Mathematics of Deep Learning: Lecture 3 – More on depth separation.

Transcribed by Paxton Turner (edited by Asad Lodhia, Elchanan Mossel and Matthew Brennan)

In the Boolean circuits setup, there is a separation result based on depth of the following form:

(Rossman, Servedio, Tan): There exists a function \(f: \{0,1\}^n \to \mathbb{R} \) that can be computed by a linear size depth \( d \) formula using AND, OR, and NOT gates such that any depth \(d – 1\) circuit that agrees with \(f\) on \(1/2 + o(1) \) inputs is of size \(\exp(n^{\Omega(1/d)}) \)

Question: Can we prove a similar result for deep nets?

Last time we discussed the following separation result.

(Eldan, Shamir): There exists a probability measure on \( \mathbb{R}^d \) and \(g: \mathbb{R}^d \to [-2,2] \) which is:

  • supported on \( \{x: ||x||^2 \leq Cd \} \) and expressible by a 3-layer network of width \( \text{poly}(d) \).
  • for all \( f\) supported by a 2-layer network of width \( \exp(\epsilon d) \) it holds that
    \[
    \mathbb{E}_\mu[ |f – g|^2 ] \geq \epsilon.
    \]

Today we will discuss a more elegant treatment of the 2 vs. 3 separation result which yields slightly stronger results.

Danielly’s Model

Danielly’s Model is formed as follows. Let \( x, x’ \) chosen uniformly (i.e., with respect to the Haar measure) from \(S^{d-1}\) and let \(f(x, x’) = g(\langle x, x’ \rangle) \). The nets of depth \(\ell\) are of the form:
\begin{eqnarray*}
\ell = 2 \qquad &w_2^{\text{T}} \sigma( W_1 x + W_2 x’ + b_1) + b_2 \\
\ell = 3 \qquad &w_3^{\text{T}} \sigma (W_2 \sigma(W_1 x + W_2 x’ + b_1) + b_2) + b_3 \\
\vdots \qquad &\vdots
\end{eqnarray*}
on the input \((x, x’)\). Here, we prove the following theorem.

If \( g:[-1,1] \to [-1,1] \) is \(L\)-Lipschitz, then there exists a 3-layer representation \(F\) of \(f\) with width \( O(d^2 L / \epsilon) \) and weights bounded by \(O(1 + L) \) and \( ||f – F||_{\infty} \leq \epsilon \).

This illustrates the “magic of 3.” Intuitively, we see the power of quadratic models over linear ones, at least for this specific context.

We begin with the following lemma which is of the flavor of universality of depth 1 nets:

If \(f: [-R, R]\) is \(L\)-Lipschitz for all \( \epsilon > 0 \), then \(f\) can be approximated with \(\beta_i \leq R, \alpha_i \leq 2L\), and \(\gamma_i \in \{-1, 1\} \) as follows:
\[
\left\|f(x) – f(0) – \sum_{i=1}^m \alpha_i \sigma(\gamma_i x – \beta_i) \right\|_{\infty} \leq \epsilon, \, m \leq \frac{2RL}{\epsilon}
\]

Proof of Lemma (Sketch): Repartition the interval, and note that the Lipschitz condition implies the graph doesn’t deviate too far from a straight line.

Proof of the Theorem: We will prove the theorem in the following section and then show a corresponding lower bound.

Upper bound

We first show that the function can be approximated well by a depth 3 net:
\[
\langle x, x’ \rangle = \Bigg\| \frac{x+x’}{2} \Bigg\|_2^2 – \frac{1}{2} = \sum_{i=1}^d \left( \frac{x + x’}{2} \right)^2,
\]

which implies that a 2-layer network approximates \( \langle x, x’ \rangle \) well. We can also compute \( \sigma( \langle x, x’ \rangle) \) as a linear combination of the previous layers. This in combination with the lemma implies the desired result.

Lower bound

We want to show that a depth two net that approximates the function well is
very wide. For the analysis it will be useful to answer the following question:

Question: What is the distribution of \( \langle x, x’ \rangle \)?

Approximation: Rotation invariance implies that

\[
D(\langle x, x’ \rangle) = D'( \langle x, r \rangle ) = D( \langle x, e_1 \rangle) = D(x_1)
\]
Note that this is also true if \( x’ \) is deterministic and \( x \) is random.

The individual components \(\langle x, e_k \rangle\) of a uniform random vector in \( x \in S^{d-1}\) approaches a Gaussian as \(d \to \infty \). Indeed,
\begin{eqnarray*}
d \mu_d(x) &= \text{Vol}\left( (1 – x^2)^{1/2} S^{d-2} \right) \\
&= \frac{\Gamma(d/2)}{\sqrt{\pi} \Gamma((d-1)/2)} \left(1 – x^2 \right)^{\frac{d-3}{2}}
\end{eqnarray*}
which is a t-statistic distribution (and as \(d \to \infty\) converges to a gaussian). The same reasoning applies also if one of the vectors x, x’ is fixed. Therefore, if \(g = \psi ( \langle v, x \rangle, \langle v’, x’ \rangle ) \), then
\[
||g||_2^2 = ||\psi||_{L^2(\mu_d \times \mu_d)}^2,
\]
and if \(f = \phi( \langle x, x’ \rangle ) \), then
\[
||f||_2^2 = ||\phi||_{L^2(\mu_d)}^2 = \int_{S^{d-1} \times S^{d-1}} f^2(x, x’) \, d(m \times m),
\]
where \(d(m \times m)\) indicates the Haar measure.

Orthogonal Polynomials: Define \( P_0 = 1, P_1 = x,\) and
\[P_n = \frac{2n + d – 4}{n + d – 3} \cdot P_{n-1}(x) – \frac{n – 1}{n + d – 3} \cdot P_{n-2}(x)\]
We observe that
\[
||P_n||_2^2 = 1/N_{n,d} \quad \text{where} \quad N_{n,d} = \frac{(2n + d – 2) (n + d – 3)!}{n! (d – 2)!}
\]
and that \( ||P_n||_{\infty} = 1\). Now define
\[
h_n(x, x’) = \frac{P_n(\langle x, x’ \rangle)}{\sqrt{N_{n,d}}} \quad \text{and} \quad L_n^x(x) = h_n(x,x).
\]
The key fact we use is that
\[
\langle L_i^v, L_j^v \rangle = \mathbb{E}[ L_i^v(x) L_j^{v’}(x)] = \mathbb{E}[h_i(\langle v, x \rangle) h_j(\langle v’, x \rangle) ] = \delta_{i,j} P_i(\langle v, v’ \rangle),
\]
which implies
\[
\mathbb{E}[h_n(x, x’) L_i^v(x) L_j^{v’}(x)] = 1(i = j = n) N_{n_1}^{-1/2} P_n(\langle v, v’ \rangle).
\]
To flesh out the above more carefully,
\begin{align*}
\mathbb{E}[h_n(x,x’) L_i^v(x) L_j^{v’}(x’)] &= \mathbb{E}_x\left[L_i^v(x) \mathbb{E}_{x’}[h_n(x,x’) L_j^{v’}(x’)]\right] \\
&= \mathbb{E}_x[L_i^v \delta_{j = n} P_n(v’, x)] \\
&= \mathbb{E}_x\left[L_i^v(x) \frac{L_j^{v’}(x)}{\sqrt{N_{n,d}}}\right]
\end{align*}

Lower Bound Idea: We now outline how to obtain the desired lower bound. Expand \(g(x, x’) = \sum \alpha_i h_i(x,x’) \) and choose \(f\) that is “noise sensitive”, i.e., has a lot of mass at higher Fourier levels. For the rest of the argument assume that \( f = h_n \) for a large \(n\) (though \( h_n \) is not a Lip function, but we can find a Lip function that has a lot of weight on high levels).

We want \( ||g – \sum_{j=1}^m g_j ||_2^2 \) to be large when \(g_j = b_j \sigma( \langle v_j, x \rangle) \geq c \) unless either \(b_j\) is large. Otherwise, there exists \(j\) such that \( \langle g, g_j \rangle \geq \frac{c}{m} \). Note that if \(h_n\) is close to \(\sum_{j = 1}^m g_j \), then we need to have at least one \(j\) where \( \langle g_j, f \rangle \geq \frac{||f||_2^2}{4m}. \)

We now compute \( \langle g_j , f \rangle \). Observe that
\[
g_j = \sum \beta_{k, \ell} L_k^v(x) L_{\ell}^{v’}(x’)
\]
\[
g_j(x,x’) = \psi(\langle v, x \rangle, \langle v’, x’ \rangle)
\]
\[
\sum \beta_{k, \ell}^2 = ||g_j||_2^2.
\]
Therefore,
\[
|\langle g_j, h_n \rangle | = N_{n,d}^{-1/2} |\beta_{n,n} P_n( \langle v, v’ \rangle )| \leq N_{n,d}^{-1/2} |\beta_{n,n}|
\]
so \(m\) is very large or \(||g_j^2||_2^2\) is very large. Informally, we get a bad approximation of \(h_n(\langle x, x’ \rangle) \) unless


  • Width \(\geq N_{n,d}^{\Omega(1)}\) or
  • Max-weight \(\geq N_{n,d}^{\Omega(1)} \)

Open Questions: 1.Can it really be that large weights solve the approximation problem? 2. Instead of working with the Haar measure, can this computation be performed in Gaussian space directly? We lose rotation invariance but it is a bit strange we work with measure that are in some sense getting close to Gaussian measure and cannot prove it directly in Gaussian space.

Kane-Williams Model

Threshold gates are given by

\[
\mathbf{1}\left\{\sum w_i x_i \geq t\right\}.
\]

There exists a function \( A_n: \{0,1\}^n \to \{0,1\} \) that can be computed

  • Using a 3-layer threshold circuit with \(O(n)\) gates,
  • but requires at least \( \Omega(n^{3/2}) \) gates to be computed correctly on \(.5 + \epsilon \) of inputs by a 2-layer circuit.

The basic unit threshold is
\[
f(x) = \text{sgn}(w_0 + \sum_{i = 1}^n w_i x_i).
\]

We assume a distribution that is uniform over \(\{0,1\}^n\).

Meta-Question: Can we use reduction between models (eg, the previous result) to get better separation in the binary case?

Partial answer: Due to complexity-theoretic issues, better than \( \text{poly}(n) \) isn’t possible.

Question: Can we replace new activations into the previous result (or this one)?

Answer: Isn’t written down, but should be straightforward.

First define the function \(A_n = A_{2^{k+1}}\).
\[
M(x_1, \ldots, x_{2^k}, a_1, \ldots, a_k) = x_{a_1} \ldots x_{a_k}
\]
is a Mux function generalization. Next,
\[
A_n(x_1, \ldots, x_{2^k}, a_1, \ldots, a_{2^k}) = M(x_1, \ldots, x_{2^k}, \oplus_{i=1}^{2^k/k} a_i, \oplus_{i=2^k/k + 1}^{2 \cdot 2^k/k} , \ldots),
\]
where \(\oplus_{i=1}^{2^k/k} a_i\) indicates the parity of the slice of the string.

Claim 1: \(M\) can be computed depth 2 using \(O(n \log n)\) gates.

Claim 2: Parity can be computed using threshold gates.

Proof of Claim 2: Let

\[
\sum x_i \geq 1 \to b_1
\]

\[
\vdots
\]

\[
\sum x_i \geq n \to b_n.
\]

It can be shown that there exist weights \( w_1, \ldots w_n\) such that

\[
\text{sgn}(\sum w_i b_i) = \oplus x_i.
\]

In fact, this can be done with \(1 \leq w \leq \text{poly}(k)\), though we don’t show this here.

Lower bound ingredients: We now move onto the lower bound.

Lemma: For all but \( \epsilon \) of \(n\)-bit Boolean functions \(f\), there is no depth 2 network of size less than \( o(\epsilon^2 2^n/n) \) that agrees with \(f\) on more than \(.5 + \epsilon\) of the inputs.

Proof sketch of lemma: The number of threshold functions is less than \(2^{O(n^2)}\), while the total number of functions is \(2^{2^n}\). That is, there aren’t too many threshold functions in general. Also, for most Boolean functions \(f\) on \(\{0,1\}^n\), a 2-layer network of size \(O(\frac{\xi 2^n}{n^2})\) does not approximate \(f\) well.

Returning to the proof of the lemma, the idea is to generate a random function, allowing us to use the lemma’s statement that most functions do not approximate \(f\) well. The construction is:


  • Pick \(x\) randomly.
  • In each block of \(a_i\), fix all coordinates but one randomly.
  • This generates a random function on \(k\) variables.

Question: How does a 2-layer look like after fixing randomly almost all bits?

Intuitive answer: Once enough bits are fixed, most gates become constant and die. Then we’re left with a smaller network which can’t approximate \(f\) well.

Main Lemma: A random restriction of a threshold gate will result in a constant function with probability at least
\[
1 – O(\log n / n^{1/2}).
\]
Thus, the number of gates that remain is \( O(n/\log^2(n)) \), which is not enough to represent a random function.

PAC Learning

We close this lecture with the definition of PAC-learning . We have a space \(X_n\) and are interested in functions \( f: X_n \to \{0,1\} \). Let \( \mathcal{C} \) be the class of such functions (eg, 2-layer networks).

The class \(\mathcal{C}\) is PAC-learnable if for all \( \epsilon, \delta \) and for all distributions \(D\) on \(X_n\), there exists an algorithm \(A\) such that given \(\text{poly}(n, 1/\epsilon, 1/\delta)\) samples
\[(x_i, y_i); \, x \sim D ; \, y_i = f(x_i); \, f \in \mathcal{C},
\]
with probability larger than \(1 – \delta\), \(A\) returns a function \(h: X_n \to \{0,1\}\) such that
\[
\mathbb{P}[h(x) \neq f(x)] \leq \epsilon
\]
and \(A\) runs in \(\text{poly}(n, 1/\epsilon, 1/\delta)\).

Here, \(\delta\) is interpreted as a global parameter , and \(\epsilon\) is the probabilistic error .

Mathematics of Deep Learning: Lecture 2 – Depth Separation.

Transcribed by Julien Edward Clancy (edited by Asad Lodhia, Elchanan Mossel and Matthew Brennan)

Last lecture we saw the most basic theorem about neural network approximation: that for most activation functions, including any \( \sigma \colon \mathbb{R} \to \mathbb{R} \) such that \( \sigma(x) \to 0 \) as \( x \to -\infty \) and \( \sigma(x) \to 1 \) as \( x \to \infty \), and also including the ReLU function, we can represent almost any function given enough neurons, with small error in any reasonable norm. One of the proofs relied on Fourier analysis. The simpler one was to observe that by scaling and translating to \( \sigma(\lambda x + b) \) for large \( \lambda \) and some \( b \) we can approximate the function
\[
f(x) = \chi_{[a, \infty)}(x)
\]
for every \( a \). By subtracting one of these from another we can approximate the indicator of any interval, and by standard arguments we can: 1) approximate any continuous function in the uniform sense, and 2) approximate any \( L^2 \) function in the \( L^2 \) sense. The last observation can also be adapted to Sobolev spaces, or even more detailed function classes.

An argument can be made that the result above has little to do with the practice of neural networks. Deep networks have gained currency because of their expressive power relative to their size and number of parameters — though the networks used in practice have millions of nodes, they dramatically outpace shallow networks of the same size. The first order of business of this lecture is to take a step towards explaining this.

Telgarsky’s Depth Separation: 0-1 Loss

For the result we’re about to state (from a paper of Telgarsky that can be found here ) let’s consider ReLU networks. Let \( m(x) \) be the piecewise linear function that is \( 2x \) if \( x \in [0, 0.5] \), \( 2-2x \) if \( x \in [0.5, 1] \), and \( 0 \) otherwise. This looks like:


lect2-fig1

This can be computed exactly by a two- or three-layer network, depending on your notational preferences, by the expression
\[
\sigma( 2 \sigma(x) – 4 \sigma(x – 0.5) )
\]
which is a diamond-shaped network:


diamond_network

or in terms of the activation function:


lecture2-fig_3

where the function \( m \) is slightly offset for clarity. Since we’re looking at deep networks, what do iterates of this function look like? Here are the first four:


m_iterates

Each iteration replaces the triangle with two triangles that are half as wide (but of the same height), so \( m^{(n)}(x) \) has \( 2^n \) “teeth”. Moreover, by chaining these networks, we can obviously represent this function with \( 3n+1 \) nodes in a deep network. However, it should take very many nodes to do this in a shallow network. Intuitively, depth increases the number of oscillations multiplicatively (which is why we have an exponential number of teeth) whereas width can only do so additively. The precise (and fairly strong) version of this intuition states:

Given a function \( f \) and data points \( (x_i, y_i) \), define its classification error by
\[
R(f) = \frac{1}{n} \sum_i \chi\{\tilde{f}(x_i) \neq y_i\}
\]
where \( \tilde{f}(x) = \mathbf{1}\{ f(x_i) > 1/2 \} \) is the “sign-rounding” of \( f \). Let \( x_i = i/2^n \) and \( y_i = m^{(n)}(x_i)\) where \(y = (0,1,0,1,\ldots) \). Notice that \( R(m^{(n)}) = 0 \). If a network \( g \) has \( \ell \) layers and width \( w < 2^{(n-k)/\ell-1} \), then \( R(g) > \frac{1}{2} – \frac{1}{3 \cdot 2^{k-1}} \).
The right way to quantify the oscillatory quality of \( m^{(n)} \), for the purposes of this proof, is by how many affine pieces it has. We say a function \( f \) is \( t \)-sawtooth if it is piecewise affine with \( t \) pieces. Clearly \( m \) is \( 4 \)-sawtooth, and \( \sigma \) is \( 2 \)-sawtooth. Let’s first show a quick lemma:

If \(f\) is \(a\)-sawtooth and \(g\) is \(b\)-sawtooth, then \(f + g\) is at most \(a + b\)-sawtooth and \(f \circ g\) is at most \(ab\)-sawtooth.
[of the Lemma] Addition is simple: the “joints” of \(f\) must fall within intervals on which \(g\) is affine, and adding an affine function does not change sawtooth degree. For composition, each interval on which \(f\) is affine has as its domain at most the range of \(g\), and therefore on this interval \(f(g)\) looks like (a subset of) \(ag + b\), which is \(b\)-sawtooth. Since there are \(a\) such intervals by assumption, \(f \circ g\) is \(ab\)-sawtooth.

From this result it is clear that a depth-\(\ell\) and width-\(w\) ReLU network must produce an output that is at most \((2w)^\ell\)-sawtooth. If \(w < 2^{(n-k)/\ell-1}\) then the network computes a \(r \leq 2^{n-k}\)-sawtooth function, whereas \(m^{(n)}\) is precisely \(2^n\)-sawtooth. Now, let's examine each interval on which \(g\) is affine. If the interval contains \(p\) data points, then since \(g\) is affine on this interval it can have the correct sign for at most \(\frac{p+3}{2}\) of them. There are \(2^n\) data points overall, and if we partition them into the \(2^{n-k}\) intervals where \(g\) is constant, then \(g\) can have the correct sign for at most \(2^{n-1} + 3 \cdot 2^{n-k+1}\) of them. This means that \(R(g) > \frac{1}{2} – \frac{1}{3 \cdot 2^{k-1}} \)

It is an instructive exercise to the reader to reproduce the example for the three-class, three-output extension (where the assigned class is taken to be the maximum of the three output values).

Depth Separation: Squared Loss

More generally we are interested to approximate a function \(f \colon [0,1]^d \to \mathbb{R}\) by a ReLU network \(\tilde{f}\) of depth \(\ell\) and width \(w\), where the output level is linear and we used the squared loss to evaluate network performance. The general theme is to find structural properties of \(f\) that make it hard to approximate by shallow networks — the theorem above was the first such result we’ll see, for a certain notion of oscillatory behavior for function in one dimension. Since deep nets are most used for high dimensional problems, it is interesting to try and classify how well do networks approximated such functions.

Quadratic Functions

Recalling that shallow nets cannot have too many saw-teeth, we first obtain a more general result for quadratic functions:

Let \(p(x) = p_2 x^2 + p_1 x + p_0\). If \(g\) is \(n\)-sawtooth then there is some constant \(C > 0\) such that
\[
\lVert p – g \rVert_2^2 \geq C p_2^2 / n^4
\]
The decay looks bad, but in the end it is practically optimal. This should not discourage the reader, however, because what is important for our purposes is the relative rate of deep vs. shallow networks — this result says that shallow networks can get at best an \(n^4\) approximation rate, while deep networks can get about \(2^{4n}\).
Since \(g\) is piecewise affine let’s look at the \(L^2\) error on each of its pieces separately. On each interval \([b, b+2a]\) the error looks like
\[
\int_b^{b+2a} \left| p_2 x^2 + p_1 x + p_0 – cx – d \right|^2 \, dx
\]
where \(g_{[b,b+2a]} = cx+d\). The trick is to alter the function \(g\) on each piece individually and look at the minimum error over all affine functions; adding these over all intervals will give the lower bound we want. The minimum of the above expression is
\[
\min_{c,d} \int_b^{b+2a} \left| p_2 x^2 – cx – d \right|^2 \, dx = \min_{c,d} \int_{-a}^{a} \left| p_2 x^2 – cx – d \right|^2 \, dx
\]
since translating does not change the top coefficient of a quadratic polynomial (and everything else has been absorbed into the minimum). Scaling the integration variable by \(a\) we get
\[
\min_{c,d} a\int_{-1}^{1} \left| p_2 a^2x^2 – cax – d \right|^2 \, dx
\]
and taking out the top coefficient (again absorbing everything else) yields
\[
\min_{c,d} a^5 p_2^2 \int_{-1}^{1} \left| x^2 – cx – d \right|^2 \, dx = C a^5 p_2^2
\]
Now let \(\{[a_i,b_i]\}\) be the intervals where \(g\) is affine, and let their lengths be denoted as \(\{\ell_i\}\). Then
\[
\lVert p – g \rVert_2^2 \geq \sum C p_2^2 \ell_i^5 \geq C \frac{p_2^2}{n^4}
\]
since \(1 \leq \sum \ell_i \leq \left( \sum \ell_i^5 \right)^{1/5} \left( \sum 1 \right)^{4/5}\) so \(1/n^4 \leq \sum \ell_i^5\) by Minkowski’s inequality.

One nice interpretation of the above result is that the coefficient of the \(x^2\) term is the curvature of the function, so we’ve bounded the approximation capabilities of a neural network in terms of the curvature of the function being approximated.

Generalizing to Strongly Convex Functions

That result covers quadratic functions. Using the theory of orthogonal polynomials we can extend the result to the case of arbitrary degree polynomials, or analytic functions, or really any \( L^2 \) function. The Legendre polynomials are one such family, starting out as
\[
1, x, \frac{1}{2} \left( 3x^2 – 1 \right), \ldots
\]
This sequence can be generated in many different ways, but one simple way is to do Gram-Schmidt on the set \(\{1, x, x^2, x^3, \ldots\}\) (with the \(L^2([-1,1])\) inner product, with which we will be working for the foreseeable future).

The theory of orthogonal polynomials is a rich one but for the theorem above it suffices to work with the Legendre polynomials. Another common family is the Hermite polynomials, defined among other ways as (the normalizations of)
\[
H_n(x) = (-1)^n e^{x^2} \frac{d^n}{dx^n} e^{-x^2}
\]
or
\[
H_{n+1}(x) = 2x H_n(x) – H_n'(x)
\]
starting at \(H_0 = 1\), or implicitly by
\[
e^{2xt – t^2} = \sum_{n=0}^\infty H_n(x) \frac{t^n}{n!}
\]
or as the function \(e^{x^2}\) multiplied by the solutions to the eigenvalue problem
\[
\mathcal{F} f = \lambda f
\]
where \(\mathcal{F}\) is the Fourier transform, or many more definitions. The Legendre polynomials satisfy similarly many interesting relations. The interested reader is encouraged to learn more about these families and their applications.

Back to the topic at hand. Since the Legendre polynomials \(h_i\) are an orthonormal family, and \(\{h_0, \ldots, h_n\}\) spans exactly the space of polynomials of degree at most \(n\), the minimum from before becomes the length of an orthogonal projection onto the space of polynomials orthogonal to linear polynomials, and kills exactly the terms in an expansion corresponding to \(h_1\) and \(h_0\), the constant and linear terms. More precisely, if \(f = \sum a_i h_i\), then
\[
\min_{c, d} \int_{-1}^1 \left| f – cx – d \right|^2 = \sum_{i \geq 2} |a_i|^2 \geq |a_2|^2
\]
This gives the same curvature-type result as above, in an \(L^2\) sense. However we can relate this to more canonical properties of a function as follows. We say that a function \(f\) is \(\lambda\)-strongly convex (concave) in \(I\) if \(f”(x) \geq \lambda\) (\( \leq \lambda\)) throughout \(I\). This condition easily implies that \(f'(t) \geq f'(0) + \lambda t\) for \(t \geq 0\), and \(f'(t) \leq f'(0) + \lambda t\) for \(t \leq 0\). For any \(f\) we have
\[
a_2 = \langle f, h_2 \rangle \sim \int_{-1}^1 f(t) (3t^2 – 1) \, dt = – \int_{-1}^1 (t^3 – t) f'(t) \, dt
\]
and if \(f\) is \(\lambda\)-strongly convex this is
\[
– \int_0^1 (t – t^3) f'(t) \, dt + \int_{-1}^0 (t – t^3) f'(t) \, dt
\]
The first integral is at most \(\int_0^1 (t -t^3) (f'(0) + \lambda t) \, dt\), which by some elementary manipulation is at most \(C \lambda\). The other integral is treated identically. Therefore the \(L^2\) bound gives
\[
\lVert f – g \rVert^2 \geq C \frac{\lambda^2}{n^4}
\]

What about higher dimensions? Let \(f \in \mathcal{C}^2([0,1]^d)\). If there is a unit vector \(v\) and a connected set \(U\) such that \(\langle v, H(f)(x) v \rangle \geq \lambda\) on \(U\), where by \(H_x(f)\) we mean the Hessian matrix, then projecting along \(v\) gives strong convexity, and integrating along these one-dimensional slices gives
\[
\lVert f – g \rVert \geq C \frac{\lambda^2}{n^4} \text{Vol}(U)
\]
There’s something a little unsatisfying about this condition, however, which is that in truly high dimensions we expect functions to be intrinsically more complex than in lower dimensions, and their approximation theory should scale accordingly. Neural networks excel at high-dimensional classification in practice, and this success is surely not limited to the type of one-dimensional phenomenon described above. However, let us keep in mind that while we think of the space of, say, \(1024 \times 1024\) images as a million-dimensional vector space, it really is just a space of functions operating on two dimensions: the dimensionality of a sampling of a scene, which is exactly what a camera does, is a different notion than the dimensionality of the space on which the true scene, as a function, operates. This should be contrasted with (say) the case of DNA, where the “true” space actually does have nearly a hundred thousand dimensions.

Other Depth Separation Results

The remainder of the lecture is an overview of a few results, with the ideas of proofs sketched or absent.

Circuits

One potential model for approximability of functions (which is not exactly a neural network) is a logical circuit, with function complexity measured in circuit complexity. The idea is that we have a bit string input (in \(\{0,1\}^n\)) as the “input layer”, and at each layer we can pass the bits from the previous layer through the unary NOT or binary AND or OR gates; we call such circuits Boolean. The very strong result we want to compare to is:

( Rossman, Servedio, Tan ) For any fixed \(d\), there is a function \(f \colon \{0,1\}^n \to \{0,1\}\) computed by a linear circuit of depth \(d\) and size \(O(n)\) such that any depth-\(d-1\) circuit computing a function that agrees with \(f\) on a fraction \(1/2 + \delta\) of the possible inputs, for \(\delta > 0\), has size \(O(e^{n^{\Omega (1/d)}})\).

This result follows a long line of work in theoretical computer science, in particular work by Hastad. We note, in comparison, that the separation results obtained by Telgarsky do not provide a depth separation between depth \( d \) and \( d – 1 \) like the result above.

Eldan-Shamir

As a step in this direction we will discuss a result of Eldan and Shamir showing separation between depths 2 and 3 for pretty general activation functions. More precisely, suppose \(\sigma\) satisfies
\[
|\sigma(x)| \leq C(1 + |x|^\alpha)
\]
for some \(C\) and \(\alpha\), and assume it is sufficiently expressive in the sense that for all Lipschitz \(f \colon \mathbb{R} \to \mathbb{R}\), constant outside \([-R, R]\), and \(\delta > 0\), there exist \(a, \alpha_i, \beta_i, \gamma_i\) such that
\[
\lVert f – a – \sum_i \alpha_i \sigma(\beta_i x – \gamma_i) \rVert_\infty < \delta \]

( Eldan and Shamir ) There exists universal constants \(c, C\) with the following properties. For all \(d\) there exists a probability measure \(\mathbb{P}\) on \(\mathbb{R}^d\) and a function \(g \colon \mathbb{R}^d \to [-1,1]\) supported in \(B_{C \sqrt{d}}(0)\) expressible by a \(3\)-layer network with width \(\sim d^5\), such for any \(f\) expressible by a two-layer network with width \(c e^{c d}\) we must have
\[
\mathbb{E}\left[|f – g|^2\right] \geq c
\]

Informally, there is a universal lower bound on the approximability of \(3\)-layer networks by \(2\)-layer networks.

[Sketch] In the end the function \(g\) will be radial: \(g(x) = G(\lVert x \rVert^2)\). Equivalently, its Fourier transform will be radial. In any bounded domain \(x_i^2\) is Lipschitz so we can approximate it in one network layer by the expressivity assumption, therefore we can represent \(\lVert x \rVert^2\) in one layer as well (if we wanted to output this directly we would need another layer for summation, but since we’re using as input to the next layer this summation is build into the neuron inputs). The next layer will be used to compute \(g\). In Fourier space, choose \(\mu\) to have density \(|\phi|^2\), where \(\phi(\xi) = \mathcal{F}(\chi_{B_1(0)})(\xi)\). Clearly this is radial, and
\[
\mathbb{E}_\mu |f – g|^2 = \int |f – g|^2 |\phi|^2 \, dm = \lVert \widehat{f \phi} – \widehat{g \phi} \rVert_2^2
\]
Since \(f\) (the function we’re comparing \(g\) to) is two-layered, we have
\[
f(x) = \sum_i f_i(\langle v_i, x \rangle)
\]
where here \(f_i(x) = \sigma(x – \beta_i)\), but actually there’s enough freedom in the proof to choose different activations at each neuron in each layer provided they satisfy the properties assumed with uniform constants. By examining \(\widehat{f \phi}\) as a convolution, \(\widehat{f \phi}\) is supported in
\[
T = \cup \left(\text{span} (v_i) + B_1(0) \right)
\]
The essence is to design \(g\) such that exponentially many of these “tubes” are required to cover its support. Intuitively, we want \(\widehat{g \phi}\) to have a lot of mass away from the origin. The construction uses a randomized \(\tilde{g} = \sum_i c_i \epsilon_i g\), where \(g_i\) is radial and essentially \(\chi\{\lVert x \rVert \in \Delta_i\}\), \(\Delta_i\) is basically an interval of width \(1/N\) about \(1/d\), and \(\epsilon_i\) is Rademacher. Then \(\tilde{g} \phi \sim \sum c_i \epsilon_i g\) so \(\widehat{\tilde{g} \phi} \sim \sum \epsilon_i c_i \hat{g}_i\). It’s possible to show with not too much difficulty that \(\hat{g}_i\) has mass far away from the origin, but does \(\widehat{g \phi}\) (with some nonzero probability)? What about cancellations? Let \(P\) be the projection onto the part of the spectrum near the boundary of the ball. Then it turns out
\[
\mathbb{E}_\epsilon \left\lVert P \sum c_i \epsilon_i \hat{g}_i \right\rVert^2 = \sum_i \lVert P c_i \hat{g}_i \rVert^2
\]
which gives that there is indeed mass far from zero.