Brennan Gebotys Machine Learning, Statistics, and All Things Cool

Going with the Flow: An Introduction to Normalizing Flows

Photo Link

alt text

Normalizing Flows (NFs) (Rezende & Mohamed, 2015) learn an invertible mapping \(f: X \rightarrow Z\), where \(X\) is our data distribution and \(Z\) is a chosen latent-distribution.

Normalizing Flows are part of the generative model family, which includes Variational Autoencoders (VAEs) (Kingma & Welling, 2013), and Generative Adversarial Networks (GANs) (Goodfellow et al., 2014). Once we learn the mapping \(f\), we generate data by sampling \(z \sim p_Z\) and then applying the inverse transformation, \(f^{-1}(z) = x_{gen}\).

Note: \(p_Z(z)\) is the probability density of sampling \(z\) under the distribution \(Z\).

In this blog to understand normalizing flows better, we will cover the algorithm’s theory and implement a flow model in PyTorch. But first, let us flow through the advantages and disadvantages of normalizing flows.

Note: If you are not interested in the comparison between generative models you can skip to ‘How Normalizing Flows Work’

Why Normalizing Flows

With the amazing results shown by VAEs and GANs, why would you want to use Normalizing flows? We list the advantages below

Note: Most advantages are from the GLOW paper (Kingma & Dhariwal, 2018)

  • NFs optimize the exact log-likelihood of the data, log(\(p_X\))
    • VAEs optimize the lower bound (ELBO)
    • GANs learn to fool a discriminator network
  • NFs infer exact latent-variable values \(z\), which are useful for downstream tasks
    • The VAE infers a distribution over latent-variable values
    • GANs do not have a latent-distribution
  • Potential for memory savings, with NFs gradient computations scaling constant to their depth
    • Both VAE’s and GAN’s gradient computations scale linearly to their depth
  • NFs require only an encoder to be learned
    • VAEs require encoder and decoder networks
    • GANs require generative and discriminative networks

But remember what mother says, “There ain’t no such thing as a free lunch”.

Some of the downsides of normalizing flows are as follows,

  • The requirements of invertibility and efficient Jacobian calculations restrict model architecture
    • more on this later…
  • Less resources/research on NFs compared to other generative models
    • The reason for this blog!
  • NFs generative results are still behind VAEs and GANs

Now let us get dirty in some theory!

How Normalizing Flows Work

In this section, we understand the heart of Normalizing Flows.

Probability Distribution Change of Variables

Consider a random variable \(X \in \mathbb{R}^d\) (our data distribution) and an invertable transformation \(f: \mathbb{R}^d \mapsto \mathbb{R}^d\)

Then there is a random variable \(Z \in \mathbb{R}^d\) which \(f\) maps \(X\) to.

Furthermore,

\[P(X = x) = P(f(X) = f(x)) = P(Z = z)\tag{0}\]

Now consider some interval \(\beta\) over \(X\). Then there exists some interval \(\beta^{\prime}\) over \(Z\) such that,

\[P(X \in \beta) = P(Z \in \beta^{\prime})\tag{1}\] \[\int_{\beta} p_X dx = \int_{\beta^{\prime}} p_Z dz\tag{2}\]

For the sake of simplicity, we consider a single region.

\[dx \cdot p_X(x) = dz \cdot p_Z(z) \tag{3}\] \[p_X(x) = \mid\dfrac{dz}{dx}\mid \cdot p_Z(z) \tag{4}\]

Note: We apply the absolute value to maintain the equality since by the probability axioms \(p_X\) and \(p_Z\) will always be positive.

\[p_X(x) = \mid\dfrac{df(x)}{dx}\mid \cdot p_Z(f(x)) \tag{5}\] \[p_X(x) = \mid det(\dfrac{df}{dx}) \mid \cdot p_Z(f(x)) \tag{6}\]

Note: We use the determinant to generalize to the multivariate case (\(d > 1\))

\[\log(p_X(x)) = \log(\mid det(\dfrac{df}{dx}) \mid) + \log(p_Z(f(x))) \tag{7}\]

Tada! To model our random variable \(X\), we need to maximize the right-hand side of equation (7).

Breaking the equation down:

  • \(\log(\mid det(\dfrac{df}{dx}) \mid)\) is the amount of stretch/change \(f\) applies to the probability distribution \(p_X\).
    • This term is the log determinant of the Jacobian matrix (\(\dfrac{df}{dx}\)). We refer to the determinant of the Jacobian matrix as the Jacobian.
  • \(\log(p_Z(f(x)))\) constrains \(f\) to transform \(x\) to the distribution \(p_Z\).

Since there are no constraints on \(Z\) we can choose \(p_Z\)! Usually, we choose \(p_Z\) to be gaussian.

Now I know what your thinking, as a reader of this blog you strive for greatness and say,

‘Brennan, a single function does not satisfy me. I have a hunger for more.’

Applying multiple functions sequentially

Fear not my readers! I will show you how we can sequentially apply multiple functions.

Let \(z_n\) be the result of sequentially applying \(n\) functions to \(x \sim p_X\).

\[z_n = f_n \circ \dots \circ f_1(x) \tag{8}\] \[f = f_n \circ \dots \circ f_1 \tag{9}\]

Using the handy dandy chain rule, we can modify equation (7) with equation (8) to get equation (10) as follows.

\[\log(p_X(x)) = \log(\mid det(\dfrac{df}{dx}) \mid) + \log(p_Z(f(x))) \tag{7}\] \[\log(p_X(x)) = \log(\prod_{i=1}^{n} \mid det(\dfrac{dz_i}{dz_{i-1}}) \mid) + \log(p_Z(f(x)))\tag{10}\]

Where \(x \triangleq z_0\) for conciseness.

\[\log(p_X(x)) = \sum_{i=1}^{n} \log(\mid det(\dfrac{dz_i}{dz_{i-1}}) \mid) + \log(p_Z(f(x))) \tag{11}\]

We want the Jacobian term to be easy to compute since we will need to compute it \(n\) times.

To efficiently compute the Jacobian, the functions \(f_i\) (corresponding to \(z_i\)) are chosen to have a lower or upper triangular Jacobian matrix. Since the determinant of a triangular matrix is the product of its diagonal, which is easy to compute.

Now that you understand the general theory of Normalizing flows, lets flow through some PyTorch code.

The Family of Flows

For this post we will be focusing on, real-valued non-volume preserving flows (R-NVP) (Dinh et al., 2016).

Though there are many other flow functions out and about such as NICE (Dinh et al., 2014), and GLOW (Kingma & Dhariwal, 2018). For keeners wanting to learn more, I will show you to the ‘More Resources’ section at the bottom of this post which includes blog posts with more flows which may interest you.

R-NVP Flows

We consider a single R-NVP function \(f: \mathbb{R}^d \rightarrow \mathbb{R}^d\), with input \(\mathbf{x} \in \mathbb{R}^d\) and output \(\mathbf{z} \in \mathbb{R}^d\).

To quickly recap, in order to optimize our function \(f\) to model our data distribution \(p_X\), we want to know the forward pass \(f\), and the Jacobian \(\mid det(\dfrac{df}{dx}) \mid\).

We then will want to know the inverse of our function \(f^{-1}\) so we can transform a sampled latent-value \(z \sim p_Z\) to our data distribution \(p_X\), generating new samples!

Forward Pass

\[f(\mathbf{x}) = \mathbf{z}\tag{12}\]

The forward pass is a combination of copying values while stretching and shifting the others. First we choose some arbitrary value \(k\) which satisfies \(0 < k < d\) to split our input.

R-NVPs forward pass is then the following

\[\mathbf{z}_{1:k} = \mathbf{x}_{1:k} \tag{13}\] \[\mathbf{z}_{k+1:d} = \mathbf{x}_{k+1:d} \odot \exp(\sigma(\mathbf{x}_{1:k})) + \mu(\mathbf{x}_{1:k})\tag{14}\]

Where \(\sigma, \mu: \mathbb{R}^k \rightarrow \mathbb{R}^{d-k}\) and are any arbitrary functions. Hence, we will choose \(\sigma\) and \(\mu\) to both be deep neural networks. Below is PyTorch code of a simple implementation.

Log Jacobian

The Jacobian matrix \(\dfrac{df}{d\mathbf{x}}\) of this function will be

\[\begin{bmatrix}I_d & 0 \\ \frac{d z_{k+1:d}}{d \mathbf{x}_{1:k}} & \text{diag}(\exp[\sigma(\mathbf{x}_{1:k})]) \end{bmatrix} \tag{15}\]

The log determinant of such a Jacobian Matrix will be

\[\log(\det(\dfrac{df}{d\mathbf{x}})) = \log(\prod_{i=1}^{d-k} \mid\exp[\sigma_i(\mathbf{x}_{1:k})]\mid) \tag{16}\] \[\log(\mid\det(\dfrac{df}{d\mathbf{x}})\mid) = \sum_{i=1}^{d-k} \log(\exp[\sigma_i(\mathbf{x}_{1:k})]) \tag{17}\] \[\log(\mid\det(\dfrac{df}{d\mathbf{x}})\mid) = \sum_{i=1}^{d-k} \sigma_i(\mathbf{x}_{1:k}) \tag{18}\]

Inverse

\[f^{-1}(\mathbf{z}) = \mathbf{x}\tag{19}\]

One of the benefits of R-NVPs compared to other flows is the ease of inverting \(f\) into \(f^{-1}\), which we formulate below using the forward pass of equation (14)

\[\mathbf{x}_{1:k} = \mathbf{z}_{1:k} \tag{20}\] \[\mathbf{x}_{k+1:d} = (\mathbf{z}_{k+1:d} - \mu(\mathbf{x}_{1:k})) \odot \exp(-\sigma(\mathbf{x}_{1:k})) \tag{21}\] \[\Leftrightarrow \mathbf{x}_{k+1:d} = (\mathbf{z}_{k+1:d} - \mu(\mathbf{z}_{1:k})) \odot \exp(-\sigma(\mathbf{z}_{1:k})) \tag{22}\]

Summary

And voilà, the recipe for R-NVP is complete!

To summarize we now know how to compute \(f(\mathbf{x})\), \(\log(\mid\det(\dfrac{df}{d\mathbf{x}})\mid)\), and \(f^{-1}(\mathbf{z})\).

Below is the full jupyter notebook with PyTorch code for model optimization and data generation.

Jupyter Notebook

Note: In the notebook the multilayer R-NVP flips the input before a forward/inverse pass for a more expressive model.

Optimizing Model

\[\log(p_X(x)) = \log(\mid det(\dfrac{df}{dx}) \mid) + \log(p_Z(f(x)))\] \[\log(p_X(x)) = \sum_{i=1}^{n} \log(\mid det(\dfrac{dz_i}{dz_{i-1}}) \mid) + \log(p_Z(f(x)))\]

Generating Data from Model

\[z \sim p_Z\] \[x_{gen} = f^{-1}(z)\]

Conclusion

In summary, we learned how to model a data distribution to a chosen latent-distribution using an invertible function \(f\). We used the change of variables formula to discover that to model our data we must maximize the Jacobian of \(f\) while also constraining \(f\) to our latent-distribution. We then extended this notion to sequentially applying multiple functions \(f_n \circ \dots \circ f_1(x)\). Lastly, we learned about the theory and implementation of the R-NVP flow.

Thanks for reading!

Question? Criticism? Phrase? Advice? Topic you want to be covered? Leave a comment in the section below!

Want more content? Follow me on Twitter!

References

  1. Rezende, D. J., & Mohamed, S. (2015). Variational inference with normalizing flows. ArXiv Preprint ArXiv:1505.05770.
  2. Kingma, D. P., & Welling, M. (2013). Auto-Encoding Variational Bayes.
  3. Goodfellow, I., Pouget-Abadie, J., Mirza, M., Xu, B., Warde-Farley, D., Ozair, S., Courville, A., & Bengio, Y. (2014). Generative Adversarial Nets. In Z. Ghahramani, M. Welling, C. Cortes, N. D. Lawrence, & K. Q. Weinberger (Eds.), Advances in Neural Information Processing Systems 27 (pp. 2672–2680). Curran Associates, Inc. http://papers.nips.cc/paper/5423-generative-adversarial-nets.pdf
  4. Kingma, D. P., & Dhariwal, P. (2018). Glow: Generative flow with invertible 1x1 convolutions. Advances in Neural Information Processing Systems, 10215–10224.
  5. Dinh, L., Sohl-Dickstein, J., & Bengio, S. (2016). Density estimation using real nvp. ArXiv Preprint ArXiv:1605.08803.
  6. Dinh, L., Krueger, D., & Bengio, Y. (2014). Nice: Non-linear independent components estimation. ArXiv Preprint ArXiv:1410.8516.

More Resources