Andrew Chan


Contents

Diffusion Models

Notes on the theory behind models like Stable Diffusion and their applications.

I spent 2022 learning to draw and was blindsided by the rise of AI artThere is lots to say about this, whether it has been or will be a good thing for artists and society in the long run. I hope to write about it in another post. models like Stable Diffusion. Suddenly, the computer was a better artist than I could ever hope to be.

It's been two years, and image generation with diffusion is better than ever. It's also led to breakthroughs in animation, video generation, 3D modeling, protein structure prediction, and even robot trajectory planning. Where did it come from, how does it work and where is it going?

This post collects my notes on the theory of diffusion and applications to image generation and other tasks. Readers should know some probability theory (Bayes' rule, Gaussian distributions). Examples and code using PyTorch are provided.

1. Generative modeling

The basic problem of generative modeling is: given a set of samples from an unknown distribution \( \mathbf{x} \sim p(\mathbf{x}) \), we want to generate new samples from that distribution.

Generative adversarial networks treat this as a game: a generator model taking a random seed is trained to fool a discriminator, which is simultaneously trained to tell real samples from the dataset from fake. GANs can synthesize amazing images but are notoriously hard to train. They do not explicitly model \( p(\mathbf{x}) \) and in practice end up incapable of generating substantial subsets of it. In the extreme case we get mode collapse, where the generator learns to cycle between a small subset of possible outputs to fool the discriminator.

A more explicit approach is to learn a deterministic, reversible mapping from the samples we have to a distribution which we know how to sample from, like the unit gaussian. Then we can sample a point from the known distribution and apply the inverse mapping to get a sample from \( p(\mathbf{x}) \). This is conceptually attractive and is called normalizing flows. Flows have also been used for images: OpenAI's 2018 Glow generated realistic images of faces with a semantically meaningful latent space.

Normalizing flow Normalizing flow

Hover to play. Image via Eric Jang's blog. A normalizing flow learns a deterministic, probability-density-preserving mapping between the normal distribution and a 2D dataset.

1.1 Denoising diffusion models

What if instead of mapping data points to a normal distribution deterministically, we mapped points stochastically, by blending random noise into them?

This seems weird at first. Technically this mapping wouldn't be reversible, because a given data point could map to any point in the target space.

But suppose we were to do this over many steps, where we start with a clean data point, then blend in a small amount of noise, repeating many times until we have something that looks like pure noise.

This is like the physical process of diffusion, where a drop of ink slowly diffuses out to fill a tank by the random motion of individual ink particles.

A 2D dataset being mapped to the unit gaussian over 50 noising steps. Adjust the slider or click the previews below to see it in action.

Left: our 2D dataset with noise added at the current step. Right: the expected direction over all the directions a noisy point might have come from in the previous step.

Forward noising step Forward noising step drift

Why might this stochastic mapping work better than the deterministic one that we get from normalizing flows? One answer is that in practice, the invertibility requirement for flows is highly limiting. Not only does each layer of the flow network need to be invertible, but the determinant of the Jacobian for each layer must be fast to compute.Computing the determinant of an arbitrary \(N \times N\) Jacobian is \( O(N^3) \), which is unacceptably slow. Much research focuses on finding specific functions for which this can be faster. This limits what you can express with a given model size, which could be why flows weren't the first model type to scale to Stable Diffusion levels of fidelity. In contrast, denoising diffusion models only need to learn a mapping that goes in one direction.

Training works by adding random noise to each data point in our training set, having the model predict the noise, then minimizing the L2 loss between the prediction and the actual noise direction via gradient descent.

There are a few ways to sample from a pre-trained model. They boil down to:

  1. Start with a pure noise image.
  2. Predict the noise in it, and subtract a predefined fraction of it.
  3. Repeat (2) many times (10-1000 depending on the sampler), get a noise-free image.

If you're like me, you may be wondering a few things:

2. DDPM

Let's take a look at the original approach, Denoising Diffusion Probabilistic Models. Newer advances build on the language and math of this paper.

2.1 Noising and de-noising

Given an input image \( \mathbf{x}_0 \), we map it to a point in the unit normal distribution by iteratively blending noise to it in a forward diffusion process over \(t=1,2,…,T\) timesteps. Each timestep generates a new image by blending in a small amount of random noise to the previous one: $$ \mathbf{x}_t = \sqrt{\alpha_t}\mathbf{x}_{t-1} + \sqrt{1-\alpha_t}\epsilon $$ where:

We can write the probability density of the forward step as: $$ q(\mathbf{x}_t | \mathbf{x}_{t-1}) := \mathcal{N}(\sqrt{\alpha_t}\mathbf{x}_{t-1}, (1 - \alpha_t)\mathbf{I}) $$

Recurrence property

Each step depends only on the last timestep, and the noise blended in is independent of all previous noise samples. So we can expand the recurrence and derive an equation to obtain \(\mathbf{x}_t\) in one step from \(\mathbf{x}_0\) by blending in a single gaussian noise vector, since sums of independent gaussians are also gaussian: $$ \mathbf{x}_t = \sqrt{\bar\alpha_t}\mathbf{x}_0 + \sqrt{1-\bar\alpha_t}\epsilon $$ where \(\bar\alpha_t = \prod_{i=1}^t \alpha_i\) and \(\epsilon \sim \mathcal{N}(0, \mathbf{I})\). This is used to derive the reverse process which we want to learn, and the training objective where we predict the noise that we add to images.

Noising and denoising processes in DDPM

Image via .

Now consider the reverse process. Given a noisy image \( \mathbf{x}_t \), what's the distribution of the previous, less-noisy version of it \(q(\mathbf{x}_{t-1} | \mathbf{x}_t)\)?

This is easier if we know the original image \( \mathbf{x}_0 \). By Bayes' rule, we have: $$ q(\mathbf{x}_{t-1} | \mathbf{x}_t, \mathbf{x}_0) = \frac{q(\mathbf{x}_t | \mathbf{x}_{t-1}) q(\mathbf{x}_{t-1} | \mathbf{x}_0) q(\mathbf{x}_0)}{q(\mathbf{x}_t | \mathbf{x}_0) q(\mathbf{x}_0)} $$ Subbing in the distribution formulas and doing the algebra we get... $$ q(\mathbf{x}_{t-1} | \mathbf{x}_t, \mathbf{x}_0) = \mathcal{N}(\mu(\mathbf{x}_t, \mathbf{x}_0), \Sigma(t)\mathbf{I}) $$ where $$ \mathbf{\mu}(\mathbf{x}_t, \mathbf{x}_0) = \frac{\sqrt{\alpha_t}(1-\bar{\alpha}_{t-1})\mathbf{x}_t + \sqrt{\bar{\alpha}_{t-1}}(1-\alpha_t)\mathbf{x}_0}{1-\bar{\alpha}_t} \\ \Sigma(t) = \frac{(1-\alpha_t)(1-\bar{\alpha}_{t-1})}{1-\bar{\alpha}_t} $$ That is, given a noisy image and the known original image, the distribution of the previous, less-noisy version of it is gaussian.

What can we do with this information? When we're de-noising a noisy image we won't know the original corresponding to it. We want \( q(\mathbf{x}_{t-1} | \mathbf{x}_t) \).

Since we have a closed form solution for \(q(\mathbf{x}_{t-1} | \mathbf{x}_t, \mathbf{x}_0)\), if we could use the entire dataset at generation time, we could use the law of total probability to compute \(q(\mathbf{x}_{t-1} | \mathbf{x}_t)\) as a mixture of gaussians, but we can't (billions of images!) and moreover that would not give us the novelty we want, since if we followed it for all timesteps, we would just end up recovering the training samples. We want to learn some underlying distribution function which gives us novelty in generated samples by compressing the dataset.

2.2 Learning to de-noise

It turns out that \(q(\mathbf{x}_{t-1} | \mathbf{x}_t)\) is approximately gaussian for very small amounts of noise. This is an old result from statistical physics. This gives us a way to learn a reverse distribution: we can estimate the parameters \(\mu_\theta, \Sigma_\theta\) of a gaussian, and take the KL divergence to all of the distributions \(q(\mathbf{x}_{t-1} | \mathbf{x}_t, \mathbf{x}_0)\) for every training example \(\mathbf{x}_0\).

Recall that the KL divergence is a metric measuring the difference between two probability distributions. It's easy to compute for us because we are computing it between two gaussians with known parameters, so it has a closed formFor arbitrary continuous distributions, the KL divergence requires taking an integral. This is a special case. See the formula and a short proof here.. And as it turns out, minimizing this gives us a distribution which is most likely to generate all our training samples.

The reverse distributions q conditioned on training samples, and the distribution p that we learn.

The reverse distributions \(q(\mathbf{x}_{t-1} | \mathbf{x}_t, \mathbf{x}_0^{(1)})\) and \(q(\mathbf{x}_{t-1} | \mathbf{x}_t, \mathbf{x}_0^{(2)})\) conditioned on training samples \(\mathbf{x}_0^{(1)},\mathbf{x}_0^{(2)}\), and the distribution \(p_\theta\) that we learn by minimizing KL divergence to them.

👉 We can prove that minimizing \( L \) maximizes the likelihood of generating the dataset because it optimizes a lower bound for the same, through a process called variational inference.

For a proof, see the derivation of \(L_\text{VLB}\) on Lilian Weng's blog.

Concretely, let our training objective be: $$ L = \mathbb{E}_{\mathbf{x}_{0:T} \sim q}[\sum_{t=1}^TD_{KL}(q(\mathbf{x}_{t-1}|\mathbf{x}_t, \mathbf{x}_0) || p_\theta(\mathbf{x}_{t-1}|\mathbf{x}_t))] $$ where \(D_{KL}(q || p_\theta)\) is an expressionNote the KL divergence is asymmetric, so minimizing \(D_{KL}(q || p_\theta)\) over \(p_\theta\) (which squeezes \(q\) under \(p_\theta\)) gives a different result than \(D_{KL}(p_\theta || q)\) (which does the opposite). But as we see next this doesn't ultimately matter. involving the variances \(\Sigma_\theta,\Sigma(t)\) and means \(\mu_\theta,\mu(\mathbf{x}_t,\mathbf{x}_0)\) of the two gaussians.

Ho 2020 fixed the \(\Sigma_\theta\) to be equal to \(\Sigma(t)\), since they found that trying to learn it made training too unstable, and this gave good results. So in practice we only learn the means \(\mu_\theta\). After substituting in the KL divergence formula for gaussians, we end up with an objective to minimize the L2 distance between estimated and actual means: $$ L = \sum_{t=1}^T\mathbb{E}_{\mathbf{x}_{0:T} \sim q}[\frac{1}{2\Sigma(t)}||\mu(\mathbf{x}_t, \mathbf{x}_0) - \mu_\theta(\mathbf{x}_t)||^2] $$

We can simplify further and take advantage of the fact that \(\mathbf{x}_t\) can be written as a blending of \(\mathbf{x}_0\) with gaussian noise \(\epsilon\).

This means we can rewriteMuch thanks to Calvin Luo's blog for providing detailed derivations. I learned while writing this post that I like seeing detailed proofs only a little more than I dislike math. $$ \mathbf{\mu}(\mathbf{x}_t, \mathbf{x}_0) = \frac{1}{\sqrt{\alpha_t}}\mathbf{x}_t - \frac{1-\alpha_t}{\sqrt{1-\bar\alpha_t}\sqrt{\alpha_t}}\epsilon $$ And we can define \(\mu_\theta(\mathbf{x}_t)\) in terms of an estimator \(\epsilon_\theta\) to match: $$ \mathbf{\mu}_\theta(\mathbf{x}_t) = \frac{1}{\sqrt{\alpha_t}}\mathbf{x}_t - \frac{1-\alpha_t}{\sqrt{1-\bar\alpha_t}\sqrt{\alpha_t}}\epsilon_\theta(\mathbf{x}_t, t) $$

Plugging this in turns our mean prediction problem into a noise prediction problem: $$ L = \sum_{t=1}^T\mathbb{E}_{\mathbf{x}_{0} \sim q,\epsilon}[\frac{(1-\alpha_t)^2}{2\Sigma(t)\alpha_t(1-\bar{\alpha}_t)}||\epsilon-\epsilon_\theta(\sqrt{\bar\alpha_t}\mathbf{x}_0 + \sqrt{1-\bar\alpha_t}\epsilon,t)||^2] $$

It turns out ignoring the weighting improves the quality of results. You could view this as down-weighting loss terms at small \(t\) so that the network focuses on learning the more difficult problem of denoising images with lots of noise. So the final loss function is $$ L_\text{simple} = \mathbb{E}_{t \sim [1, T], \mathbf{x}_{0} \sim q,\epsilon}[||\epsilon-\epsilon_\theta(\sqrt{\bar\alpha_t}\mathbf{x}_0 + \sqrt{1-\bar\alpha_t}\epsilon,t)||^2] $$ In code, our training loop is:

            
def train(model, train_data, alpha_min=0.98, alpha_max=0.999, T=1000, n_epochs=5):
    opt = torch.optim.SGD([model.parameters()], lr=0.1)
    alpha = torch.linspace(alpha_max, alpha_min, T)
    alpha_bar = torch.cumprod(alpha, dim=-1)

    for _ in range(n_epochs):
        for x0s in train_data:
            eps = torch.randn_like(x0s)
            t = torch.randint(T, (x0s.shape[0],))

            xts = alpha_bar[t].sqrt() * x0s +  (1.-alpha_bar[t]).sqrt() * eps
            eps_pred = model(xts, t)

            loss = torch.nn.functional.mse_loss(eps_pred, eps)
            loss.backward()
            opt.step()
            opt.zero_grad()
            
        

2.3 Sampling

Once we've learned a noise estimation model \( \epsilon_\theta(\mathbf{x}_t, t) \), we've effectively learned the reverse process. Then we can use this learned model to sample an image \( \mathbf{x}_0 \) from the image distribution by:

  1. Sampling a random noise image \(x_T \sim \mathcal{N}(0, \mathbf{I})\).
  2. For timesteps \(t\) from \(T\) to \(1\):

    1. Predict the noise \(\hat\epsilon_t = \epsilon_\theta(\mathbf{x}_t, t)\).
    2. Sample the de-noised image \(\mathbf{x}_{t-1} \sim \mathcal{N}(\frac{1}{\sqrt{\alpha_t}}(\mathbf{x}_t - \frac{1 - \alpha_t}{\sqrt{1 - \bar\alpha_t}}\hat\epsilon_t), \Sigma_\theta)\).

In code:

            
def sample(model, img_size, alpha, alpha_bar):
    xt = torch.randn(img_size)
    for t in reversed(range(T)):
        with torch.no_grad():
            eps_pred = model(xt, t)

        alpha_bar_t = alpha_bar[t]
        alpha_bar_t1 = alpha_bar[t-1] if t > 0 else 1.
        sigma = ((1.-alpha[t])*(1.-alpha_bar_t1)/(1.-alpha_bar_t)).sqrt()
        z = torch.randn(img_size)
        
        mu_pred = (xt - (1.-alpha[t])/(1.-alpha_bar[t]).sqrt()*eps_pred)/alpha[t].sqrt()
        xt = mu_pred + sigma*z
    return xt
            
        

2.4 Summary and example

Let's summarize what we've learned about DDPM:

Let's train a DDPM network on a 2D dataset. We will use the Datasaurus datasetInspired by tanelp's tiny-diffusion. of 142 points, plotted below. Follow along via Colab: Open In Colab

Datasaurus

The neural network will be a function from \(\mathbb{R}^2 \mapsto \mathbb{R}^2\). We'll start with a bog-standard MLP with 3 hidden layers of size 64 with ReLU activations. This architecture has 12,000+ parameters, so one might think there is a high chance of memorizing the dataset (284 numbers), but as we'll see, the distribution we learn will be pretty good: it will not only fit the training samples but will have high diversity.

After training, we can sample 1000 points to see how well it learned the distribution:

Datasaurus

Oh no! That doesn't look anything like the dinosaur we wanted. What happened?

One problem is that we're not passing any timestep information to the model. The noise drift vectors look pretty different at higher timesteps compared to lower timesteps. Let's try passing the timestep \(t=0,...,50\) normalized to between \(0\) and \(1\) to our model, which now map \(\mathbb{R}^3 \mapsto \mathbb{R}^2\).

Datasaurus

That's much better. But we can do better by using input encodings. These are fixed functions that transform the input before feeding them to the neural network, and they can make a big difference. We will use a fourier encoding, since we know the distribution underlying our data is like an image - a high-frequency signal in a low-dimensional (2D) space.

For an input \(D\)-dimensional point \( \mathbf{x} \), we will encode it as: $$ \text{FourierEncoding}(\mathbf{x}) = \left[ \cos(2\pi\mathbf{Bx}), \sin(2\pi\mathbf{Bx}) \right]^T $$ here \(\mathbf{B}\) is a random \(L \times D\) Gaussian matrix, where each entry is drawn independently from a normal distribution. What we are doing is transforming the input space into a \(L\)-dimensional space of random frequency features. We'll set the hyperparameter \(L\) to 32.

Datasaurus

Nice! Our distribution is looking pretty good. One more thing we can do is tweak our noising schedule. This can be crucial for performance.

Noised images at different resolution with the same noise level

Image via . The same amount of noise in different resolution images yields very different looking results, with low-res images looking much noisier than high-res ones.

Let's adjust our schedule so that the model trains on more high-signal examples. This improves performance on lower-dimensional data while doing the opposite for higher-dimensional data. It gets us our best dinosaur yet:

Left: our original and new \(\bar\alpha_t\) schedules. Right: 1000 samples from the trained model.

The original schedule already didn't take us to pure noise, with \(\bar\alpha_T \approx 0.28 \). The new schedule ends at where the old schedule was halfway, at \(0.6\).

Datasaurus Datasaurus

3. Advances


3.1 Faster generation

A major disadvantage of diffusion when it was first invented was the generation speed due to the DDPM assumption that the reverse distribution is gaussian, which is only true for large \(T\). Since then, many techniques to speed up generation have been developed, some of which can be used out-of-the-box on models pre-trained using the DDPM objective, while others require new models to be trained.

Score matching and faster samplers

Diffusion has a remarkable connection to differential equations, which enabled many faster samplers to be created as we were able to tap into the rich literature of the latter.

First, it turns out that the noise direction that we learn to estimate given a noisy input \(\mathbf{x}_t\) is equivalent For a proof, I like this video from Jia-Bin Huang or blog post from Calvin Luo. to the gradient of the log-likelihood of the forward process generating \(\mathbf{x}_t\) (also known as the score of \(\mathbf{x}_t\)) up to a constant which depends on timestep: $$ \nabla_{\mathbf{x}_t} \log q(\mathbf{x}_t) = -\frac{1}{\sqrt{1-\bar\alpha_t}}\hat\epsilon_\theta(\mathbf{x}_t, t) $$ This is interesting by itself. To see why, ignore the forward process for a second and assume that we have learned the score for \(\mathbf{x}_0\). If we imagine that \(\mathbf{x}_0\) has nonzero probability everywhere in image space, then the score would provide a vector field over the entire space that would tell us in what direction we should walk if we want to move towards the modes of the distribution. But in real life \(\mathbf{x}_0\) does not have nonzero probability everywhere. If we add noise to it, we can spread density out to where there is none, but keep the modes the same.

Sampling by following the score in a mixture of gaussians

From Calvin Luo's blog: Sampling by following the score function in a mixture of gaussians. These sampling trajectories all start from the center and have noise injected at each step. In the context of DDPMs, the noise is needed to model the reverse distribution correctly, while in the context of score-based models, the noise is needed to avoid having sampling just converge on a mode.

This formed the basis for noise-conditioning score networks, which learned the score of a progressively noised dataset and generated new samples by iteratively following the score field. If that sounds familiar, that's because it is basically the same as diffusion!

Second, it turns out that the forward diffusion process can be described by something called a stochastic differential equation (SDE) which tells us how the data distribution evolves over time as we add noise to it. And here is the magic part: there exists an ODE that describes a deterministic process whose time-dependent distributions are exactly the same as the stochastic process at each timestep, with a simple closed form involving the score function from aboveSee this introduction from Eric Ma for details.!

Comparison of reverse SDE and ODE trajectories in diffusion sampling

Comparison of reverse SDE and ODE trajectories in a diffusion of a 1-dimensional dataset. The x-axis represents the timestep \(t\), while the y-axis represents the value of \(\mathbf{x}_t\). The color is the probability density of that value at that timestep. Notice how much straighter the ODE trajectory is, which suggests a way to "speed" sampling by stepping in larger increments.

Not only does this mean that there exists a fully deterministicE.g. without injecting noise way to sample from any given pretrained diffusion model, but it also means we can use off-the-shelf ODE solvers to do the sampling for us. Whereas DDPM can take up to 1000 steps to sample a high-quality result in Stable Diffusion, a sampler based on the Euler method of solving ODEs can yield high quality results in as little as 10 steps. Karras 2022 (video) provide a great overview of the tradeoffs of these and how stochasticity of samplers like DDPM can still be important in some cases.

Distillation

Progressive distillation

A visualization of two distillation steps from where a teacher mapping noise to samples \(\mathbf{x}\) in 4 deterministic steps is distilled into a model doing the same in a single step.

In progressive distillation, given a pre-trained teacher which can sampleGenerally using a deterministic sampling algorithm. in 1000 steps, we train a student to do the same thing in 500 steps by having the student predict the output of every 2 steps of the teacher. This can be repeated by making the student the new teacher, halving the steps each time.

Progressive distillation is lossy in practice, and too much can lead to blurry or unrealistic samples. This can be mitigated by adversarial distillation, in which a discriminator is jointly trained to ensure the student samples are realistic. However, as seen in GANs, this trades off sample diversityThis begs the question: why use diffusion instead of GANs? One answer is that diffusion plus distillation allows finer control over the tradeoff between speed and diversity.. This method was used to train Stable Diffusion XL Turbo, which is capable of generating high-quality images in a single step.

3.2 Conditional generation

Given a model trained on animal images, how do I generate only cats?

In principle, it's possible to model any type of conditional probability distribution \(p(\mathbf{x} | y)\) by training a diffusion model \(\epsilon_\theta(\mathbf{x}_t, t, y)\) with pairs \(\mathbf{x}_0, y\) from the dataset. This was done by Ho 2021, who trained a class-conditional diffusion model on ImageNet. The label \(y\) can also be a text embedding, a segmentation mask, or any other conditioning information.

Class-conditioned ImageNet generations

Class-conditional generations for ImageNet from .

However, the label can sometimes lead to samples that are not realistic or lack diversity if the model has not seen enough samples from \(p(\mathbf{x} | y)\) for a particular \(y\). So we often want to tune how much the model “follows” the label during generation. This leads to the concept of guidance.

Classifier guidance

Given an image \(\mathbf{x}_0\), a classifier gives a probability distribution \(p_\phi(y|\mathbf{x}_0)\) that it lies in some class \(y\). If we take the gradient of that with respect to the input, we get a vector \(\nabla_{\mathbf{x}_0}p_\phi(y|\mathbf{x}_0)\) which we can use to push the image This is similar to how Google's DeepDream worked back in the day. towards our class \(y\).

What if each sampling step, we added the classifier gradient with respect to \(\mathbf{x}_t\) to our estimated mean? Hopefully, the diffusion will ensure the sample will land in some plausible region of image space. To ensure our classifier knows what to do with the (potentially very noisy) image \(\mathbf{x}_t\), we'll train it on noisy images.

This turns out really well experimentally and mathematically. For DDPM, if we set our reverse step estimated mean to $$ \mu_{\theta,\phi}=\mu_\theta + \sigma_t^2 * \nabla_{\mathbf{x}_t}\log p_\phi(y | \mathbf{x}_t)|_{\mathbf{x}_t=\mu_\theta} $$ Then it can be shown to a first order approximation that we're sampling from the distribution $$ p_{\theta,\phi}(\mathbf{x}_{t}|\mathbf{x}_{t+1}, y) \propto p_\theta(\mathbf{x}_{t}|\mathbf{x}_{t+1})p_\phi(y|\mathbf{x}_t) $$

The classifier used doesn't need to be particularly high-quality. For example, here are classifier-guided examples for the "T-shirt" class on Fashion-MNIST using a classifier with 40% accuracy:

Classifier-guided examples for the 'T-shirt' class on Fashion-MNIST

The level of guidance parameter cg scales the classifier gradient. More guidance leads to stronger class characteristics but possibly less realism.

Classifier-free guidance

Training a classifier takes extra work. Can we do guidance without one? Let's apply Bayes' rule to our class gradient: $$ \nabla_{\mathbf{x}_t}\log p(y | \mathbf{x}_t) = \nabla_{\mathbf{x}_t}\log p(\mathbf{x}_t | y) - \nabla_{\mathbf{x}_t}\log p(\mathbf{x}_t) $$ We have turned our class gradient into two score (§3.1) functions:

  1. \(\nabla_{\mathbf{x}_t}\log p(\mathbf{x}_t | y)\) is the score of the data \(\mathbf{x}_t\) conditioned on class \(y\).
  2. \(\nabla_{\mathbf{x}_t}\log p(\mathbf{x}_t)\) is the score of all the data \(\mathbf{x}_t\).

We have seen that denoising diffusion models learn the score of their training data, so this gives us an approach for guidance without a classifier:

  1. Train a single diffusion model on every training sample \(\mathbf{x}_0\) twice: once paired with its class label \(y\), and once paired with a null class label.
  2. When sampling from the model, call it twice: once with the desired class label and once without, then take the difference and use that as our guidance vector.

Image conditioning

Image-to-image

Basic image-to-image doesn't require retraining a model. Instead, given an input image, we can add noise to the image according to the desired strength of the conditioning image (less noise for stronger conditioning), then de-noise it. This is called SDEdit and will result in the same overall shapes as the input image. The drawback is that we cannot specify what exactly the input image is controlling; the method will make all features of the result look like the inputAt least as far as staying within the training data manifold allows. In practice, models like Stable Diffusion have a bias towards detailed paintings or photographs. based on the strength. So for example if we put in a sketch, we have to choose between the result either looking like a sketch or not following the specified shape close enough.

Attempting to use SDEdit on a sketch

Image via . SDEdit allows generating new images that look like the input, but does not allow fine-grained control as increasing conditioning strength makes all features of the output look like the input. Here, the method struggles to add color and texture at high strength (low \(t\)).

How do we specify exactly what our input image will control? Training a conditional model \(\epsilon_\theta(\mathbf{x}_t, t, \mathbf{y})\) from scratch with conditioning images \(\mathbf{y}\) in the desired modality works but is expensive.

A better idea is to use something like classifier guidance. Suppose we want to condition our generation with sketch images \(\mathbf{y}\). Given a noisy image \(\mathbf{x}_t\), we can train a model to predict the sketch lines \(\hat{\mathbf{y}}=\mathcal{F}(\mathbf{x}_t)\). We then guide each sampling step with the "sketch loss" gradient \(\nabla_{\mathbf{x}_t} ||\mathcal{F}(\mathbf{x}_t) - \mathbf{y}||^2\). This is the idea behind Sketch-Guided Diffusion.

Another idea is to fine-tune our model: take our denoiser, make some architecture changes so it takes a conditioning image \(\mathbf{y}\), then re-train it with the usual denoising objective on pairs \((\mathbf{x}_0, \mathbf{y})\). But naive fine-tuning can result in issues like overfitting and catastrophic forgetting. Rather than risk the hard-won weights of our base model \(\epsilon_\theta\), we make a copy of them \(\epsilon_{\theta_2}\).

ControlNet

Image of ControlNet via . We fine-tune a model on pairs of output images and conditions \((\mathbf{x}_0, \mathbf{c})\) which has been augmented with a copy of the original model taking in \(\mathcal{C}(\mathbf{c}) + \mathbf{x}_t\). "Zero convolution" denotes a 1x1 convolution with weights initialized to zero before training.

We'll call this the control-net, and it will take a combination of the conditioning image and noisy image \(\mathcal{C}_{\phi_1}(\mathbf{x}_t) + \mathbf{y}\), where \(\mathcal{C}_{\phi_1}(.)\) denotes a 1x1 convolution with learnable weights \(\phi_1\). Then we will combine the output with our original denoiser: \(\epsilon_\theta(\mathbf{x}_t, t) + \mathcal{C}_{\phi_2}(\epsilon_{\theta_2}(\mathcal{C}_{\phi_1}(\mathbf{x}_t) + \mathbf{y}, t))\). We'll freeze the original model but learn the parameters \(\theta_2, \phi_1, \phi_2\) via fine-tuning, with the convolutions \(\phi_1, \phi_2\) initialized to zero at first so we gradually learn a delta to the denoising step.

ControlNet examples

ControlNet provides a general, efficient procedure to add conditional controls like Canny edges or human pose to a diffusion model via fine-tuning.

The ControlNet method in practice applies at a per-block level rather than on the level of the entire denoising model. Based on human evaluations, it performs better than alternatives like Sketch-Guided Diffusion.

Inpainting

Inpainting is filling in a masked part of an image. One idea to implement this would be via image-to-image: rather than adding noise to the whole image, we just add it to the masked part. But this doesn't work because at any \(t > 0\), the denoising model doesn't know what to do with the non-noisy parts of the image.

Image via .

Instead, what works is to add noise to both the masked and un-masked parts of the image, and pass that in as \(\mathbf{x}_T\). Then at each subsequent sampling step \(t\), given \(\mathbf{x}_t\), we copy the un-masked parts of the original image, noise them according to \(t\), then place them over \(\mathbf{x}_t\) and use that as input into the denoiser.

Text-to-image

Imagen at a high level

Imagen at a high-level.

Text-to-image is conditional generation with text embedding labels. OpenAI's Dall-E trained an encoding model called CLIP to project both images and text into the same space, but a multimodal embedding space is not strictly required. Google's Imagen model used the T5 large language model to encode text into embeddings. As long as the embeddings are a rich enough representation, any can be used.

3.3 Data

While not specific to diffusion, no discussion of generative models is complete without mentioning the data they were trained on. This section will cover data used for image generation models.

Searching for 'cat' in LAION

Searching for 'cat' in LAION.

Searching for 'cat' in LAION aesthetic

Searching for 'cat' in LAION-aesthetic.

A major component of the AI art backlash is the ethics of collecting art for datasets like LAION and training image generation models on them without the consent of the artists, especially since image models can pose a direct threat to the livelihoods of those artists. However, there have been efforts to train competitive image generation models more ethicallySimon Willison calls these "vegan" models.. For example, Adobe Firefly is supposed toExcept for a recent scandal where Firefly was trained on some Midjourney images. be trained only on licensed content, such as Adobe Stock, and public domain content where copyright has expired. Additionally, Stable Diffusion 3 allowed artists to opt-out of having their images be used for training, with over 80 million images removed as a result.

Data poisoning

Nightshade

Nightshade is an example of a data poisoning attack against image generation models which received attention during the AI art backlash. Models are trained on billions of images, but for a given concept there might only be dozens. The idea of Nightshade is to poison data on a concept-specific basis.

The authors demonstrate an attack against Stable Diffusion XL using 50 images modified to cause the model to output a cow for every mention of “car” in its prompts. The modification is engineered to be as un-noticeable to the human eye as possible, by optimizing a multi-objective function involving perceptual loss.

An initial attack requires access to a model's feature extractor. The authors then examine how an attack based on 1 of 4 models performs on all the others, and say the results show their attack will generalize to models besides the initial model.

3.4 Higher Resolution

Latent space of an auto-encoder Latent diffusion

Images of latent diffusion from blog post by Ignacio Aristimuño.

An early approach allowing higher resolution images was Cascaded Diffusion, which trained a diffusion model to do initial generation at a low resolution, then a series of super-resolution diffusion models to upscale the image.

Stable Diffusion uses an approach called latent diffusion, which generates images via diffusion in the latent space of an auto-encoder, then decodes the latent to get a high-resolution image. While there had been previous attempts at simultaneously training auto-encoders and diffusion models together, the latent diffusion authors found that simply training an auto-encoder to compress the image data first, then separately training a diffusion model on the encoded latents worked best.

The above methods use a backbone diffusion model with multiple other models to scale up generated images. Progress in single-model resolution has involved various training tricks like multi-scale loss and the use of different backbone architectures like transformers.

4. Beyond images


4.1 Audio, video, and 3D

Riffusion was an early music generation model capable of generating twelve-second long songs, notable because it was made by fine-tuning Stable Diffusion to output spectrogram images. Sonauto is a more recent and controllable model built on diffusion transformers, capable of generating 1:35-long songs with coherent lyrics.

From left to right: scaling compute 1x, 4x, and 32x with Sora.

OpenAI's Sora and Google's Veo are diffusion transformer video generation models capable of generating minute-long 1080p video clips from text prompts. At a high level, Sora works by decomposing videos into spacetime patches, then learning to denoise patches.

A key insight of the Sora technical report is that diffusion transformers scale for video generation, and that performance scales with computeOpenAI did not clarify what "compute" means in this context (dataset size, model size, or training time).. Both models support various video editing tasks such as masked editing, creating perfectly looping video, animating static images, extending videos forwards or backwards in time, etc. They build on past video diffusion work like Imagen Video (2022). Autoregressive models like VideoPoet (2024) are an alternative to diffusion in this space.

Stable Video 3D

One remarkable aspect of 2D diffusion models is that they implicitly learn some 3D features like correspondences. DreamFusion (2022) exploited this to generate 3D models from text by using a text-to-image diffusion model as a prior to guide a gradient-descent based 3D reconstruction algorithmThey propose something called Score Distillation Sampling to allow the image model to provide a loss for a differentiable renderer. Surprisingly, dumber techniques like generating multiple views by clever prompting of the text-to-image model can also yield decent, though lower-quality outputs.. Stable Video 3D (2024) is a more recent work which uses video diffusion for improved multi-view consistency. Such models still rely on 3D reconstruction algorithms like photogrammetry, 3D gaussian splatting, or neural radiance fields to generate the 3D representation, possibly due to the relative sparsity of 3D dataFrom Twitter, 3D artists have learned from 2D artists how important ownership of their data is, so if this is going to change, it must do so in a more creator-friendly way..

4.2 Life sciences

Diffusion models are finding many applications in medicine and biology. For example, performing partial CT and MRI scans greatly reduces patient exposure to radiation and increases comfort, but is challenging because it requires reconstructing full scans from partial data. Diffusion models have advanced the state-of-the-art in medical image reconstruction, providing superior performance and generalization to supervised methods.

A structure predicted by AlphaFold 3. Ground truth shown in gray.

Diffusion is also state-of-the-art in protein structure prediction, with DeepMind's AlphaFold 3 using a diffusion-based architecture and showing significant improvements over both previous versions and specialized tools. Given an input list of molecules, AlphaFold 3 reveals how they fit together by generating their joint 3D structure, starting with a cloud of atoms and iteratively refining to a final molecular structure.

Beyond AlphaFold, other applications of diffusion in computational biology include single-cell data analysis, drug and small molecule design, and protein-ligand interaction.

4.3 Robotics

Video by Toyota Research Institute on how diffusion is enabling breakthroughs in robotics. See their blog post for more.

To interact with the real world, robots must be capable of a huge range of physical behaviors. Traditional approaches to get robots to do things like open doors or tie shoelaces involves explicitly programming numerous edge cases and ways to recover from them. While this works for controlled settings like factories, it does not scale.

Policy learning from demonstration is a more scalable approach where robots are instead taught how to perform tasks via human demonstrations, usually done by a human controlling the robot motors via teleoperation.

This may require anywhere from a dozen to hundreds of demonstrations, after which the robot is able to learn how to generate actions conditioned on sensor observations and possibly natural language prompts. Diffusion models are state-of-the-art policy generation models, showing substantial improvements over previous techniques, with powerful advantages like gracefully handling multimodal action distributions, being suitable for high-dimensional action spaces, and exhibiting impressive training stability.

Acknowledgements

Thanks to Luciano Vinas, Danni Zhang, and Jeff Shaw for help reviewing this article.