Skip to content

Latest commit

 

History

History
107 lines (83 loc) · 7.64 KB

physicalloss.md

File metadata and controls

107 lines (83 loc) · 7.64 KB

Physical Loss Terms

The supervised setting of the previous sections can quickly yield approximate solutions with a fairly simple training process. However, what's quite sad to see here is that we only use physical models and numerical methods as an "external" tool to produce a big pile of data 😢.

We as humans have a lot of knowledge about how to describe physical processes mathematically. As the following chapters will show, we can improve the training process by guiding it with our human knowledge of physics.

---
height: 220px
name: physloss-overview
---
Physical losses typically combine a supervised loss with a combination of derivatives from the neural network.

Using physical models

Given a PDE for $\mathbf{u}(\mathbf{x},t)$ with a time evolution, we can typically express it in terms of a function $\mathcal F$ of the derivatives of $\mathbf{u}$ via

$$ \mathbf{u}t = \mathcal F ( \mathbf{u}{x}, \mathbf{u}{xx}, ... \mathbf{u}{xx...x} ) , $$

where the $_{\mathbf{x}}$ subscripts denote spatial derivatives with respect to one of the spatial dimensions of higher and higher order (this can of course also include mixed derivatives with respect to different axes). $\mathbf{u}_t$ denotes the changes over time.

In this context, we can approximate the unknown $\mathbf{u}$ itself with a neural network. If the approximation, which we call $\tilde{\mathbf{u}}$, is accurate, the PDE should be satisfied naturally. In other words, the residual R should be equal to zero:

$$ R = \mathbf{u}t - \mathcal F ( \mathbf{u}{x}, \mathbf{u}{xx}, ... \mathbf{u}{xx...x} ) = 0 . $$

This nicely integrates with the objective for training a neural network: we can train for minimizing this residual in combination with direct loss terms. Similar to before, we can make use of sample solutions $[x_0,y_0], ...[x_n,y_n]$ for $\mathbf{u}$ with $\mathbf{u}(\mathbf{x})=y$. This is typically important, as most practical PDEs we encounter do not have unique solutions unless initial and boundary conditions are specified. Hence, if we only consider $R$ we might get solutions with random offset or other undesirable components. The supervised sample points therefore help to pin down the solution in certain places. Now our training objective becomes

$$ \text{arg min}_{\theta} \ \alpha_0 \sum_i \big( f(x_i ; \theta)-y_i \big)^2 + \alpha_1 R(x_i) , $$ (physloss-training)

where $\alpha_{0,1}$ denote hyperparameters that scale the contribution of the supervised term and the residual term, respectively. We could of course add additional residual terms with suitable scaling factors here.

It is instructive to note what the two different terms in equation {eq}physloss-training mean: The first term is a conventional, supervised L2-loss. If we were to optimize only this loss, our network would learn to approximate the training samples well, but might average multiple modes in the solutions, and do poorly in regions in between the sample points. If we, instead, were to optimize only the second term (the physical residual), our neural network might be able to locally satisfy the PDE, but still could produce solutions that are still far away from our training data. This can happen due to "null spaces" in the solutions, i.e., different solutions that all satisfy the residuals. Therefore, we optimize both objectives simultaneously such that, in the best case, the network learns to approximate the specific solutions of the training data while still capturing knowledge about the underlying PDE.

Note that, similar to the data samples used for supervised training, we have no guarantees that the residual terms $R$ will actually reach zero during training. The non-linear optimization of the training process will minimize the supervised and residual terms as much as possible, but there is no guarantee. Large, non-zero residual contributions can remain. We'll look at this in more detail in the upcoming code example, for now it's important to remember that physical constraints in this way only represent soft-constraints, without guarantees of minimizing these constraints.

Neural network derivatives

In order to compute the residuals at training time, it would be possible to store the unknowns of $\mathbf{u}$ on a computational mesh, e.g., a grid, and discretize the equations of $R$ there. This has a fairly long "tradition" in DL, and was proposed by Tompson et al. {cite}tompson2017 early on.

A popular variant of employing physical soft-constraints {cite}raissi2018hiddenphys instead uses fully connected NNs to represent $\mathbf{u}$. This has some interesting pros and cons that we'll outline in the following, and we will also focus on it in the following code examples and comparisons.

The central idea here is that the aforementioned general function $f$ that we're after in our learning problems can also be used to obtain a representation of a physical field, e.g., a field $\mathbf{u}$ that satisfies $R=0$. This means $\mathbf{u}(\mathbf{x})$ will be turned into $\mathbf{u}(\mathbf{x}, \theta)$ where we choose the NN parameters $\theta$ such that a desired $\mathbf{u}$ is represented as precisely as possible.

One nice side effect of this viewpoint is that NN representations inherently support the calculation of derivatives. The derivative $\partial f / \partial \theta$ was a key building block for learning via gradient descent, as explained in {doc}overview. Now, we can use the same tools to compute spatial derivatives such as $\partial \mathbf{u} / \partial x$, Note that above for $R$ we've written this derivative in the shortened notation as $\mathbf{u}{x}$. For functions over time this of course also works for $\partial \mathbf{u} / \partial t$, i.e. $\mathbf{u}{t}$ in the notation above.

Thus, for some generic $R$, made up of $\mathbf{u}t$ and $\mathbf{u}{x}$ terms, we can rely on the backpropagation algorithm of DL frameworks to compute these derivatives once we have a NN that represents $\mathbf{u}$. Essentially, this gives us a function (the NN) that receives space and time coordinates to produce a solution for $\mathbf{u}$. Hence, the input is typically quite low-dimensional, e.g., 3+1 values for a 3D case over time, and often produces a scalar value or a spatial vector. Due to the lack of explicit spatial sampling points, an MLP, i.e., fully-connected NN is the architecture of choice here.

To pick a simple example, Burgers equation in 1D, $\frac{\partial u}{\partial{t}} + u \nabla u = \nu \nabla \cdot \nabla u $ , we can directly formulate a loss term $R = \frac{\partial u}{\partial t} + u \frac{\partial u}{\partial x} - \nu \frac{\partial^2 u}{\partial x^2} u$ that should be minimized as much as possible at training time. For each of the terms, e.g. $\frac{\partial u}{\partial x}$, we can simply query the DL framework that realizes $u$ to obtain the corresponding derivative. For higher order derivatives, such as $\frac{\partial^2 u}{\partial x^2}$, we can simply query the derivative function of the framework multiple times. In the following section, we'll give a specific example of how that works in tensorflow.

Summary so far

The approach above gives us a method to include physical equations into DL learning as a soft-constraint: the residual loss. Typically, this setup is suitable for inverse problems, where we have certain measurements or observations for which we want to find a PDE solution. Because of the high cost of the reconstruction (to be demonstrated in the following), the solution manifold shouldn't be overly complex. E.g., it is not possible to capture a wide range of solutions, such as with the previous supervised airfoil example, with such a physical residual loss.