Note to Self: Gaussian Hypercontractivity

The purpose of this note is to describe the Gaussian hypercontractivity inequality. As an application, we’ll obtain a weaker version of the Hanson–Wright inequality.

The Noise Operator

We begin our discussion with the following question:

Let f : \real^n \to \real be a function. What happens to f, on average, if we perturb its inputs by a small amount of Gaussian noise?

Let’s be more specific about our noise model. Let x \in \real^n be an input to the function f and fix a parameter 0 \le \varrho \le 1 (think of \varrho as close to 1). We’ll define the noise corruption of x to be

(1)   \[\tilde{x}_\varrho = \varrho \cdot x + \sqrt{1-\varrho^2} \cdot g, \quad \text{where } g\sim \operatorname{Normal}(0,I). \]

Here, \operatorname{Normal}(0,I) is the standard multivariate Gaussian distribution. In our definition of \tilde{x}_\varrho, we both add Gaussian noise \sqrt{1-\varrho^2} \cdot g and shrink the vector x by a factor \varrho. In particular, we highlight two extreme cases:

  • No noise. If \varrho = 1, then there is no noise and \tilde{x}_\varrho = x.
  • All noise. If \varrho = 0, then there is all noise and \tilde{x}_\varrho = g. The influence of the original vector x has been washed away completely.

The noise corruption (1) immediately gives rise to the noise operator1The noise operator is often called the Hermite operator. The noise operator is related to the Ornstein–Uhlenbeck semigroup operator P_t by a change of variables, P_t = U_{\e^{-t}}. T_\varrho. Let f : \real^n \to \real be a function. The noise operator T_\varrho is defined to be:

(2)   \[(T_\varrho f)(x) = \expect[f(\tilde{x}_\varrho)] = \expect_{g\sim \operatorname{Normal}(0,I)}[f( \varrho \cdot x + \sqrt{1-\varrho^2}\cdot g)]. \]

The noise operator computes the average value of f when evaluated at the noisy input \tilde{x}_\varrho. Observe that the noise operator maps a function f : \real^n \to \real to another function T_\varrho f : \real^n \to \real. Going forward, we will write T_\varrho f(x) to denote (T_\varrho f)(x).

To understand how the noise operator acts on a function f, we can write the expectation in the definition (2) as an integral:

    \[T_\varrho f(x) = \int_{\real^d} f(\varrho x + y) \frac{1}{(2\pi (1-\varrho^2))^{d/2}}\e^{-\frac{|y|^2}{2(1-\varrho^2)}} \, \mathrm{d} y.\]

Here, |y| denotes the (Euclidean) length of y\in\real^d. We see that T_\varrho f is the convolution of f(\varrho x) with a Gaussian density. Thus, T_\varrho acts to smooth the function f.

See below for an illustration. The red solid curve is a function f, and the blue dashed curve is T_{0.95}f.

As we decrease \varrho from 1 to 0, the function T_\varrho f is smoothed more and more. When we finally reach \varrho = 0, T_\varrho f has been smoothed all the way into a constant.

Random Inputs

The noise operator converts a function f to another function T_\varrho f. We can evaluate these two functions at a Gaussian random vector x\sim \operatorname{Normal}(0,I), resulting in two random variables f(x) and T_\varrho f(x).

We can think of T_\varrho f(x) as a modification of the random variable f(x) where “a 1-\varrho^2 fraction of the variance of x has been averaged out”. We again highlight the two extreme cases:

  • No noise. If \varrho = 1, T_\varrho f(x) = f(x). None of the variance of x has been averaged out.
  • All noise. If \varrho=0,T_\varrho f(x) = \expect_{g\sim\operatorname{Normal}(0,I)}[f(g)] is a constant random variable. All of the variance of x has been averaged out.

Just as decreasing \varrho smoothes the function T_\varrho f until it reaches a constant function at \varrho = 0, decreasing \varrho makes the random variable T_\varrho f(x) more and more “well-behaved” until it becomes a constant random variable at \varrho = 0. This “well-behavingness” property of the noise operator is made precise by the Gaussian hypercontractivity theorem.

Moments and Tails

In order to describe the “well-behavingness” properties of the noise operator, we must answer the question:

How can we measure how well-behaved a random variable is?

There are many answers to this question. For this post, we will quantify the well-behavedness of a random variable by using the L_p norm.2Using norms is a common way of measuring the niceness of a function or random variable in applied math. For instance, we can use Sobolev norms or reproducing kernel Hilbert space norms to measure the smoothness of a function in approximation theory, as I’ve discussed before on this blog.

The L_p norm of a (\real-valued) random variable y is defined to be

(3)   \[\norm{y}_p \coloneqq \left( \expect[|y|^p] \right)^{1/p}.\]

The pth power of the L_p norm \norm{y}_p^p is sometimes known as the pth absolute moment of y.

The L_p norms of random variables control the tails of a random variable—that is, the probability that a random variable is large in magnitude. A random variables with small tails is typically thought of as a “nice” or “well-behaved” random variable. Random quantities with small tails are usually desirable in applications, as they are more predictable—unlikely to take large values.

The connection between tails and L_p norms can be derived as follows. First, write the tail probability \prob \{|y| \ge t\} for t > 0 using pth powers:

    \[\prob \{|y| \ge t\} = \prob\{ |y|^p \ge t^p \}.\]

Then, we apply Markov’s inequality, obtaining

(4)   \[\prob \{|y| \ge t\} = \prob \{ |y|^p \ge t^p \} \le \frac{\expect [|y|^p]}{t^p} = \frac{\norm{y}_p^p}{t^p}.\]

We conclude that a random variable with finite L_p norm (i.e., \norm{y}_p < +\infty) has tails that decay at at a rate 1/t^p or faster.

Gaussian Contractivity

Before we introduce the Gaussian hypercontractivity theorem, let’s establish a weaker property of the noise operator, contractivity.

Proposition 1 (Gaussian contractivity). Choose a noise level 0 < \varrho \le 1 and a power p\ge 1, and let x\sim \operatorname{Normal}(0,I) be a Gaussian random vector. Then T_\varrho contracts the L_p norm of f(x):

    \[\norm{T_\varrho f(x)}_p \le \norm{f(x)}_p.\]

This result shows that the noise operator makes the random variable T_\varrho f(x) no less nice than f(x) was.

Gaussian contractivity is easy to prove. Begin using the definition of the noise operator (2) and L_p norm (3):

    \[\norm{T_\varrho f(x)}_p^p = \expect_{x\sim \operatorname{Normal}(0,I)} \left[ \left|\expect_{g\sim \operatorname{Normal}(0,I)}[f(\varrho x + \sqrt{1-\varrho^2}\cdot g)]\right|^p\right]\]

Now, we can apply Jensen’s inequality to the convex function t\mapsto t^p, obtaining

    \[\norm{T_\varrho f(x)}_p^p \le \expect_{x,g\sim \operatorname{Normal}(0,I)} \left[ \left|f(\varrho x + \sqrt{1-\varrho^2}\cdot g)\right|^p\right].\]

Finally, realize that for the independent normal random vectorsx,g\sim \operatorname{Normal}(0,I), we have

    \[\varrho x + \sqrt{1-\varrho^2}\cdot g \sim \operatorname{Normal}(0,I).\]

Thus, \varrho x + \sqrt{1-\varrho^2}\cdot g has the same distribution as x. Thus, using x in place of \varrho x + \sqrt{1-\varrho^2}\cdot g, we obtain

    \[\norm{T_\varrho f(x)}_p^p \le \expect_{x\sim \operatorname{Normal}(0,I)} \left[ \left|f(x)\right|^p\right] = \norm{f(x)}_p^p.\]

Gaussian contractivity (Proposition 1) is proven.

Gaussian Hypercontractivity

The Gaussian contractivity theorem shows that T_\varrho f(x) is no less well-behaved than f(x) is. In fact, T_\varrho f(x) is more well-behaved than f is. This is the content of the Gaussian hypercontractivity theorem:

Theorem 2 (Gaussian hypercontractivity): Choose a noise level 0 < \varrho \le 1 and a power p\ge 1, and let x\sim \operatorname{Normal}(0,I) be a Gaussian random vector. Then

    \[\norm{T_\varrho f(x)}_{1+(p-1)/\varrho^2} \le \norm{f(x)}_p.\]

In particular, for p=2,

    \[\norm{T_\varrho f(x)}_{1+\varrho^{-2}} \le \norm{f(x)}_2.\]

We have highlighted the p=2 case because it is the most useful in practice.

This result shows that as we take \varrho smaller, the random variable T_\varrho f(x) becomes more and more well-behaved, with tails decreasing at a rate

    \[\prob \{ |T_\varrho f(x)| \ge t \} \le \frac{\norm{T_\varrho f(x)}_{1+(p-1)/\varrho^2}}{t^{1 + (p-1)/\varrho^2}} \le \frac{\norm{f(x)}_p}{t^{1 + (p-1)/\varrho^2}}.\]

The rate of tail decrease becomes faster and faster as \varrho becomes closer to zero.

We will prove the Gaussian hypercontractivity at the bottom of this post. For now, we will focus on applying this result.

Multilinear Polynomials

A multilinear polynomial f(x) = f(x_1,\ldots,x_d) is a multivariate polynomial in the variables x_1,\ldots,x_d in which none of the variables x_1,\ldots,x_d is raised to a power higher than one. So,

(5)   \[1+x_1x_2\]

is multilinear, but

    \[1+x_1+x_1x_2^\]

is not multilinear (since x_2 is squared).

For multilinear polynomials, we have the following very powerful corollary of Gaussian hypercontractivity:

Corollary 3 (Absolute moments of a multilinear polynomial of Gaussians). Let f be a multilinear polynomial of degree k. (That is, at most k variables x_{i_1}\cdots x_{i_k} occur in any monomial of f.) Then, for a Gaussian random vector x\sim \operatorname{Normal}(0,I) and for all q \ge 2,

    \[\norm{f(x)}_q \le (q-1)^{k/2} \norm{f(x)}_2.\]

Let’s prove this corollary. The first observation is that the noise operator has a particularly convenient form when applied to a multilinear polynomial. Let’s test it out on our example (5) from above. For

    \[f(x) = 1+x_1x_2,\]

we have

    \begin{align*}T_\varrho f(x) &= \expect_{g_1,g_2 \sim \operatorname{Normal}(0,1)} \left[1+ (\varrho x_1 + \sqrt{1-\varrho^2}\cdot g_1)(\varrho x_2 + \sqrt{1-\varrho^2}\cdot g_2)\right].\\&= 1 + \expect[\varrho x_1 + \sqrt{1-\varrho^2}\cdot g_1]\expect[\varrho x_2 + \sqrt{1-\varrho^2}\cdot g_2]\\&= 1+ (\varrho x_1)(\varrho x_2) \\&= f(\varrho x).\end{align*}

We see that the expectation applies to each variable separately, resulting in each x_i replaced by \varrho x_i. This trend holds in general:

Proposition 4 (noise operator on multilinear polynomials). For any multilinear polynomial f, T_\varrho f(x) = f(\varrho x).

We can use Proposition 4 to obtain bounds on the L_p norms of multilinear polynomials of a Gaussian random variable. Indeed, observe that

    \[f(x) = f(\varrho \cdot x/\varrho) = T_\varrho f(x/\varrho).\]

Thus, by Gaussian hypercontractivity, we have

    \[\norm{f(x)}_{1+\varrho^{-2}}=\norm{T_\varrho f(x/\varrho)}_{1+\varrho^{-2}} \le \norm{f(x/\varrho)}_2.\]

The final step of our argument will be to compute \norm{f(x/\varrho)}_2. Write f as

    \[f(x) = \sum_{i_1,\ldots,i_s} a_{i_1,\ldots,i_s} x_{i_1} \cdots x_{i_s}.\]

Since f is multilinear, i_j\ne i_\ell for j\ne \ell. Since f is degree-k, s\le k. The multilinear monomials x_{i_1}\cdots x_{i_s} are orthonormal with respect to the L_2 inner product:

    \[\expect[(x_{i_1}\cdots x_{i_s}) \cdot (x_{i_1'}\cdots x_{i_s'})] = \begin{cases} 0 &\text{if } \{i_1,\ldots,i_s\} \ne \{i_1',\ldots,i_{s'}\}, \\1, & \text{otherwise}.\end{cases}\]

(See if you can see why!) Thus, by the Pythagorean theorem, we have

    \[\norm{f(x)}_2^2 = \sum_{i_1,\ldots,i_s} a_{i_1,\ldots,i_s}^2.\]

Similarly, the coefficients of f(x/\varrho) are a_{i_1,\ldots,i_s} / \varrho. Thus,

    \[\norm{f(x/\varrho)}_2^2 = \sum_{i_1,\ldots,i_s} \varrho^{-2s} a_{i_1,\ldots,i_s}^2 \le \varrho^{-2k} \sum_{i_1,\ldots,i_s} a_{i_1,\ldots,i_s}^2 = \varrho^{-2k}\norm{f(x)}_2^2.\]

Thus, putting all of the ingredients together, we have

    \[\norm{f(x)}_{1+\varrho^{-2}}=\norm{T_\varrho f(x/\varrho)}_p \le \norm{f(x/\varrho)}_2 \le \varrho^{-k} \norm{f(x)}_2.\]

Setting q = 1+\varrho^{-2} (equivalently \varrho = 1/\sqrt{q-1}), Corollary 3 follows.

Hanson–Wright Inequality

To see the power of the machinery we have developed, let’s prove a version of the Hanson–Wright inequality.

Theorem 5 (suboptimal Hanson–Wright). Let A be a symmetric matrix with zero on its diagonal and x\sim \operatorname{Normal}(0,I) be a Gaussian random vector. Then

    \[\prob \{|x^\top A x| \ge t \} \le \exp\left(- \frac{t}{\sqrt{2}\mathrm{e}\norm{A}_{\rm F}} \right) \quad \text{for } t\ge \sqrt{2}\mathrm{e}\norm{A}_{\rm F}.\]

Hanson–Wright has all sorts of applications in computational mathematics and data science. One direct application is to obtain probabilistic error bounds for the error incurred by a stochastic trace estimation formulas.

This version of Hanson–Wright is not perfect. In particular, it does not capture the Bernstein-type tail behavior of the classical Hanson–Wright inequality

    \[\prob\{|x^\top Ax| \ge t\} \le 2\exp \left( -\frac{t^2}{4\norm{A}_{\rm F}^2+4\norm{A}t} \right).\]

But our suboptimal Hanson–Wright inequality is still pretty good, and it requires essentially no work to prove using the hypercontractivity machinery. The hypercontractivity technique also generalizes to settings where some of the proofs of Hanson–Wright fail, such as multilinear polynomials of degree higher than two.

Let’s prove our suboptimal Hanson–Wright inequality. Set f(x) = x^\top Ax. Since A has zero on its diagonal, f is a multilinear polynomial of degree two in the entries of x. The random variable f(x) is mean-zero, and a short calculation shows its L_2 norm is

    \[\norm{f(x)}_2 = \sqrt{\Var(f(x))} = \sqrt{2} \norm{A}_{\rm F}.\]

Thus, by Corollary 3,

(6)   \[\norm{f(x)}_q \le (q-1) \norm{f(x)}_2 \le \sqrt{2} q \norm{A}_{\rm F} \quad \text{for every } q\ge 2. \]

In fact, since the L_q norms are monotone, (6) holds for 1\le q\le 2 as well. Therefore, the standard tail bound for L_q norms (4) gives

(7)   \[\prob \{|x^\top A x| \ge t \} \le \frac{\norm{f(x)}_q^q}{t^q} \le \left( \frac{\sqrt{2}q\norm{A}_{\rm F}}{t} \right)^q\quad \text{for }q\ge 1.\]

Now, we must optimize the value of q to obtain the sharpest possible bound. To make this optimization more convenient, introduce a parameter

    \[\alpha \coloneqq \frac{\sqrt{2}q\norm{A}_{\rm F}}{t}.\]

In terms of the \alpha parameter, the bound (7) reads

    \[\prob \{|x^\top A x| \ge t \} \le \exp\left(- \frac{t}{\sqrt{2}\norm{A}_{\rm F}} \alpha \ln \frac{1}{\alpha} \right) \quad \text{for } t\ge \frac{\sqrt{2}\norm{A}_{\rm F}}{\alpha}.\]

The tail bound is minimized by taking \alpha = 1/\mathrm{e}, yielding the claimed result

    \[\prob \{|x^\top A x| \ge t \} \le \exp\left(- \frac{t}{\sqrt{2}\mathrm{e}\norm{A}_{\rm F}} \right) \quad \text{for } t\ge \sqrt{2}\mathrm{e}\norm{A}_{\rm F}.\]

Proof of Gaussian Hypercontractivity

Let’s prove the Gaussian hypercontractivity theorem. For simplicity, we will stick with the d = 1 case, but the higher-dimensional generalizations follow along similar lines. The key ingredient will be the Gaussian Jensen inequality, which made a prominent appearance in a previous blog post of mine. Here, we will only need the following version:

Theorem 6 (Gaussian Jensen). Let b : \real^2 \to \real be a twice differentiable function and let (x,\tilde{x}) \sim \operatorname{Normal}(0,\Sigma) be jointly Gaussian random variables with covariance matrix \Sigma. Then

(8)   \[b(\expect[h_1(x)], \expect[h_2(\tilde{x})]) \ge \expect [b(h_1(x),h_2(\tilde{x}))]\]

holds for all test functions h_1,h_2 : \real \to \real if, and only if,

(9)   \[\Sigma \circ \nabla^2 b \quad\text{is negative semidefinite on all of $\real^2$}.\]

Here, \circ denotes the entrywise product of matrices and \nabla^2 b : \real^2\to \real^{2\times 2} is the Hessian matrix of the function b.

To me, this proof of Gaussian hypercontractivity using Gaussian Jensen (adapted from Paata Ivanishvili‘s excellent post) is amazing. First, we reformulate the Gaussian hypercontractivity property a couple of times using some functional analysis tricks. Then we do a short calculation, invoke Gaussian Jensen, and the theorem is proved, almost as if by magic.

Part 1: Tricks

Let’s begin with “tricks” part of the argument.

Trick 1. To prove Gaussian hypercontractivity holds for all functions f, it is sufficient to prove for all nonnegative functions f\ge 0.

Indeed, suppose Gaussian hypercontractivity holds for all nonnegative functions f. Then, for any function f, apply Jensen’s inequality to conclude

    \begin{align*} T_\varrho |f|(x) &= \expect_{g\sim \operatorname{Normal}(0,1)} \left| f(\varrho x+\sqrt{1-\varrho^2}\cdot g)\right| \\&\ge \left| \expect_{g\sim \operatorname{Normal}(0,1)} f(\varrho x+\sqrt{1-\varrho^2}\cdot g)\right| \\&= |T_\varrho f(x)|.\end{align*}

Thus, assuming hypercontractivity holds for the nonnegative function |f|, we have

    \[\norm{T_\varrho f(x)}_{1+(p-1)/\varrho^2} \le \norm{T_\varrho |f|(x)}_{1+(p-1)/\varrho^2} \le \norm{|f|(x)}_p = \norm{f}_p.\]

Thus, the conclusion of the hypercontractivity theorem holds for f as well, and the Trick 1 is proven.

Trick 2. To prove Gaussian hypercontractivity for all f\ge 0, it is sufficient to prove the following “bilinearized” Gaussian hypercontractivity result:

    \[\expect[g(x) \cdot T_\varrho f(x)]\le \norm{g(x)}_{q'} \norm{f(x)}_p\]

holds for all g\ge 0 with \norm{g(x)}_{q'} < +\infty. Here, q'=q/(q-1) is the Hölder conjugate to q = 1+(p-1)/\varrho^2.

Indeed, this follows3This argument may be more clear to parse if we view f and g as functions on \real equipped with the standard Gaussian measure \gamma. This result is just duality for the L_q(\gamma) norm. from the dual characterization of the norm of T_\varrho f(x):

    \[\norm{T_\varrho f(x)}_q = \sup_{\substack{\norm{g(x)} < +\infty \\ g\ge 0}} \frac{\expect[g(x) \cdot T_\varrho f(x)]}{\norm{g(x)}_{q'}}.\]

Trick 2 is proven.

Trick 3. Let x,\tilde{x} be a pair of standard Gaussian random variables with correlation \rho. Then the bilinearized Gaussian hypercontractivity statement is equivalent to

    \[\expect[g(x) f(\tilde{x})]\le (\expect[(g(x)^{q'})])^{1/q'} (\expect[(f(\tilde{x})^{p})])^{1/p}.\]

Indeed, define \tilde{x} = \varrho x + \sqrt{1-\varrho^2} \cdot g for the random variable in the definition of the noise operator T_\varrho. The random variable \tilde{x} is standard Gaussian and has correlation \varrho with f, concluding the proof of Trick 3.

Finally, we apply a change of variables as our last trick:

Trick 4. Make the change of variables u \coloneqq f^p and v \coloneqq g^{q'}, yielding the final equivalent version of Gaussian hypercontractivity:

    \[\expect[v(x)^{1/q'} u(\tilde{x})^{1/p}]\le (\expect[v(x)])^{1/q'} (\expect[u(\tilde{x}))])^{1/p}\]

for all functions u and v (in the appropriate spaces).

Part 2: Calculation

We recognize this fourth equivalent version of Gaussian hypercontractivity as the conclusion (8) to Gaussian Jensen with

    \[b(u,v) = u^{1/p}v^{1/q'}\]

. Thus, to prove Gaussian hypercontractivity, we just need to check the hypothesis (9) of the Gaussian Jensen inequality (Theorem 6).

We now enter the calculation part of the proof. First, we compute the Hessian of b:

    \[\nabla^2 b(u,v) = u^{1/p}v^{1/q'}\cdot\begin{bmatrix} - \frac{1}{pp'} u^{-2} & \frac{1}{pq'} u^{-1}v^{-1} \\ \frac{1}{pq'} u^{-1}v^{-1} & - \frac{1}{qq'} v^{-2}\end{bmatrix}.\]

We have written p' for the Hölder conjugate to p. By Gaussian Jensen, to prove Gaussian hypercontractivity, it suffices to show that

    \[\nabla^2 b(u,v)\circ \twobytwo{1}{\varrho}{\varrho}{1}= u^{1/p}v^{1/q'}\cdot\begin{bmatrix} - \frac{1}{pp'} u^{-2} & \frac{\varrho}{pq'} u^{-1}v^{-1} \\ \frac{\varrho}{pq'} u^{-1}v^{-1} & - \frac{1}{qq'} v^{-2}\end{bmatrix}\]

is negative semidefinite for all u,v\ge 0. There are a few ways we can make our lives easier. Write this matrix as

    \[\nabla^2 b(u,v)\circ \twobytwo{1}{\varrho}{\varrho}{1}= u^{1/p}v^{1/q'}\cdot B^\top\begin{bmatrix} - \frac{p}{p'} & \varrho \\ \varrho & - \frac{q'}{q} \end{bmatrix}B \quad \text{for } B = \operatorname{diag}(p^{-1}u^{-1},(q')^{-1}v^{-1}).\]

Scaling A\mapsto \alpha \cdot A by nonnegative \alpha and conjugation A\mapsto B^\top A B both preserve negative semidefiniteness, so it is sufficient to prove

    \[H = \begin{bmatrix} - \frac{p}{p'} & \varrho \\ \varrho & - \frac{q'}{q} \end{bmatrix} \quad \text{is negative semidefinite}.\]

Since the diagonal entries of H are negative, at least one of H‘s eigenvalues is negative.4Indeed, by the Rayleigh–Ritz variational principle, the smallest eigenvalue of a symmetric matrix H is \lambda_{\rm min}(H) = \min_{\norm{x}=1} x^\top Hx. Taking x = e_i for i=1,2,\ldots to be each of the standard basis vectors, shows that the smallest eigenvalue of A is smaller than the smallest diagonal entry of H. Therefore, to prove H is negative semidefinite, we can prove that its determinant (= product of its eigenvalues) is nonnegative. We compute

    \[\det H = \frac{pq'}{p'q} - \varrho^2 .\]

Now, just plug in the values for p'=p/(p-1), q=1+(p-1)/\varrho^2, q'=q/(q-1):

    \[\det H = \frac{pq'}{p'q} - \varrho^2 = \frac{p-1}{q-1} - \varrho^2 = \frac{p-1}{(p-1)/\varrho^2} - \varrho^2 = 0.\]

Thus, \det H \ge 0. We conclude H is negative semidefinite, proving the Gaussian hypercontractivity theorem.

Leave a Reply

Your email address will not be published. Required fields are marked *