Yes but what is a Gaussian process? or, Once, twice, three times a definition; or A descent into madness

Gaussian processes. As narrated by an increasingly deranged man during a day of torrential rain.

Gaussian processes
Fundamentals
Theory
Deep dive
Author

Dan Simpson

Published

November 3, 2021

I guess I’m going to talk about Gaussian processes now. This wasn’t the plan but who really expected a) there to be a plan or b) me to stick to the plan. I feel like writing about Gaussian processes and so I shall! It will be grand.

What is a Gaussian process?

Well I could tell you that a Gaussian process is defined by its joint distribution \[ u \sim N(\mu, \Sigma), \] where \(u_i = u(s_i)\), \(\mu_i = \mu(s_i)\) and \(\Sigma_{ij} = c(s_i, s_j)\) for some positive definite covariance (or kernel) function \(c(\cdot, \cdot)\).

But that would be about as useful as presenting you with a dog that can bark “she’s a grand old flag”: perhaps good enough for a novelty hit, but there’s just no longevity in it.

To understand a Gaussian process you need to feel it deep down within you where the fear and the detailed mathematical concepts live.

So let’s try again.

We’re gonna have a … you know what. I’m not gonna do that. But I am going to define this stuff three times. Once for mum, once for dad, and once for the country.

You’ve got to wonder why anyone would introduce something three ways. There are some reasons. The first is, of course, that each definition gives you a different insight into different aspects of Gaussian processes (the operational, the boundless generality, the functional). And the second is because I’ve had to use all three of these ideas (and several more) over the years in order to understand how Gaussian processes work.

I learnt about GPs from several sources (listed not in order):

  • A Swede1 (so I will rant about random fields in the footnotes eventually);

  • A book2 that was introducing GPs in a very general way because they needed the concept in outrageous generality to answer questions about the distribution of the maximum of a Gaussian process;

  • A book3 written by a Russian who’s really only into measure theory and doesn’t believe anything is real if it isn’t at least happening on a Frechet space;

  • And a book4 by a different Russian who’s really only into generalised Markov properties and needed to work with Gaussian processes that are defined over functions.

Of these, the most relevant is probably the first one. I was primarily taught this stuff by Finn Lindgren, who had the misfortune of having the office next to mine when we worked together in Trondheim a very long time ago. (We both had a lot more hair then.)

One of the things that I learnt from him is that Gaussian processes can appear in all kinds of contexts, which means you need to understand them as a model for an unknown function rather than as a tool to be used in a specific context (like for Gaussian process regression or Gaussian process classification).

It’s some effort to really get a good grip on the whole “Gaussian processes as a model for an unknown function” thing but once you relax into it5, it stops being alarming to see models where you are observing things that aren’t just \(u(s_k)\). It is not alarming when you are observing integrals of the GP over regions, or derivatives. And you (or your methods) don’t fall apart when presented with complex non-linear functions on the GP (as happens if you look at
Bayesian inverse problems literature6).

What is a Gaussian process? (Version 1)

I’m going to start with the most common definition of a Gaussian process7. This is the definition that was alluded to in the first section and it’s also the definition operationalised in books like Rasmussen and Williams’8, which is a bread and butter reference for most machine learners interested in GPs, use.

The idea is pretty straightforward: I need to define a stochastic model for an unknown function \(u(s)\) and I want it to be, in some sense, Gaussian. So how do I go about doing this?

Firstly, I probably don’t care too much about the function as an abstract object. For example, if I’m using the Gaussian process to model something like temperature, I am only going to observe it at a fairly small number of places (even though I could choose any set of places I want). This means that for some arbitrary set set of \(K\) locations \(s_1, s_2, \ldots, s_K\), I am most interested9 in understanding the joint distribution10 \[ (u(s_1), \dots, u(s_K))^T. \]

So how would we model the joint distribution? If we want the model to be tractable, we probably want a nice distribution. This is where the Gaussian part comes in. The Gaussian distribution is an extremely tractable11 distribution in medium-to-high dimensions. So the choice to model our joint distribution (which could be any size \(K\)) as \[ (u(s_1), \dots, u(s_K))^T \sim N\left(\mu_{s_1, \ldots, s_K}, \Sigma_{s_1, \ldots, s_K}\right), \] makes sense from a purely mercenary position12.

So how do we choose the mean and the covariance function? We will see that the mean can be selected as \([\mu_{s_1, \ldots, s_K}]_{k} = \mu(s_k)\) for pretty much any function13 \(\mu(\cdot)\), but, when we come to write \[ [\Sigma_{s_1, \ldots, s_K}]_{ij} = c(s_i, s_j), \] there will be some very strong restrictions on the covariance function \(c(\cdot, \cdot)\).

So where do these restrictions come from?

Oh those (gay) Russians!

As with all things in probability, all the good shit comes from the Soviets. Kolmogorov14 was a leading light in the Soviet push to formalise probability and one of his many many many contributions is something called the Kolmogorov extension theorem, which gives the exact conditions under which we can go from declaring that the distributions of \((u(s_1), \ldots, u(s_K))^T\) (these are called finite dimensional distributions) are Gaussian to describing a legitimate random function \(u(s)\).

There are essentially two conditions:

  1. The order of the observations doesn’t matter in a material way. In our case changing the order just permutes the rows and columns of the mean vector and covariance matrix, which is perfectly ok.
  2. There is a consistent way to map between the distributions of \((u(s_1), \ldots, u(s_K), u(s_{K+1}))^T\) and \((u(s_1), \ldots, u(s_K))^T\). This is the condition that puts a strong restriction on the covariance function.

Essentially, we need to make sure that we have a consistent way to add rows and columns to our covariance matrix while ensuring that stays positive definite (that is, while all of the eigenvalues stay non-negative, which is the condition required for a multivariate normal distribution15). The condition—which is really gross—is that for every positive integer \(K\) and every set of points \(s_1, \ldots, s_k\), and for every \(a_1, \ldots, a_K\) not all equal to zero, we require that \[ \sum_{i=1}^K \sum_{j = 1}^K a_ia_j c(s_i, s_j) \geq 0. \]

This condition is obviously very difficult to check. This is why people typically choose their covariance function from a very short list16 that is typically found in a book on Gaussian processes.

But Kolmogorov said a little bit more

There’s a weird thing in grad school in North America where they insist on teaching measure theoretic probability theory and then never ever ever ever ever using any of the subtleties. But Gaussian processes (and, in general, stochastic processes on uncountable index spaces) are a great example of when you need these details.

Why? Because unlike discrete probability (where the set of events that we can compute the probability of is obvious) or even continuous random variables (where the events that we can’t compute the probability of are so weird we can truly just ignore them unless we are doing something truly exotic), for Gaussian processes,17 the set of allowable events is considerably smaller than the set of all things you might want probabilities of.

The gist of it is that we have built up a random function \(u(s)\) from a bunch of finite random vectors. This means that we can only assign probabilities to events that can be built up from events on finite random vectors. The resulting set of events (or \(\sigma\)-algebra to use the adult term) is called the cylindrical18 \(\sigma\)-algebra and can be roughly19 thought of as the set of all events that can be evaluated by evaluating \(u(\cdot)\) at most a countable number of times.

Things that aren’t measurable

This will potentially become a problem if, for instance, you are working with a Gaussian process in a model that uses a Gaussian process in a weird way. When this happens, it is not guaranteed that, for instance, your likelihood is a measurable function, which would mean that you can’t normalise your probability distribution! (I mean, don’t worry. Unless you’re doing something fairly wild it will be, but it has come up especially in the inverse problems literature!)

This limited set of measurable events even seems to preclude well studied “events” like “\(u\) is continuous” or “\(u\) is twice continuously differentiable” or “\(u\) has a finite supremum”. All things that we a) want to know about a Gaussian process and b) things people frequently say about Gaussian processes. It is common for people to say that “Brownian motion is continuous” and similar things.

As with all of mathematics, there are a lot of work arounds that we can use. For those three statements in particular, there is some really elegant mathematical work (due, again, to Kolmogorov and extended greatly by others). The idea is that we can build another function \(\tilde u(s)\) such that \(\Pr(u(s) = \tilde u(s)) = 1\) for all20 \(s\) such that \(\tilde u(s)\) is continuous (or differentiable or bounded).

In the language of stochastic processes, \(\tilde u(s)\) is called a version of \(u(s)\) and the more correct, temperate language (aka the one least likely to find in the literature) is that \(u(s)\) has a continuous/differentiable/bounded version.

If you’re interested in seeing how a differentiable version of a Gaussian process is constructed, you basically have to dick around with dyads for a while. Martin Hairer’s lecture notes21 is a nice clear example.

Where are the limitations of this definition?

There are a few. These are, of course, in the eye of the beer holder. The definition is workable in a lot of situations and, with some explanation can be broadened out a bit more. It’s less of a great definition when you’re trying to manipulate Gaussian processes as mathematical objects, but that’s what the next one is for.

The first limitation is maybe not so much a limit of the definition as a bit of work you have to do to make it applicable. And that is: what happens if I am observing (or my likelihood depends on) averages like \[ \left(\int_S \ell_1(s) u(s)\,ds, \ldots, \int_S \ell_K(s) u(s)\,ds\right)^T \] instead of simple point evaluations22.

This might seem like a massively different problem, until we remember that integrals are just sums dressed up for Halloween, so we can approximate the integrals arbitrarily well by sums23. In fact, if we squint24 a bit, we can see that the above vector will also be multivariate Gaussian with mean vector \[ [\mu]_k = \int_S \ell_k(s) \mu(s)\,ds \] and covariance matrix with entries \[ [\Sigma]_{ij} = \int_{S \times S} \ell_i(s) \ell_j(s')c(s, s')\,dsds'. \] Similar formulas hold for derivative observations.

Probably the bigger limitation is that in this way of seeing things, your view is tied very tightly to the covariance function. While it is a natural object for defining Gaussian processes, it is fucking inconvenient if you want to understand things like how well approximate Gaussian processes work.

And let’s face it, a big chunk of Gaussian processes we see in practice are approximate because the computational burden on large data sets is too big to do anything but approximate.

(Fun fact, when I was much much younger I wrote a paper that was a better title than a paper25 called26 In order to make spatial statistics computationally feasible, we need to abandon the covariance function. I copped a lot of shit for it at the time [partly because the title was better than the paper, but partly because some people are dicks], but I think the subsequent 10 years largely proved me (or at least my title) right27.)

The focus on the covariance function also hides the strong similarity between Gaussian process literature and the smoothing splines literature starting from Grace Wahba in the 1970s. It’s not that nobody notices this, but it’s work to get there!

In a similar way, it hides the fundamental role the reproducing kernel Hilbert space (or Cameron-Martin space) is doing and the ways that Gaussian process regression is (and is not) like kernel smoothing in RKHSs. This, again, isn’t a secret per se—you can find this information if you want it—but it’s confusing to people and the lack of clarity leads to people missing useful connections (or sometimes leads to them drawing mistaken parallels).

How many times have you seen someone say that realisations of a Gaussian process are in the RKHS associated with the covariance function? They are not. In fact, every realisation of a Gaussian process is rougher than any function in the RKHS (with probability 1)! Unfortunately, this means that your reason for choosing the kernel in a RKHS regression and for choosing the covariance function in a Gaussian process prior need to be subtly different. Or, to put it differently, a penalty is not a log-prior and interpreting the maximum a penalised likelihood is, in high dimensions, a very distant activity from interpreting a posterior distribution (even when the penalty is the log of the prior).

What is a Gaussian process? (Version 2)

Ok. Let’s do this again. This definition lives in a considerably more mathematical space and while I’m gonna try to explain the key terms, I will fail. But hey. Who doesn’t like googling weird terms?

A Gaussian process is a collection of random variables \(u(s)\), where \(s \in S\) and \(S\) is some set of things that isn’t too topologically disastrous28.

But what makes it Gaussian? Here’s the general definition.

A stochastic process/random field is Gaussian if and only if every continuous linear functional has a univariate Gaussian distribution.

Well that’s very useful Daniel. What the hell is a linear functional?

Great question angry man who lives inside my head! It is any function \(\ell(\cdot)\) that takes the Gaussian process \(u(s)\) and an input and spits out a real number that is is

  1. Linear. Aka \(\alpha \ell(u) + \beta\ell(v) = \ell(\alpha u + \beta v)\)
  2. Bounded29.

Great. Love a definition. Shall we try something more concrete?

Point evaluation \(u(s_j)\) (aka evaluating the function at a point) is a linear functional (\((u + v)(s)_j = u(s_j) + v(s_j)\)). As is a definite integral over a set \(\int_A u(s)\,ds\).

It’s a fun little exercise to convince yourself that this all implies that for any collection \(\ell_1(\cdot), \ldots, \ell_J(\cdot)\) of continuous linear functionals, then \(u(s)\) is a Gaussian process means that the vector \[ (\ell_1(u), \ldots \ell_J(u))^T \] is multivariate Gaussian.

Your idea of fun is not my idea of fun. Anyway. Keep talking.

If \(u\) lives in a Banach space30 \(B\), then the set of all continuous/bounded linear functionals on \(B\) is called the dual space and is denoted \(B^*\).

I mean, cool I guess but where the merry hell is the covariance function

In this context, the most important thing about \(B^*\) is it does double duty: it is both a space of linear functionals and a space that can be identified with random variables.

How the fuck do you do that?

Well, the trick is to remember the definition! If \(\ell \in B^*\), then \(\ell(u)\) is a Gaussian. Similarly, if we have two functionals \(\ell, \ell' \in B^*\) we consider the covariance of their associated random variables \[ C_u(\ell, \ell') = \mathbb{E}(\ell(u)\ell'(u)). \]

\(C_u(\ell, \ell')\) is a symmetric, positive definite bilinear form (aka good candidate for an inner product)!

We can use this to add more functions to \(B^*\), particularly for any sequence \(b_n \in B^*\) that is Cauchy with respect to the norm \(\|\ell\|_{R_u} = \sqrt{C_u(\ell, \ell)}\) we append the limit to \(B^*\) to complete the space. Once we take equivalence classes, we end up with a Hilbert space \(R_u\) that, very unfortunately, probabilists have a tendency to call the reproducing kernel Hilbert space associated with \(u\).

Why is this unfortunate? Well primarily because it’s not the exact same space that machine learners call the reproducing kernel Hilbert space, which is, to put it mildly, confusing. But we can build the machine learner’s RKHS (known to probabilists as the Cameron-Martin space).

Why are you even telling me this? Is this a digression?

Honestly. Yes. But regardless the space \(R_u\) is quite useful to understand what’s going on. To start off, let’s do one example that shows just how different a Gaussian process is from a multivariate normal random vector. We will show that if we multiply a GP by a constant, we completely change its support31! Many a computational and inferential ship have come to grief on these sharp rocks.

To do this, though, we need32 to make an assumption on \(B\): We assume that \(B\) is separable33. This isn’t an vacuous assumption, but in a lot of cases of practical interest, this is basically the same thing as assuming the set \(S\) is a nice bounded domain or a friendly compact manifold (and not something like \(\mathbb{R}^d\))34.

So. How do we use \(R_u\) to show that Gaussian processes are evil? Well we begin by noting that \(R_u\) is a separable35 Hilbert space it contains an orthonormal basis \(e_n\), \(n=1, \ldots, \infty\) (that is \(\|e_n\|_{R_u} = 1\) and \(\langle e_n, e_m\rangle_{R_u} = 0\) if \(n\neq m\)). We can use this basis to show some really really weird stuff about \(u(s)\).

In particular, consider another Gaussian process \(v(s) = c u(s)\), where \(c\) is a non-zero constant. For this process we can build \(R_v\) in an analogous way. The \(e_n\) are still orthogonal in \(R_v\) but now \(\|e_n\|_{R_v} = c^2\).

Now consider the functional \(X_K(\cdot) = K^{-1}\sum_{k = 1}^Ke_i(\cdot)^2\). We are going to use this function to break stuff! To do this, we are going to define two disjoint sets of functions \(A_1 = \{u: \lim_{K\rightarrow \infty} X_K(u) = 1\}\) and \(A_2 = \{u: \lim_{K\rightarrow \infty} X_K(u) = c^2\}\). Clearly \(A_1\) and \(A_2\) are disjoint if \(|c|\neq 1\).

Because \(e_n(\cdot)\) are orthonormal in \(R_u\), it follows that that \(u_n = e_n(u) \sim N(0,1)\) are iid. Similarly, \(v_n = e_n(v) \sim N(0, c^2)\) are also independent. Hence it follows from the properties of \(\chi^2\) random variables (aka the mean plus the strong law of large numbers) that \(X_K(u) \rightarrow 1\) and hence \(\Pr(u \in A_1) = 1\). On the other hand, \(X_K(v) \rightarrow c^2\), so \(\Pr(v \in A_2) = 1\). As \(A_1\) and \(A_2\) are disjoint, this means that unless \(|c|=1\), the processes \(u\) and \(v\) are mutually singular (aka they have no overlapping support).

What does this mean? This means the distributions of \(u\) and \(v\) (which remember is just \(u\) multiplied by a constant) are as different from each other as a normal distribution truncated to \((-\infty, 1)\) and another normal distribution truncated to \((1, \infty)\)! Or, more realistically36, as disjoint as a distribution over \(2\mathbb{Z}\) and \((2\mathbb{Z} - 1)\).

This is an example of the most annoying phenomena in Gaussian processes37: the slightest change in a Gaussian process can lead to a mutually singular process. In fact, this is not a particularly strange example. It can be shown that Gaussian processes over uncountable index spaces are either absolutely continuous or mutually singular. There is no half-arsing it!

This has a lot of implications when it comes to computing38, setting priors on the parameters that control the properties of the covariance function39, and just generally inference40.

Yes but where’s our reproducing kernel Hilbert space

We just saw that if \(u\) is a Gaussian process than \(c u\) will be a singular GP if \(|c| \neq 1\). What happens if we add things? Well, a result known as the Cameron-Martin theorem says that, for a deterministic \(h(s) \in B\), \(u(s) + h(s)\) is absolutely continuous wrt \(u(s)\) if and only if \(h(s)\) is in the Cameron-Martin space \(H_u\) (this is the one that machine learners call the RKHS!).

But how do we find this mythical space? I find this quite stressful!

Like, honey I do not know. But when a probabilist is in distress, we can calm them by screaming characteristic function at the top of our lungs right into their ear. Try it. It definitely works. You won’t be arrested.

So let’s do that. The characteristic function of a univariate random variable \(X\) is \[ \phi_X(t) = \mathbb{E}\left(e^{itX}\right), \] which doesn’t seem like it’s going to be an amazingly useful thing, but it actually is. It’s how you prove the central limit theorem41, and a few other shiny things.

When we are dealing with more complex random things, like random vectors and Gaussian processes, we can use characteristic functions, but we need to extend beyond the fact that they’re currently only defined for univariate random variables. Conveniently, we have some lying around. In particular, if \(\ell \in B^*\), we have the associated random variable \(\ell(u)\) and we can compute its characteristic function42, which leads to the definition of a characteristic function of a stochastic process on \(B\) \[ \phi_u(\ell) = \mathbb{E}(e^{i\ell(u)}), \quad \ell \in B^*. \]

Now this feels quite different. It’s no longer a function of some real number \(t\) but is instead a function of a linear functional \(\ell\), which feels weird but isn’t.

Characteristic functions are immensely useful because if two Gaussian processes have same characteristic function they have the same distribution43.

Because \(u(s)\) is a Gaussian process, we can compute its characteristic function! We know that \(\ell(u)\) is Gaussian so we can look up its characteristic function on Wikipedia and get that \[ \mathbb{E}(e^{i\ell(u)}) = \exp\left[{i \mu(\ell) - \frac{\sigma^2(\ell)}{2}}\right], \] where \(\mu(\ell) = \mathbb{E}(\ell(u))\) and \(\sigma^2(\ell) = \mathbb{E}(\ell(u) - \mu(\ell))^2\).

We know that \[ \mu(\ell) = \mathbb{E}(\ell(u)) \] and \[ \sigma^2(\ell) = \mathbb{E}\left[(\ell(u) - \mu(\ell)^2\right], \] the latter of which can be extended naturally to the aforementioned positive definite quadratic form \[ C_u(\ell, \ell') = \mathbb{E}\left[(\ell(u) - \mu(\ell)(\ell'(u) - \mu(\ell'))\right], \quad \ell, \ell' \in B^*. \]

This leads to the exact form of the characteristic function and to this theorem, which is true.

Theorem 1 A stochastic process \(u(\cdot)\) is a Gaussian process if and only if \[ \phi_u(\ell) = \exp\left[i\mu(\ell) - \frac{1}{2}C_u(\ell, \ell)\right]. \]

So Alf is back. In pog form.

Yes.

In this case, we can define the covariance operator \(C_u: B^* \rightarrow B\) as44 \[ (C_u \ell) (\ell') = \mathbb{E}\left[(\ell(u) - \mu(\ell)(\ell'(u) - \mu(\ell'))\right]. \] The definition is cleaner when \(\mu(\ell) = 0\) (which is why people tend to assume that when writing this shit down45), in which case we get \[ C_u\ell = \mathbb{E}(u\ell(u)) \] and \[ C_u(\ell, \ell') = \ell'(C_u\ell) \]

Great gowns, beautiful gowns.

Wow. Shady.

Anyway, the whole reason to introduce this is the following:

Theorem 2 Let \(v = x + h\). Then \[ \phi_v(\ell) = e^{i\ell(h)}\phi_u(\ell). \]

This does not not help us answer the question of whether or not \(v\) has the same support as \(u\). To do this, we construct a variable that is absolutely continuous with respect to \(u\) (we guarantee this because we specify its density46 wrt \(u\)).

To this end, take some \(g \in R_u\) and define a stochastic process \(w\) with density wrt47 u \[ \rho(u) = \exp\left[iC_u(g, u) - \frac{1}{2}C_u(g,g)\right]. \]

From this, we can compute48 the characteristic function of \(w\) \[\begin{align*} \phi_w(\ell) &= \mathbb{E}_w\left(e^{i\ell(w)}\right) \\ &= \mathbb{E}_u\left(\rho(u) e^{i\ell(u)}\right) \\ &= \exp\left[iC_u(g,\ell) + i \mu(\ell) - \frac{1}{2}C_u(\ell, \ell)\right] \end{align*}\]

So we are fine if we can find some \(h \in B\) such that \[ C_u(g, \ell) = \ell(h). \]
To do this, we note that \[ C_u(g, \ell) = \ell(C_u g), \] so for any \(g\) we can find a \(h \in B\) such that \(h = C_ug\) and for such a \(h\) \(v(s) = u(s) + h(s)\) is absolutely continuous with respect to \(u(s)\).

This gives us our definition of the Cameron-Martin space (aka the RKHS) associated with \(u\).

Definition 1 The Cameron-Martin space (or reproducing kernel Hilbert space, if you must) associated with a Gaussian process \(u\) is the Hilbert space \(H_u = \{h\in B: h = C_uh^* \text{ for some } h^* \in R_u\}\) equipped with the inner product \[ \langle h, h'\rangle_{H_u} = C_u(h^*, (h')^*) \]

A fun note is that the reason the probabilists don’t call the Cameron-Martin space the reproducing kernel Hilbert space is that there is no earthly reason to think that point evaluation will be bounded in general. So it become a problematique name. (And no, I don’t know why they’re ok with calling \(R_u\) that some things are just mysterious.)

Lord in heaven. Any chance of being a bit more concrete?

Sure! Let’s consider the case where \(u \in \mathbb{R}^n\) is a Gaussian random vector \[ u \sim N(\mu, \Sigma). \] While all of this is horribly over-powered for this case, it does help get a grip on what the inner product on \(H_u\) is.

In this case, \(B^*\) is row vectors like \(f^T\), \(f\in \mathbb{R}^n\) and \[ C_u(f^T, g^T) = \operatorname{Cov}(f^Tu, g^Tu) = f^T\Sigma g. \]

Furthermore, the operator \(C_u = \Sigma f\) satisfies \(g^T(\Sigma f) = C_u(f^T,g^T)\).

So what is \(H_u\)? Well, every \(n\) dimensional vector space can be represented as an \(n\)-dimensional vector, so what we really need to do is identify \(h^*\) from \(h\). To do this, we use the relationship \(C(h^*, \ell) = \ell(h)\) for all \(\ell \in B^*\). Translating that to our finite dimensional case we get that \[ (h^*)^T\Sigma g = h^T g,\qquad g \in \mathbb{R}^n, \] from which it follows that \(h^* = \Sigma^{-1}h\). Hence we get the inner product between \(h, k \in H_u\) \[\begin{align*} \langle h, k\rangle_{H_u} &= \langle h^*, k^*\rangle_{R_h} \\ &= (\Sigma^{-1} h)^T \Sigma (\Sigma^{-1 k}) \\ &= h^T \Sigma^{-1} k. \end{align*}\]

Ok! That’s cool!

Yes! And the same thing holds in general, if you squint49. Just replace the covariance matrix \(\Sigma\) with the covariance operator \[ (\mathcal{C}f)(s) = \int_S c(s, s') f(s') \, ds'. \]

This operator has (in a suitable sense) a symmetric50 non-negative definite (left) (closed) inverse operator \(\mathcal{Q}\), which defines the RKHS inner product by \[ \langle f, g \rangle_{H_u} = \int_{S} f(s) (\mathcal{Q} g)(s) \,ds, \] where \(f\) and \(g\) are smooth enough functions for this to make sense. In general, \(\mathcal{Q}\) will be a (very) singular integral operator, but when \(u(s)\) has the Markov property, \(\mathcal{Q}\) is a differential operator. In all of these cases the RKHS is the set of functions that are smooth enough that \(\langle f, f \rangle_{H_u} < \infty\).

We sometimes call the operator \(\mathcal{Q}\) the precision operator and it’s fundamental to thin plate spline theory as well as some nice ways to approximate GPs in 1-4 dimensions. I will blog about this later, probably, but for now if you’re interested Finn Lindgren, Håvard Rue, and David Bolin just released a really nice survey paper about the technique.

Tell me some things about the Cameron-Martin space

Now that we’ve gone to the effort of finding it, I should probably tell you why it’s so important. So here are a collection of facts!

Fact 1: The Cameron-Martin space (the set of functions and the inner product) determines a51 Gaussian process, in that if two Gaussian processes have the same mean and the the same Cameron-Martin space, they have the same distribution. In fact, the next definition of a Gaussian process is going to show this constructively.

This is nice because it means you can define a Gaussian process without needing to specify its covariance function. You just (just!) need to specify a Hilbert space. It turns out that this is a considerably easier task than trying to find a positive definite covariance function if the domain \(S\) is weird.

Fact 2: \(u(s)\) is never in the RKHS. That is, \(\Pr(u \in H_u) = 0\). But52 if, for any \(\epsilon>0\), \(A_\epsilon \subset B\) is any measurable set of functions with \(\Pr(u \in A) = \epsilon\), then \(\Pr(u \in A_\epsilon + H_u) = 1\), where \(A_\epsilon+H_u = \{a + h \in B: a\in A_\epsilon, h \in H_u\}\). Or to say it in words, although \(u\) is never in \(H_u\), if you find a set \(A_\epsilon\) that \(u\) could be in (even if it’s extremely unlikely to be there), then \(u\) is almost surely made up of a function in \(A_\epsilon\) plus a function in \(H_u\).

This is wild. It means that while \(u(\cdot)\) is never in the RKHS, all you need to do is add a bit of rough to get all of the stuff out. Another characterisation of the RKHS that are related to this is that it is the intersection of all subsets of \(B\) that have full measure under \(u\) (aka all sets \(A\subset B\) such that \(\Pr(u \in A) = 1\)).

Fact 3: If we observe some data \(y = N(Tu, \Sigma_y)\), where \(Tu = (\ell_1(u),\ldots, \ell_n(u))^T\) is some observation vector, then the posterior mean \(\mathbb{E}(u \mid y)\) is in the RKHS and that posterior distribution of \(u\mid y\) is a Gaussian process that’s absolutely continuous with respect to the prior GP u(s). This means that the posterior mean, which is our best point prediction under squared error loss, is always smoother than any of the posterior draws.

This kinda makes sense: averaging things smooths out the rough edges. And so when we average a Gaussian process in this way, we make it smoother. But this is a thing that we need to be aware of! Our algorithms, our reasoning for choosing a kernel, and our interpretations of the posterior need to be aware that the space of posterior realizations \(B\) is rougher than the space that contains the posterior mean.

Frequentists / people who penalise likelihoods don’t have to worry about this shit.

So what have we learnt?

So so so so so so so much notation and weird maths shit.

But there are three take aways here:

  1. The importance of the Fourier transform (aka the characteristic function) when it comes to understanding Gaussian processes.
  2. The maths buys us understanding of some of the more delicate properties of a Gaussian process as a random object (in particular it’s joint properties)
  3. You can define a Gaussian process exclusively using the RKHS inner product. (You can also do all of the computations that way too, but we’ll cover that later). So you do not need to explicitly specify a covariance function. Grace Wahba started doing this with thin plate splines (and \(L\)-splines) in 1974 and it worked out pretty well for her.

So to finish off this post, let’s show one more way of constructing a Gaussian process. This time we will explicitly start from the RKHS.

What is a Gaussian process? (Version 3)

Our final Gaussian process definition is going to centre the RKHS53 as the fundamental object. This construction, which is known as an abstract Wiener space54 is less general55 than our previous definition, but it covers most of the processes we are going to encounter in applications.

This construction is by far the most abstract of the three (it is in the name after all). So buckle up.

The jumping off point here is a separable Hilbert space \(H\). This has an inner-product \(\langle\cdot, \cdot \rangle_H\) on it, and the associated notion of orthogonality and an orthogonal projector. Consider an \(n\)-dimensional subspace \(V_n \subset H_u\). We can, without any trouble, define a Gaussian process on \(V_n\) \(\tilde u_n\) with characteristic function \[ \phi_{\tilde u_n}(h) = \exp\left(-\frac{1}{2}\langle h,h\rangle_H\right). \] We hit no mathematical problems because \(V_n\) is finite dimensional and nothing weird happens to Gaussians in finite dimensions.

The thing is, we can do this for any finite dimensional subspace \(V_n\) and, in particular, if we have a sequence of subspace \(V_1 \subset V_2 \subset \ldots\), where \(\operatorname{dim}(V_n) =n\), then we can build a sequence of finite dimensional Gaussian processes \(\tilde u_n\) that are each supported in their respective \(V_n\).

The question is: can we construct a Gaussian process \(\tilde{u}\) supported56 on \(H\) such that \(P_n \tilde u \stackrel{d}{=} \tilde u_n\), where \(P_n\) is the orthogonal projector from \(H\) to \(V_n\)?

You would think the answer is yes. It is not. In fact, Komolgorov’s extension theorem says that we can build a Gaussian process this way, but it does not guarantee that the process will be supported on \(H\). And it is not.

To see why this is, we need to look a bit more carefully at the covariance operator of a Gaussian process on a separable Hilbert space. The key mathematical feature of a separable Hilbert space is that it has an57 orthonormal58 basis \(e_n\). We can use the orthonormal basis to do a tonne of things, but the one we need right now is the idea of a trace59 \[ \operatorname{tr}(C_u) = \sum_{n = 1}^\infty C_u(e_i, e_i). \]

For a (zero mean) Gaussian process \(u\) supported on \(H\), we can see that \[\begin{align*} \operatorname{tr}(C_u) &= \sum_{n = 1}^\infty \mathbb{E}\left[(\langle e_n, u\rangle)^2\right] \\ &= \mathbb{E}\left[ \sum_{n = 1}^\infty\langle e_n, u\rangle_H^2\right] \\ &= \mathbb{E}\left[\langle u, u\rangle_H\right] < \infty, \end{align*}\] where the second line is just true because I say it is and the third line is Pythagoras’ theorem writ large (and is finite because Gaussian processes have a lot of moments60!).

If we were to say this in words, we would say that the covariance operator of a Gaussian process supported on a separable Hilbert space is a trace-class operator (or has a finite trace).

And this is where we rejoin the main narrative. You see, if \(\tilde{u}\) was a stochastic process on \(H\), then its characteristic function would be \[ \phi_{\tilde u}(h) = \exp\left(-\frac{1}{2}\langle h, h \rangle_H\right). \] But it can’t be! Because \(H\) is infinite dimensional and the proposed covariance operator is the identity on \(H\), which is not trace class (its trace is clearly infinite).

So whatever \(\tilde u\) is61, it is emphatically not a Gaussian process on \(H\).

That doesn’t seem like a very useful trip through abstract land

Well, while we did not successful make a Gaussian process on \(H\) we did actually build the guts of a Gaussian process on a different space. The trick is to use the same idea in reverse. We showed that \(\tilde u\) was not a Gaussian process because its covariance operator wasn’t on trace class. It turns out that the reverse also holds: if \[ \phi_u(h) = \exp\left(-\frac{1}{2}\langle C_uh, h\rangle_{H'}\right) \] and \(C_u\) is trace class on \(H'\), then \(u\) is a Gaussian process supported on \(H'\).

The hard part is going to be finding another Hilbert space \(H' \supset H\).

To do this, we need to recall a definition of a separable Hilbert space \(H\) with orthonormal basis \(e_n\), \(n=1, \ldots, \infty\): \[ H = \left\{\sum_{n=1}^\infty a_n e_n: \sum_{n=1}^\infty a_n^2 < \infty\right\}. \] From this, we can build a larger separable Hilbert space \(H'\) as \[ H' = \left\{\sum_{n=1}^\infty a_n e_n: \sum_{n=1}^\infty \frac{a_n^2}{n^2} < \infty\right\}. \] This is larger because there are sequences of \(a_n\)s that are admissible for \(H'\) that aren’t admissible for \(H\) (for example62, \(a_n = \sqrt{n}\)).

We let \(j:H \rightarrow H'\) be the linear embedding that we get by considering an element \(h \in H\) as an element of \(H'\). If we let \(e_n'\) be an orthonormal basis on \(H'\) (note: this is not the same as \(e_n\) as it needs to be re-scaled to have unit norm in \(H'\)), then we get \[ j\left(\sum_{n=1}^\infty \alpha_n e_n\right) = \sum_{n=1}^\infty \frac{\alpha_n}{n} e_n'. \] Why? Because \(\|e_n\|_{H'} = n^{-1}\) which means that \(e_n' = n e_n\) is an orthonormal basis for \(H'\). This means we have to divide the coefficients by \(n\) when we move from \(H\) to \(H'\), otherwise we wouldn’t be representing the same function.

With this machinery set up, we can ask if \(\tilde u\) is a Gaussian process on \(H'\). Or, more accurately, we can ask if \(u = j(\tilde u)\) is a Gaussian process on \(H\).

Well.

Let’s compute its characteristic function. \[\begin{align*} \phi_u(h') &= \mathbb{E}\left(e^{i\left\langle u, h' \right\rangle_{H'}}\right) \\ &= \mathbb{E}\left[\exp\left(i\left\langle \sum_{n=1}^\infty \frac{\langle \tilde u, e_n\rangle_H}{n}e_n', \sum_{n=1}^\infty h_n'e_n' \right\rangle_{H'}\right) \right] \\ &= \mathbb{E}\left[\exp\left(i \sum_{n=1}^\infty \frac{\langle \tilde u, e_n\rangle}{n} h_n\right) \right] \\ &= \exp\left(-\frac{1}{2} \sum_{n=1}^\infty \frac{h_n^2}{n^2}\right). \end{align*}\] It follows that \(\phi_u(e_n') = e^{-1/(2n^2)}\) and so63 \[ \operatorname{tr}(C_u) = -2\sum_{n=1}^\infty \log \phi_u(e_n') = \sum_{n=1}^\infty \frac{1}{n^2} < \infty, \] \(C_u\) is a trace class operator on \(H'\) and, therefore, \(u\) is a Gaussian process on \(u\).

But wait, there is more! To do the calculation above, we identified elements of \(H'\) as infinite sequences \(h' = (h'_1, h'_2, \ldots)\) that satisfy \(\sum_{n=1}^\infty n^{-2}h_n^2 < \infty\). In this case the covariance operator is \(C_{u}\) is diagonal, so the \(n\)th entry of \(C_u h' = n^{-2}h'_n\). From this, and the reasoning in the previous section, we see that the Cameron-Martin space can be thought of as a subset of \(H'\). The Cameron-Martin inner product can be constructed from the inverse of \(C_u\), which gives \[ \langle a, b\rangle_{H_u} = \sum_{i=1}^\infty n^2 a_n b_n. \] Clearly, this will not be finite unless we put much much stronger restrictions on \(a_n\) and \(b_n\) than that \(\sum_{n\geq 1} n^{-2}a_n^2 < \infty\).

The Cameron Marin space is the subspace of \(H'\) consisting of all functions \(h' = \sum_{n=1}^\infty a_n e_n'\) such that \[ \sum_{n=1}^\infty n^2a_n^2 < \infty. \] This is (isomorphic to) \(H\)!

To see this, we note that the condition is only going to hold if \(a_n = n^{-1}\alpha_n\) for some sequence \(\alpha_n\) such that \(\sum_{n\geq 1} \alpha_n^2 < \infty\). Remembering that \(e_n' = n e_n\), it follows that \(h \in H_u\) if and only if \[\begin{align*} h &= \sum_{n=1}^n\frac{\alpha_n}{n} e_n' \\ &=\sum_{n=1}^n\frac{\alpha_n}{n} n e_n \\ &=\sum_{n=1}^n \alpha_n e_n, \end{align*}\] which is exactly the definition of \(H\).

Are you actually trying to kill me?

Yes.

So let’s recap what we just did: We took a separable Hilbert space \(H\) and used it to construct a Gaussian process on a larger space \(H'\) with \(H\) as its Cameron-Martin space. And we did all of this without ever touching a covariance function. This is an abstract Wiener space construction of a Gaussian process.

The thing is that this construction is a lot more general than this. The following is a (simplified64) version of the abstract Wiener space theorem.

Theorem 3 Let \(H\) be a separable Hilbert space and let \(B\) be a separable Banach space. Furthermore, we assume that \(H\) is dense in \(B\). Then there is a unique Gaussian process \(u\) with \(\Pr(u \in B) = 1\) and \(H_u = H\). It can be constructed from the canonical cylindrical Gaussian process \(\tilde u\) on \(H\) by \(u = j(\tilde u)\), where \(j:H \rightarrow E\) is the natural embedding.

Was there any point to doing that?

I mean, probably not. The main thing we did here was see that you can take the RKHS as the primal object when building a Gaussian process. Why that may be a useful observation was not covered.

We also saw that there are some restrictions required on the covariance operator to ensure that a Gaussian process is a proper stochastic process on a given space. (For the tech-heads, the problem with \(\tilde u\) is that it’s associated probability measure is not countably additive. That is a bad thing, so we do not allow it.)

The restrictions are very clear for covariance operators on separable Hilbert spaces (they must be trace class). Unfortunately, there isn’t any clean characterization of all allowable covariance operators on more complex spaces like Banach spaces65.

Where do we go now but nowhere

And with that I have finished my task. I have defined Gaussian processes three different ways and if anyone is still reading at this point: you’re a fucking champion.

I probably want to talk about other stuff eventually:

  • Using all this technology to work out what happens to a posterior when we approximate a Gaussian process (which we usually do for computational reasons)
  • Understanding how singularity/absolute continuity of Gaussian measures can help you set priors for the parameters in a covariance function
  • The Markov property in space: what is it and how do you use it
  • Show how we can use methods for solving PDEs to approximate Gaussian processes.

The last one has gotten a lot less urgent because Finn, David and Håvard just released a lovely survey paper.

Maybe by the time I am finished with these things (if that ever happens, I don’t rate my chances), I will have justified all of this technicality. But for now, I am done.

Footnotes

  1. Not the root vegetable.↩︎

  2. The first 5 chapters of Adler and Taylor’s masterpiece s are glorious↩︎

  3. Not gonna lie. Bogachev’s Gaussian Measures is only recommended if you believe in intercessory prayer.↩︎

  4. Rozanov’s Markov Random Fields, which is freely available from that link and is so beautiful you will cry when it turns the whole question into one about function space embeddings. It will be a moist old time. Bring tissues.↩︎

  5. I recommend some video head cleaner↩︎

  6. which spent an embarrassing amount of time essentially divorced from the mainstream statistical literature↩︎

  7. This is the nomenclature that machine learners thrust upon us and it’s annoying and I hate it. Traditionally, a stochastic process was indexed by time (so in this case it would be a one-dimensional Gaussian process and when it was indexed by any other set it was referred to as a random field. So I would much rather be talking about Gaussian random fields. Why? Because there’s a bunch of shit that is only true in 1D and I’m not interested in talking about that)↩︎

  8. Great book. Great reference. No shade whatsoever↩︎

  9. Maybe? But maybe I’m most interested in the average temperature over a region. This is why we are going to need to think about things more general than just evaluating Gaussian processes at a location.↩︎

  10. Why the joint? Well because it’s likely that nearby temperature measurements will be similar, while measurements that are far apart are more likely to be (almost) independent (maybe after adjusting for season, time of day, etc).↩︎

  11. in the sense that we have formulas for almost everything we want to have formulas for↩︎

  12. We can also play games with multivariate Gaussians (like building deep Gaussian processes or putting stochastic models on the covariance structure) that markedly increase their flexibility.↩︎

  13. Usually this involves covariates!↩︎

  14. Wikipedia edit war aside (have a gander, it’s a blast), there’s evidence that Kolmogorov had a long-term relationship with Aleksandrov that was a) known at the time and b) used by the Soviets to blackmail them. So that’s fun.↩︎

  15. to be proper on some subspace. We are allowed zero eigenvalues for technical reasons and it actually turns out to be useful later, making things like thin plate splines a type of Gaussian process. Grace Wahba had to do all this without Google.↩︎

  16. Exponential, Mat'{e}rn, and squared-exponential are the common ones on \(\mathbb{R}^d\). After that shit gets exotic.↩︎

  17. and any other process built from Kolmogorov’s extension theorem↩︎

  18. so named because a set \(A = \{u(\cdot): u(s_1, \ldots, u(s_K)) \in B;\; B\in \mathcal{B}(\mathbb{R}^K)\}\) is called a cylinder set↩︎

  19. the exact definition is the smallest \(\sigma\)-algebra for which all continuous linear functionals are measurable↩︎

  20. The all part is important here. Consider the function \(u(s) = 1_{A}(s)\) where \(A\) is uniformly distributed on \(S\) and \(1_A(s)\) is 1 when \(s=A\) and zero otherwise. This function is equal to \(0\) for almost every \(s\) (rather than for every \(s\)), but the random function \(u(s)\) is definitely not the zero function (it is always non-zero at exactly one point).↩︎

  21. Bottom of page 12 through page 14 here↩︎

  22. This is a straight generalisation. If \(\ell_k(s) = \delta_{s_k}(s)\) then it’s the exact situation we were in before.↩︎

  23. In the technical language of the next section, the set of delta functions is dense in the space of bounded linear functionals↩︎

  24. aka replace integrals with sums, compute the joint distribution of the sums, and then send everything to infinity, which is ok when \(\ell_k\) are bounded↩︎

  25. The paper is pretty good and I think it’s a nice contribution. But the title was perfect.↩︎

  26. Sorry for the pay wall. It’s from so long ago it’s not on arXiv.↩︎

  27. Yes. NN-GPs, Vecchia approximations, fixed-rank Kriging, variational GPs, and all of the methods I haven’t specifically done work on, all abandon some or all of the covariance function. Whether the people who work on those methods think they’re abandoning the covariance function is between them an Cher.↩︎

  28. I say this, but you can make this work over pretty bonkers spaces. If we want to be general, if \(E\) is a linear space and \(F\) is a space of functionals on \(E\) that separates the points of \(F\), then \(u(s)\) is defined as a Gaussian process (wrt the appropriate cylindrical \(\sigma\)-algebra) if \(f(u)\) is Gaussian for all \(f\in F\). Which is fairly general but also, like, at this point I am just really showing off my maths degree.↩︎

  29. It is very convenient that continuous linear functionals and bounded linear functionals are the same thing.↩︎

  30. it’s the one with a norm↩︎

  31. The support of \(u\) is a set \(A \subset B\) such that \(\Pr(u \in A) = 1\).↩︎

  32. need is a big word here. We don’t need to do this, but not doing it makes things more technical. The assumption we are about to make let’s us breeze past a lot of edge cases as we sail from the unfettered Chapter 2 of Bogachev to the more staid and calm Chapter 3 of Bogachev.↩︎

  33. That it contains a countable dense set. Somewhat surprisingly, this implies that the Gaussian process is separable (or alternatively that it’s law is a Radon measure), which is a wildly technical condition that just makes everything about 80% less technical↩︎

  34. There are separable spaces on the whole space too, but, like, leave me alone.↩︎

  35. I’ve made a↩︎

  36. These sets don’t overlap, but they’re probably not very far apart from each other? Honestly I can’t be arsed checking but this is my feeling.↩︎

  37. and continuously index stochastic processes/random fields in general↩︎

  38. see Simon Cotter and Friends↩︎

  39. see Geir-Arne Fuglstad and friends↩︎

  40. Zhang, H. (2004). Inconsistent estimation and asymptotically equal interpolations in model-based geostatistics. Journal of the American Statistical Association, 99(465):250–261.↩︎

  41. Everyone who’s ever suffered through that inexplicable grad-level measure-valued probability course that builds up this really fucking intense mathematical system and then essentially never uses it to do anything interesting should be well aware of the many many many many ways to prove the central limit theorem.↩︎

  42. well, the characteristic function when \(t=1\) because if \(\ell \in B^*\), \(t\ell \in B^*\).↩︎

  43. In a locally convex space, this is true as measures over the cylindrical \(\sigma\)-algebra, but for separable spaces it’s true over the Borel \(\sigma\)-algebra (aka all open sets), which is an enormous improvement. (This happens because for separable spaces these two \(\sigma\)-algebras coincide.) That we have to make these sorts of distinctions (between Baire and Borel measures) at all is an important example of when you really need the measure theoretic machinery to do probability theory. Unfortunately, this is beyond the machinery that’s typically covered in that useless fucking grad probability course.↩︎

  44. It is not at all clear that the range of this operator is contained in \(B\). It should be mapping to \(B^{**}\), but that separability really really helps! Check out the Hairer notes.↩︎

  45. Or they define it on the set \(\{\ell - \mu(\ell): \ell \in B^*\}\), the completion of which is the general definition of \(R_u\).↩︎

  46. Radon-Nikodym derivative↩︎

  47. A true pain in the arse when working on infinite dimensional spaces is that there’s no natural equivalent of a Lebesgue measure, so we don’t have a universal default measure to take the density against. So we have to take it against an existing probability measure. In this case, the most convenient one is the distribution of \(u\). In finite dimensions, the density \(\rho(u)\) would satisfy \(p_w(x) = \rho(x)p_u(x)\) where \(p_w(\cdot)\) is the density of \(w\).↩︎

  48. I’m skipping the actual computation because I’m lazy.↩︎

  49. or if you’re working on a separable Hilbert space↩︎

  50. self-adjoint↩︎

  51. separable↩︎

  52. This next thing is a consequence of Borel’s inequality: \(\mathbb{B}(t, H_u)\) is the \(t\) ball in \(H_u\) and a \(A\) is any measurable subset of \(B\) with \(\Pr(u \in A) = \Phi(\alpha)\), then \(\Pr(u \in A + \mathbb{B}(t, H_u)) \geq \Phi(\alpha + t)\), where \(\Phi\) is the CDF of the standard normal distribution. Just take \(t\rightarrow \infty\).↩︎

  53. At some point while writing this I’ve started using RKHS and Cameron-Martin space interchangeably for the one that is a subset of \(B\). We’re all just gonna have to be ok with that.↩︎

  54. You can get references for this from the Bogachev book, but I actually quite like this survey from Jan van Neervaen, even though it’s almost comically general.↩︎

  55. Although I made the big, ugly assumption that \(B\) was separable halfway through the last definition, almost everything is true without that. Just with more caveats. Whereas, the abstract Wiener space construction really fundamentally uses the separability of \(H_u\) and \(B\) as a place to start.↩︎

  56. ie with \(\Pr(\tilde u \in H) = 1\)↩︎

  57. It has lots of them but everything we’re about to talk about is independent of the choice of orthonormal basis.↩︎

  58. \(\|e_n\|_H = 1\) and \(\langle e_n, e_m \rangle = 0\) for \(m\neq n\).↩︎

  59. This is the trace of the operator \(C_u\) and I would usually write this as \(\sum_{n\geq 1} \langle Ce_i, e_i\rangle\), but it makes no difference here.↩︎

  60. There’s a result called Fernique’s theorem that implies that Gaussian processes have all polynomial and exponential moments.↩︎

  61. It’s called an iso-normal process and is strongly related to the idea of white noise and I’ll probably talk about that at some point. But the key thing is it is definitely not a Gaussian process in the ordinary sense on \(H\). We typically call it a generalized Gaussian process or a Generalized Gaussian random field and it is a Gaussian process indexed by \(H\). Life is pain.↩︎

  62. chaos_reins.gif↩︎

  63. You can convince yourself this is true. I’m not doing all the work for you↩︎

  64. If you want more, read Bogachev or that Radonification paper↩︎

  65. The best reference I have is this survey↩︎

Reuse

Citation

BibTeX citation:
@online{simpson2021,
  author = {Dan Simpson},
  editor = {},
  title = {Yes but What Is a {Gaussian} Process? Or, {Once,} Twice,
    Three Times a Definition; or {A} Descent into Madness},
  date = {2021-11-03},
  url = {https://dansblog.netlify.app/yes-but-what-is-a-gaussian-process-or-once-twice-three-times-a-definition-or-a-descent-into-madness},
  langid = {en}
}
For attribution, please cite this work as:
Dan Simpson. 2021. “Yes but What Is a Gaussian Process? Or, Once, Twice, Three Times a Definition; or A Descent into Madness.” November 3, 2021. https://dansblog.netlify.app/yes-but-what-is-a-gaussian-process-or-once-twice-three-times-a-definition-or-a-descent-into-madness.