Skip to content
New issue

Have a question about this project? Sign up for a free GitHub account to open an issue and contact its maintainers and the community.

By clicking “Sign up for GitHub”, you agree to our terms of service and privacy statement. We’ll occasionally send you account related emails.

Already on GitHub? Sign in to your account

Blueprint for Plancharel's theorem #6

Merged
merged 4 commits into from
Apr 25, 2024
Merged
Show file tree
Hide file tree
Changes from all commits
Commits
File filter

Filter by extension

Filter by extension

Conversations
Failed to load comments.
Loading
Jump to
Jump to file
Failed to load files.
Loading
Diff view
Diff view
34 changes: 30 additions & 4 deletions BonnAnalysis/Plancherel.lean
Original file line number Diff line number Diff line change
Expand Up @@ -37,16 +37,42 @@ theorem snorm_fourierIntegralInv {f : V → E} (hf : Integrable f) (h2f : Memℒ
scoped[MeasureTheory] notation:25 α " →₁₂[" μ "] " E =>
((α →₁[μ] E) ⊓ (α →₂[μ] E) : AddSubgroup (α →ₘ[μ] E))


/- Note: `AddSubgroup.normedAddCommGroup` is almost this, but not quite. -/
instance : NormedAddCommGroup (V →₁₂[volume] E) :=
AddGroupNorm.toNormedAddCommGroup sorry
AddGroupNorm.toNormedAddCommGroup {
toFun := fun ⟨f,_⟩ ↦ ENNReal.toReal <| snorm f 2 volume
map_zero' := by simp [snorm_congr_ae AEEqFun.coeFn_zero, snorm_zero]
add_le' := fun ⟨f, _, hf⟩ ⟨g, _, hg⟩ ↦ ENNReal.toReal_le_add (by
simp [snorm_congr_ae (AEEqFun.coeFn_add f g),
snorm_add_le ((Lp.mem_Lp_iff_memℒp.1 hf).1) ((Lp.mem_Lp_iff_memℒp.1 hg).1)])
((Lp.mem_Lp_iff_snorm_lt_top.1 hf).ne) ((Lp.mem_Lp_iff_snorm_lt_top.1 hg).ne)
neg' := by simp[snorm_congr_ae (AEEqFun.coeFn_neg _)]
eq_zero_of_map_eq_zero' := by
intro ⟨f, _, hf⟩ h
simp [ENNReal.toReal_eq_zero_iff] at h
rcases h with h | h; swap
· absurd h; exact (Lp.mem_Lp_iff_snorm_lt_top.1 hf).ne
ext
apply ae_eq_trans <| (snorm_eq_zero_iff ((Lp.mem_Lp_iff_memℒp.1 hf).1) (by simp)).1 h
apply ae_eq_trans (Lp.coeFn_zero E 2 volume).symm; rfl
}

instance : NormedSpace ℝ (V →₁₂[volume] E) := sorry


/- The Fourier integral as a continuous linear map `L^1(V, E) ∩ L^2(V, E) → L^2(V, E)`. -/
def fourierIntegralL2OfL12 : (V →₁₂[volume] E) →L[ℝ] (V →₂[volume] E) :=
sorry
def fourierIntegralL2OfL12Fun : (V →₁₂[volume] E) → (V →₂[volume] E) :=
fun ⟨f,hf,hf2⟩ ↦ (memℒp_fourierIntegral (memℒp_one_iff_integrable.1 <|
Lp.mem_Lp_iff_memℒp.1 hf) (Lp.mem_Lp_iff_memℒp.1 hf2)).toLp <| 𝓕 f

def fourierIntegralL2OfL12 : (V →₁₂[volume] E) →L[ℝ] (V →₂[volume] E) := sorry
/-have : IsBoundedLinearMap ℝ fourierIntegralL2OfL12Fun := {
map_add := sorry
map_smul := sorry
bound := sorry
}
IsBoundedLinearMap.toContinuousLinearMap this-/



/- The Fourier integral as a continuous linear map `L^2(V, E) → L^2(V, E)`. -/
Expand Down
293 changes: 283 additions & 10 deletions blueprint/src/chapter/plancherel.tex
Original file line number Diff line number Diff line change
@@ -1,30 +1,303 @@
\chapter{Plancherel's Theorem}
\label{chap:plancherel}
\section{Basic Properties of the Fourier Transform}
In this section we record, mostly without proofs, basic statements about the Fourier transform on
$L^1$ functions. Most of these are already formalized in mathlib.

Let $(V,\mu),(W,\nu)$ be vector spaces over $\R$ with a $\sigma$-finite measure,
$E,F$ be normed spaces over $\C$ and let $L:V\times W\to \R$, $M:E\times F\to\C$
be bilinear maps.
\begin{definition}
\label{def:fourier-transform}
\uses{}
\lean{VectorFourier.fourierIntegral, Fourier.fourierIntegral, Real.fourierIntegral,
Real.fourierIntegralInv}
\leanok
Let $f\in L^1(V,E)$. Its \emph{Fourier transform} (w.r.t. $L$) is the function
$\mathcal Ff=\widehat f:W\to E$ given by
$$\mathcal Ff(w):=\widehat f(w):=\int_V e^{-2\pi i L(v,w)}f(v)\,d\mu.$$
The \emph{inverse Fourier transform} (w.r.t. $L$) is similarly defined as
$$\mathcal F^{-1}f(w)=\check f(w):=\int_Ve^{2\pi iL(v,w)}f(v)\,d\mu.$$
\end{definition}

\begin{lemma}
\label{lem:fourier-bounded}
\uses{def:fourier-transform}
\lean{VectorFourier.fourierIntegral_convergent_iff,
VectorFourier.norm_fourierIntegral_le_integral_norm}
\leanok
Let $f\in L^1(V,E)$. Then its Fourier transform $\widehat f$ is well-defined and bounded.
In particular, the Fourier transform defines a map $\mathcal F:L^1(V,E)\to L^\infty(V,E)$.
\end{lemma}

From now on assume that $V$ and $W$ are equipped with second-countable topologies such that
$L$ is continuous.
\begin{lemma}
\label{lem:fourier-cont}
\uses{def:fourier-transform}
\lean{VectorFourier.fourierIntegral_continuous}
\leanok
Let $f\in L^1(V,E)$. Then $\widehat f$ is continuous.
\end{lemma}

\begin{lemma}[Multiplication formula]
\label{lem:fourier-multiplication}
\uses{lem:fourier-cont}
\lean{VectorFourier.integral_bilin_fourierIntegral_eq_flip}
\leanok
Let $f,g\in L^1(V,E)$. Then
$$\int_WM(\widehat f(w),g(w))\,d\nu(w)=\int_VM(f(v),\widehat g(v))\,d\mu(v).$$
\end{lemma}

\begin{lemma}
\label{lem:fourier-prop}
\uses{def:fourier-transform}
\lean{}
% \leanok
Lef $f,g\in L^1(V,E)$, $t\in\R$ and $a,b\in\C$. The Fourier transform satisfies the following elementary properties:
\begin{enumerate}
\item[(i)] $\mathcal F(af+bg)=a\mathcal Ff+b\mathcal Fg$\hfill(Linearity)
\item[(ii)] $\mathcal F(f(x-t)) = e^{-2\pi i ty}\mathcal F f(y)$\hfill (Shifting)
\item[(iii)] $\mathcal F(f(tx)) = \frac1{|t|}\mathcal F f(\frac yt)$\hfill (Scaling)
\item[(iv)] $\mathcal F(\overline{f(x)}) = \overline{\mathcal Ff(-y)}$\hfill (Conjugation)
\item[(v)] Assume $V,E$ are inner product spaces. Then $\mathcal F(f\ast g) =\mathcal Ff\cdot\mathcal Fg$ \hfill (Convolution)
\end{enumerate}
\end{lemma}
From now on, let $V$ be a finite-dimensional inner product space.
We denote this product as ordinary multiplication, and the induced norm as $|\cdot|.$

We now study a family of functions which is useful for later proofs.
\begin{lemma}
\label{lem:fourier-gaussian}
\uses{def:fourier-transform, lem:fourier-prop}
\lean{fourierIntegral_gaussian_innerProductSpace'}
\leanok
Let $x\in V$ and $\delta>0$. Define the \emph{modulated Gaussian}
$$u_{x,\delta}(y):V\to\C,\quad y\mapsto e^{-\pi|y|^2}e^{2\pi i x y}.$$
Its Fourier transform (w.r.t. the inner product) is given by
$$\widehat{u_{x,\delta}}(z)=\delta^{-n/2}e^{-\pi|x-y|^2/\delta}=:K_\delta(x-z).$$
\end{lemma}
\begin{proof}
\leanok
By choosing an orthonormal basis, wlog we may assume $V=\R^n$.
First note $\widehat{u_{x,\delta}}(z-x)=\widehat{u_{0,\delta}}(z)$, so it is enough to consider $x=0$.
Next, $$\widehat{u_{0,\delta}}(z)=\int_{\R^n}e^{-\pi\delta|y|^2-2\pi iyz}\,dy=
\prod_{i=1}^n\int_{\R^n}e^{-\pi\delta y_i^2-2\pi iy_iz_i}dy_i\quad\text{and}\quad
K_\delta(-z)=\prod_{i=1}^n\delta^{-1/2}e^{-\pi y_i^2/\delta},$$ hence we may assume $n=1$.
The change of variables $w=\delta^{1/2}y+iz/\delta^{1/2}$ results in
$$\widehat{u_{0,\delta}}(z)=\int_{\R} e^{-\pi\delta y^2-2\pi i y z}\,dy=
\delta^{-1/2}e^{-\pi z^2/\delta}\int_{Im(w)=-z/\delta^{1/2}}e^{-\pi w^2}\,dw.$$
Contour integration along the rectangle with vertices $(\pm R,0),(\pm R,-iz/\delta^{1/2})$, together
with the bound $$\left|\int_{\pm R}^{\pm R-iz/\delta^{1/2}}e^{-\pi w^2}\right|\leq
\frac{|z|}{\delta^{1/2}}\sup_{w\in[R,R-iz/\delta^{1/2}]}|e^{-\pi w^2}|=
\frac{|z|}{\delta^{1/2}}e^{-(R^2+z^2/\delta)}\xrightarrow{R\to\infty}0$$
yields
$$\int_{Im(w)=-z/\delta^{1/2}}e^{-\pi w^2}\,dw=
\int_{\R} e^{-\pi w^2}\,dw=1,$$ finishing the proof.
\end{proof}

\begin{lemma}
\label{lem:weierstrass-kernel}
\uses{}
\lean{}
% \leanok
Let $K_\delta(v)=\delta^{-n/2}e^{-\pi|v|^2/\delta}$
as in Lemma \ref{lem:fourier-gaussian}. This is a
\emph{good kernel}, called the \emph{Weierstrass kernel}, satisfying
$$\int_{V}K_\delta(x)\,dx= 1\quad\text{and}\quad
\int_{|x|>\eta}K_\delta(x)\,dx\xrightarrow{\delta\to0}0\quad\text{for all }\eta>0.$$
Furthermore, it satisfies the stronger bounds
$$K_\delta(x)\leq \delta^{-n/2}\quad\text{and}\quad K_\delta(x)\leq B\delta^{1/2}|x|^{-n-1}$$ for
some constant $B$ independent of $\delta$.
\end{lemma}
\begin{proof}
% \leanok
By choosing an orthonormal basis, wlog we may assume $V=\R^n$. Then these are all straight-forward calculations:
$$\int_{\R^n}e^{-\pi|x|^2/\delta}\,dx=\delta^{n/2}\int_{\R^n}e^{-\pi|x|^2}\,dx=1.$$
$$\int_{|x|>\eta}\delta^{-n/2}e^{-\pi|x|^2/\delta}\,dx=\int_{|x|>\eta/\delta^{1/2}}e^{-\pi|x|^2}\xrightarrow{\delta\to0}0.$$
The first upper bound is trivial. For the second one, consider for $r,z>0$ the inequality
$$\Gamma(r+1)=\int_0^\infty e^{-y}y^r\,dy\geq\int_z^\infty e^{-y}y^r\,dy\geq z^r\int_z^\infty e^{-y}\,dy=z^re^{-z}.$$
Applied to $z=\pi|x|^2/\delta$ and $r=(n+1)/2$, this gives
$$|x|^{n+1}=\delta^{n+1/2}\frac{|x|^{n+1}}{\delta^{(n+1)/2}}\leq
\underbrace{\frac{\Gamma((n+3)/2)}{\pi^{(n+1)/2}}}_{=:B}\delta^{(n+1)/2}e^{\pi|x|^2/\delta},$$
which is equivalent to the second upper bound of the lemma.
\end{proof}

The following technical theorem is used in the proofs of both the inversion formula and Plancharel's theorem.
\begin{theorem}
\label{lem:plancherel}
\uses{}
\label{thm:kernel-approximation}
\uses{lem:weierstrass-kernel}
\lean{}
% \leanok
Let $f:V\to E$ be integrable. Let $K_\delta$ be the Weierstrass kernel from \ref{lem:weierstrass-kernel},
or indeed any family of functions satisfying the conditions of lemma \ref{lem:weierstrass-kernel}. Then
$$(K_\delta\ast f)(x):=\int_V K_\delta(y)f(x-y)\,d\mu(y)\xrightarrow{\delta\to0}f(x)$$ in the $L^1$-norm.
If $f$ is continuous, the convergence also holds pointwise.
\end{theorem}
\begin{proof}
% \leanok
Again we may assume $V=\R^n$. Consider the difference $$\Delta_\delta(x):=(K_\delta\ast f)(x)-f(x)=
\int_{\R^n}(f(x-y)-f(x))K_\delta(y)\,dy.$$ We prove $L^1$-convergence first:
Take $L^1$-norms and use Fubini's theorem to conclude
$$\|\Delta_\delta\|_1\leq\int_{\R^n}\|f(x-y)-f(x)\|_1K_\delta(y)\,dy.$$ For $\varepsilon>0$ find $\eta>0$
small enough so that $\|f(x-y)-f(x)\|_1<\varepsilon$ when $|y|\eta$. Thus
$$\|\Delta_\delta\|_1\leq\varepsilon+\int_{|y|>\eta}\|f(x-y)-f(x)\|K_\delta(y)\,dy\leq\varepsilon+2\|f\|_1\int_{\|y\|>\eta}K_\delta(y)\,dy.$$
By one of the properties in lemma \ref{lem:weierstrass-kernel}, we can choose $\delta$ small enough so that the second integral
is less than $\varepsilon$, which finishes the proof in this case.

Now assume that $f$ is continuous.
Let $d=\delta^{1/2}$ and shorten $g_\delta(x,y)=|f(x-y)-f(x)|K_\delta(y)$.Then
$$|\Delta_\delta(x)|\leq\int_{0<|y|<d}g_\delta(x,y)\,dy+\sum_{k\in\mathbb N}\int_{2^kd<|y|<2^{k+1}d}g_\delta(x,y)\,dy.$$
To bound these integrals, consider
$$\varphi(r)=\frac1{r^n}\int_{|y|<r}|f(x-y)-f(x)|\,dy.$$
It is easy to see that $\varphi$ is continuous, bounded, and approaches $0$ for $r\to0$, by continuity of $f$.
Now $$\int_{0<|y|<d}g_\delta(x,y)\,dy\overset{(\ast)}\leq d^{-n}\int_{0<|y|<d}|f(x-y)-f(x)|\,dy=\varphi(d)$$ and
$$\int_{2^kd<|y|<2^{k+1}d}g\,dy\overset{(\ast)}\leq
\frac{Bd}{(2^kd)^{n+1}}\int_{2^kd<|y|<2^{k+1}d}|f(x-y)-f(x)|\,dy\leq2^{n-k}B\varphi(2^{k+1}d),$$
where for the inequalities labeled $(\ast)$ we used the upper bounds from lemma \ref{lem:weierstrass-kernel}.
Together, we find
$$|\Delta_\delta(x)|\leq\varphi(d)+C\sum_{k\in\mathbb N}2^{-k}\varphi(2^{k+1}d)$$ for
$C=\frac{2^n\Gamma((n+3)/2)}{\pi^{(n+1)/2}}$. Say $\varphi$ is bounded by $M\in\R$ and let $\varepsilon>0$.
Take $N$ large enough such that $\sum_{k\geq N}2^{-k}<\varepsilon$. Choose $\delta$ small enough that
$A(2^kd)<\varepsilon/N$ for all $k<N$. Then
$$|\Delta_\delta(x)|\leq\varepsilon/N+(N-1)C\varepsilon/N+C\varepsilon M\leq \varepsilon C(M+1).$$
\end{proof}
\begin{remark}
One can drop the continuity assumption and still get pointwise convergence almost everywhere. The proof stays the same,
but one focuses on Lebesgue points of $f$. It takes slightly more work to argue that $\varphi$ behaves nicely,
but the rest of the proof stays the same.
\end{remark}

\begin{theorem}[Inversion formula]
\label{thm:fourier-inversion}
\uses{thm:kernel-approximation, lem:fourier-multiplication}
\lean{MeasureTheory.Integrable.fourier_inversion, Continuous.fourier_inversion}
Let $f:V\to E$ be integrable and continuous. Assume $\widehat f$ is integrable as well. Then
$$\mathcal F^{-1}\mathcal F f=f.$$
\leanok
\end{theorem}
\begin{proof}
\leanok
Apply the multiplication formula \ref{lem:fourier-multiplication} to $u_{x,\delta}$ and $f$, and conclude with
theorem \ref{thm:kernel-approximation}.
\end{proof}

\begin{remark}
Note that both assumptions are necessary, since $\mathcal F^{-1}\mathcal Ff$ is continuous, and
only defined if $\mathcal Ff$ is integrable.
\end{remark}

\section{\texorpdfstring{Plancharel's Theorem and the Fourier Transform on $L^2$}
{Plancharel's Theorem and the Fourier Transform on L2}}
Again let $V$ be a finite-dimensional inner product space over $\R$ and let $E$ be a normed space over $\C$.
%TODO: The proof needs convolutions of functions into E. Is there another proof? Or assume E has products?
\begin{theorem}
\label{thm:plancherel}
\uses{thm:kernel-approximation, lem:fourier-multiplication, lem:fourier-gaussian, lem:fourier-prop}
\lean{MeasureTheory.memℒp_fourierIntegral, MeasureTheory.snorm_fourierIntegral} % the full Lean name this corresponds to (can be a comma-separated list)
\leanok % the *statement* of the lemma is formalized
Let $V$ be a finite-dimensional inner product space over $\R$ and let $E$ be a normed space over $\C$. Suppose that $f : V \to C$ is in $L^1(V,E)\cap L^2(V,E)$ and let $\widehat{f}$ be the Fourier transform of $f$. Then $\widehat{f}\in L^2(V,E)$ and
\[\|\widehat{f}\|_{L^2} = \|f\|_{L^2}.\]
Suppose that $f : V \to E$ is in $L^1(V,E)\cap L^2(V,E)$ and let $\widehat{f}$ be the Fourier transform of $f$. Then $\widehat{f},\check{f}\in L^2(V,E)$ and
\[\|\widehat{f}\|_{L^2} = \|f\|_{L^2}=\|\check f\|_{L^2}.\]
\end{theorem}
\begin{proof}
% \uses{} % Put any results used in the proof but not in the statement here
% \leanok % uncomment if the lemma has been proven
Write proof here
Let $g(x)=\overline{f(-x)}$ and apply the multiplication formula \ref{lem:fourier-multiplication}
to $f\ast g$ and $u_{0,\delta}$: $$\int_V\widehat{f\ast g}\cdot u_{0,\delta}(x)\,dx=\int_V(f\ast g)(x)K_\delta(-x)\,dx
\overset{\delta\to0}\to(f\ast g)(0)=\int_Vf(x)\overline{f(x)}\,dx=\| f\|_2^2$$ by theorem \ref{thm:kernel-approximation}.
On the other hand, by lemma \ref{lem:fourier-prop} the left hand side simplifies to
$$\int_V|\widehat f(x)|^2e^{-\delta\pi|x|^2}\xrightarrow{\delta\to0}\|\widehat f\|_2^2$$ by dominated convergence.

Since $\check f(x)=\widehat f(-x)$, the corresponding statements for $\check f$ follow immediately from the ones for $\widehat f$.
\end{proof}

We now want to extend the Fourier transform to $L^2(V,E)$. For this, take a sequence of functions $(f_n)_n\subset L^1(V,E)\cap L^2(V,E)$
such that $f_n\xrightarrow[L^2]{}f$. Such sequences exist:
\begin{lemma}
\label{lem:L12-dense}
\uses{MeasureTheory.Memℒp.exists_hasCompactSupport_snorm_sub_le}
\lean{}
% \leanok
$L^1(V,E)\cap L^2(V,E)$ is dense in $L^2(V,E)$.
\end{lemma}
\begin{proof}
%\leanok
It is well-known that the space of compactly supported continuous functions is dense in every $L^p(V,E)$.
Since those are contained in $L^1(V,E)\cap L^2(V,E)$, the claim follows.
\end{proof}

Let $f\in L^2(V,E)$. Plancharel's theorem lets us now approximate a potential $\widehat f$:
\begin{lemma}
\label{lem:fourier12-cauchy}
\uses{lem:L12-dense, thm:plancharel, lem:fourier-prop}
\lean{}
% \leanok
Let $f\in L^2(V,E)$ and $(f_n)_n\subset L^1(V,E)\cap L^2(V,E)$ a sequence with $f_n\xrightarrow[L^2]{}f$. Then $(\widehat f_n)_n$ is a Cauchy sequence,
hence converges in $L^2(V,E)$.
\end{lemma}
\begin{proof}
% \leanok
$$\|\widehat f_n-\widehat f_m\|_2=\|\widehat{f_n-f_m}\|_2\overset{\text{Plancharel}}=\|f_n-f_m\|_2$$ goes to $0$ for $n,m$ large, as $(f_n)_n$ is
convergent, hence Cauchy. Since $L^2(V,E)$ is complete, $(\widehat f_n)_n$ converges.
\end{proof}
\begin{definition}
\label{def:fourier-L2}
\uses{lem:fourier12-cauchy}
\lean{}
% \leanok
Let $f\in L^2(V,E)$ and take a sequence $(f_n)_n\subset L^1(V,E)\cap L^2(V,E)$ with $f_n\xrightarrow[L^2]{}f$. Set
$$\mathcal Ff:=\widehat f:=\lim_{n\to\infty}\widehat{f_n},$$ the limit taken in the $L^2$-sense.
\end{definition}
\begin{lemma}
\label{lem:fourier2-welldef}
\uses{def:fourier-L2, lem:fourier12-cauchy, thm:plancharel}
\lean{}
% \leanok
This is well-defined: By \ref{lem:fourier12-cauchy}, the limit exists. Further it does not depend
on the choice of sequence $(f_n)_n$. If $f\in L^1(V,E)\cap L^2(V,E)$, this definition agrees with
the Fourier transform on $L^1(V,E)$.
\end{lemma}
\begin{proof}
% \leanok
Let $(g_n)_n$ be another sequence approximating $f$. Then
$$\|\widehat f_n-\widehat g_n\|_2=\|f_n-g_n\|\leq\|f_n-f\|+\|g_n-g\|\to0.$$
If $f\in L^1(V,E)\cap L^2(V,E)$, we can choose the constant sequence $(f_n)_n=(f)_n$.
\end{proof}

\begin{definition}
\label{def:invFourier-L2}
\uses{lem:fourier12-cauchy, thm:plancharel}
\lean{}
% \leanok
Define analogously $\mathcal F^{-1}f:=\check f:=\lim_{n\to\infty}\check f_n$, if $f_n\xrightarrow[L^2]{}f\in L^2(V,E)$
with $(f_n)_n\subset L^1(V,E)\cap L^2(V,E)$. By the same arguments as above, this is well-defined.
\end{definition}

\begin{corollary}
\label{thm:fourier2-properties}
\uses{def:fourier-L2, def:invFourier-L2, thm:plancharel, thm:fourier-inversion, lem:fourier-prop}
\lean{}
% \leanok
Plancharel's Theorem, the inversion formula, and the properties of lemma \ref{lem:fourier-prop} hold
for the Fourier transform on $L^2(V,E)$ as well.
\end{corollary}
\begin{proof}
% \leanok
All of these follow immediately from the definition and the oberservation, that all operations (norms, sums, conjugation, $\ldots$) are continuous.
For example, let $f\in L^2(V,E)$ and take an approximating sequence $(f_n)_n$ as before. Then
$$\|\widehat f\|_2=\|\lim_{n\to\infty}\widehat f_n\|_2=\|\lim_{n\to\infty}\|\widehat f_n\|_2=\lim_{n\to\infty}\|f_n\|_2
=\|\lim_{n\to\infty}f_n\|_2=\|f\|_2.$$
\end{proof}

\begin{corollary}
\label{lem:fourier-is-l2-linear}
\uses{lem:plancherel} % the (list of) LaTeX labels of the lemmas that are used for this result
\label{thm:fourier-is-l2-linear}
\uses{thm:fourier2-properties} % the (list of) LaTeX labels of the lemmas that are used for this result
\lean{MeasureTheory.fourierIntegralL2} % the full Lean name this corresponds to (can be a comma-separated list)
\leanok % the *statement* of the lemma is formalized
Let $V$ be a finite-dimensional inner product space over $\R$ and let $E$ be a normed space over $\C$. The Fourier transform induces a continuous linear map $L^2(V,E) \to L^2(V,E)$.
The Fourier transform induces a continuous linear map $L^2(V,E) \to L^2(V,E)$.
\end{corollary}
\begin{proof}
% \uses{} % Put any results used in the proof but not in the statement here
% \leanok % uncomment if the lemma has been proven
Write proof here
This follows immediately from \ref{thm:fourier2-properties}: Linearity from the $L^2$-version of
\ref{lem:fourier-prop}, and continuity and well-definedness from the $L^2$-version of Plancharel's theorem.
\end{proof}
Loading