ML Interview Q Series: Why might you use an energy-based model’s cost function (e.g., contrastive divergence) in certain generative tasks, and how does it compare to typical likelihood-based losses?
📚 Browse the full ML Interview series here.
Hint: Energy-based models focus on energy landscapes rather than normalized probability distributions.
Comprehensive Explanation
Energy-based models (EBMs) are a class of models that assign an energy value to a configuration of variables instead of explicitly requiring a normalized probability distribution. A lower energy value indicates a more likely configuration, while a higher energy value indicates a less likely configuration. These models can be used for generative modeling by interpreting the negative energy as an unnormalized log-probability. However, unlike typical likelihood-based models, EBMs may sidestep the necessity of computing or approximating partition functions (normalizing constants) directly in a traditional way.
Core Mathematical Formulation
A common way to link an energy function E(x; theta) to a probability model is through the unnormalized Boltzmann distribution:
Here E(x; theta) is the energy function that maps configurations of data x to a scalar energy value (lower is more favorable). theta represents the model parameters. Z(theta) is the partition function that sums or integrates over all possible configurations x, ensuring the distribution is normalized.
In typical likelihood-based approaches such as a variational autoencoder (VAE) or a normalizing flow, we explicitly model p(x) in a normalized form and optimize log p(x) via maximum likelihood. In contrast, EBMs characterize data through the energy function E(x; theta). To train these models, one often uses cost functions such as contrastive divergence (CD) that avoid the direct (and often intractable) computation of Z(theta).
How Contrastive Divergence Works
The intuition behind contrastive divergence is to minimize the difference between the model’s distribution and the empirical data distribution, but it does so without computing the partition function explicitly. The algorithm typically works by:
• Drawing a sample x from the data distribution (i.e., from the training set). • Sampling from the model (often using Markov Chain Monte Carlo, or MCMC) to obtain a “negative sample” x'. • Updating theta based on the difference between statistics under the data distribution and statistics under the model distribution.
By repeatedly doing this, the energy surface is shaped such that real data points settle into low-energy regions, and spurious or non-data-like points are pushed into higher-energy regions. The great advantage here is that the training procedure does not require the exact computation of the partition function; instead, it uses an iterative approximation (e.g., short-run Gibbs sampling) to obtain negative samples.
When to Use Energy-Based Models
EBMs are particularly attractive in cases where we care more about the learned structure in feature space rather than having a strictly normalized probability distribution. They are also useful when the data distribution is highly complex and we want flexibility in how we represent it. EBMs can capture multi-modal distributions and complicated dependencies between variables since they rely on an energy surface that can be shaped quite flexibly.
Comparison with Likelihood-Based Losses
• Normalized Probability vs. Unnormalized Score Typical likelihood-based generative models aim to estimate a normalized probability distribution p(x) over the data, and training usually involves maximizing log p(x). EBMs, on the other hand, define an unnormalized log-probability via the energy. This potentially gives more flexibility since one does not have to maintain a tractable partition function.
• Computational Tractability In models like VAEs or autoregressive models, the likelihood is factorized in a way that is (relatively) easy to compute. EBMs do not typically factorize in such a manner. Instead, the training might rely on MCMC approaches. Though MCMC-based training is often more challenging and can be computationally expensive, it circumvents explicit normalization.
• Training Stability Likelihood-based methods can exhibit numerical instability if the parameterization is poorly chosen or if the distribution has heavy tails. EBMs may offer some robustness because they focus on pushing real data to low energy rather than explicitly modeling a normalized distribution. However, EBMs also face challenges such as mode collapse or slow mixing in MCMC if not carefully implemented.
• Expressive Power In principle, EBMs can represent highly complex data distributions. They do not require an explicit factorization of the data distribution, making them more expressive, though at the cost of more difficult and specialized training algorithms (e.g., contrastive divergence, Persistent Contrastive Divergence, Score Matching, or Noise Contrastive Estimation).
Follow-up Questions
How do we interpret the energy function in an EBM?
The energy function E(x; theta) can be seen as a scalar value that indicates how “compatible” or “plausible” the data point x is under the parameters theta. If E(x; theta) is small (low energy), that means x is likely to be observed under the current parameters. If E(x; theta) is large (high energy), the model is essentially penalizing x as unlikely.
One way to think about it is to draw parallels with physics, where an energy function describes how stable or unstable a particular configuration of a physical system is. In an energy-based model, stable or likely data points correspond to low-energy states.
Why is the partition function in an EBM often intractable?
The partition function Z(theta) is typically the sum or integral of exp(-E(x; theta)) over all possible configurations x. For high-dimensional data (like images or text), the space of x is extremely large. Thus, computing this integral or sum exactly is generally impossible in practice.
This intractability is a central reason why EBMs rely on training approaches such as contrastive divergence. These approaches circumvent the direct calculation of Z(theta) by using approximate sampling methods (e.g., Gibbs sampling or Langevin dynamics) to obtain negative samples from the current model distribution, thereby approximating the gradient update without normalizing the distribution.
What are the main challenges of training an EBM with contrastive divergence?
Contrastive divergence can suffer from several issues in practice:
• Mixing and Mode Coverage MCMC methods may get stuck in local minima, failing to explore the full variety of modes in complex distributions. This can lead to poor negative samples and partial coverage of the data distribution.
• Hyperparameter Tuning Learning rates, number of MCMC steps, and other parameters (such as noise levels in Langevin sampling) are crucial in shaping stable learning. Poorly chosen hyperparameters can lead to instability or convergence issues.
• Mode Collapse If the MCMC chain does not move far from the initial data point, the model may collapse to generating samples similar to the training examples without capturing the underlying general data structure.
When should one prefer an EBM over a traditional likelihood-based model?
EBMs are often favored when you need a flexible representation of complex, high-dimensional data and you are willing to pay the price of more advanced sampling procedures. They are beneficial in scenarios like:
• Implicit Generation When you do not need a direct sampling mechanism (e.g., an explicit invertible function), but you care more about the relative energy assigned to different configurations.
• Multi-modal Data EBMs can adapt to multi-modal distributions without explicitly segmenting data modes.
• Flexible Architectures If you want to embed a variety of deep feature extractors (such as convolutional or transformer networks) without worrying about ensuring a tractable likelihood factorization.
Can we combine the advantages of EBMs with other generative modeling techniques?
Yes. There have been research works exploring hybrids. For example, one can use VAEs or normalizing flows within an EBM framework, or vice versa:
• VAEs with Energy Regularization Sometimes, additional energy-like terms are added to VAE objectives to shape latent representations.
• EBMs as Priors EBMs can serve as a learned prior in hierarchical models, where the EBM is used in combination with other inference models that provide approximate likelihoods.
These combinations strive to harness the best of both worlds: the flexible expressiveness of EBMs and the tractable sampling/inference of likelihood-based models.
Implementation Example in Python
Below is a simplified illustration of how one might implement a toy contrastive divergence procedure for an EBM:
import torch
import torch.nn as nn
import torch.optim as optim
class SimpleEnergyModel(nn.Module):
def __init__(self, input_dim, hidden_dim):
super(SimpleEnergyModel, self).__init__()
self.fc = nn.Sequential(
nn.Linear(input_dim, hidden_dim),
nn.ReLU(),
nn.Linear(hidden_dim, 1)
)
def forward(self, x):
# Outputs a scalar energy for each input
return self.fc(x)
def contrastive_divergence_step(model, x_data, lr=1e-3, k_steps=1):
# x_data: real samples
# k_steps: MCMC steps for negative sample generation
# This is a toy example using naive random noise as negative samples
optimizer.zero_grad()
# Compute energy of real data
energy_real = model(x_data).mean()
# Generate negative samples (naive random approach for illustration)
x_neg = torch.randn_like(x_data)
for _ in range(k_steps):
# A more real approach would do gradient-based MCMC updates
x_neg.requires_grad_(True)
energy_neg = model(x_neg).sum()
grad = torch.autograd.grad(energy_neg, [x_neg], create_graph=True)[0]
x_neg = x_neg - lr * grad # gradient descent step to go to lower energy
x_neg = x_neg.detach()
energy_generated = model(x_neg).mean()
# Contrastive divergence objective
# Typically it's something like: d = E_data[energy] - E_model[energy]
# Minimizing this leads to data having lower energy than generated samples
loss = energy_real - energy_generated
loss.backward()
optimizer.step()
return loss.item()
# Usage
input_dim = 2
hidden_dim = 10
model = SimpleEnergyModel(input_dim, hidden_dim)
optimizer = optim.Adam(model.parameters(), lr=1e-3)
# Fake data for demonstration
x_data = torch.randn(64, input_dim)
# Perform one CD update step
loss_value = contrastive_divergence_step(model, x_data)
print("CD loss:", loss_value)
While this code is highly simplified and uses random initialization for negative samples, it highlights the idea behind contrastive divergence: push energies of real data down while pushing energies of generated samples up (or vice versa, depending on the sign convention).
In real-world applications, you would use more sophisticated approaches for negative sample generation (e.g., persistent chains in Persistent Contrastive Divergence, or advanced MCMC steps), and you would carefully tune hyperparameters.
Additional Considerations
Energy-based methods can be more difficult to train compared to direct likelihood-based approaches. However, because they allow great flexibility and do not require a tractable normalization factor, they can model complex data distributions effectively. The choice between EBMs and typical likelihood-based losses ultimately depends on the problem constraints, computational budget, and whether the extra flexibility of EBMs is advantageous enough to justify the complexity.
Below are additional follow-up questions
How do you effectively initialize an EBM, and how does poor initialization affect training performance?
Proper initialization can be the difference between smooth, stable training and a model that fails to converge or suffers from mode collapse. In EBMs, the energy function often involves deep neural networks, which can be sensitive to the choice of initial weights. Unlike models that optimize a supervised loss with abundant labeled data, EBMs must simultaneously learn to map real data to low-energy basins and push non-data configurations toward higher energy. If the initialization places parameters in a regime that induces extremely large or small energy values, the model might generate gradients that are too large to handle (leading to exploding gradients) or too close to zero (leading to vanishing gradients).
One approach to good initialization is using standard heuristics from deep learning, such as Xavier or Kaiming initialization, to maintain stable layer-wise variance. In some implementations, people start with smaller weights or even pretrained feature extractors to avoid randomizing the entire network from scratch. Another tactic is to initialize the final output layer so that the energy values are moderate. For instance, you might want to ensure that before training, real data points have energies that are neither too large nor too small, allowing the contrastive divergence updates to adjust parameters effectively.
A potential pitfall arises if the initialization is too “flat,” meaning the model assigns almost the same energy to all data points. Early learning signals in this scenario can be excessively noisy, creating an unstable training phase. Conversely, if the initialization is overly “sharp,” the model might develop unbalanced energy regions that fail to capture the global structure, which can prompt the optimizer to get stuck in narrow local minima.
Edge cases:
• Overly large initial weights can push nearly all real data points to very high energy, resulting in strong gradients that fail to converge. • Overly small weights can make it difficult to differentiate between real and generated samples, slowing convergence to a crawl. • For higher-dimensional data like images or text embeddings, slight miscalibration in initialization can be amplified across layers, so it’s important to monitor energy histograms or distributions in the early phase of training.
What are practical strategies to mitigate mode collapse in EBMs?
Mode collapse occurs when the model learns to produce only a subset of the possible modes or fails to distribute energy properly across the true data manifold. In EBMs, this often manifests as the energy function focusing too narrowly on certain training examples while ignoring other modes of the data distribution.
Several strategies can help mitigate this:
• Multi-Chain Sampling Instead of relying on a single Markov chain to explore all modes, one can maintain multiple chains initialized at diverse locations in data space. This improves coverage because the different chains can explore different modes simultaneously.
• Longer and Smarter MCMC Short-run MCMC can lead to poor exploration if the step size or the number of steps is insufficient. Employing more sophisticated sampling methods (e.g., Hamiltonian Monte Carlo or adaptive step-size methods) can enhance the search of the distribution’s broader regions.
• Persistent Contrastive Divergence PCD retains a set of “persistent” negative samples across training iterations. This technique typically explores the model’s distribution more thoroughly than restarting MCMC chains from scratch at each update.
• Regularization Techniques Adding certain penalty terms to the loss function (e.g., weight decay or gradient penalties) can promote smoother, more evenly distributed energy surfaces, dissuading the model from collapsing onto small energy basins.
• Careful Hyperparameter Tuning Parameters such as learning rate, batch size, and step size for MCMC updates heavily influence how well the EBM learns. Aggressive hyperparameters can lead to fast but unstable training, often causing partial coverage of modes. Conservative tuning can be safer, though potentially slower.
A subtle issue is that EBMs inherently do not force coverage of all modes unless the sampling procedure thoroughly explores them. Without careful sampling, the model might not “see” or pay sufficient attention to underrepresented or subtle modes in the data. Monitoring sample diversity and distribution coverage is essential to detect emerging mode collapse.
In what ways can EBMs be used for out-of-distribution detection or anomaly detection, and how do they compare to typical approaches?
Energy-based models naturally provide a scalar energy value, which can serve as a measure of how “normal” or “plausible” a sample is under the learned distribution. For data x, a high energy might indicate that x is not well-supported by the model, whereas a low energy indicates that the sample is recognized as typical.
To leverage EBMs for out-of-distribution (OOD) detection or anomaly detection, one can compute the energy for incoming samples. If the energy surpasses a certain threshold, the sample is flagged as OOD or anomalous. This is conceptually similar to many probabilistic methods where likelihood or density is used to determine abnormality.
However, EBMs differ in that:
• They do not require a fully normalized probability distribution, potentially offering more flexibility when data distributions are very complex. • They can be robust to the “over-confidence” issue sometimes seen in likelihood-based models, where a model might assign surprisingly high likelihoods to OOD data. The EBM, by virtue of its unnormalized nature, could be less prone to systematically underestimating the energy for OOD samples.
A potential pitfall is that the energy landscape might still place relatively low energy on certain OOD regions if the model fails to push them away during training. Also, some EBMs might be less stable if MCMC sampling for negative examples doesn’t thoroughly explore the data space. As with other OOD methods, the choice of detection threshold and the thoroughness of data coverage during training are critical for reliability.
Edge cases include:
• Highly adversarial OOD samples that lie in ambiguous regions of data space. An EBM might require extensive negative sampling to push up energy values in those areas. • Situations where normal data contains multiple, highly distinct modes. If the EBM fails to capture some of these modes accurately, it could misclassify legitimate but rare samples as anomalies.
What is persistent contrastive divergence (PCD), and how is it different from standard contrastive divergence?
Persistent Contrastive Divergence is a variant of the contrastive divergence algorithm designed to improve the quality and stability of negative samples. In standard contrastive divergence, one typically starts Markov chains at real data points every iteration (often referred to as CD-k, where k is the number of MCMC steps). This can sometimes limit the exploration of the energy landscape if each chain only runs a few steps from its initial point.
In PCD, rather than restarting Markov chains from data samples every time, one maintains a pool (or “persistent set”) of negative particles (i.e., the negative samples). After each update, these particles are not discarded or re-initialized. Instead, the chains continue from their previous state in subsequent iterations. This approach allows the chains to wander further from the data manifold and explore a broader portion of the distribution’s support.
Benefits of PCD:
• More thorough coverage of the model distribution, potentially reducing mode collapse because the chains can traverse various modes over time. • Greater consistency in the negative samples from iteration to iteration, leading to more stable gradients.
Drawbacks:
• Requires additional memory to store the persistent particles. • If the sampling method doesn’t mix well, the chains might remain stuck in certain regions, leading to biased sampling. • Tuning is more involved; one must find a balance for MCMC step size, number of steps, and how many persistent chains to maintain.
A subtlety is that while PCD generally improves coverage compared to standard CD, it does not magically solve all exploration problems. The fundamental mixing challenges of Markov chains in high-dimensional spaces remain. Nevertheless, it is often considered a more robust practice than resetting all chains from data examples at each iteration.
How do you handle continuous vs. discrete variables in EBMs, and what are the implications for MCMC-based sampling methods?
EBMs can be applied to both continuous and discrete data, but the choice of sampling procedure differs significantly:
• Continuous Variables Langevin dynamics and Hamiltonian Monte Carlo are common for continuous variables. The update steps typically involve gradient-based adjustments to each variable, leveraging partial derivatives of the energy function with respect to continuous inputs.
• Discrete Variables When variables are discrete (e.g., binary or categorical), gradient-based methods are no longer straightforward because we cannot compute partial derivatives with respect to discrete states in the same way. Instead, techniques like Gibbs sampling or Metropolis-Hastings are employed. Gibbs sampling iteratively updates each discrete variable (or block of variables) given the rest. However, if the state space is large, these methods can be very slow to mix.
Implications for training:
• For continuous domains, gradient-based MCMC tends to be more efficient but can be sensitive to step size and require careful tuning. • For discrete variables, the combinatorial explosion of states makes mixing and coverage more challenging, especially if the distribution is highly multi-modal. • One often needs to incorporate problem-specific structures or constraints. For instance, in language modeling with discrete tokens, specialized samplers or approximations may be necessary.
A potential pitfall is that naive sampling for complex discrete problems can lead to extremely slow updates and partial coverage of the state space. One must then consider advanced sampling algorithms or approximations such as block Gibbs sampling or specialized transformations that can mimic gradient-based updates in discrete spaces.
How can we interpret or visualize the learned energy surface to gain insights into the model behavior or training progress?
Visualizing the energy landscape can be challenging for high-dimensional data, but it provides crucial insights into how the model views different regions of input space. In lower-dimensional scenarios (e.g., 2D or 3D synthetic datasets), one can systematically grid the input space, compute the energy at each point, and create contour plots or heatmaps.
For higher-dimensional data:
• Dimensionality Reduction Techniques like t-SNE or UMAP can project data and generated samples into a 2D embedding. While this doesn’t directly plot energy, it might reveal whether generated samples cluster around real data or diverge in separate regions.
• Energy Histograms One practical monitoring strategy is to keep track of histograms of energy values for both real data and negative samples. If there is significant overlap, it might imply that the model is not sufficiently distinguishing between real and synthetic data. If the energies for real data drift too low compared to negative samples, this might indicate an overconfident model that could be prone to mode collapse.
• Latent Space Traversals If the EBM is combined with a latent variable model (such as in certain hybrid architectures), walking through the latent space can reveal how energy changes along interpolation paths.
Pitfalls include:
• Projection methods can distort the true geometry of the data, so visualizations may be misleading. • Observing only the energy for a small set of negative samples might not represent the full breadth of the distribution, especially if the sampling procedure is biased or not well mixed.
Are there any special architecture considerations for EBMs, and how do they help with gradient-based MCMC sampling?
Architectural choices can significantly influence an EBM’s ability to produce rich yet stable energy landscapes. Some key considerations include:
• Residual Connections Residual or skip connections can help stabilize deep networks that produce energy values. These connections often mitigate issues with vanishing or exploding gradients by allowing direct signal flow from earlier to later layers.
• Normalization Layers Batch normalization or layer normalization can help control the scale of outputs, preventing extremely large or small energies. However, one must be careful because normalization layers can introduce complexities in MCMC sampling if the statistics used by the normalization layers depend on batch data.
• Activation Functions Smooth activations like ReLU, ELU, or Softplus are typically used. In energy-based settings, saturating or discontinuous activations might create challenging high-gradient regions, complicating the sampling process.
• Symmetry or Invariance If the data has known symmetries (e.g., translation invariance in images), architectures like convolutional networks can naturally encode these. This can make the energy landscape more consistent for shifted or transformed versions of the same sample, improving generalization.
• Decoupling Feature Extraction from Energy Computation In many designs, a large neural network is used to extract features, while a smaller network or linear layer maps those features to an energy value. Separating representation learning from final energy scoring can increase interpretability and modularity.
Pitfalls arise if the architecture introduces heavy nonlinearity or discontinuities that hamper gradient-based sampling. Also, certain normalizations (like batch normalization) can lead to subtle issues if the negative samples in a batch have statistics that differ widely from real data, potentially distorting the energy estimates. Hence, some implementations avoid batchnorm for the parts of the network used to compute energies, or they use specialized variants that handle MCMC samples more gracefully.