\Phi: D \rightarrow E where DRkD \subset \mathbb{R}^k is compact. Say DD is the parameter domain of Φ\Phi and denote its points as (u1,,uk)(u_1, …, u_k). (intuitively, a function with kk arguments, defines a surface that is of dimension kk)
  3. [\wedge] The \wedge means exterior product; can think of it as a cross product. Basically, there is a direction involved.
  4. [kk-form in EE] Let EE be an open set in Rn\mathbb{R}^n. A differential form of order k1k \geq 1 in EE is a function ω:\omega: kk-surface R\rightarrow \mathbb{R}. Symbolically, ω=Σai1ik(x)dxi1dxik\omega = \Sigma a_{i_1 … i_k}(x) dx_{i_1} \wedge … \wedge dx_{i_k} and ω(Φ)=Φω=DΣai1,,ik(Φ(u))(xi1,,xik)(u1,,uk)du\omega (\Phi) = \int_{\Phi} \omega = \int_{D} \Sigma a_{i_1, …, i_k}(\Phi(u)) \frac{\partial (x_{i_1}, …, x_{i_k})}{\partial (u_1, …, u_k)} du. (intuitively, it's the sum of functions coupled with the exterior product of kk differentials)
    1. uDRku \in D \subset \mathbb{R}^k, Φ(u)ERn\Phi(u) \in E \subset \mathbb{R}^n
    2. ai1,,ik:ERnRa_{i_1, …, i_k} : E \subset \mathbb{R}^n \rightarrow \mathbb{R} continuous; don't be scared about the index, they are just to say there are (nk)n \choose k functions.
    3. (xi1,,xik)(u1,,uk)\frac{\partial (x_{i_1}, …, x_{i_k})}{\partial (u_1, …, u_k)} is just the Jacobian
    4. dudu means integrate over the uu-cell
    5. Intuitively, saying that for each infinitesimal volume in DD, we take its value after being mapped to EE, then weighted by the volume in EE (that's what the Jacobian calculates).
  5. [Properties of kk-form]
    1. Φcω=cΦω\int_{\Phi} c \omega = c \int_{\Phi} \omega
    2. Φω1+ω2=Φω1+Φω2\int_{\Phi} \omega_1 + \omega_2 = \int_{\Phi} \omega_1 + \int_{_\Phi} \omega_2
    3. Φω=Φω\int_{\Phi} -\omega = - \int_{\Phi} \omega
    4. dxidxj=dxjdxidx_i \wedge dx_j = - dx_j \wedge dx_i
    5. dxidxi=0dx_i \wedge dx_i = 0
    6. dxj1dxjk=ϵ(j1,,jk)dxJdx_{j_1} \wedge … \wedge dx_{j_k} = \epsilon (j_1, …, j_k) dx_J (i.e. every kk-form can be represented in terms of basic kk-forms.
  6. [Basic kk-form] Say dxI=dxi1dxikdx_I = dx_{i_1} \wedge … \wedge dx_{i_k} is a basic kk-form in Rn\mathbb{R}^n if I={i1,,ik}I = \{i_1, …, i_k\} is an increasing tuple. (intuitively, this is just to standardize the permutations since due to the anti-commutative relation, most are the same)
  7. [Standard Presentation of kk-form] ω=ΣIbI(x)dxI\omega = \Sigma_{I} b_I(x) dx_I where the summation is over all increasing kk-index II.
  8. [Product of basic kk-forms] Suppose I={i1,,ip}I = \{i_1, …, i_p\} and J={j1,,jq}J = \{j_1, …, j_q\}, then the product of dxIdx_I and dxJdx_J in Rn\mathbb{R}^n is dxIdxJ=dxi1dxipdxj1dxjqdx_I \wedge dx_J = dx_{i_1} \wedge … \wedge dx_{i_p} \wedge dx_{j_1} \wedge … \wedge dx_{j_q}, a (p+q)(p+q)-form in Rn\mathbb{R}^n.
  9. [Multiplication of kk-forms] Let ω=ΣIbI(x)dxI\omega = \Sigma_{I} b_I(x)dx_I and λ=ΣJcJ(x)dxJ\lambda = \Sigma_{J} c_J(x)dx_J. Then ωλ=ΣI,JbI(x)cJ(x)dxIdxJ\omega \wedge \lambda = \Sigma_{I,J} b_I(x)c_J(x) dx_I \wedge dx_J, a (p+q)(p + q)-form in EE.
  10. [Differentiation] Let ω=ΣbI(x)dxI\omega = \Sigma b_I(x) dx_I be the standard presentation of a kk-form ω\omega, and bIC(E)b_I \in \mathcal{C}'(E) for each increasing kk-index II, then define dω=ΣI(dbI)dxId\omega = \Sigma_I (db_I) \wedge dx_I. (intuition: give a kk-form, get back a k+1k + 1 form; as expected in differential analysis)
    1. Example: 00-form ff is just a real function. df=Σi=1nfxi(x)dxidf = \Sigma_{i=1}^n \frac{\partial f}{\partial x_i}(x) dx_i
  11. [Change of Variables] Let ERnE \subset \mathbb{R}^n be open, VRmV \subset \mathbb{R}^m be open and T:EVT:E \rightarrow V. Let T(x)=(t1(x),t2(x),,tm(x))=(y1,y2,,ym)=yT(x) = (t_1(x), t_2(x), …, t_m(x)) = (y_1, y_2, …, y_m) = y. Let ω=ΣIbI(y)dyI\omega = \Sigma_I b_I(y) dy_I be a kk-form in VV, then the corresponding kk-form in EE is ωT=ΣIbI(T(x))dti1dtik\omega_T = \Sigma_I b_I(T(x)) dt_{i_1} \wedge … \wedge dt_{i_k}.
  12. [Oriented affine kk-simplex] σ=[p0,p1,,pk]\sigma = [p_0, p_1, …, p_k] is the kk-surface in Rn\mathbb{R}^n with parameter domain QkQ^k given by σ(α1e1++αkek)=p0+Σi=1kαi(pip0)\sigma(\alpha_1 e_1 + … + \alpha_k e_k) = p_0 + \Sigma_{i=1}^k \alpha_i (p_i - p_0). (intuitively, just the normal standard simplex, but defined somewhere else in space, can be distorted a bit)
    1. σ(0)=p0\sigma(0) = p_0
    2. σ(ei)=pi\sigma(e_i) = p_i
    3. σ(u)=p0+Au\sigma(u) = p_0 + Au where Aei=pip0Ae_i = p_i - p_0
  13. [Integration over a simplex] A zero simplex is just [p0][p_0]. Define σf=ϵf(p0)\int_{\sigma} f = \epsilon f(p_0), where ϵ=±1\epsilon = \pm 1 depending on the orientation of the simplex.
  14. [Affine kk-chain] An affine kk-chain Γ\Gamma in an open set ERnE \subset \mathbb{R}^n is a collection of finitely many oriented affine kk-simplexes {σ1,,σr}\{\sigma_1, …, \sigma_r \} in EE (need not be distinct; i.e. simplexes can occur with multiplicity). Write as Γ=σ1++σr\Gamma = \sigma_1 + … + \sigma_r. (Γ\Gamma is still a kk-surface)
  15. [Integration of kk-chain] Define Γω=Σi=1rσiω\int_{\Gamma}\omega = \Sigma_{i=1}^{r} \int_{\sigma_i} \omega (intuition: just sum over all the individual small simplexes)
  16. [Boundary of affine kk-simplex] For k1k \geq 1, the boundary of the oriented affine kk-simplex σ=[p0,p1,,pk]\sigma = [p_0, p_1, …, p_k] is defined to be the affine (k1)(k-1)-chain σ=Σj=0k(1)j[p0,p1,,pj1,pj+1,,pk]\partial \sigma = \Sigma_{j=0}^k (-1)^j [p_0, p_1, …, p_{j-1}, p_{j+1}, …, p_k] (intuition: a solid becomes a surface; lowers dimension by 11)
  17. [Differentiable simplex] Let TC:ERnVRmT \in \mathcal{C“}: E \subset \mathbb{R}^n \rightarrow V \subset \mathbb{R}^m, not necessarily 1-1. Let σ\sigma be an oriented affine kk-simplex in EE. Say Φ=Tσ\Phi = T \circ \sigma is an oriented kk-simplex of class C\mathcal{C”}. Note that Φ=Tσ\Phi = T \circ \sigma is a kk-surface in VV.
  18. [Differentiable chain] Say Ψ\Psi is a kk-chain of class C\mathcal{C“} in VV if it is a finite collection of oriented kk-simplex Φ1\Phi_1, …, Φr\Phi_r of class C\mathcal{C”}. Write Ψ=ΣΦi\Psi = \Sigma \Phi_i. For kk-form ω\omega, define Ψω=Σi=1rΦiω\int_{\Psi}\omega = \Sigma_{i=1}^{r} \int_{\Phi_i} \omega.
  19. [Boundary of oriented kk-simplex] Φ=T(σ)\partial \Phi = T(\partial \sigma)
  20. [Boundary of kk-chain] Ψ=ΣΦi\partial \Psi = \Sigma \partial \Phi_i
  21. [Positively oriented boundaries]
    1. For QnQ^n (the standard simplex), let σ0\sigma_0 be the identity mapping with domain QnQ^n. Then σ0\partial \sigma_0 (an affine (n1)(n-1)-chain) is the positively oriented boundary of set QnQ^n.
    2. For E=T(Qn)E = T(Q^n), the positively oriented boundary of set EE, denoted by E\partial E, is the (n1)(n-1)-chain T=T(σ0\partial T = T(\partial \sigma_0
    3. For Ω=E1Er\Omega = E_1 \cup … \cup E_r where Ei=Ti(Qn)E_i = T_i(Q^n) pairwise disjoint, then Ω=T1++Tr\partial \Omega = \partial T_1 + … + \partial T_r is the positively oriented boundary of Ω\Omega.

Definitions (Fourier)

<in progress>

Theorems and Lemmas (Lebesgue)

1. [Tao 7.1.1]

2. [Tao 7.4.4] Properties of measurable sets

  1. If EE measurable, then Rn\E\mathbb{R}^n \RM E also measurable.
  2. If EE measurable and xRnx \in \mathbb{R}^n, then x+Ex+E measurable with m(x+E)=m(E)m(x+E)=m(E).
  3. If E1,E2E_1, E_2 measurable, then union and intersection also measurable.
  4. Similarly fo of measurable sets.
  5. Every open box and every closed box is measurable.
  6. Any set EE with m(E)=0m^*(E) = 0 is measurable.

3. [Pugh 2.5] The collection MM of measurable sets wrt any other measure on any set MM is a σ\sigma-algebra and the outer measure restricted to this σ\sigma-algebra is countably additive.

4. [Tao 7.5.2] Let ΩRn\Omega \subset \mathbb{R}^n and f:ΩRmf:\Omega → \mathbb{R}^m be continuous. Then ff is measurable.

5. [Tao 7.5.3] ff is measurable if and only if f1(B)f^{-1}(B) is measurable for every open box BB.

Proof (I came up with this by myself; please forgive any errors) By Lemma 7.4.10, every open set can be written as a countable or finite union of open boxes.

If B\forall B open box, f1(B)f^{-1}(B) is measurable, then VRm\forall V \subset \mathbb{R}^m open, VV can be expressed as the union of countab of boxes, say (Bn)n(B_n)_n. Then, since f1(B)f^{-1}(B) are all measurable by assumption, the sequence of sets (f1(Bn))n(f^{-1}(B_n))_n are all measurable, hence the union of all of these, which is exactly f1(V)f^{-1}(V), is measurable.

On the other hand, if ff is measurable, then VRn\forall V \in \mathbb{R}^n, f1(V)f^{-1}(V) is measurable. Then simply just take VV as any open box. (LOL)

6. [Pugh 6.11] Each measurable set EE can be sandwiched between an FσF_\sigma-set (say FF) and a GδG_\delta-set (say GG) (i.e. FEGF \subset E \subset G), such that m(G\F)=0m^*(G \RM F) = 0 (i.e. G\FG \RM F is a zero set). Conversely, if GGδ,FFσ\exists G \in G_\delta, F \in F_\sigma s.t. FEGF \subset E \subset G, then EE is measurable.

Corollary: A bounded subset ERnE \subset \mathbb{R}^n is measurable if and only if it has a regularity sandwich FEGF \subset E \subset G s.t. FFσF \in F_\sigma, GGδG \in G_\delta and m(E)=m(G)m(E) = m(G).

Corollary: All measurable sets are FσF_\sigma-sets and GδG_\delta sets modulo measure zero sets.

Proof of Unbounded Case TODO

7. [Pugh 6.15] An affine motion T:RnRnT:\mathbb{R}^n → \mathbb{R}^n is a meseomorphism and multiplies measure by det(T)|det(T)|.

8. If ABA \subset B where BB is a box, then m(B)=m(A)+m(B\A)m(B) = m_*(A) + m^*(B \RM A).

9. If ABRnA \subset B \subset \mathbb{R}^n and BB is a box, then AA is measurable if and only if it divides BB cleanly (i.e. m(A)=m(AB)+m(A\B)m(A) = m(A \cap B) + m(A \RM B))

10. [Measurable Product Theorem] If ARnA \subset \mathbb{R}^n and BRkB \subset \mathbb{R}^k are measurable, then A×BA \times B is measurable and m(A×B)=m(A)m(B)m(A \times B) = m(A)\cdot m(B). We treat 0=00\cdot \infty = 0.

11. [Zero Slice Theorem] If ERn×RkE \subset \mathbb{R}^n \times \mathbb{R}^k is measurable, then EE is a zero set if and only if almost every slice of EE is a (slice) zero set.

12. [Measure Continuity Theorem] Let ω\omega be any outer measure. If {Ek}\{E_k\} and {Fk}\{F_k\} are sequences of measurable sets, then upward measure continuity yields EkEω(Ek)ω(E)E_k \uparrow E \Rightarrow \omega(E_k) \uparrow \omega(E); downward measure continuity yields FkFF_k \downarrow F and ω(F1)<\omega(F_1) < \infty ω(Fk)ω(F)\Rightarrow \omega(F_k) \downarrow \omega(F). (Notation: EkEE_k \uparrow E means E1E2E_1 \subset E_2 \subset … and E=EkE = \bigcup E_k. FkFF_k \downarrow F means F1F2F_1 \supset F_2 \supset … and F=FkF = \bigcap F_k)

Proof TODO Key Idea Disjointize

Trinity of Theorems (MCT, DCT, FL)

12. [Monotone Convergence Theorem] Assume that (fn)n(f_n)_n is a sequence of measurable functions with fn:R[0,)f_n:\mathbb{R} → [0, \infty) and fnff_n \uparrow f as nn → \infty, then fnf\int f_n \uparrow \int f

PROOF TODO!!

12.1 [Corollary] If (fn)n(f_n)_n is a sequence of integrable functions that converges monotonically downward to a limit function ff almost everywhere, then fnf\int f_n \downarrow \int f.

5. [Monotone Convergence Theorem] Let ΩRn\Omega \subset \mathbb{R}^n be a measurable set and (fn)n(f_n)_n be a sequence of nonnegative measurable functions, where fi:ΩRf_i:\Omega → \mathbb{R} are increasing s.t. fi+1f_{i+1} majorizes fif_i. Then:

  • 0Ωf1Ωf20 \leq \int_{\Omega} f_1 \leq \int_{\Omega} f_2 \leq …
  • Ωsupnfn=supnΩfn\int_{\Omega} \text{sup}_n f_n = \text{sup}_n \int_{\Omega} f_n

13. [Dominated Convergence Theorem] If n\forall n fn:R[0,)f_n: \mathbb{R} → [0, \infty) is a measurable functions such that (fn)nf(f_n)_n → f (what is a.e.) and if g:R[0,)\exists g:\mathbb{R} → [0, \infty) whose integral is finite and is an upper bound fn\forall f_n, then ff is integrable and fnf\int f_n → \int f as nn → \infty.

PROOF TODO!!

14. [Fatou's Lemma] Let Ω\Omega be a measurable set. If fn:R[0,)f_n:\mathbb{R} → [0, \infty) is a sequence of nonnegative measurable functions then lim inffnlim inffn\int \text{lim inf} f_n \leq \text{lim inf} \int f_n.

PROOF TODO!!

15. [Borel-Cantelli Lemma] Let Ω1,Ω2,,\Omega_1, \Omega_2, …, be measurable subsets of Rn\mathbb{R}^n such that Σn=1m(Ωn)\Sigma_{n=1}^{\infty} m (\Omega_n) is finite. Then the set {xRn:xΩn for infinitely many n}\{x \in \mathbb{R}^n: x \in \Omega_n \text{ for infinitely many } n\} is a set of measure 00. (i.e. almost every point belongs to only finitely many Ωn\Omega_n. (Proof: see below)

16. [Fubini] Let f:RnRf:\mathbb{R}^n → \mathbb{R} be an absolutely integrable function. Then exists absolutely integrable functions F:RRF:\mathbb{R} → \mathbb{R} and G:RRG:\mathbb{R} → \mathbb{R} s.t. for almost every xx, f(x,y)f(x,y) is absolutely integrable in yy with F(x)=Rf(x,y)dyF(x) = \int_{\mathbb{R}} f(x,y) dy and for almost every yy, f(x,y)f(x,y) is absolutely integrable in xx with G(y)=Rf(x,y)dxG(y) = \int_{\mathbb{R}} f(x,y) dx. We also have: RF(x)dx=R2f=RG(y)dy\int_{\mathbb{R}} F(x) dx = \int_{\mathbb{R}^2} f = \int_{\mathbb{R}} G(y) dy.

Theorems and Lemmas (Multivariable)

  1. [Contraction Principle]
  2. [Inverse Function Theorem]
  3. [Implicit Function Theorem]
  4. [Rank Theorem]
  5. [Differentiation of Integrals]
  6. [Primitive Mapping]
  7. [Partitions of Unity]
  8. [Change of Variables]
  9. [Differential Forms]
  10. [Uniqueness] Suppose ω=ΣIbI(x)dxI\omega = \Sigma_I b_I(x) dx_I is the standard presentation of a kk-form ω\omega in an open set ERnE \subset \mathbb{R}^n. If ω=0\omega = 0 in EE, then I\forall I increasing kk-index and xE\forall x \in E, bI(x)=0b_I(x) = 0.

Proof: Assume on the contrary that vERn\exists v \in E \subset \mathbb{R}^n s.t. bJ(v)>0b_J(v) > 0 for some increasing kk-index J={j1,,jk}J = \{j_1, …, j_k\}.

Since bJb_J continuous (by definition of kk-form), h>0\exists h > 0 s.t. for all xRnx \in \mathbb{R}^n s.t. xivi<h|x_i - v_i| < h, bJ(x)>0b_J(x) > 0. (i.e. inside the cube of dimension 2h2h centered at vv, bJ>0b_J > 0) Let DRkD' \subset \mathbb{R}^k s.t. uDurhu \in D' \Leftrightarrow |u_r| \leq h.

Define Φ:DRkRn\Phi:D' \subset \mathbb{R}^k \rightarrow \mathbb{R}^n s.t. Φ(u)=v+Σr=1kurejr\Phi(u) = v + \Sigma_{r=1}^k u_r e_{j_r}. (i.e. Φ\Phi is a kk-surface in EE with parameter domain DD' and bJ(Φ(u))>0uDb_J(\Phi(u)) > 0 \forall u \in D' since (Φ(u))ivi<h|(\Phi(u))_i - v_i| < h)

Note that Φω=Db{j1,,jk}(Φ(u))(xj1,,xjk)u1,,ukdu=DbJ(Φ(u))det(I)du=DbJ(Φ(u))du>0\int_{\Phi} \omega = \int_{D'} b_{\{j_1, …, j_k\}}(\Phi(u)) \frac{\partial(x_{j_1}, …, x_{j_k})}{u_1, …, u_k} du = \int_{D'} b_J(\Phi(u)) det(\mathbb{I}) du = \int_{D'} b_J(\Phi(u)) du > 0. Hence, contradiction! (since \exists increasing JJ s.t. bJ(x)0ω0b_J(x) \neq 0 \Rightarrow \omega \neq 0 in EE)

  1. Let ω\omega and λ\lambda be kk- and mm-forms of class C\mathcal{C}' in EE, then d(ωλ)=(dω)λ+(1)kωdλd(\omega \wedge \lambda) = (d\omega) \wedge \lambda + (-1)^k \omega \wedge d\lambda.
  2. If ω\omega is of class C\mathcal{C}'' in EE, then d2ω=0d^2\omega = 0.
  3. [Change of Variables] Let ERnE \subset \mathbb{R}^n be open, TT be a C\mathcal{C}'-mapping of EE into VRmV \subset \mathbb{R}^m open and ω,λ\omega, \lambda be kk- and mm-forms in VV. Then:
    1. (ω+λ)T=ωT+λT(\omega + \lambda)_T = \omega_T + \lambda_T
    2. (ωλ)T=ωTλT(\omega \wedge \lambda)_T = \omega_T \wedge \lambda_T
    3. d(ωT)=d(ω)Td(\omega_T) = d(\omega)_T
  4. Let ω\omega be a kk-form in open ERnE \subset \mathbb{R}^n, Φ\Phi be a kk-surface in EE (parameter domain DRkED \subset \mathbb{R}^k \rightarrow E) and Δ\Delta be kk-surface in Rk\mathbb{R}^k (parameter domain DRkRkD \subset \mathbb{R}^k \rightarrow \mathbb{R}^k) defined by Δ(u)=u\Delta(u) = u (i.e. just maps the vector to itself). Then Φω=ΔωΦ\int_{\Phi} \omega = \int_{\Delta} \omega_{\Phi}

Proof: Suffices to consider ω=a(x)dxi1dxik\omega = a(x) dx_{i_1} \wedge … \wedge dx_{i_k}, since if we can prove so, then every other term will be equal.

Note ωΦ=a(Φ(u))dϕi1dϕi2dϕik\omega_{\Phi} = a(\Phi(u)) d\phi_{i_1} \wedge d\phi_{i_2} \wedge … \wedge d\phi_{i_k}.

Define α(p,q)=ϕipxq(u)\alpha(p, q) = \frac{\partial \phi_{i_p}}{\partial x_q}(u). Then dϕip=Σqα(p,q)duqd\phi_{i_p} = \Sigma_{q} \alpha(p, q) du_q. The k×kk \times k matrix with entries α(i,j)\alpha(i, j) has determinant J(u)J(u)

Hence, dϕi1dϕik=Σα(1,q1)α(k,qk)duq1duqk=Σα(1,q1)α(k,qk)s(q1,,qk)du1duk=du1dukΣα(1,q1)α(k,qk)s(q1,,qk)=du1dukJ(u)=J(u)du1dukd\phi_{i_1} \wedge … \wedge d\phi_{i_k} = \Sigma \alpha(1, q_1)\cdot … \cdot \alpha(k, q_k) du_{q_1} \wedge … \wedge du_{q_k} = \Sigma \alpha(1, q_1)\cdot … \cdot \alpha(k, q_k) s(q_1, …, q_k) du_{1} \wedge … \wedge du_{k} = du_{1} \wedge … \wedge du_{k} \Sigma \alpha(1, q_1)\cdot … \cdot \alpha(k, q_k) s(q_1, …, q_k) = du_{1} \wedge … \wedge du_{k} J(u) = J(u) du_{1} \wedge … \wedge du_{k}

Hence, Φω=Da(Φ(u))J(u)du=Da(Φ(u))J(u)du1duk=Da(Φ(u))dϕi1dϕik=ΔωΦ\int_{\Phi} \omega = \int_{D} a(\Phi(u))J(u)du = \int_{D} a(\Phi(u))J(u) du_1 \wedge … du_k = \int_{D} a(\Phi(u)) d\phi_{i_1} \wedge … \wedge d\phi_{i_k} = \int_{\Delta} \omega_{\Phi}

  1. [Composition Theorem] Note the name composition is made up by me. Suppose TT is a continuous differentiable mapping of open ERnVRmE \subset \mathbb{R}^n \rightarrow V \subset \mathbb{R}^m open. Let Φ\Phi be a kk-surface in EE (i.e. Φ:DRkERn\Phi: D \subset \mathbb{R}^k \rightarrow E \subset \mathbb{R}^n) and thus TΦT\Phi is also a kk-form but in VV instead (TΦ:DRkVRmT\Phi: D \subset \mathbb{R}^k \rightarrow V \subset \mathbb{R}^m). Let ω\omega be a kk-form in VV. Then TΦω=ΦωT\int_{T\Phi} \omega = \int_{\Phi} \omega_T.

Proof: TΦω=ΔωTΦ=Δ(ωT)Φ=ΦωT\int_{T\Phi} \omega = \int_{\Delta} \omega_{T\Phi} = \int_{\Delta} (\omega_T)_{\Phi} = \int_{\Phi} \omega_T

  1. [Simplexes and Chains]
  2. If σ\sigma is an oriented kk-simplex in open ERnE \subset \mathbb{R}^n and if σˉ=ϵσ\bar{\sigma} = \epsilon \sigma, then σˉω=ϵσω\int_{\bar{\sigma}} \omega = \epsilon \int_{\sigma} \omega for every kk-form ω\omega in EE.
  3. [Stokes Theorem]
  4. [Closed Forms and Exact Forms]

Theorems and Lemmas (Fourier)

Discussion Problems

Discussion Link Remarks
1 disc01.pdf (TODO: latter part of 4 and 5)
4 disc04.pdf Proved Lemma 7.4.4 (b) to (e) using alternate definition of measurability
?

Homework

Brief Description of Lectures

Date Brief Description
01/18 Introduction; Outer Measure
01/20 Measurable set
01/25 Lemmas
01/27 Lemmas and Alternative “Measurability” Def
02/01 Abstract measure theory, Regularity (Fσ,Gδ)F_\sigma, G_\delta)
02/03 Slices
02/08 Monotone Convergence Theorem, Dominated Convergence Theorem (Pugh)
02/10 Fatou's Lemma; Mesometry (Pugh)
02/15 Measurable functions; Simple functions (Tao)
02/17 Monotone Convergence Theorem (Tao)
02/22
02/24
03/01
03/03
03/08
03/10
03/15
03/17 Implicit Function Theorem (Samuel)
03/22 🍃 Spring Break 🍃
03/24 🍃 Spring Break 🍃
03/29 Inverse Function Theorem (Chloe)
03/31 Differential Forms
04/05 Stokes' Theorem
04/07 Poincaré's Lemma
04/12
04/14
04/19
04/21

Zero Slice Theorem (Lecture 6) Proof

help X.X

Theorem: Let ERn×RkE \subset \mathbb{R}^n \times \mathbb{R}^k be measurable. Denote by Ex:=E({x}×Rk){x}×RkRkE_x := E \cap (\{x\} \times \mathbb{R}^k) \subset \{x\} \times \mathbb{R}^k \approx \mathbb{R}^k. Let Z={xRn:mRk(Ex)0}Z = \{x \in \mathbb{R}^n : m_{\mathbb{R}^k}(E_x) \neq 0\} be a measure 00 set in Rn\mathbb{R}^n. Then m(E)=0m(E) = 0.

Proof: (I took down Prof Peng's proof in class and tried to fill in some missing steps which I didn't catch and some of my own commentary)

Let E~=E\(Z×Rk)\tilde{E} = E \RM (Z \times \mathbb{R}^k). Then, since m(Z×Rk)=0m(Z \times \mathbb{R}^k) = 0, so m(E~)=m(E)m (\tilde{E}) = m(E), because these two differ by a measure 00 set. Suffices to show m(E~)=0m (\tilde{E}) = 0.

WLOG, we can assume EE in place of E~\tilde{E} with Z=ϕZ = \phi. Hence, now, mRk(Ex)=0m_{\mathbb{R}^k}(E_x) = 0 x\forall x (as Z=ϕZ = \phi).

Assume EE is bounded and n=1,k=1n = 1, k = 1 (i.e. ER×RE \subset \mathbb{R} \times \mathbb{R}). Further assume E[0,1]2E \subset [0, 1]^2 (i.e. EE is in the unit square; we make this because we can always scale our ϵ\epsilon later).

Know mR(Ex)=0m_{\mathbb{R}}(E_x) = 0 x\forall x. Want to show ϵ>0,mR×R(E)<ϵ\forall \epsilon > 0, m_{\mathbb{R}\times\mathbb{R}}(E) < \epsilon.

By inner-regularity of EE (since EE is measurable), we can always find closed KEK \subset E such that m(E\K)ϵ/2m(E \RM K) \leq \epsilon/2. Then, KK is bounded, so KK is compact. Also, mR×R(Kx)mR×R(Ex)=0m_{\mathbb{R} \times \mathbb{R}}(K_x) \leq m_{\mathbb{R} \times \mathbb{R}}(E_x) = 0 by monotonicity, hence mR×R(Kx)=0m_{\mathbb{R} \times \mathbb{R}}(K_x) = 0.

Claim: We can cover KK by boxes of total volume (actually, area in this case since n=k=1n = k = 1) ϵ/2\leq \epsilon/2

Proof of Claim: xR\forall x \in \mathbb{R}, if KxϕK_x \neq \phi, then we can find an open set V(x)RV(x) \in \mathbb{R} s.t. mR(V(x))ϵ/2m_{\mathbb{R}}(V(x)) \leq \epsilon/2 and V(x)V(x) contains KxK_x.

Lemma: \exists an enlargement U(x)RU(x) \subset \mathbb{R} with xU(x)x \in U(x) s.t. π1(U(x))KU(x)×V(x)\pi^{-1}(U(x)) \cap K \subset U(x) \times V(x) (i.e. x~U(x),VxKx~\forall \tilde{x} \in U(x), V_x \supset K_{\tilde{x}}). Note that here π:R2R\pi : \mathbb{R}^2 → \mathbb{R}, so what the claim is saying is that we can always find an interval containing xx such that the intersection of the vertical strip with KK is entirely contained in the Cartesian product U(x)×V(x)U(x) \times V(x). Another way of putting it is: suppose KK is a potato. Then a slice with width as the orange interval across Rn\mathbb{R}^n in the direction of Rk\mathbb{R^k} cuts the potato. Then the potato strip is entirely contained in U(x)×V(x)U(x) \times V(x).

Proof of Lemma: Suppose not true, i.e. does not exist U(x)RU(x) \subset \mathbb{R} containing xx with this property. That means for every interval (x1n,x+1n),nZ(x - \frac{1}{n}, x + \frac{1}{n}), n \in \mathbb{Z} containing xx, x~n(x1n,x+1n)\exists \tilde{x}_n \in (x - \frac{1}{n}, x + \frac{1}{n}) s.t. Kx~K_{\tilde{x}} is not strictly contained in V(x)V(x). Then \exists a sequence (x~n,y~n)nK(\tilde{x}_n, \tilde{y}_n)_n \in K s.t. limnx~n=x\text{lim}_{n → \infty} \tilde{x}_n = x and y~nV(x)\tilde{y}_n \notin V(x). Since this sequence exists in a compact set, \exists a subsequence that converges, i.e. (x~nk,y~nk)kK\exists (\tilde{x}_{n_k}, \tilde{y}_{n_k})_k \in K that converges to, say, (x,y)K(x, y) \in K. But y~nV(x)cyV(x)c\tilde{y}_n \in V(x)^c \Rightarrow y \in V(x)^c. (Contradiction! since by assumption KxV(x)K_x \in V(x), so yV(x)y \in V(x))

Thus, xR\forall x \in \mathbb{R}, V(x)Kx\exists V(x) \supset K_x s.t. V(x)V(x) is open and mR(V(x))<ϵ/2m_{\mathbb{R}}(V(x)) < \epsilon/2. Also, U(x)R\exists U(x) \subset \mathbb{R} containing xx s.t. U(x)×V(x)π1(U(x))KU(x) \times V(x) \supset \pi^{-1}(U(x)) \cap K.

Hence, KxR(V(x)×U(x))K \subset \bigcup_{x \in \mathbb{R}} (V(x) \times U(x)). Since KK is compact, we can obtain a finite subcover by Heine-Borel theorem K(V(x1)×U(x1))(V(x2)×U(x2))(V(xN)×U(xN))\Rightarrow K \subset (V(x_1) \times U(x_1)) \cup (V(x_2) \times U(x_2)) \cup … \cup (V(x_N) \times U(x_N)), NNN \in \mathbb{N}.

Now define Ui=U(xi)\(k=1i1U(xk))U'_i = U(x_i) \RM (\bigcup^{i-1}_{k=1} U(x_{k})). The motivation is to disjointize KK. Then K(V(x1)×U1)(V(x2)×U2)(V(xN)×UN)K \subset (V(x_1) \times U'_{1}) \sqcup (V(x_2) \times U'_{2}) \sqcup … \sqcup (V(x_N) \times U'_{N}). This is possible because for any point x(V(x1)×U(x1))(V(xN)×U(xN))x \in (V(x_1) \times U(x_1)) \cup … \cup (V(x_N) \times U(x_N)), suppose xx first appeared in (V(xi)×U(xi))(V(x_i) \times U(x_i)) for some ii, then xx will only appear in (V(xi)×Ui)(V(x_i) \times U'_i).

The union of UiU'_i are contained in the unit interval. Hence, Σim(V(xi)×Ui)=Σim(V(xi))×m(Ui)ϵ2Σim(Ui)ϵ2\Sigma_i m(V(x_i) \times U'_i) = \Sigma_i m(V(x_i)) \times m(U'_i) \leq \frac{\epsilon}{2} \Sigma_i m(U'_i) \leq \frac{\epsilon}{2}. (The last inequality is due to the disjointness of UiU'_i). Hence m(K)<ϵ/2m(K) < \epsilon/2.

Finally, since we obtained m(K)<ϵ2m(K) < \frac{\epsilon}{2}, we have m(E)m(K)+m(E\K)=ϵ2+ϵ2=ϵm(E) \leq m(K) + m(E \RM K) = \frac{\epsilon}{2} + \frac{\epsilon}{2} = \epsilon. Hence, m(E)=0m(E) = 0 as desired.

Key Ideas: Approximate EE from the inside by compact set KK. Generate an open covering of KK by slices along Rn\mathbb{R}^n in the direction of Rk\mathbb{R}^k such that these slices approximate KK really well. Apply Heine-Borel and disjointness.

Monotone Convergence Theorem (Lecture 7) Proof

In progress

Dominated Convergence Theorem (Lecture 7) Proof

Theorem: Let (fn)n(f_n)_n be a sequence of measurable functions with fn:R[0,)f_n : \mathbb{R} → [0, \infty) and (fn)nf(f_n)_n → f almost everywhere. Suppose g:R[0,)\exists g:\mathbb{R} → [0, \infty) s.t, fn(x)g(x)|f_n(x)| \leq g(x) almost everywhere and g<\int g < \infty. Then ff is integrable and fnf\int f_n → \int f as nn → \infty.

Proof: (in my own words)

Firstly, fnˉg\bar{f_n} \leq g almost everywhere, thus U(fn)U(g)m(U(fn))m(U(g))=g<U(f_n) \subset U(g) \Rightarrow m(U(f_n)) \leq m(U(g)) = \int g < \infty by assumption. Hence, fnf_n are all integrable, since m(U(fn))<m(U(f_n)) < \infty.

Consider fn\underline{f_n} and fnˉ\bar{f_n}. Since (fn)n(\underline{f_n})_n is a sequence of measurable functions that converges to ff almost everywhere and fn+1fn\underline{f_{n+1}} \geq \underline{f_{n}} (infimum of a smaller set), fnf\underline{f_n} \uparrow f. Similarly, (fnˉ)n(\bar{f_n})_n is a sequence of measurable functions that converges to ff almost everywhere and fn+1ˉfnˉ\bar{f_{n+1}} \leq \bar{f_{n}} (supremum of a smaller set), so fnˉf\bar{f_n} \downarrow f.

Applying Monotone Convergence Theorem, fnf\int \underline{f_n} \uparrow \int f and fnˉf\int \bar{f_n} \downarrow \int f.

We also have: U(fn)U(fn)U^(fnˉ)U(\underline{f_n}) \subset U(f_n) \subset \hat{U}(\bar{f_n}). This is because fn(x)=inf{fm(x):mn}fn(x)\underline{f_n}(x) = \text{inf}\{f_m(x): m \geq n\} \leq f_n(x), hence (x,y)U(fn)\forall (x,y) \in U(\underline{f_n}), (x,y)U(fn)(x, y) \in U(f_n). Similarly, fnˉ(x)=sup{fm(x):mn}fn(x)\bar{f_n}(x) = \text{sup}\{f_m(x): m \geq n\} \geq f_n(x), hence (x,y)U(fn)\forall (x,y) \in U(f_n), (x,y)U^(fnˉ)(x, y) \in \hat{U}(\bar{f_n}).

Since U(fn)U(fn)U^(fnˉ)U(\underline{f_n}) \subset U(f_n) \subset \hat{U}(\bar{f_n}) and fnf\int \underline{f_n} \uparrow \int f and fnˉf\int \bar{f_n} \downarrow \int f, fnf\int f_n → \int f.

A website which I found during my search for DCT: https://www.math3ma.com/blog/dominated-convergence-theorem

Density Theorem (03/01 Class Presentation)

Definition

  • Let ERnE \subset \mathbb{R}^n be measurable. For pRnp \in \mathbb{R}^n, define the density of EE at pp as: δ(p,E)=limQpm(EQ)m(Q)\delta(p, E) = \text{lim}_{Q \downarrow p} \frac{m(E \cap Q)}{m(Q)}
  • QpQ \downarrow p denotes QQ as a cube that shrinks down to pp. I visualize it as a sequence of cubes (Qn)n(Q_n)_n s.t. pQnnp \in Q_n \forall n and limnQn=0\text{lim}_{n \rightarrow \infty} |Q_n| = 0. Note that the cubes need not be centered at pp; see diagram below.
  • Define the lower density δ(p,E)=lim infQpm(EQ)m(Q)\underline{\delta}(p, E) = \text{lim inf}_{Q \downarrow p} \frac{m(E \cap Q)}{m(Q)}.
  • Say pp is a density point of EE if pEp \in E and δ(p,E)=1\delta(p, E) = 1. Pugh denoted the set of density points as dp(E)dp(E).

Properties

  • 0δ10 \leq \delta \leq 1
  • δ(p,E)\delta(p, E) exists \Rightarrow ϵ>0\forall \epsilon > 0, l>0\exists l > 0 s.t. if QQ is a cube containing pp with side length less than ll, then m(EQ)m(Q)δ(p,E)<ϵ|\frac{m(E \cap Q)}{m(Q)} - \delta(p, E)| < \epsilon.

Suppose otherwise, then ϵ>0\exists \epsilon > 0 s.t. 1i\forall \frac{1}{i} with iNi \in \mathbb{N}, exists a cube QiQ_i with side length less than 1i\frac{1}{i} s.t. m(EQi)m(Qi)δ(p,E)ϵ|\frac{m(E \cap Q_i)}{m(Q_i)} - \delta(p, E)| \geq \epsilon. Then, this sequence of cubes converge to pp since the side length converges to 00, but limi(m(EQi)m(Qi)δ(p,E))=0>ϵ\text{lim}_{i \rightarrow \infty} (\frac{m(E \cap Q_i)}{m(Q_i)} - \delta(p, E)) = 0 > \epsilon. (contradiction!)

Theorem [Lebesgue Density]: If EE is measurable, then almost every pEp \in E is a density point of EE.

Examples:

  1. Every interior point of EE is a density point: Let pp be an interior point. Then r>0\exists r > 0 s.t. Br(p)EB_r(p) \subset E. Then, for all cube QQ containing pp with side l<r2nl < \frac{r}{2\sqrt{n}}, QEQ \subset E. Hence m(QE)m(Q)=m(Q)m(Q)=1p\frac{m(Q \cap E)}{m(Q)} = \frac{m(Q)}{m(Q)} = 1 \Rightarrow p is density point.
  2. The irrationals: Let Qc\mathbb{Q}^c denote the set of irrational numbers and pQcp \in \mathbb{Q}^c. Then, for every cube QQ s.t. pQp \in Q, m(QQc)m(Q)=m(Q)m(QQ)m(Q)=m(Q)m(Q)=1\frac{m(Q \cap \mathbb{Q}^c)}{m(Q)} = \frac{m(Q) - m(Q \cap \mathbb{Q})}{m(Q)} = \frac{m(Q)}{m(Q)} = 1. Thus, every point is a density point.

Proof: WLOG, assume EE is bounded (will justify this later, don't worry). Take aa s.t. 0a<10 \leq a < 1 and consider Ea={pEδ(p,E)<a}E_{a} = \{p \in E | \underline{\delta}(p, E) < a\}. Want to show m(Ea)=0m^{*}(E_a) = 0. (The reason why we use outer measure is because we do not know a priori that EaE_a is measurable.)

For every pEap \in E_a, δ(p,E)<a\underline{\delta}(p, E) < a \Rightarrow exists arbitrarily small cubes {Qα}\{Q_\alpha\} in which m(EQα)m(Qα<a\frac{m(E \cap Q_{\alpha})}{m(Q_{\alpha}} < a. (This is the reason why EaE_a was constructed with strict inequality, otherwise we are unable to have strict inequality here.) Let SS be the set of all such cubes p\forall p. Then SS is a Vitali covering of EaE_a, since pEa,r>0,QS\forall p \in E_a, \forall r > 0, \exists Q \in S s.t. pQBr(p)p \in Q \subset B_r(p).

By Vitali Covering Lemma, for an arbitrarily chosen ϵ>0\epsilon > 0, we can select a countable collection Q1,Q2,Q3,Q_1, Q_2, Q_3, … from SS s.t.: (1) QiQj=ϕQ_i \cap Q_j = \phi for iji \neq j, (2) m(i=1m(Qi))=Σi=1m(Qi)m(Ea)+ϵm (\bigcup_{i=1}^{\infty}m(Q_i)) = \Sigma_{i=1}^{\infty} m(Q_i) \leq m^{*}(E_a) + \epsilon and (3) Ea\(i=1Qi)E_a \RM (\bigcup_{i=1}^{\infty} Q_i) is a null set.

Thus, we have the following: m(Ea)Σi=1m(EaQi)Σi=1m(EQi)=Σi=1m(EQk)m(Qk)m(Qk)<aΣi=1m(Qi)a(m(Ea)+ϵ)m^{*}(E_a) \leq \Sigma_{i=1}^{\infty} m^{*}(E_a \cap Q_i) \leq \Sigma_{i=1}^{\infty} m(E \cap Q_i) = \Sigma_{i=1}^{\infty} \frac{m(E \cap Q_k)}{m(Q_k)}\cdot m(Q_k) < a \Sigma_{i=1}^{\infty} m(Q_i) \leq a(m^{*}(E_a) + \epsilon)

Note that in the above, the first inequality is due to sub-additivity of outer measure. The second is due to monotonicity of outer measure, where we used m(EQi)=m(EQi)m^{*}(E \cap Q_i) = m(E \cap Q_i) since EQiE \cap Q_i measurable.

Hence, m(Ea)<aϵ1am(Ea)=0m^*(E_a) < \frac{a \epsilon}{1 - a} \Rightarrow m^*(E_a) = 0 since ϵ\epsilon is chosen arbitrarily. We also demonstrated that EaE_a is in fact measurable, since it has outer measure 00.

Now consider the sequence (an)n(a_n)_n where an=11/na_n = 1 - 1/n. Then m(Ean)=0m(E_{a_n}) = 0 n\forall n as discussed above. Since EaiEai+1E_{a_i} \subset E_{a_{i+1}}, the sequence (m(Ean))n(m(E_{a_n}))_n converges m(nEan)m (\bigcup_n E_{a_n}) by upward measure continuity m(nEan)=0\Rightarrow m (\bigcup_n E_{a_n}) = 0.

Consider E\(nEan)E \RM (\bigcup_n E_{a_n}). Then lim infQpm(EQ)m(Q)=1limQpm(EQ)m(Q)=1δ(p,E)=1\text{lim inf}_{Q \downarrow p} \frac{m(E \cap Q)}{m(Q)} = 1 \Rightarrow \text{lim}_{Q \downarrow p} \frac{m(E \cap Q)}{m(Q)} = 1 \Rightarrow \delta(p, E) = 1 pE\(nEan)\forall p \in E \RM (\bigcup_n E_{a_n}). Thus, almost every point of EE is a density point of EE.

We assumed EE is bounded, because we can always divide EE into countable disjoint sets. Then, for each set, the measure of non-density points is 00, thus the union of all non-density points is also 00. Hence, almost all pEp \in E is a density point of EE.

Corollary: If EE is measurable, then for almost every pRnp \in \mathbb{R}^n, we have: χE(p)=limQpm(EQ)m(Q)\chi_{E}(p) = \text{lim}_{Q \downarrow p} \frac{m(E \cap Q)}{m(Q)}. Here χE(p)\chi_{E}(p) refers to the indicator function that returns 11 if pEp \in E and 00 else.

Proof: From above, for almost every pEp \in E, we have limQpm(EQ)m(Q)=1=χE(p)\text{lim}_{Q \downarrow p} \frac{m(E \cap Q)}{m(Q)} = 1 = \chi_{E}(p). Since EE measurable \Rightarrow EcE^c measurable, for almost every pEcp \in E^c, limQpm(EcQ)m(Q)=1\text{lim}_{Q \downarrow p} \frac{m(E^c \cap Q)}{m(Q)} = 1. The measurability of EE implies m(QE)+m(QEc)=m(Q)m(QE)m(Q)+m(QEc)m(Q)=1m(Q \cap E) + m(Q \cap E^c) = m(Q) \Rightarrow \frac{m(Q \cap E)}{m(Q)} + \frac{m(Q \cap E^c)}{m(Q)} = 1. Hence, m(QE)m(Q)=1m(QEc)m(Q)=0=χE(p)\frac{m(Q \cap E)}{m(Q)} = 1 - \frac{m(Q \cap E^c)}{m(Q)} = 0 = \chi_{E}(p). And, we are done.

The General Stokes Formula Part I: Review (04/05 Class Presentation)

In this presentation, I will cover Stokes' Theorem including closed forms and exact forms. If I can present the following properly and well, I would be really satisfied. =)

Update: My presentation notes in PDF stokes_presentation.pdf

Review of concepts from differential forms:

  1. Differential kk-form: A function that takes in a kk-dimensional surface and sends it to R\mathbb{R}.
    1. It is of the form fi1,,ikdxi1dxikf_{i_1, …, i_k} dx_{i_1}…dx_{i_k}. In particular, it has kk differentials.
    2. (fi1,,ikdxi1dxik)(C)=Cfi1,,ikdxi1dxik(f_{i_1, …, i_k} dx_{i_1}…dx_{i_k})(C) = \int_{C} f_{i_1, …, i_k} dx_{i_1}…dx_{i_k} i.e. evaluate this integral over the kk-dimensional surface CRnC \in \mathbb{R}^n. You can think of fi1,,ikf_{i_1, …, i_k} as a weight function.
  2. A kk-cell in Rn\mathbb{R}^n is a smooth map ϕ:[0,1]kRn\phi: [0, 1]^k \rightarrow \mathbb{R}^n i.e. think of it as the transformation of the kk-dimensional cube to nn dimensional space.
    1. We can now evaluate kk-forms over kk-cells via change of variables: (fi1,,ikdxi1dxik)(C)=Cfi1,,ikdxi1dxik=[0,1]kfi1,,ikϕIudu=0101fi1,,ik(ϕi1,,ϕik)(u1,,uk)du1du2duk(f_{i_1, …, i_k} dx_{i_1}…dx_{i_k})(C) = \int_{C} f_{i_1, …, i_k} dx_{i_1}…dx_{i_k} = \int_{[0,1]^k} f_{i_1, …, i_k} \frac{\partial \phi_I}{\partial u} du = \int_{0}^{1} … \int_{0}^{1} f_{i_1, …, i_k} \frac{\partial(\phi_{i_1},…,\phi_{i_k})}{\partial(u_1, …, u_k)} du_1 du_2 … du_k

x=(x1,,xk)Ikx = (x_1, …, x_k) \in I^k, y=(y1,,yn)Rny = (y_1, …, y_n) \in \mathbb{R}^n. ϕ:[0,1]kRn\phi:[0,1]^k \rightarrow \mathbb{R}^n is a kk-cell in Rn\mathbb{R}^n. y=ϕ(x)y = \phi(x). ϕi(x)=ϕi(x1,,xk)=yi\phi_i(x) = \phi_i(x_1, …, x_k) = y_i.

Geometric interpretation

  1. II-shadow area: dyi1,,ik:ϕ[0,1]kϕIudu=0101(ϕi1,,ϕik)(u1,,uk)du1du2dukdy_{i_1, …, i_k}: \phi \rightarrow \int_{[0,1]^k} \frac{\partial \phi_I}{\partial u} du = \int_{0}^{1} … \int_{0}^{1} \frac{\partial(\phi_{i_1},…,\phi_{i_k})}{\partial(u_1, …, u_k)} du_1 du_2 … du_k
  2. Weighted II-shadow area: fdyi1,,ik:ϕ[0,1]kf(ϕ(u))ϕIudu=0101f(ϕ(u))(ϕi1,,ϕik)(u1,,uk)du1du2dukf dy_{i_1, …, i_k}: \phi \rightarrow \int_{[0,1]^k} f(\phi(u)) \frac{\partial \phi_I}{\partial u} du = \int_{0}^{1} … \int_{0}^{1} f(\phi(u)) \frac{\partial(\phi_{i_1},…,\phi_{i_k})}{\partial(u_1, …, u_k)} du_1 du_2 … du_k

Theorem [Pugh 36]: Reparametrization of kk-cell produces the same answer up to a sign change.

Proof: ϕTfdyi1,,ik=[0,1]kf(ϕT(u))(ϕT)i1,,ikvTudu=[0,1]kf(ϕ(v))ϕi1,,ikv(1)sdv=(1)sϕω\int_{\phi \circ T} f dy_{i_1, …, i_k} = \int_{[0,1]^k} f(\phi \circ T(u)) \frac{\partial(\phi \circ T)_{i_1, …, i_k}}{\partial v} \frac{\partial T}{\partial u} du = \int_{[0,1]^k} f(\phi(v)) \frac{\partial \phi_{i_1, …, i_k}}{\partial v} (-1)^s dv = (-1)^s \int_{\phi} \omega

Proposition [Pugh 37]: Each kk-form ω\omega has a unique expression as a sum of simple kk-forms with ascending kk-tuple indices, ω=ΣfAdyA\omega = \Sigma f_A dy_A.

Proof: For every kk tuple (i1,,ik)(i_1, …, i_k), it can be arranged such that it ascends. Hence, existence is guaranteed.

Let A=(i1,,ik)A = (i_1, …, i_k) be ascending. Fix yRny \in \mathbb{R}^n and L:RkRnL:\mathbb{R}^k \rightarrow \mathbb{R}^n be s.t. L(u)=u1ei1++ukeikL(u) = u_1 e_{i_1} + … + u_k e_{i_k}. For r>0r > 0, define function gr,y(u)=y+rL(u)g_{r, y}(u) = y + rL(u), i.e. gr,yg_{r, y} sends [0,1]k[0, 1]^k to a kk-dimensional cube of side length rr at yy.

fA(y)=limr01rkω(g)f_A(y) = lim_{r \rightarrow 0} \frac{1}{r^k} \omega(g)

Wedge Products α=Σi1,,ikαi1,,ikdyi1,,ik\alpha = \Sigma_{i_1, …, i_k} \alpha_{i_1, …, i_k} dy_{i_1, …, i_k} β=Σj1,,jlβj1,,jldyj1,,jl\beta = \Sigma_{j_1, …, j_l} \beta_{j_1, …, j_l} dy_{j_1, …, j_l}

αβ=Σi1,,ik,j1,,jlαi1,,ikβj1,,jldyi1,,ik,j1,,jl\alpha \wedge \beta = \Sigma_{i_1, …, i_k, j_1, …, j_l} \alpha_{i_1, …, i_k} \beta_{j_1, …, j_l} dy_{i_1, …, i_k, j_1, …, j_l}

kk-form wedge product with ll-form gives a (k+l)(k+l)-form

Exterior derivative: If ff is kk-form. Then dfdf is (k+1)(k+1)-form. Think of it as one more differential. df=fx1dx1++fxndxndf = \frac{\partial f}{\partial x_1} dx_1 + … + \frac{\partial f}{\partial x_n} dx_n

For a kk-form ω=Σfi1,,ikdyi1,,ik\omega = \Sigma f_{i_1, …, i_k} dy_{i_1, …, i_k} dω=Σdfi1,,ikdyi1,,ikd\omega = \Sigma df_{i_1, …, i_k} \wedge dy_{i_1, …, i_k} Intuitively, it is how the coefficient fi1,,ikf_{i_1, …, i_k} changes. Example d(fdx+gdy)=fydydx+gxdxdy=(gxfy)dxdyd(f dx + g dy) = f_y dy \wedge dx + g_x dx \wedge dy = (g_x - f_y) dx \wedge dy d(dΩ)=0ωd(d\Omega) = 0 \forall \omega kk-form on Rn\mathbb{R}^n.

Pushforward and Pullback Consider T:RnRmT: \mathbb{R}^n \rightarrow \mathbb{R}^m be smooth. Then since ϕ:[0,1]kRn\phi: [0,1]^k \rightarrow \mathbb{R}^n is a kk-cell in Rn\mathbb{R}^n, we have an induced kk-cell Tϕ:[0,1]kRmT \circ \phi: [0, 1]^k \rightarrow \mathbb{R}^m. This inducing function, we call it T:ϕTϕT_*: \phi \rightarrow T \circ \phi is a transformation of kk-cells.

Pullback is the dual of pushforward. Let α\alpha be a kk-form on Rm\mathbb{R}^m, then its pullback to Rn\mathbb{R}^n is a kk-form on Rn\mathbb{R}^n, denoted by T(α)T^*(\alpha) that sends each kk-cell ϕ\phi to α(Tϕ)\alpha(T \circ \phi)

i.e. (T(α))(ϕ)=α(Tϕ)=α(T(ϕ))(T^*(\alpha))(\phi) = \alpha(T \circ \phi) = \alpha(T_*(\phi))

The General Stokes Formula Part II: Mechanics (04/05 Class Presentation)

Definition: A kk-chain is a linear combination of kk-cells. i.e. Φ=Σj=1Najϕj\Phi = \Sigma_{j=1}^N a_j \phi_j, where aiRa_i \in \mathbb{R}. Φω=Σj=1Najϕjω\int_{\Phi} \omega = \Sigma_{j=1}^N a_j \int_{\phi_j} \omega (i.e. just integrate separately)

Definition: The boundary of a (k+1)(k+1)-cell ϕ\phi is the kk-chain ϕ=Σj=1k+1(1)j+1(ϕij,1ϕij,0)\partial \phi = \Sigma_{j=1}^{k+1} (-1)^{j+1} (\phi \circ i^{j, 1} - \phi \circ i^{j, 0}), where ij,0(u1,,uk)(u1,,uj1,0,uj,,uk)i^{j, 0}(u_1, …, u_k) \rightarrow (u_1, …, u_{j - 1}, 0, u_j, …, u_k) and ij,1(u1,,uk)(u1,,uj1,1,uj,,uk)i^{j, 1}(u_1, …, u_k) \rightarrow (u_1, …, u_{j - 1}, 1, u_j, …, u_k)

Here, ij,0(u1,,uk)=(u1,,uj1,0,uj,,uk)i^{j,0}(u_1, …, u_k) = (u_1, …, u_{j-1}, 0, u_j, …, u_k) and ij,1(u1,,uk)=(u1,,uj1,1,uj,,uk)i^{j,1}(u_1, …, u_k) = (u_1, …, u_{j-1}, 1, u_j, …, u_k). They are kk-cells in Rk+1\mathbb{R}^{k+1} because they map the unit cube in Rk\mathbb{R}^k i.e. [0,1]k[0,1]^k to Rk+1\mathbb{R}^{k+1}.

Geometrically, they map to the faces of the unit cube in Rk+1\mathbb{R}^{k+1}, which is why ϕ\partial \phi is called the boundary.

Definition: The jjth dipole of ϕ\phi to be δjϕ=ϕij,1ϕij,0\delta^j \phi = \phi \circ i^{j, 1} - \phi \circ i^{j, 0}, i.e. the jjth term or the function representing the jjth faces, so ϕ=Σj=1k+1(1)j+1δjϕ\partial \phi = \Sigma_{j=1}^{k+1} (-1)^{j+1} \delta^j \phi.

Claim: δjϕ\delta^j \phi is the pushforward of δji\delta^j i; i.e. δjϕ=ϕ(δji)\delta^j \phi = \phi_*(\delta^j i)

Proof: δjϕ=ϕij,1ϕij,0=ϕ(ij,1ij,0)=ϕ(ij,1ij,0)=ϕ(iij,1iij,0)=ϕ(δji)\delta^j \phi = \phi \circ i^{j,1} - \phi \circ i^{j,0} = \phi \circ (i^{j,1} - i^{j,0}) = \phi_*(i^{j,1} - i^{j,0}) = \phi_*(i \circ i^{j,1} - i \circ i^{j,0}) = \phi_*(\delta^j i)

Stoke's Formula for Cubes

If ω\omega is a (n1)(n - 1)-form in Rn\mathbb{R}^n (i.e. taking k=n1k = n - 1) and i:[0,1]nRni: [0, 1]^n \rightarrow \mathbb{R}^n is the identity inclusion nn-cell in Rn\mathbb{R}^n, then idω=iω\int_i d\omega = \int_{\partial i} \omega.

Proof: Write ω=Σi=1nfi(x)dx1dxi^dxn\omega = \Sigma_{i=1}^{n} f_i(x) dx_1 \wedge … \wedge \hat{dx_i} \wedge … \wedge dx_n in standard, ascending form (this is why ω\omega can be written as the sum of nn terms). The exterior derivative is then dω=Σi=1ndfi(x)dx1dxi^dxn=Σi=1nfixidxidx1dxi^dxn=Σi=1n(1)i1fixidx1dxn=(Σi=1n(1)i1fixi)dx1dxnd\omega = \Sigma_{i=1}^{n} df_i(x) \wedge dx_1 \wedge … \wedge \hat{dx_i} \wedge … \wedge dx_n = \Sigma_{i=1}^{n} \frac{\partial f_i}{\partial x_i} dx_i \wedge dx_1 \wedge … \wedge \hat{dx_i} \wedge … \wedge dx_n = \Sigma_{i=1}^{n} (-1)^{i - 1} \frac{\partial f_i}{\partial x_i} dx_1 \wedge … \wedge dx_n = (\Sigma_{i=1}^{n} (-1)^{i - 1} \frac{\partial f_i}{\partial x_i}) dx_1 \wedge … \wedge dx_n.

Hence idω=Σi=1n(1)i1[0,1]nfixi1dx1dxn\int_{i} d\omega = \Sigma_{i=1}^n (-1)^{i-1} \int_{[0,1]^n} \frac{\partial f_i}{\partial x_i} \cdot 1 \cdot dx_1…dx_n. It's worth emphasizing that the Jacobian is the identity matrix since it is the inclusion map, so determinant is 11.

We have now settled one side of the equation. For the other side:

iω=Σj=1n(1)j+1δjiω=Σj=1n(1)j+1δjiω\int_{\partial i} \omega = \int_{\Sigma_{j=1}^{n} (-1)^{j+1} \delta^j i} \omega = \Sigma_{j=1}^{n} (-1)^{j+1} \int_{\delta^j i} \omega

Note that δji=iij,1iij,0=ij,1ij,0\delta^j i = i \circ i^{j,1} - i \circ i^{j,0} = i^{j,1} - i^{j,0} and ij,0,ij,1:Rn1Rni^{j,0}, i^{j,1}: \mathbb{R}^{n-1} \rightarrow \mathbb{R}^n

Then when we calculate the Jacobians: (ij,0)x1,,xi^,,xn(u1,,un1)=1\frac{\partial(i^{j,0})_{x_1, …, \hat{x_i}, …, x_n}}{\partial(u_1, …, u_{n-1})} = 1 if i=ji=j and 00 otherwise. Similarly, (ij,1)x1,,xi^,,xn(u1,,un1)=1\frac{\partial(i^{j,1})_{x_1, …, \hat{x_i}, …, x_n}}{\partial(u_1, …, u_{n-1})} = 1 if i=ji = j and 00 otherwise. This is because for the jjth face (which is represented by the jjth dipole), the jjth coordinate is fixed at either 00 or 11. Deleting the iith component for iji \neq j will still result in the jjth component being constant, so Jacobian =0= 0. Otherwise, since it is inclusion, the Jacobian is 11.

δjiω=δjiΣl=1nfl(x)dx1dxl^dxn=δji(fj(ij,1(u))fj(ij,0(u)))du1dun1=δji(fj(u1,,uj1,1,uj,,un1)fj(u1,,uj1,0,uj,,un1))du1dun1=δjifj(u1,,uj1,y,uj,,un1)ydydu1dun1=δjifjxjdx1dxn\int_{\delta^j i} \omega = \int_{\delta^j i} \Sigma_{l=1}^n f_l(x) dx_1 \wedge … \wedge \hat{dx_l} \wedge … \wedge dx_n = \int_{\delta^j i} (f_j(i^{j, 1}(u)) - f_j(i^{j, 0}(u))) du_1 \cdot … \cdot du_{n-1} = \int_{\delta^j i} (f_j(u_1, …, u_{j-1}, 1, u_j, …, u_{n-1}) - f_j(u_1, …, u_{j-1}, 0, u_j, …, u_{n-1})) du_1 \cdot … \cdot du_{n-1} = \int_{\delta^j i} \frac{\partial f_j(u_1, …, u_{j-1}, y, u_{j}, …, u_{n-1})}{\partial y} dy \cdot du_1 \cdot … \cdot du_{n-1} = \int_{\delta^j i} \frac{\partial f_j}{\partial x_j} dx_1 … dx_n

In the second last equality, we used the fundamental theorem of calculus to split up the difference of fj(u1,,uj1,1,uj,,un1)fj(u1,,uj1,0,uj,,un1)=01f(u1,,uj1,y,uj,,un1)ydyf_j(u_1, …, u_{j-1}, 1, u_j, …, u_{n-1}) - f_j(u_1, …, u_{j-1}, 0, u_j, …, u_{n-1}) = \int_{0}^{1} \frac{\partial f(u_1, …, u_{j-1}, y, u_j, …, u_{n-1})}{\partial y} dy where yy refers to jjth coordinate.

In the last equality, we used Fubini's Theorem i.e. order of integration in ordinary multiple integration is irrelevant. Also, y,u1,,un1y, u_1, …, u_{n-1} are just arbitrary variables that can be looked at as x1,,xnx_1, …, x_n

Hence, iω=Σj=1n(1)j1δjiω=Σj=1n(1)j10101fjxjdx1dxn=Σj=1n(1)j1[0,1]nfjxjdx1dxn\int_{\partial i} \omega = \Sigma_{j=1}^n (-1)^{j-1} \int_{\delta^j i} \omega = \Sigma_{j=1}^n (-1)^{j-1} \int_{0}^{1} … \int_{0}^{1} \frac{\partial f_j}{\partial x_j} dx_1 … dx_n = \Sigma_{j=1}^n (-1)^{j-1} \int_{[0,1]^n} \frac{\partial f_j}{\partial x_j} dx_1 … dx_n, exactly the same as what we've got before.

So, iω=idω\int_{\partial i} \omega = \int_{i} d\omega for cubes.

Stoke's Formula for General Cell: Let ω\omega be a (n1)(n-1)-form in Rm\mathbb{R}^m and ϕ\phi an nn-cell in Rm\mathbb{R}^m, then ϕdω=ϕω\int_{\phi} d\omega = \int_{\partial \phi} \omega.

Proof: ϕdω=ϕidω=iϕdω=idϕω=iϕω=ϕiω=ϕω\int_{\phi} d\omega = \int_{\phi \circ i} d\omega = \int_i \phi^* d\omega = \int_i d\phi^*\omega = \int_{\partial i} \phi^* \omega = \int_{\phi_* \partial i} \omega = \int_{\partial \phi} \omega

The first equality is because ϕ=ϕi\phi = \phi \circ i since ii is the identity inclusion map from [0,1]nRn[0,1]^n \rightarrow \mathbb{R}^n. Second equality follows from 43d since Tϕα=ϕT(α)\int_{T \circ \phi} \alpha = \int_{\phi} T^*(\alpha). Third equality follows from 43c commutativity property of exterior derivative and pullback operator ϕdω=dϕω\phi^* d\omega = d\phi^* \omega. The fourth equality follows from Stoke's Formula for Cubes. The fifth equality follows from duality equation T(α)(ϕ)=α(T(ϕ))T^*(\alpha)(\phi) = \alpha(T_*(\phi)), so ϕTα=T(ϕ)α\int_{\phi} T^*\alpha = \int_{T_*(\phi)} \alpha.

The final equality follows from the fact that ϕ(δi)=ϕ(Σj=1k+1(1)j+1(iii,1iii,0))=ϕ(Σj=1k+1(1)j+1(ij,1ij,0))=ϕ(Σj=1k+1(1)j+1(ij,1ij,0))=Σj=1k+1(1)j+1(ϕij,1ϕij,0)=ϕ\phi_*(\delta i) = \phi_*(\Sigma_{j=1}^{k+1} (-1)^{j+1} (i \circ i^{i,1} - i \circ i^{i,0})) = \phi_*(\Sigma_{j=1}^{k+1} (-1)^{j+1} (i^{j,1} - i^{j,0})) = \phi \circ (\Sigma_{j=1}^{k+1} (-1)^{j+1} (i^{j,1} - i^{j,0})) = \Sigma_{j=1}^{k+1} (-1)^{j+1} (\phi \circ i^{j,1} - \phi \circ i^{j,0}) = \partial \phi

Stoke's Formula for Manifolds: Let ω\omega be a (m1)(m-1)-form in Rn\mathbb{R}^n and MRnM \subset \mathbb{R}^n divides into mm-cells diffeomorphic to [0,1]m[0,1]^m and its boundary divides into (m1)(m-1)-cells diffeomorphic to [0,1]m1[0,1]^{m-1}, then Mdω=Mω\int_{M} d\omega = \int_{\partial M} \omega.

The General Stokes Formula Part III: Applications (04/05 Class Presentation)

Stoke's Theorem: Mdω=Mω\int_{M} d\omega = \int_{\partial M} \omega

  • Fundamental Theorem of Calculus

Take M=[a,b]RM = [a, b] \subset \mathbb{R} and ω=f\omega = f (i.e. a zero-form). Then Mω=f(b)f(a)\int_{\partial M} \omega = f(b) - f(a) (since it is the evaluation of ff on the boundaries of MM which is just {a,b}\{a, b\}). dω=fdxd\omega = f' dx. Hence, Mdω=abf(x)dx=f(b)f(a)\int_{M} d\omega = \int_a^b f'(x) dx = f(b) - f(a) as we know from FTC.

Let f:R2Rf: \mathbb{R}^2 \rightarrow \mathbb{R} be smooth. Then df=fxdx+fydydf = f_x dx + f_y dy. Let MM represent a path in R2\mathbb{R}^2 from point pp to qq (i.e. MM is a 11-cell in R2\mathbb{R}^2), then M=p,q\partial M = {p, q}. By Stoke's Theorem, Mdf=Mf=f(q)f(p)\int_{M} df = \int_{\partial M} f = f(q) - f(p). For any two paths, their boundaries are the same. Hence, we can say df=fxdx+fydydf = f_x dx + f_y dy is path independent, which corresponds with what we know.

  • Green's Theorem

Take ω=fdx+gdy\omega = f dx + g dy. Take M=DM = D a region. Then M=C\partial M = C, the curve bounding DD. Then:

Cfdx+gdy=Cω=Mω=Mdω=Dd(fdx+gdy)=Dfydydx+gxdxdy=D(gxfy)dxdy\int_C f dx + g dy = \int_C \omega = \int_{\partial M} \omega = \int_{M} d\omega = \int_{D} d(f dx + g dy) = \int_{D} f_y dy \wedge dx + g_x dx \wedge dy = \int_{D} (g_x - f_y) dx dy

  • Divergence Theorem

Let F=(f,g,h)F = (f, g, h) be a smooth vector field on UR3U \subset \mathbb{R}^3. Then, take M=DM = D, then M=S\partial M = S. Consider the 22-form ω=fdydz+gdzdx+hdxdy\omega = f dy \wedge dz + g dz \wedge dx + h dx \wedge dy.

Sfdydz+gdzdx+hdxdy=Mω=Md(fdydz+gdzdx+hdxdy)=M(fxdxdydz+gydydzdx+hzdzdxdy)=M(fx+gy+hz)dxdydz=DF\int_{S} f dy \wedge dz + g dz \wedge dx + h dx \wedge dy = \int_{\partial M} \omega = \int_{M} d(f dy \wedge dz + g dz \wedge dx + h dx \wedge dy) = \int_{M} (f_x dx \wedge dy \wedge dz + g_y dy \wedge dz \wedge dx + h_z dz \wedge dx \wedge dy) = \int_{M} (f_x + g_y + h_z) dx \wedge dy \wedge dz = \int_{D} \nabla \cdot F

We call LHS the flux. Intuitively, it is a statement about the conservation of flow. For an surface, the total flow out of the surface (i.e. flux) is equal to the integral of the divergence (measure of source / sink) across the whole enclosed surface.

  • Stoke's Curl Theorem

Take ω=fdx+gdy+hdz\omega = f dx + g dy + h dz and M=SM = S (a surface), then M=C\partial M = C a curve (the boundary of the surface).

Cfdx+gdy+hdz=Mω=Mdω=Sd(fdx+gdy+hdz)=Sfydydx+fzdzdx+gxdxdy+gzdzdy+hxdxdz+hydydz=S(hygz)dydz+(fzhx)dzdx+(gxfy)dxdy\int_{C} f dx + g dy + h dz = \int_{\partial M} \omega = \int_{M} d\omega = \int_{S} d(f dx + g dy + h dz) = \int_{S} f_y dy \wedge dx + f_z dz \wedge dx + g_x dx \wedge dy + g_z dz \wedge dy + h_x dx \wedge dz + h_y dy \wedge dz = \int_{S} (h_y - g_z) dy \wedge d_z + (f_z - h_x) dz \wedge dx + (g_x - f_y) dx \wedge dy as desired.

Intuitively, it is saying that the line integral of a loop in a vector field is equal to the flux of the curl of the vector field through the enclosed surface.

The General Stokes Formula Part IV: Closed and Exact Forms (04/05 Class Presentation)

Definition: Say a kk-form ω\omega is closed if its exterior derivative is 00. (i.e. dω=0d\omega = 0)

Definition: Say a kk-form ω\omega is exact if it is the exterior derivative of a (k1)(k-1)-form. (i.e. α\exists \alpha s.t. d(α)=ωd(\alpha) = \omega)

Proposition: Every exact form is closed.

Proof: Let ω\omega be an exact form, then α\exists \alpha s.t. d(α)=ωd(\alpha) = \omega. Since d2=0d^2 = 0, dω=d(d(α))=0d\omega = d(d(\alpha)) = 0 as desired.

Motivation: Exact always implies closed. But when does closed imply exact? When can we find an anti-derivative for a closed form ω\omega? i.e. find an α\alpha s.t. ω=dα\omega = d\alpha?

Ans: If forms are defined on Rn\mathbb{R}^n, then the answer is always, by Poincaré Lemma.

Poincaré Lemma: If ω\omega is a closed kk-form (k>0k > 0) on Rn\mathbb{R}^n, then it is exact.

Motivation: We can show for a kk-form ω\omega in Rn\mathbb{R}^n, we have an integral operator LkL_k such that (Lk+1d+dLk)(ω)=ω(L_{k+1}d + dL_k)(\omega) = \omega. Hence, if ω\omega is closed, then dω=0d\omega = 0, hence dLk(ω)=d(Lk(ω))=ωdL_k(\omega) = d(L_k(\omega)) = \omega, so ω\omega has an anti-derivative. Suffices to show the construction of such LkL_k.

Claim: Lk\exists L_k that maps kk-form on Rn\mathbb{R}^n to (k1)(k-1)-form on Rn\mathbb{R}^n with the property that ω\forall \omega s.t. ω\omega is a kk-form, (Lk+1d+dLk)(ω)=ω(L_{k+1}d + dL_k)(\omega) = \omega.

If such function LkL_k exists, then since ω\omega is closed, dω=0d\omega = 0, so ω=(Lk+1d+dLk)(ω)=Lk+1dω+dLkω=d(Lkω)\omega = (L_{k+1}d + dL_k)(\omega) = L_{k+1}d\omega + dL_k\omega = d(L_k\omega), in particular, ω\omega has an anti-derivative, so it is exact.

Proof of Claim:

Let β\beta be a kk-form on Rn+1\mathbb{R}^{n+1} instead. Let (x,t)(x, t) be the representation of a point in Rn+1\mathbb{R}^{n+1}, where xRnx \in \mathbb{R}^n and tRt \in \mathbb{R}. Then β=Σi1,,ikfi1,,ikdxi1,,ik+Σj1,,jk1dtdxj1,,jk1\beta = \Sigma_{i_1, …, i_k} f_{i_1, …, i_k} dx_{i_1, …, i_k} + \Sigma_{j_1, …, j_{k-1}} dt \wedge dx_{j_1, …, j_{k-1}}.

Then dβ=Σi1,,ik,lfi1,,ikxldxldxi1,,ik+Σi1,,ikfi1,,iktdtdxi1,,ik+Σj1,,jk1,lgj1,,jk1xldxldtdxj1,,jk1d\beta = \Sigma_{i_1, …, i_k, l} \frac{f_{i_1, …, i_k}}{\partial x_l} dx_l \wedge dx_{i_1, …, i_k} + \Sigma_{i_1, …, i_k} \frac{\partial f_{i_1, …, i_k}}{\partial t} dt \wedge dx_{i_1, …, i_k} + \Sigma_{j_1, …, j_{k-1}, l} \frac{\partial g_{j_1, …, j_{k-1}}}{\partial x_l} dx_l \wedge dt \wedge dx_{j_1, …, j_{k-1}} where 1ln1 \leq l \leq n.

Define operators N:Ωk(Rn+1)Ωk1(Rn)N: \Omega^k(\mathbb{R}^{n+1}) \rightarrow \Omega^{k-1}(\mathbb{R}^n) (i.e. NN maps kk-forms in Rn+1\mathbb{R}^{n+1} to k1k-1-forms in Rn\mathbb{R}^{n}) as the following:

N(β)=Σj1,,jn1(01gj1,,jn1(x,t)dt)dxj1,,jn1N(\beta) = \Sigma_{j_1, …, j_{n-1}}(\int_{0}^{1} g_{j_1, …, j_{n-1}}(x, t)dt) dx_{j_1, …, j_{n-1}} i.e. NN is defined to ignore terms which dtdt doesn't appear. In the terms that dtdt appears, it integrates them and reduces the form by 11 differential. Here, β\beta, a kk-form in Rn+1\mathbb{R}^{n+1} is reduced to a (k1)(k-1)-form in Rn\mathbb{R}^n.

Subclaim: βΩk(Rn+1)\forall \beta \in \Omega^k(\mathbb{R}^{n+1}), (dN+Nd)(β)=Σi1,,ik(fi1,,ik(x,1)fi1,,ik(x,0))dxi1,,ik(dN + Nd)(\beta) = \Sigma_{i_1, …, i_k} (f_{i_1, …, i_k}(x, 1) - f_{i_1, …, i_k}(x, 0)) dx_{i_1, …, i_k}

Proof of Subclaim:

N(dβ)=N(Σi1,,ik,lfi1,,ikxldxldxi1,,ik+Σi1,,ikfi1,,iktdtdxi1,,ik+Σj1,,jk1,lgj1,,jk1xldxldtdxj1,,jk1)=Σi1,,ik(01fi1,,iktdt)dxi1,,ikΣj1,,jk1,l(01gj1,,jk1xldt)dxldxj1,,jk1N(d\beta) = N(\Sigma_{i_1, …, i_k, l} \frac{f_{i_1, …, i_k}}{\partial x_l} dx_l \wedge dx_{i_1, …, i_k} + \Sigma_{i_1, …, i_k} \frac{\partial f_{i_1, …, i_k}}{\partial t} dt \wedge dx_{i_1, …, i_k} + \Sigma_{j_1, …, j_{k-1}, l} \frac{\partial g_{j_1, …, j_{k-1}}}{\partial x_l} dx_l \wedge dt \wedge dx_{j_1, …, j_{k-1}}) = \Sigma_{i_1, …, i_k}(\int_{0}^{1} \frac{\partial f_{i_1, …, i_k}}{\partial t} dt) dx_{i_1, …, i_k} - \Sigma_{j_1, …, j_{k-1}, l}(\int_{0}^{1} \frac{\partial g_{j_1, …, j_{k-1}}}{\partial x_l} dt) dx_l \wedge dx_{j_1, …, j_{k-1}}

The first term in dβd\beta is ignore since it has no dtdt term. The second and third term contain dtdt, hence are integrated. The - sign appears due to the shifting of dtdt forward by one position.

On the other side,

dN(β)=d(N(β))=d(Σj1,,jk1(01gj1,,jk1(x,t)dt)dxj1,,jk1)=Σj1,,jk1,l(01gj1,,jk1(x,t)dt)xldxldxj1,,jk1=Σj1,,jk1,l(01gj1,,jk1xldt)dxldxj1,,jk1dN(\beta) = d(N(\beta)) = d(\Sigma_{j_1, …, j_{k-1}}(\int_{0}^{1} g_{j_1, …, j_{k-1}}(x, t)dt) dx_{j_1, …, j_{k-1}}) = \Sigma_{j_1, …, j_{k-1},l} \frac{\partial(\int_{0}^{1} g_{j_1, …, j_{k-1}}(x, t)dt)}{\partial x_l} dx_l \wedge dx_{j_1, …, j_{k-1}} = \Sigma_{j_1, …, j_{k-1}, l}(\int_{0}^{1} \frac{\partial g_{j_1, …, j_{k-1}}}{\partial x_l} dt) dx_l \wedge dx_{j_1, …, j_{k-1}}

In the last equality, we interchanged the integration and differentiation, because gg is smooth so it and its partial derivatives are continuous.

Hence: (dN+Nd)(β)=Σi1,,ik(01fItdt)dxi1,,ik=Σi1,,ik(fi1,,ik(x,1)fi1,,ik(x,0))dxi1,,ik(dN + Nd)(\beta) = \Sigma_{i_1, …, i_k}(\int_{0}^{1} \frac{\partial f_I}{\partial t} dt) dx_{i_1, …, i_k} = \Sigma_{i_1, …, i_k} (f_{i_1, …, i_k}(x, 1) - f_{i_1, …, i_k}(x, 0)) dx_{i_1, …, i_k} as desired, where the last inequality comes from the Fundamental Theorem of Calculus. (whew!)

Now we define the cone map ρ(x,t)=tx\rho(x, t) = tx. ρ:Rn+1Rn\rho: \mathbb{R}^{n+1} \rightarrow \mathbb{R}^n.

Claim: L=NρL = N \circ \rho^* is our desired construction i.e. (Ld+dL)ω=ω(Ld + dL)\omega = \omega

Proof of Claim:

Ld+dL=Nρd+dNρ=Ndρ+dNρ=(Nd+dN)ρLd + dL = N\rho^* d + dN \rho^* = Nd\rho^* + dN\rho^* = (Nd + dN)\rho^* where the second equality comes from the commutativity of ρ\rho^* (a pullback) and dd.

Let ω=hdxi1,,ik\omega = h dx_{i_1, …, i_k} be a simple kk-form in Rn\mathbb{R}^n.

ρ(ω)=ρ(hdxi1,,ik)=(ρh)ρi1,,ikidk=h(tx)dρi1,,ik=h(tx)d(txi1)d(txik)\rho^*(\omega) = \rho^*(h dx_{i_1, …, i_k}) = (\rho^*h)\frac{\partial \rho_{i_1, …, i_k}}{idk} = h(tx) d\rho_{i_1, …, i_k} = h(tx) d(tx_{i_1}) \wedge … \wedge d(tx_{i_k})

=h(tx)((tdxi1+xi1dt)(tdxik+xikdt))=h(tx)(tkdxi1,,ik)+terms with dt= h(tx)( (t dx_{i_1} + x_{i_1} dt) \wedge … \wedge (t dx_{i_k} + x_{i_k}dt) ) = h(tx)(t^k dx_{i_1, …, i_k}) + \text{terms with dt}

I'm not sure about the second equality. I think it's due to my lack of understanding of pushforward and pullback.

Hence (Nd+dN)ρ(hdxi1,,ik)=(Nd+dN)(h(tx)(tkdxi1,,ik)+terms with dt)=(Nd+dN)(h(tx)(tkdxi1,,ik))=(h(1x)1kh(0x)0k)dxi1,,ik=hdxi1,,ik(Nd + dN) \circ \rho^*(h dx_{i_1, …, i_k}) = (Nd + dN)(h(tx)(t^k dx_{i_1, …, i_k}) + \text{terms with dt}) = (Nd + dN)(h(tx)(t^k dx_{i_1, …, i_k})) = (h(1\cdot x)1^k - h(0\cdot x)0^k)dx_{i_1, …, i_k} = h dx_{i_1, …, i_k}.

Hence (Ld+dL)(hdxi1,,ik)=(Nd+dN)ρ(hdxi1,,ik)=hdxi1,,ik(Ld+dL)(h dx_{i_1, …, i_k}) = (Nd + dN) \circ \rho^*(h dx_{i_1, …, i_k}) = h dx_{i_1, …, i_k}.

Since L,dL, d are linear, we have (Ld+dL)(ω)=ω(Ld+dL)(\omega) = \omega, hence closed forms on Rn\mathbb{R}^n are exact. Reminder that L=NρL = N \circ \rho^*.

I am very grateful to Professor Morgan Weiler for her YouTube video on Poincaré Lemma, in which she proved Poincaré Lemma inductively and helped me visualize the key steps in the proof: https://www.youtube.com/watch?v=C-LhZ9Tkc2g (she's also previously from Berkeley! =) )

The General Stokes Formula Part V: Final Touches (04/05 Class Presentation)

Corollary 47: If UU diffeomorphic to Rn\mathbb{R}^n, then all exact forms on UU are exact.

Proof of Corollary 47: Let T:URnT:U \rightarrow \mathbb{R}^n be a diffeomorphism and assume ω\omega closed kk-form on UU.

Let α=(T1)ω\alpha = (T^{-1})^*\omega be a kk-form on Rn\mathbb{R}^n. Since ω\omega closed, dω=0d\omega = 0, hence dα=d((T1)ω)=(T1)dω=(T1)0=0d\alpha = d( (T^{-1})^*\omega) = (T^{-1})^* d\omega = (T^{-1})^*0 = 0, so α\alpha is closed. By Poincaré Lemma, (k1)\exists (k-1)-form μ\mu in Rn\mathbb{R}^n with α=dμ\alpha = d\mu.

Then dTμ=Tdμ=Tα=T(T1)ω=(T1T)ω=ωdT^*\mu = T^*d\mu = T^*\alpha = T^* \circ (T^{-1})^*\omega = (T^{-1} \circ T)^*\omega = \omega, hence ω\omega has an antiderivative TμT^*\mu.

Hence, ω\omega is exact.

Corollary 48: Locally, closed forms defined on open subsets of Rn\mathbb{R}^n are exact.

Proof of Corollary 48: Locally, an open subset of Rn\mathbb{R}^n is diffeomorphic to Rn\mathbb{R}^n. (Why?)

Corollary 49: If URnU \subset \mathbb{R}^n is open and star-like (in particular, if UU is convex), then closed forms on UU are exact. A starlike set URnU \subset \mathbb{R}^n contains a point pp s.t. the line segment from each qUq \in U to pp lies in UU.

Proof of Corollary 49: Every starlike open sets in Rn\mathbb{R}^n is diffeomorphic to Rn\mathbb{R}^n. (Why?)

Corollary 50: A smooth vector field FF on R3\mathbb{R}^3 (or on an open set diffeomorphic to R3\mathbb{R}^3 is the gradient of a scalar function if and only if its curl is everywhere zero.

Proof of Corollary 50:

(\Leftarrow)

Let F=(f,g,h)F = (f, g, h). Define ω=fdx+gdy+hdz\omega = f dx + g dy + h dz. Then dω=fydydx+fzdzdx+gxdxdy+gzdzdy+hxdxdz+hydydz=(gxfy)dxdy+(hygz)dydz+(fzhx)dzdx=0d\omega = f_y dy \wedge dx + f_z dz \wedge dx + g_x dx \wedge dy + g_z dz \wedge dy + h_x dx \wedge dz + h_y dy \wedge dz = (g_x - f_y) dx \wedge dy + (h_y - g_z) dy \wedge dz + (f_z - h_x) dz \wedge d_x = 0

The last equality follows since ×F=0\nabla \times F = 0, so gxfy=0g_x - f_y = 0, hygz=0h_y - g_z = 0 and fzhx=0f_z - h_x = 0. Hence, ω\omega is closed and therefore exact, i.e. it has an antiderivative ϕ\phi. Note ϕ=ωϕxdx+ϕydy+ϕzdz=fdx+gdy+hdz\nabla \phi = \omega \Leftrightarrow \phi_x dx + \phi_y dy + \phi_z dz = f dx + g dy + h dz, hence ϕ=F\nabla \phi = F as desired.

(\Rightarrow)

For completeness, let F=ϕ=[ϕx,ϕy,ϕz]TF = \nabla \phi = [\phi_x, \phi_y, \phi_z]^T. Then ×F=[ϕzyϕyz,ϕxzϕzx,ϕyxϕxy]=0\nabla \times F = [\phi_zy - \phi_yz, \phi_xz - \phi_zx, \phi_yx - \phi_xy] = 0 where the last equality holds by equivalence of partial derivatives.

Corollary 51: A smooth vector field on R3\mathbb{R}^3 (or on an open set diffeomorphic to R3\mathbb{R}^3) has everywhere zero divergence iff it is the curl of some other vector field.

Proof of Corollary 51:

(\Leftarrow)

Let G=(A,B,C)G = (A, B, C). Define ω=Adydz+Bdzdx+Cdxdy\omega = A dy \wedge dz + B dz \wedge dx + C dx \wedge dy. Since G=Ax+By+Cz=0\nabla \cdot G = A_x + B_y + C_z = 0, dω=Axdxdydz+Bydydzdx+Czdzdxdy=(Ax+By+Cz)dxdydz=0d\omega = A_x dx \wedge dy \wedge dz + B_y dy \wedge dz \wedge dx + C_z dz \wedge dx \wedge dy = (A_x + B_y + C_z) dx \wedge dy \wedge dz = 0, hence ω\omega closed and therefore exact. Hence, exists anti-derivative α=fdx+gdy+hdz\alpha = f dx + g dy + h dz with dα=fydydx+fzdzdx+gxdxdy+gzdzdy+hxdxdz+hydydz=(gxfy)dxdy+(hygz)dydz+(fzhx)dzdx=Adydz+Bdzdx+Cdxdy=ωd\alpha = f_y dy \wedge dx + f_z dz \wedge dx + g_x dx \wedge dy + g_z dz \wedge dy + h_x dx \wedge dz + h_y dy \wedge dz = (g_x - f_y) dx \wedge dy + (h_y - g_z) dy \wedge dz + (f_z - h_x) dz \wedge dx = A dy \wedge dz + B dz \wedge dx + C dx \wedge dy = \omega.

So, A=(hygz)A = (h_y - g_z), B=(fzhx)B = (f_z - h_x) and C=(gxfy)C = (g_x - f_y). Define F=(f,g,h)F = (f, g, h), then ×F=[hygz,fzhx,gxfy]T=G\nabla \times F = [h_y - g_z, f_z - h_x, g_x - f_y]^T = G.

(\Rightarrow)

For completeness, let F=(f,g,h)F = (f, g, h) and G=×FG = \nabla \times F. Then G=[hygz,fzhx,gxfy]TG = [h_y - g_z, f_z - h_x, g_x - f_y]^T so G=hyxgzx+fzyhxy+gxzfyz=0\nabla \cdot G = h_{yx} - g_{zx} + f_{zy} - h_{xy} + g_{xz} - f_{yz} = 0.

FFT, Wavelet Transformation, Uncertainty Principle (The Last Presentation)

We start with developing further the theory of Fourier series for functions of a fixed period LL, so as to (hopefully) motivate the Discrete Fourier Transform. Then, I will introduce FFT. If there is enough time, I will go through Wavelet Transform and Uncertainty Principle.

Fourier series for functions of fixed period LL

A good example is to solve Tao 5.5.6.

Let L>0L > 0 and f:RCf:\mathbb{R} \rightarrow \mathbb{C} be continuous, LL-periodic. Define cn=1L[0,L]f(x)e2πinxLdxc_n = \frac{1}{L} \int_{[0,L]} f(x)e^{\frac{2 \pi inx}{L}} dx.

(a)

Define g(x)=f(Lx)g(x) = f(Lx). Note that g(x)C(R/Z;C)g(x) \in C(\mathbb{R}/\mathbb{Z}; \mathbb{C}). Hence, by Fourier's Theorem, limNgΣn=Nn=Ng^(n)en2=0limNgΣn=Nn=Ng^(n)en22=0limN[0,1]g(x)Σn=NNg^(n)en(x)2dx=0\text{lim}_{N \rightarrow \infty} \| g - \Sigma_{n=-N}^{n=N} \hat{g}(n)e_n \|_2 = 0 \Rightarrow \text{lim}_{N \rightarrow \infty} \| g - \Sigma_{n=-N}^{n=N} \hat{g}(n)e_n \|_2^2 = 0 \Rightarrow \text{lim}_{N \rightarrow \infty} \int_{[0,1]} |g(x) - \Sigma_{n=-N}^{N}\hat{g}(n)e_n(x)|^2 dx = 0.

Applying a change of variable xxLx \rightarrow \frac{x}{L}, we have: [0,1]g(x)Σn=NNg^(n)en(x)2dx=[0,L]g(xL)Σn=NNg^(n)en(xL)2dxL=1L[0,L]f(x)Σn=NNg^(n)en(xL)2dx\int_{[0,1]} |g(x) - \Sigma_{n=-N}^{N}\hat{g}(n)e_n(x)|^2 dx = \int_{[0,L]} |g(\frac{x}{L}) - \Sigma_{n=-N}^{N}\hat{g}(n)e_n(\frac{x}{L})|^2 \frac{dx}{L} = \frac{1}{L}\int_{[0,L]} |f(x) - \Sigma_{n=-N}^{N}\hat{g}(n)e_n(\frac{x}{L})|^2 dx

g^(n)=g,en=[0,1]g(x)e2πinxdx=[0,L]g(x/L)e2πinxLdxL=1L[0,L]f(x)e2πinxLdx=cn\hat{g}(n) = \langle g, e_n \rangle = \int_{[0, 1]} g(x)e^{-2\pi inx} dx = \int_{[0, L]} g(x/L)e^{\frac{-2\pi inx}{L}} \frac{dx}{L} = \frac{1}{L} \int_{[0, L]} f(x)e^{\frac{-2\pi inx}{L}} dx = c_n

Hence, limN[0,1]g(x)Σn=NNg^(n)en(x)2dx=1LlimN[0,L]f(x)Σn=NNcne2πinxL2dx=0\text{lim}_{N \rightarrow \infty} \int_{[0,1]} |g(x) - \Sigma_{n=-N}^{N}\hat{g}(n)e_n(x)|^2 dx = \frac{1}{L} \text{lim}_{N \rightarrow \infty} \int_{[0,L]} |f(x) - \Sigma_{n=-N}^{N}c_n e^{\frac{2\pi inx}{L}}|^2 dx = 0

This is exactly what we want to show, that Σn=cne2πinxL\Sigma_{n=-\infty}^{\infty} c_n e^{\frac{2 \pi inx}{L}} converges in L2L_2 to ff.

(b, c) Subsequently, the proof for these two parts are exactly similar to the proof in Tao with the change of variables.

Discrete Fourier Transform

The discrete Fourier transform (DFT) converts a sequence of NN numbers (yj)j=0N1=(y0,,yN1)C(y_j)_{j=0}^{N-1} = (y_0, …, y_{N-1}) \in \mathbb{C} to a new sequence of NN numbers c0,,cN1Cc_0, …, c_{N-1} \in \mathbb{C}, given by: ck=Σj=0N1yje2πijkNc_k = \Sigma_{j=0}^{N-1} y_j e^{\frac{-2 \pi ijk}{N}}

Think of (yj)(y_j) as the values of a function or signal at equally spaced times x=0,,N1x = 0, …, N - 1. The output ckc_k encodes the amplitude and phase of e2πikxNe^{\frac{2 \pi ikx}{N}}. Key idea: approximate the signals by a linear combination of waves that has wavelength that are an integer factor of NN i.e. all such waves has wavelength that are an integer

Normally, we know the wave equation as y(x)=cos(kx)+isin(kx)=eikxy(x) = cos(kx) + i\cdot sin(kx) = e^{ikx} where k=2πλk = \frac{2\pi}{\lambda} is the wavenumber; λ\lambda is the wavelength.

Hence, the wave corresponding to ckc_k, which is e2πikxNe^{\frac{2 \pi ikx}{N}} has wavenumber k=2πikNk = \frac{2\pi ik}{N}, thus it has a wavelength of λ=Nk\lambda = \frac{N}{k}. Suffices to compute ckc_k to find the coefficients of an approximation of the original signal (yj)j(y_j)_j by a linear combination of the waves

For the encoding of signals, suffices to solve a system of linear equations. Hence, introduce the Vandermonde Matrix for roots of unity

Suffices to matrix multiply the Vandermonde matrix for roots of unity with the signal vector.

Remark: This problem is also equivalent to evaluating a polynomial at the roots of unity, the so-called changing from coefficient representation to value representation.

Fast Fourier Transform

A naive implementation of the Discrete Fourier Transform takes O(N2)O(N^2). However, Fast Fourier Transform performs it in O(NlogN)O(N log N) by exploiting the symmetry of the roots of unity.

An illustration of idea is: break down the polynomial into even and odd terms. We can recycle many computations. Use 11 and 1-1 as an example.

Inverse Fourier Transform

yj=1NΣk=0N1cke2πijkNy_j = \frac{1}{N} \Sigma_{k=0}^{N-1} c_k e^{\frac{2 \pi ijk}{N}}

The trick here is to notice that the inverse matrix is just another Vandermonde matrix with ωˉ\bar{\omega} instead of ω\omega.

Applications

Many applications come from the ability to encode a signal:

  1. Digital files can be shrunk by eliminating contributions from the least important waves in the combination.
  2. Comparing different sound files by comparing coefficients XkX_k of DFT.
  3. De-noising radio waves

Wavelet Transformation and the Uncertainty Principle

The problem with Fourier Transform is that it produces a periodic function over the whole space. What if we just want a localized wave? This is so called a “wavelet”.

A function ψ:RCL2(R\psi:\mathbb{R} \rightarrow \mathbb{C} \in L^2(\mathbb{R} (square integrable functions i.e. ψ2<\int_{-\infty}^{\infty} |\psi|^2 < \infty) is said to be an orthonormal wavelet if it can be used to define a Hilbert basis.

The Heisenberg Uncertainty Principle from Physics states: ΔEΔt4π\Delta E \Delta t \geq \frac{\hbar}{4\pi}. Translated to signal processing, that becomes ΔωΔt12\Delta \omega \Delta t \geq \frac{1}{2}. Intuitively, it means we cannot measure and clearly resolve both the frequency and the time to a very large degree.

References

Remnants of Rudin Chapter 10

Reminder of definitions to self:

  1. kk-surface in EE: a C\mathcal{C}' mapping from compact DRkERnD \subset \mathbb{R}^k \rightarrow E \subset \mathbb{R}^n.
  2. kk-form in EE: a function :k:k-surface R\rightarrow \mathbb{R}. Think of it as an integral of terms with kk differentials. For each infinitesimal volume in DRkD \subset \mathbb{R}^k, map it to the point in EE, take the value and weight it by the scaling of the volume.
  3. Affine kk-simplex: just a simplex but somewhere else in space. It's a kk-surface in Rn\mathbb{R}^n with parameter domain QkQ^k (the standard simplex)
  4. Affine kk-chain: a collection of simplexes

Motivation: why consider simplex? The partitioning of sets into simplexes is called triangulation, which plays important role in topology. Also, it allows for the introduction of boundaries.

Theorem [Stokes]: If Φ\Phi is a kk-chain of class C\mathcal{C}“ in open set VRmV \subset \mathbb{R}^m and if ω\omega is a (k1)(k-1)-form of class C\mathcal{C}' in VV, then Ψdω=Ψω\int_{\Psi} d\omega = \int_{\partial\Psi} \omega.

Remark: k=m=1k = m = 1 is Fundamental Theorem of Calculus. k=m=2k = m = 2 is Green's Theorem.

Proof (here we go):

Firstly, we perform some reductions. We know Ψ=ΣΦi\Psi = \Sigma \Phi_i and Ψ=ΣΦi\partial \Psi = \Sigma \partial \Phi_i. Suffices to show Φdω=Φω\int_{\Phi} d\omega = \int_{\partial \Phi} \omega for every oriented kk-simplex Φ\Phi of class C\mathcal{C}” in VV. This is because then Ψdω=Σi=1rΦidω=Σi=1rΦiω=Ψω\int_{\Psi} d\omega = \Sigma_{i=1}^{r} \int_{\Phi_i} d\omega = \Sigma_{i=1}^{r} \int_{\partial \Phi_i} \omega = \int_{\partial \Psi} \omega (the middle equality is due to the result).

Fix an oriented kk-simplex Φ\Phi and let σ=[0,e1,,ek]\sigma = [0, e_1, …, e_k] be the oriented affine kk-simplex with parameter domain QkQ^k which is defined by the identity mapping. (?)

Do some work on LHS: Tσdω=σ(dω)T=σd(ωT)\int_{T\sigma} d\omega = \int_{\sigma} (d\omega)_T = \int_{\sigma} d(\omega_T)

Do some work on RHS: (Tσ)ω=T(σ)ω=σd(ωT)\int_{\partial(T\sigma)} \omega = \int_{T(\partial \sigma)} \omega = \int_{\partial \sigma} d(\omega_T)

Suffices to show: σdλ=σλ\int_{\sigma} d\lambda = \int_{\partial \sigma} \lambda for the special simplex and for every (k1)(k-1)-form λ\lambda of class C\mathcal{C}' in EE.

Definition [exact]: Let ω\omega be a kk-form in open set ERnE \subset \mathbb{R}^n. If there is a (k1)(k-1)-form λE\lambda \in E s.t. ω=dλ\omega = d\lambda, then say ω\omega is exact in EE.

Definition [closed]: If ω\omega is of class C\mathcal{C}' and dω=0d\omega = 0, then ω\omega is closed.

Theorem: Suppose ERnE \subset \mathbb{R}^n convex and open, fC(E)f \in \mathcal{C}'(E), pZp \in \mathbb{Z} s.t. 1pn1 \leq p \leq n and fxj(x)=0\frac{\partial f}{\partial x_j}(x) = 0 for p<jn,xEp < j \leq n, x \in E. Then, FC(E)\exists F \in \mathcal{C}'(E) s.t. Fxp(x)=f(x)\frac{\partial F}{\partial x_p}(x) = f(x) and Fxj(x)=0\frac{\partial F}{\partial x_j}(x) = 0 for p<jn,xEp < j \leq n, x \in E.

Proof:

Theorem [Poincare's Lemma]: If ERnE \subset \mathbb{R}^n convex and open, k1k \geq 1, ω\omega is a kk-form of class C\mathcal{C}' in EE and dω=0d\omega = 0, then there is a (k1)(k-1)-form λ\lambda in EE s.t. ω=dλ\omega = d\lambda.

Closed forms are exact in convex sets.

Proof:

Theorem: Fix kk with 1kn1 \leq k \leq n. Let ERnE \subset \mathbb{R}^n be an open set in which every closed kk-form is exact. Let TT be 1-1 C\mathcal{C}“-mapping of EE onto open set URnU \subset \mathbb{R}^n whose inverse SS is of class C\mathcal{C}“. Then every closed kk-form in UU is exact in UU.

Proof:

Summary of Lebesgue Theory (HW4)

Riemann integral of a function ff is defined to be the upper Darboux integral (U(f)=inf{U(f,P)P partition of [a,b]}U(f) = \text{inf}\{U(f, P) | P \text{ partition of } [a, b]\}) or the lower Darboux integral (L(f)=sup{L(f,P)P partition of [a,b]}L(f) = \text{sup}\{L(f, P) | P \text{ partition of } [a, b]\}) (both should agree if the function is integrable). It relies heavily on the notion of a partition of an interval [a,b][a, b].

On the other hand, the al of a function ff is defined to be the measure of the undergraph of ff (i.e. m(U(f))m (\mathcal{U}(f))). It relies heavily on the concept of measure and measurability.

I feel that Lebesgue's integral is more intuitive in defining integral, because unlike Riemann integral, it does not rely heavily on an axis. In Riemann integral, we will run into problem when we take the sup\text{sup} or inf\text{inf} over an interval on an axis, which may result in the upper and lower integrals disagreeing, but for gral, we can fundamentally only care about the region under the graph, so integrals like f(x)dx=0\int f(x) dx = 0 where f(x)=1xQf(x) = 1 \forall x\in \mathbb{Q} and f(x)=0f(x) = 0 elsewhere can be defined.

I also feel that Lebesgue Theory is richer, so it feels more interesting.

Main Results of Lebesgue Measure Theory (HW6)

<to be updated after Spring Break, where I will do a formal compilation>

Littlewood's Three Principles (HW8)

  1. First Principle [Regularity]: Every measurable set is nearly a finite sum of boxes. “Nearly” here means “except for an ϵ\epsilon-set”.
  2. Second Principle [Lusin's Theorem]: A measurable function is nearly continuous i.e. given any ϵ>0\epsilon > 0, we can discard an ϵ\epsilon-set from its domain of definition and the result is a continuous function.
  3. Third Principle: Every convergent sequence of measurable functions is nearly uniformly convergent i.e. except for an ϵ\epsilon-set the convergence is uniform.
  4. Egoroff's Theorem: Almost everywhere convergence implies nearly uniform convergence.

Tao Chapter 8 Problems and Solutions

Here I complete the proofs of several lemmas which Tao left as exercises. (Some of the proofs are presented by Prof Peng in class.)

Ex 8.1.1 If ff and gg are simple functions, their images are finite. Let the image of ff be {c1,,cN}\{c_1, …, c_N\} and the image of gg be {d1,,dM}\{d_1, …, d_M\}. Then the image of f+gf+g is a subset of {ci+dj:1iN,1jM}\{c_i + d_j : 1 \leq i \leq N, 1 \leq j \leq M\} which is finite, hence f+gf+g is a simple function. Similarly, the set {cc1,,ccN}\{c \cdot c_1, …, c \cdot c_N\} is finite, so cfcf is also a simple function.

Ex 8.1.2 For each i{1,,N}i \in \{1, …, N\}, consider the open interval Ii=(ciϵ,ci+ϵ)I_i = (c_i - \epsilon, c_i + \epsilon). Then since ff is measurable, f1(Ii)f^{-1}(I_i) is measurable. Denote Ei=f1(Ii)E_i = f^{-1}(I_i). Then, note that EiEj=ϕE_i \cap E_j = \phi for iji \neq j, since if xEix \in E_i and xEjx \in E_j implies f(x)=cif(x) = c_i and f(x)=cjf(x) = c_j (contradiction!). Hence, the EiE_is are disjoint. Thus f=Σi=1NciχEif = \Sigma^{N}_{i=1} c_i \chi_{E_i} as desired.

Ex 8.1.3 Consider fn(x):=sup{j2n:jZ+{0},j2nmin(f(x),2n)}f_n(x) := \text{sup}\{\frac{j}{2^n} : j \in \mathbb{Z}^{+} \cup \{0\}, \frac{j}{2^n} \leq \text{min}(f(x), 2^n)\} (i.e. fn(x)f_n(x) is the greatest integer multiple of 2n2^{-n} that does not exceed both f(x)f(x) or 2n2^n. Clearly, fnf_n are nonnegative and increasing (since the restriction on multiple of 2n2^{-n} is relaxed as nn increases). fnf_n are also simple functions, s of possible values fn(x)f_n(x) can take on is a subset of {j2n:jZ{0},0j(2n)2}\{\frac{j}{2^n} : j \in \mathbb{Z} \cup \{0\}, 0 \leq j \leq (2^n)^2 \} For each point xΩx \in \Omega, NN\exists N \in \mathbb{N} s.t. 2N>f(x)2^N > f(x), hence n>Nn > N implies f(x)fn(x)=f(x)fn(x)2n|f(x) - f_n(x)| = f(x) - f_n(x) \leq 2^{-n}. Hence, limnf(x)fn(x)=0\text{lim}_{n→\infty} |f(x) - f_n(x)| = 0 by squeeze lemma. Thus, exists a sequence of nonnegative, increasing functions (fn)n(f_n)_n that converges pointwise to ff.

Ex 8.2.1 (a') Note that 00 as a function minorizes ff, hence 0Ωf0 \leq \int_{\Omega} f \leq \infty. If f(x)=0f(x) = 0 a.e., then for any simple function ss majorized by ff, if s=Σci1Eis = \Sigma c_i \mathbb{1}_{E_i} for some ci>0c_i > 0, then EiZE_i \subset Z where ZZ is a measure zero set (due to definition of almost everywhere). Hence EiE_i is a null set, and Ωf=0\int_{\Omega} f = 0. On the other hand, if Ωf=0\int_{\Omega} f = 0, then let EnE_n be the open pre-image of (1n,)(\frac{1}{n}, \infty), nNn \in \mathbb{N}. Then EnE_n is measurable (since ff is measurable function). Thus EnE_n must have measure 00, else, the simple function g=1n1Eng = \frac{1}{n}\mathbb{1}_{E_n} minorizes ff. Since this holds true n\forall n, the pre-image of (0,)(0, \infty) must have measure 00 (due to upward measure continuity), hence f=0f = 0 a.e.

(b') For any simple function ss that is majorized by ff, cscs is also majorized by cfcf and vice-versa. Hence, Ωcf=cΩf\int_{\Omega}cf = c\int_{\Omega}f.

(c') Any simple function ss that is majorized by ff is also majorized by gg. Hence ΩfΩg\int_{\Omega}f \leq \int_{\Omega}g simply because the latter is the sup\text{sup} of a bigger set.

(d') Let Z={xf(x)g(x)}Z = \{x | f(x) \neq g(x)\}, therefore ZZ is a zero set (by definition of a.e.). Let ZcZ^{c} denote ΩZ\Omega \setminus Z. Then s\forall s simple function, s=Zcs\int s = \int_{Z^c} s. Then f=Zcf=Zcg=g\int f = \int_{Z^c} f = \int_{Z^c} g = \int g.

(e') Similar to (c'), any simple function that is majorized by fXΩf_{X_{\Omega'}} is also majorized by ff. Hence Ωf=ΩfχΩΩf\int_{\Omega'} f = \int_{\Omega} f_{\chi_{\Omega'}} \leq \int_{\Omega} f.

Ex 8.2.2 By definition, for any simple function ss that minorizes ff and tt that minorizes gg, s+ts + t (which is also a simple function) minorizes f+gf + g. Thus, let ϵ>0\epsilon > 0, then let Ωf=A\int_{\Omega}f = A and Ωg=B\int_{\Omega}g = B, then s,t\exists s, t simple functions s.t. ss minorizes ff and tt minorizes gg and ΩsAϵ2\int_{\Omega}s \geq A - \frac{\epsilon}{2} and ΩtBϵ2\int_{\Omega}t \geq B - \frac{\epsilon}{2}. Thus, Ω(f+g)A+Bϵ\int_{\Omega}(f+g) \geq A + B - \epsilon ϵ>0\forall \epsilon>0. Thus, Ω(f+g)Ωf+Ωg\int_{\Omega}(f+g) \geq \int_{\Omega}f + \int_{\Omega}g.

Ex 8.2.3 Consider the partial sum sN=Σn=1Ngns_N = \Sigma_{n=1}^{N}g_n. Then sNs_N is nonnegative and (sn)n(s_n)_n is a sequence of increasing functions. By Monotone Convergence Theorem, Ωsupsn=supΩsn\int_{\Omega} \text{sup}s_n = \text{sup} \int_{\Omega} s_n. Note supsn=limnΣi=1ngi=Σi=1gi\text{sup}s_n = \text{lim}_{n → \infty} \Sigma_{i=1}^{n}g_i = \Sigma_{i=1}^{\infty}g_i. Also, Ωsn=ΩΣi=1ngi=Σi=1nΩgi\int_{\Omega} s_n = \int_{\Omega} \Sigma_{i=1}^{n}g_i = \Sigma_{i=1}^{n} \int_{\Omega} g_i (interchange of addition and integration), hence supΩsn=Σi=1Ωgi\text{sup} \int_{\Omega} s_n = \Sigma_{i=1}^{\infty} \int_{\Omega} g_i. Thus ΩΣi=1gi=Σi=1Ωgi\int_{\Omega} \Sigma_{i=1}^{\infty}g_i = \Sigma_{i=1}^{\infty} \int_{\Omega} g_i, as desired.

Ex 8.2.4 Consider the function Σn=1fn\Sigma_{n=1}^{\infty}f_n. The function s=χ[1,2)s = \chi_{[1, 2)} minorizes it, therefore, ΩΣn=1fn1\int_{\Omega} \Sigma_{n=1}^{\infty}f_n \geq 1. However, Rfn=0\int_{\mathbb{R}} f_n = 0 (actually, it might not even be defined in this case, since no simple function minorizes it). Thus, ΩΣn=1fnΣn=1Rfn\int_{\Omega} \Sigma_{n=1}^{\infty}f_n \neq \Sigma_{n=1}^{\infty} \int_{\mathbb{R}}f_n.

Ex 8.2.5 Firstly, note that the set E={xΩ:f(x)=+}E = \{x \in \Omega: f(x) = +\infty\} is measurable. Let EnE_n be the preimage of (n,)(n, \infty), then EnE_n are all measurable (since ff measurable). Thus, since EnEE_n \downarrow E, EE is measurable by downward measure continuity. Suppose otherwise, that the set EE has nonzero measure. Then consider the simple function sn=nχEs_n = n\cdot \chi_E. Then, nN\forall n \in \mathbb{N}, sns_n minorizes ff. Hence, ΩfEfEsn=nm(E)\int_{\Omega} f \geq \int_{E} f \geq \int_{E} s_n = n \cdot m(E). If m(E)m(E) nonzero, limnnm(E)=\text{lim}_{n → \infty} n \cdot m(E) = \infty. (contradiction!) Hence, ff must be almost finite a.e.

Ex 8.2.6 Consider the indicator functions χΩn\chi_{\Omega_n}.

Ex 8.2.7 HW5

Ex 8.2.8

Ex 8.2.9 HW5

Ex 8.2.10 HW5

Tao Chapter 8 Problems and Solutions II

Ex 8.3.1 Ωf=Ωf+ΩfΩf++Ωf=Ωf++Ωf=Ω(f++f)=Ωf|\int_{\Omega} f| = |\int_{\Omega} f^+ - \int_{\Omega} f^-| \leq |\int_{\Omega} f^+| + |\int_{\Omega} f^-| = \int_{\Omega} f^+ + \int_{\Omega} f^- = \int_{\Omega} (f^+ + f^-) = \int_{\Omega} |f| as desired.

The first equality is by definition. The first inequality is by the triangle inequality of R\mathbb{R}.

Ex 8.3.2 HW6

Ex 8.3.3 HW6

Measure Theory applied to Probability

One of the main reasons why I took this course is to gain a deeper understanding of probability, so here are some of my attempts in making this connection.

Preliminaries: The Borel sets on R\mathbb{R}, denoted as B(R)\mathcal{B}(\mathbb{R}) is the σ\sigma-field on R\mathbb{R} that is generated by all open intervals of the form (a,b)(a, b) with a<ba < b.

Let (Ω,F,P)(\Omega, \mathcal{F}, \mathbb{P}) be a measure space. F\mathcal{F} is a σ\sigma-algebra denoting a collection of subsets of Ω\Omega, i.e. F\mathcal{F} satisfies the following:

  1. ΩF\Omega \in \mathcal{F}
  2. AFAcFA \in \mathcal{F} \Rightarrow A^c \in \mathcal{F}
  3. (An)nFAnF(A_n)_n \in \mathcal{F} \Rightarrow \bigcup A_n \in \mathcal{F}

The sets in F\mathcal{F} are measurable sets. P:F[0,1]\mathbb{P}: \mathcal{F} \rightarrow [0, 1] is a probability measure which satisfies all the axioms of measure and the following:

  1. P[ϕ]=0\mathbb{P}[\phi] = 0
  2. If (An)nF(A_n)_n \in \mathcal{F} disjoint, then P[nAn]=ΣnP[An]\mathbb{P}[\bigcup_n A_n] = \Sigma_n \mathbb{P}[A_n]
  3. P[Ω]=1\mathbb{P}[\Omega] = 1

A random variable XX is a measurable function X:FRX: \mathcal{F} \rightarrow \mathbb{R}, i.e. X1(B)=ωΩ:X(ω)BFX^{-1}(B) = {\omega \in \Omega: X(\omega) \in B} \in \mathcal{F} BB(R)\forall B \in \mathcal{B}(\mathbb{R}).

Let (Xn)n(X_n)_n be a sequence of random variables. For ωΩ\omega \in \Omega, then the sequence of random variables acting on this ω\omega gives a sequence os: (Xn(ω))n(X_n(\omega))_n. Say (Xn)n(X_n)_n converges almost surely to XX if limnP[{ω:Xn(ω)=X(ω)}]=1\text{lim}_{n \leftarrow \infty} \mathbb{P}[\{\omega: X_n(\omega) = X(\omega)\}] = 1.

Borel-Cantelli Lemmas

https://www.colorado.edu/amath/sites/default/files/attached-files/almost_sure_conv.pdf

Additional Stuff

- Tao Section 3: Here I attempt to write a concise summary/key idea of Tao Section 3 without all the (why?)

Say a set AA is a coset of Q\mathbb{Q} if A=x+QA = x + \mathbb{Q} for some xRx \in \mathbb{R}.

Claim 1: x+Qx + \mathbb{Q} and y+Qy + \mathbb{Q} are either disjoint or entirely the same

  • Suppose they intersect at ww. Then w=x+q1=y+q2w = x + q_1 = y + q_2 for some q1,q2q_1, q_2. Then xy=q2q1x - y = q_2 - q_1, so yx+Qy \in x + \mathbb{Q}. Consequently, for any zx+Qz \in x + \mathbb{Q}, z=x+q3=y+q2q1+q3y+Qz = x + q_3 = y + q_2 - q_1 + q_3 \in y + \mathbb{Q}. Similarly for the opposite direction. Hence, we can conclude they are the same set if they have a nonempty intersection.

Claim 2: Every coset has a non-empty intersection with [0,1][0, 1]

  • In any interval [a,b],a<b[a, b], a < b, we can find a rational number. Hence, if A=x+QA = x + \mathbb{Q}, consider the rational number q[x,1x]q \in [-x, 1 - x]. Then, q+x[0,1]q + x \in [0, 1] as desired.

Consider the set of all cosets, and from each of them pick a representative xAx_A such that xA[0,1]x_A \in [0, 1] (by axiom of choice; there are an uncountably infinite of cosets, corresponding to an uncountably infinite number of xRx \in \mathbb{R}). Construct E={xA}E = \{x_A\}. Construct X=qQ[1,1](q+E)X = \bigcup_{q \in \mathbb{Q} \cap [-1, 1]} (q + E).

The key idea here is that, due to this construction of XX, [0,1]X[1,2][0, 1] \subset X \subset [-1, 2], hence we can bound the outer measure of XX by monotonicity (if it exists).

On one hand, any element xXx \in X can be expressed as x=q+ex = q + e where qQ[1,1]q \in \mathbb{Q} \cap [-1, 1] and eEe \in E, hence, x[1,2]x \in [-1, 2], hence m(X)<3m^*(X) < 3 by monotonicity.

On the other hand, for any x[0,1]x \in [0, 1], xx+Qx \in x + \mathbb{Q}, so yE[0,1]\exists y \in E \cap [0, 1] s.t. x=y+qx = y + q where qQq \in \mathbb{Q}. (in particular, this yy is the representative of the coset that xx belongs to) Then, since xy1|x - y| \leq 1, q[1,1]q \in [-1, 1], so xXx \in X. Hence, since [0,1]X[0, 1] \subset X, m(X)1m^*(X) \geq 1 by monotonicity.

This yields 1m(X)31 \leq m^*(X) \leq 3.

The nail in the coffin comes from the assumption that if m()m^*(\cdot) satisfies countable additivity, then since X=qQ[1,1](q+E)X = \bigcup_{q \in \mathbb{Q} \cap [-1, 1]} (q + E), and Q[1,1]\mathbb{Q} \cap [-1, 1] is countable (because Q\mathbb{Q} is countable, so is its intersection with another set) then m(X)=Σm(q+E)m^*(X) = \Sigma m^*(q + E). (*) But by the translational invariance property of outer measure, m(q+E)=m(E)m^*(q + E) = m^*(E) where qQ[1,1]q \in \mathbb{Q} \cap [-1, 1] is a subset of R\mathbb{R}. Hence, the RHS is summing the same thing m(E)m^*(E) many times. Thus, RHS takes on either the value 00 or \infty, corresponding to when m(E)=0m^*(E) = 0 or m(E)>0m^*(E) > 0 respectively. Either way, it does not fall in the bounds 1m(X)31 \leq m^*(X) \leq 3.

(*) Note that only here did we use Claim 1. It shows that for two rational numbers q1,q2[1,1]q_1, q_2 \in [-1, 1], the two sets q1+Eq_1 + E and q2+Eq_2 + E are disjoint. Suppose otherwise, then xA,xBE\exists x_A, x_B \in E s.t. q1+xA=q2+xBq_1 + x_A = q_2 + x_B, then xAxB=q2q1Qx_A - x_B = q_2 - q_1 \in \mathbb{Q} which implies xA,xBx_A, x_B belong to the same coset! This is impossible since only one representative is chosen from each coset by construction.

The failure of finite additivity follows directly (but non-trivially though, in my honest opinion). Since 1m(X)31 \leq m^*(X) \leq 3, by countable subadditivity X=qQ[1,1](q+E)Σm(q+E)=Σm(E)X = \bigcup_{q \in \mathbb{Q} \cap [-1, 1]} (q + E) \leq \Sigma m^*(q + E) = \Sigma m^*(E), we can eliminate the case where m(E)=0m^*(E) = 0, otherwise m(X)=0m^*(X) = 0.

Since we know m(X)m^*(X) now takes a finite value, say ϵ\epsilon, we can upset it by taking a finite number of rational numbers Q[1,1]\in \mathbb{Q} \cap [-1, 1], say a set JJ with 4/ϵ\lceil 4/\epsilon \rceil elements. Then, by countable subadditivity again, qJ(q+E)q[1,1](q+E)=X\bigcup_{q \in J}(q+E) \subset \bigcup_{q\in [-1,1]}(q+E) = X, so ΣqJm(q+E)m(X)\Sigma_{q \in J}m^*(q + E) \leq m^*(X). But if finite additivity works, ΣqJm(q+E)=Jm(E)(4/ϵ)ϵ=4>3m(X)\Sigma_{q \in J}m^*(q + E) = |J|m^*(E) \geq (4/\epsilon)\epsilon = 4 > 3 \geq m^*(X). (Contradiction!)

Toolbox (Lemmas, Tricks, Methods) in Analysis

  • ϵ\epsilon-room trick
  • ϵ2n\frac{\epsilon}{2^n} trick for Borel Cantelli lemma
  • [Heine-Borel] If EE is compact and EαUαE \subset \bigcup_\alpha U_\alpha for open UαU_\alpha, then EE is covered by a finite subset of the open sets.
  • [Princeton Measure Theory Theorem 1.4] Every open subset UU of Rd,d1\mathbb{R}^d, d \geq 1 can be written as a countable union of almost disjoint closed cubes.
  • Let E,FRdE, F \subset \mathbb{R}^d be such that dist(E,F)>0\text{dist}(E, F) > 0, where dist(E,F):=inf{xy:xE,yF}\text{dist}(E, F) := \text{inf}\{|x - y| : x \in E, y \in F\} is the distance between E,FE, F. Then m(EF)=m(E)+m(F)m^*(E \cup F) = m^*(E) + m^*(F)
  • [Tao 7.4.10] Every open set can be written as the countable or finite union of open boxes
  • [Princeton Measure Theory Theorem 1.4] Every open subset URdU \subset \mathbb{R}^d can be written as a countable union of almost disjoint open sets.

Proof (I followed the proof given in Princeton Measure Theory but filled in some in-between steps):

Let {Bi}\{B_i\} denote the set of boxes that are (0,1/2k)d(0, 1/2^k)^d but translated by some integer multiple of 1/2k1/2^k possibly in each of the dd coordinates. Then, for any point uUu \in U, since UU is open, r>0\exists r > 0 s.t. Br(u)UB_r(u) \subset U. We can always pick a power of 22 such that 12k<r2d\frac{1}{2^k} < \frac{r}{2\sqrt{d}}. Hence, uu is either contained in BiB_i for some ii or uu has a coordinate that is exactly an integer multiple of 12k\frac{1}{2^k}.

Let PP be the set of all planes such that one coordinate is a multiple of 12k\frac{1}{2^k}. This set is countable (since Q\mathbb{Q} is countable). We have proven before that we can cover a plane with a coun of open boxes with outer measure 00.

Just for completeness, for the plane x1=0x_1 = 0. The consider open boxes centered at 00, with each dimension other than x1x_1 being length 2k2^k and x1x_1 being dimension ϵ/(2k)d\epsilon/(2^k)^{d}. Then enumerating k=1,2,3,k = 1, 2, 3, … covers the whole plane, and the total volume of all open boxes that we used is ϵ\epsilon.

Hence, from the set of all {Bi}\{B_i\}, we choose the boxes that completely fits in the open set UU as described above, and denote it as the set BB. Then we have the following relationship: BUBPB \subset U \subset B \cup P. By monotonicity and subadditivity of outer measure, m(B)m(U)m(BP)m(B)+m(P)=m(B)m^*(B) \leq m^*(U) \leq m^*(B \cup P) \leq m^*(B) + m^*(P) = m^*(B). Hence m(U)=m(B)m^*(U) = m^*(B). This shows that every open set UU can be written as a countable union of almost disjoint open boxes.

  • [Pugh Lemma 6.17] Every open subset is a countable disjoint union of balls plus a zero set. (Prove by algorithm and previous lemma from Princeton Measure Theory)
  • [Princeton Measure Theory Lemma 3.1] If FF closed and KK compact and F,KF, K disjoint, then dist(F,K)>0\text{dist}(F, K) > 0.

Proof (I followed the proof given in Princeton Measure Theory but filled in some in-between steps):

FF is closed, so for any xKx \in K, δx\exists \delta_x s.t. d(x,F)>3δxd(x, F) > 3\delta_x. (Suppose otherwise that δx\delta_x does not exist, then for every nn, we can find fnFf_n \in F s.t. d(x,fn)<1nd(x, f_n) < \frac{1}{n}. Then since FF closed and by Bolzano Weierstrass, \exists subsequence fnf'_n s.t. limfn=xF\text{lim} f'_n = x \in F. Contradiction!)

Since xKB2δx(x)\bigcup_{x \in K} B_{2\delta_x}(x) covers KK and KK is compact, by Heine-Borel Theorem, exists a finite subcover i=1nB2δxi(xi)\bigcup_{i=1}^{n} B_{2\delta_{x_i}}(x_i).

Pick δ=minδxi\delta = \text{min} \delta_{x_i}. Then, for any xKx \in K and yFy \in F, then for any jj an index in the finite subcover, yxyxjxjx3δxj2δxj=δxjδ|y - x| \geq |y - x_j| - |x_j - x| \geq 3\cdot \delta_{x_j} - 2\cdot \delta_{x_j} = \delta_{x_j} \geq \delta. Hence, dist(F,K)δ>0\text{dist}(F, K) \geq \delta > 0.

  • [Tao Ex 7.2.3] For increasing sequence of measurable sets (Ai)(A_i) (i.e. AiAi+1A_i \subset A_{i+1}), we have m(j=1Aj)=limjm(Aj)m (\bigcup_{j=1}^{\infty}A_j) = \text{lim}_{j → \infty} m (A_j)
  • [Tao Ex 7.2.3] For decreasing sequence of measurable sets (Ai)(A_i) (i.e. AiAi+1A_i \supset A_{i+1}) with m(A1)<m(A_1) < \infty, we have m(j=1Aj)=limjm(Aj)m (\bigcap_{j=1}^{\infty}A_j) = \text{lim}_{j → \infty} m (A_j)
  • [Tao Theorem 2.1.4] Let f:XYf : X → Y be a continuous function from metric space XX to YY. Then ff is continuous at x0x_0 is equivalent to for every open set VYV \subset Y s.t. f(x0)Vf(x_0) \in V, UX\exists U \subset X open containing x0x_0 s.t. f(U)Vf(U) \subset V

Proof (I came up with this by myself; please forgive any errors)

Suppose ff is continuous at x0x_0, then for every sequence (xn)(x_n) that converges to x0x_0, we have (f(xn))nf(x0)(f(x_n))_n → f(x_0). Suppose otherwise, that for some open set VYV \subset Y s.t. f(x0)Vf(x_0) \in V but does not exist open UXU \subset X containing x0x_0 with the image f(U)Vf(U) \subset V. Since VV is open, suffices to consider a smaller set V=Br(f(x0))VV' = B_r(f(x_0)) \subset V for some rRr \in \mathbb{R}. Consider open sets Un=B1/n(x0)U_n = B_{1/n}(x_0) in XX with nNn \in \mathbb{N}. None of these balls map entirely into VV', so we can pick a point xnx'_n s.t. f(xn)Vf(x'_n) \notin V'. This gives us a sequence (xn)n(x'_n)_n. It is clear that (xn)nx0(x'_n)_n → x_0 since limn1n=0\text{lim}_{n→\infty}\frac{1}{n} = 0, hence by continuity of ff at x0x_0, limnf(xn)=f(x0)\text{lim}_{n→\infty}f(x'_n) = f(x_0). But this is impossible, since the value sequence all lies outside V=Br(f(x0))V' = B_r(f(x_0)), contradicting the assumption that ff is continuous at x0x_0.

Now let's consider the case: for every open set VYV \subset Y s.t. f(x0)Vf(x_0) \in V, UX\exists U \subset X open containing x0x_0 s.t. f(U)Vf(U) \subset V. Suppose otherwise that ff is not continuous at x0x_0. Then, it means that (xn)n\exists (x'_n)_n s.t. (xn)x0(x'_n) → x_0 but f(xn)f(x'_n) does not converge to f(x0)f(x_0). Then exists a subsequence of (xn)n(x'_n)_n, say (yn)n(y'_n)_n such that dY(f(x0),yn)=δ>0d_Y(f(x_0), y'_n) = \delta > 0 (for if otherwise, the sequence (f(xn))n(f(x'_n))_n converges to f(x0)f(x_0)). Consider the open ball Bδ/2(f(x0))B_{\delta/2}(f(x_0)). By assumption, UX\exists U \subset X containing x0x_0 such that f(U)Bδ/2(f(x0))f(U) \subset B_{\delta/2}(f(x_0)). This means none of the values yiy'_i falls in UU (for if otherwise their image falls in Bδ/2(f(x0))B_{\delta/2}(f(x_0)). Thus limyix0\text{lim}y'_i \neq x_0. Contradiction, since (yn)n(y'_n)_n is a subsequence of (xn)n(x'_n)_n which must converge to x0x_0 too.

  • [Tao Theorem 2.1.5] Let f:XYf : X → Y be a continuous function from metric space XX to YY. Then whenever VV is open (closed) set in YY, the pre-image f1(Y)f^{-1}(Y) is an open (closed) set in XX.

Resolved Thoughts

A metric space is sequentially compact if and only if it is compact.

Lemma If the metric space (X,d)(X, d) is compact and an open cover of XX is given, then δ>0\exists \delta > 0 s.t. every subset of XX with diameter less than δ\delta is contained in some member of its cover. (Wiki)

If subspace of a metric space, say KK, is compact/sequentially compact, then for any open cover {Ui}\{U_i\} of KK, δ>0\exists \delta > 0, s.t. for xK\forall x \in K, Bδ(x)UαB_\delta(x) \subset U_\alpha for some α\alpha. (Prof Peng's notes)

Key idea of proof: Prove by contradtion. Suppose otherwise, then nZ\forall n \in \mathbb{Z}, xk\exists x_k s.t. B1/nB_{1/n} is not contained within all open sets UiU_i. Consider this sequence, then apply triangle inequality at the limit point x=limkxkx = lim_{k → \infty} x_k.

- δ\delta is referred to as the of this cover. Intuitively, it's like a breathing space (LOL) for the compact set within the cover.

- Diameter is the supremum of the distance between pairs of points

If AA and BB are measurable, then A\BA \RM B is measurable.

Idea by @selenajli on Discord: BB is measurable \Rightarrow Rn\B\mathbb{R}^n \RM B is measurable (complement) \Rightarrow Rn\BA=AB\mathbb{R}^n \RM B \cap A = A - B is measurable (finite union)

✅ What are all the -morphisms? Are there intuitive pictures so that I can have a better idea? See definitions

Unresolved Thoughts

Banach-Tarski paradox (Tao 163) ✅ Lebesgue number (Pugh)

From David's non-measurable presentation rehearsal: why are these two definitions equivalent to the definition we know from Tao?

m(P)=inf{m(G)PG,G open}m^*(P) = \text{inf}\{m(G) | P \subset G, G \text{ open}\}

m(P)=sup{m(G)PG,G closed}m^*(P) = \text{sup}\{m(G) | P \supset G, G \text{ closed}\}

Lipschitz constant

Lipschitz condition

locally Lipschitz

Homeomorphism and diffeomorphism: are they connected? It seems they preserve measurability. ✅ A diffeomorphism T:UVT: U → V is a meseomorphism, i.e. preserves measurability. Prove by nn-dimensional mean value theorem (TODO!!!!) (Pugh Ex 6.23)

https://math.stackexchange.com/questions/541118/proving-that-sum-of-two-measurable-functions-is-measurable

https://adebray.github.io/lecture_notes/m381c_notes.pdf

Egoroff's Theorem (the not special case; special case was HW5 Tao 8.2.10)

Professor Wodzicki's notes on differential forms from Math 53 (!): https://math.berkeley.edu/~wodzicki/H185.S11/podrecznik/2forms.pdf

Brouer's fixed point theorem

Hairy Ball Theorem

math105-s22/s/jianzhi/start.txt · Last modified: 2022/04/27 21:47 by jianzhi