Tag Archives: optimization

Mixed-input matrix multiplication performance optimizations

AI-driven technologies are weaving themselves into the fabric of our daily routines, with the potential to enhance our access to knowledge and boost our overall productivity. The backbone of these applications lies in large language models (LLMs). LLMs are memory-intensive and typically require specialized hardware accelerators to efficiently deliver tens of exaflops of computing power. This blog post shows how we can start addressing the computational challenges by utilizing memory more effectively.

The bulk of an LLM’s memory and compute are consumed by weights in matrix multiplication operations. Using narrower data types reduces memory consumption. For example, storing weights in the 8-bit integer (i.e., U8 or S8) data type reduces the memory footprint by 4× relative to single-precision (F32) and 2× relative to half-precision (F16) or bfloat16 (BF16). Furthermore, previous work has shown that LLM models running matrix multiplications with weights in S8 and input in F16 (preserving higher precision of the user-input) is an effective method for increasing the efficiency with acceptable trade-offs in accuracy. This technique is known as weight-only quantization and requires efficient implementation of matrix multiplication with mixed-inputs, e.g., half-precision input multiplied with 8-bits integer. Hardware accelerators, including GPUs, support a fixed set of data types, and thus, mixed-input matrix multiplication requires software transformations to map to the hardware operations.

To that end, in this blog we focus on mapping mixed-input matrix multiplication onto the NVIDIA Ampere architecture. We present software techniques addressing data type conversion and layout conformance to map mixed-input matrix multiplication efficiently onto hardware-supported data types and layouts. Our results show that the overhead of additional work in software is minimal and enables performance close to the peak hardware capabilities. The software techniques described here are released in the open-source NVIDIA/CUTLASS repository.

Memory footprint for an 175B parameter LLM model with various data types formats.

The matrix-multiply-accumulate operation

Modern AI hardware accelerators such as Google’s TPU and NVIDIA’s GPU multiply matrices natively in the hardware by targeting Tensor Cores, which are specialized processing elements to accelerate matrix operations, particularly for AI workloads. In this blog, we focus on NVIDIA Ampere Tensor Cores, which provide the matrix-multiply-accumulate (mma) operation. For the rest of the blog the reference to mma is for Ampere Tensor Cores. The supported data types, shapes, and data layout of the two input matrices (called operands) for the mma operation are fixed in hardware. This means that matrix multiplications with various data types and larger shapes are implemented in the software by tiling the problem onto hardware-supported data types, shapes, and layouts.

The Tensor Core mma operation is defined by specifying two input matrices (e.g., A & B, shown below) to produce a result matrix, C. The mma operation natively supports mixed-precision. Mixed-precision Tensor Cores allow mixing input (A and B) data type with the result (C) data type. In contrast, mixed-input matrix multiplication involves mixing the input data types, and it is not supported by the hardware, so it needs to be implemented in the software.

Tensor Core operation of M-by-N-by-K on input matrix A of M-by-K and matrix B of K-by-N produces output matrix C of M-by-N.

Challenges of mixed-input matrix multiplication

To simplify the discussion, we restrict to a specific example of mixed-input matrix multiplication: F16 for user input and U8 for the model weights (written as F16 * U8). The techniques described here work for various combinations of mixed-input data types.

A GPU programmer can access a hierarchy of memory, including global memory, shared memory, and registers, which are arranged in order of decreasing capacity but increasing speed. NVIDIA Ampere Tensor Core mma operations consume input matrices from registers. Furthermore, input and output matrices are required to conform to a layout of data within a group of 32 threads known as a warp. The supported data type and layout within a warp are fixed for an mma operation, so to implement mixed-input multiplication efficiently, it is necessary to solve the challenges of data type conversion and layout conformance in software.


Data type conversion

The mma operation requires two input matrices with the same data type. Thus, mixed-input matrix multiplication, where one of the operands is stored in U8 in global memory and other in F16, requires a data type conversion from U8 to F16. The conversion will bring two operands to F16, mapping the mixed-input matrix multiplication to hardware-supported mixed-precision Tensor Cores. Given the large number of weights, there are a large number of such operations, and our techniques show how to reduce their latency and improve performance.


Layout conformance

The mma operation also requires the layout of two input matrices, within the registers of a warp, to be conformat with hardware specification. The layout for the input matrix B of U8 data type in mixed-input matrix multiplication (F16 * U8) needs to conform with the converted F16 data type. This is called layout conformance and needs to be achieved in the software.

The figure below shows an mma operation consuming matrix A and matrix B from registers to produce matrix C in registers, distributed across one warp. The thread T0 is highlighted and zoomed in to show the weight matrix B goes through data type conversion and needs a layout conformance to be able to map to the hardware-supported Tensor Core operation.

The mapping of mixed-input (F32 = F16 * U8) operation in software to natively supported warp-level Tensor Cores in hardware (F32 = F16 * F16). (Original figure source Developing CUDA kernels to push Tensor Cores to the Absolute Limit on NVIDIA A100.)

Software strategies addressing challenges

A typical data type conversion involves a sequence of operations on 32-bit registers, shown below. Each rectangular block represents a register and the adjoining text are the operations. The entire sequence shows the conversion from 4xU8 to 2x(2xF16). The sequence involves roughly 10 operations.

NumericArrayConvertor from 4xU8 to 2x(2xF16) in 32-bit registers.

There are many ways of achieving layout conformance. Two of the existing solutions are:

  1. Narrower bitwidth shared memory loads: In this approach, threads issue narrow bitwidth memory loads moving the U8 data from shared memory to registers. This results in two 32-bit registers, with each register containing 2xF16 values (shown above for the matrix B’s thread T0). The narrower shared memory load achieves layout conformance directly into registers without needing any shuffles; however, it does not utilize the full shared memory bandwidth.
  2. Pre-processing in global memory: An alternative strategy involves rearranging the data within the global memory (one level above the shared memory in memory hierarchy), allowing wider shared memory loads. This approach maximizes the shared memory bandwidth utilization and ensures that the data is loaded in a conformant layout directly in the registers. Although the rearrangement process can be executed offline prior to the LLM deployment, ensuring no impact on the application performance, it introduces an additional, non-trivial hardware-specific pre-processing step that requires an extra program to rearrange the data. NVIDIA/FasterTransformer adopts this method to effectively address layout conformance challenges.

Optimized software strategies

To further optimize and reduce the overhead of data type conversion and layout conformance, we have implemented FastNumericArrayConvertor and FragmentShuffler, respectively.

FastNumericArrayConvertor operates on 4xU8 in 32-bit registers without unpacking individual 1xU8 values. Furthermore, it uses less expensive arithmetic operations which reduces the number of instructions and increases the speed of the conversion.

The conversion sequence for U8-to-F16 is shown below. The operations use packed 32b registers, avoiding explicit unpacking and packing. FastNumericArrayConvertor uses the permute byte to rearrange bytes of 4xU8 into two registers. Additionally, FastNumericArrayConvertor does not use expensive integer to floating-point conversion instructions and employs vectorized operations to obtain the packed results in two 32-bit registers containing 2x(2xF16) values. The FastNumericArrayConvertor for U8-to-F16 approximately uses six operations, a 1.6× reduction relative to the approach shown above.

FastNumericArrayConvertor utilizes permute bytes and packed arithmetic, reducing the number of instructions in the data type conversion.

FragmentShuffler handles the layout conformance by shuffling data in a way that allows the use of wider bitwidth load operation, increasing shared memory bandwidth utilization and reducing the total number of operations.

NVIDIA Ampere architecture provides a load matrix instruction (ldmatrix). The ldmatrix is a warp-level operation, where 32 threads of a warp move the data from shared memory to registers in the shape and layout that mma matrix A and B consume. The use of ldmatrix reduces the number of load instructions and increases the memory bandwidth utilization. Since the ldmatrix instruction moves U8 data to registers, the layout after the load conforms with U8*U8 mma operation, and not with F16*F16 mma operation. We implemented FragmentShuffler to rearrange the data within registers using shuffle (shfl.sync) operations to achieve the layout conformance.

The most significant contribution of this work is to achieve layout conformance through register shuffles, avoiding offline pre-processing in global memory or narrower bitwidth shared memory loads. Furthermore, we provide implementations for FastNumericArrayConvertor covering data type conversion from U8-to-F16, S8-to-F16, U8-to-BF16, and S8-to-BF16.


Performance results

We measured the performance of eight mixed-input variants of our method (shown below in blue and red; varying the data types of matrix A and B) and two mixed-precision data types (shown in green) on an NVIDIA A100 SXM chip. The performance results are shown in FLOPS (higher is better). Notably, the first eight matrix-multipications require additional operations relative to the last two, because the mixed-precision variants directly target hardware-accelerated Tensor Core operations and do not need data type conversion and layout conformance. Even so, our approach demonstrates mixed-input matrix multiplication performance only slightly below or on par with mixed-precision.

Mixed-input matrix multiplication performance on NVIDIA A100 40GB SMX4 chip for a compute-bound matrix problem shape m=3456, n=4096, k=2048.

Acknowledgements

We would like to mention several folks who have contributed through technical brainstorming and improving the blog post including, Quentin Colombet, Jacques Pienaar, Allie Culp, Calin Cascaval, Ashish Gondimalla, Matt Walsh, Marek Kolodziej, and Aman Bhatia. We would like to thank our NVIDIA partners Rawn Henry, Pradeep Ramani, Vijay Thakkar, Haicheng Wu, Andrew Kerr, Matthew Nicely, and Vartika Singh.

Source: Google AI Blog


Summary report optimization in the Privacy Sandbox Attribution Reporting API

In recent years, the Privacy Sandbox initiative was launched to explore responsible ways for advertisers to measure the effectiveness of their campaigns, by aiming to deprecate third-party cookies (subject to resolving any competition concerns with the UK’s Competition and Markets Authority). Cookies are small pieces of data containing user preferences that websites store on a user’s device; they can be used to provide a better browsing experience (e.g., allowing users to automatically sign in) and to serve relevant content or ads. The Privacy Sandbox attempts to address concerns around the use of cookies for tracking browsing data across the web by providing a privacy-preserving alternative.

Many browsers use differential privacy (DP) to provide privacy-preserving APIs, such as the Attribution Reporting API (ARA), that don’t rely on cookies for ad conversion measurement. ARA encrypts individual user actions and collects them in an aggregated summary report, which estimates measurement goals like the number and value of conversions (useful actions on a website, such as making a purchase or signing up for a mailing list) attributed to ad campaigns.

The task of configuring API parameters, e.g., allocating a contribution budget across different conversions, is important for maximizing the utility of the summary reports. In “Summary Report Optimization in the Privacy Sandbox Attribution Reporting API”, we introduce a formal mathematical framework for modeling summary reports. Then, we formulate the problem of maximizing the utility of summary reports as an optimization problem to obtain the optimal ARA parameters. Finally, we evaluate the method using real and synthetic datasets, and demonstrate significantly improved utility compared to baseline non-optimized summary reports.


ARA summary reports

We use the following example to illustrate our notation. Imagine a fictional gift shop called Du & Penc that uses digital advertising to reach its customers. The table below captures their holiday sales, where each record contains impression features with (i) an impression ID, (ii) the campaign, and (iii) the city in which the ad was shown, as well as conversion features with (i) the number of items purchased and (ii) the total dollar value of those items.

Impression and conversion feature logs for Du & Penc.


Mathematical model

ARA summary reports can be modeled by four algorithms: (1) Contribution Vector, (2) Contribution Bounding, (3) Summary Reports, and (4) Reconstruct Values. Contribution Bounding and Summary Reports are performed by the ARA, while Contribution Vector and Reconstruct Values are performed by an AdTech provider — tools and systems that enable businesses to buy and sell digital advertising. The objective of this work is to assist AdTechs in optimizing summary report algorithms.

The Contribution Vector algorithm converts measurements into an ARA format that is discretized and scaled. Scaling needs to account for the overall contribution limit per impression. Here we propose a method that clips and performs randomized rounding. The outcome of the algorithm is a histogram of aggregatable keys and values.

Next, the Contribution Bounding algorithm runs on client devices and enforces the contribution bound on attributed reports where any further contributions exceeding the limit are dropped. The output is a histogram of attributed conversions.

The Summary Reports algorithm runs on the server side inside a trusted execution environment and returns noisy aggregate results that satisfy DP. Noise is sampled from the discrete Laplace distribution, and to enforce privacy budgeting, a report may be queried only once.

Finally, the Reconstruct Values algorithm converts measurements back to the original scale. Reconstruct Values and Contribution Vector Algorithms are designed by the AdTech, and both impact the utility received from the summary report.

Illustrative usage of ARA summary reports, which include Contribution Vector (Algorithm A), Contribution Bounding (Algorithm C), Summary Reports (Algorithm S), and Reconstruct Values (Algorithm R). Algorithms C and S are fixed in the API. The AdTech designs A and R.


Error metrics

There are several factors to consider when selecting an error metric for evaluating the quality of an approximation. To choose a particular metric, we considered the desirable properties of an error metric that further can be used as an objective function. Considering desired properties, we have chosen 𝜏-truncated root mean square relative error (RMSRE𝜏) as our error metric for its properties. See the paper for a detailed discussion and comparison to other possible metrics.


Optimization

To optimize utility as measured by RMSRE𝜏, we choose a capping parameter, C, and privacy budget, 𝛼, for each slice. The combination of both determines how an actual measurement (such as two conversions with a total value of $3) is encoded on the AdTech side and then passed to the ARA for Contribution Bounding algorithm processing. RMSRE𝜏 can be computed exactly, since it can be expressed in terms of the bias from clipping and the variance of the noise distribution. Following those steps we find out that RMSRE𝜏 for a fixed privacy budget, 𝛼, or a capping parameter, C, is convex (so the error-minimizing value for the other parameter can be obtained efficiently), while for joint variables (C, 𝛼) it becomes non-convex (so we may not always be able to select the best possible parameters). In any case, any off-the-shelf optimizer can be used to select privacy budgets and capping parameters. In our experiments, we use the SLSQP minimizer from the scipy.optimize library.


Synthetic data

Different ARA configurations can be evaluated empirically by testing them on a conversion dataset. However, access to such data can be restricted or slow due to privacy concerns, or simply unavailable. One way to address these limitations is to use synthetic data that replicates the characteristics of real data.

We present a method for generating synthetic data responsibly through statistical modeling of real-world conversion datasets. We first perform an empirical analysis of real conversion datasets to uncover relevant characteristics for ARA. We then design a pipeline that uses this distribution knowledge to create a realistic synthetic dataset that can be customized via input parameters.

The pipeline first generates impressions drawn from a power-law distribution (step 1), then for each impression it generates conversions drawn from a Poisson distribution (step 2) and finally, for each conversion, it generates conversion values drawn from a log-normal distribution (step 3). With dataset-dependent parameters, we find that these distributions closely match ad-dataset characteristics. Thus, one can learn parameters from historical or public datasets and generate synthetic datasets for experimentation.

Overall dataset generation steps with features for illustration.


Experimental evaluation

We evaluate our algorithms on three real-world datasets (Criteo, AdTech Real Estate, and AdTech Travel) and three synthetic datasets. Criteo consists of 15M clicks, Real Estate consists of 100K conversions, and Travel consists of 30K conversions. Each dataset is partitioned into a training set and a test set. The training set is used to choose contribution budgets, clipping threshold parameters, and the conversion count limit (the real-world datasets have only one conversion per click), and the error is evaluated on the test set. Each dataset is partitioned into slices using impression features. For real-world datasets, we consider three queries for each slice; for synthetic datasets, we consider two queries for each slice.

For each query we choose the RMSRE𝝉 𝜏 value to be five times the median value of the query on the training dataset. This ensures invariance of the error metric to data rescaling, and allows us to combine the errors from features of different scales by using 𝝉 per each feature.

Scatter plots of real-world datasets illustrating the probability of observing a conversion value. The fitted curves represent best log-normal distribution models that effectively capture the underlying patterns in the data.


Results

We compare our optimization-based algorithm to a simple baseline approach. For each query, the baseline uses an equal contribution budget and a fixed quantile of the training data to choose the clipping threshold. Our algorithms produce substantially lower error than baselines on both real-world and synthetic datasets. Our optimization-based approach adapts to the privacy budget and data.

RMSREτ for privacy budgets {1, 2, 4, 8, 16, 32, 64} for our algorithms and baselines on three real-world and three synthetic datasets. Our optimization-based approach consistently achieves lower error than baselines that use a fixed quantile for the clipping threshold and split the contribution budget equally among the queries.


Conclusion

We study the optimization of summary reports in the ARA, which is currently deployed on hundreds of millions of Chrome browsers. We present a rigorous formulation of the contribution budgeting optimization problem for ARA with the goal of equipping researchers with a robust abstraction that facilitates practical improvements.

Our recipe, which leverages historical data to bound and scale the contributions of future data under differential privacy, is quite general and applicable to settings beyond advertising. One approach based on this work is to use past data to learn the parameters of the data distribution, and then to apply synthetic data derived from this distribution for privacy budgeting for queries on future data. Please see the paper and accompanying code for detailed algorithms and proofs.


Acknowledgements

This work was done in collaboration with Badih Ghazi, Pritish Kamath, Ravi Kumar, Pasin Manurangsi, and Avinash Varadarajan. We thank Akash Nadan for his help.

Source: Google AI Blog


The secret to Android’s improved memory on 1B+ Devices: The latest Android Runtime update

Posted by Santiago Aboy Solanes - Software Engineer

The Android Runtime (ART) executes Dalvik bytecode produced from apps and system services written in the Java or Kotlin languages. We constantly improve ART to generate smaller and more performant code. Improving ART makes the system and user-experience better as a whole, as it is the common denominator in Android apps. In this blog post we will talk about optimizations that reduce code size without impacting performance.

Code size is one of the key metrics we look at, since smaller generated files are better for memory (both RAM and storage). With the new release of ART, we estimate saving users about 50-100MB per device. This could be just the thing you need to be able to update your favorite app, or download a new one. Since ART has been updateable from Android 12, these optimizations reach 1B+ devices for whom we are saving 47-95 petabytes (47-95 millions of GB!) globally!

All the improvements mentioned in this blog post are open source. They are available through the ART mainline update so you don’t even need a full OS update to reap the benefits. We can have our upside-down cake and eat it too!

Optimizing compiler 101

ART compiles applications from the DEX format to native code using the on-device dex2oat tool. The first step is to parse the DEX code and generate an Intermediate Representation (IR). Using the IR, dex2oat performs a number of code optimizations. The last step of the pipeline is a code generation phase where dex2oat converts the IR into native code (for example, AArch64 assembly).

The optimization pipeline has phases that execute in order that each concentrate on a particular set of optimizations. For example, Constant Folding is an optimization that tries to replace instructions with constant values like folding the addition operation 2 + 3 into a 5.

ART's optimization pipeline overview with an example showing we can combine the addition of 2 plus 3 into a 5

The IR can be printed and visualized, but is very verbose compared to Kotlin language code. For the purposes of this blog post, we will show what the optimizations do using Kotlin language code, but know that they are happening to IR code.

Code size improvements

For all code size optimizations, we tested our optimizations on over half a million APKs present in the Google Play Store and aggregated the results.

Eliminating write barriers

We have a new optimization pass that we named Write Barrier Elimination. Write barriers track modified objects since the last time they were examined by the Garbage Collector (GC) so that the GC can revisit them. For example, if we have:

Example showing that we can eliminate redundant write barriers if there a GC cannot happen between  set instructionsPreviously, we would emit a write barrier for each object modification but we only need a single write barrier because: 1) the mark will be set in o itself (and not in the inner objects), and 2) a garbage collection can't have interacted with the thread between those sets.

If an instruction may trigger a GC (for example, Invokes and SuspendChecks), we wouldn't be able to eliminate write barriers. In the following example, we can't guarantee a GC won't need to examine or modify the tracking information between the modifications:

Example showing that we can't eliminate redundant write barriers because a GC may happen between set instructionsImplementing this new pass contributes to 0.8% code size reduction.

Implicit suspend checks

Let's assume we have several threads running. Suspend checks are safepoints (represented by the houses in the image below) where we can pause the thread execution. Safepoints are used for many reasons, the most important of them being Garbage Collection. When a safepoint call is issued, the threads must go into a safepoint and are blocked until they are released.

The previous implementation was an explicit boolean check. We would load the value, test it, and branch into the safepoint if needed.

Shows the explicit suspend check (load + test + branch) when multiple threads are running

Implicit suspend checks is an optimization that eliminates the need for the test and branch instructions. Instead, we only have one load: if the thread needs to suspend, that load will trap and the signal handler will redirect the code to a suspend check handler as if the method made a call.

Shows the implicit suspend check (two loads: the first one loads null and the second one traps) when multiple threads are running

Going into a bit more detail, a reserved register rX is pre-loaded with an address within the thread where we have a pointer pointing to itself. As long as we don't need to do a suspend check, we keep that self-pointing pointer. When we need to do a suspend check, we clear the pointer and once it becomes visible to the thread the first LDR rX, [rX] will load null and the second will segfault.

The suspend request is essentially asking the thread to suspend some time soon, so the minor delay of having to wait for the second load is okay.

This optimization reduces code size by 1.8%.

Coalescing returns

It is common for compiled methods to have an entry frame. If they have it, those methods have to deconstruct it when they return, which is also known as an exit frame. If a method has several return instructions, it will generate several exit frames, one per return instruction.

By coalescing the return instructions into one, we are able to have one return point and are able to remove the extra exit frames. This is especially useful for switch cases with multiple return statements. Switch case optimized by having one return instead of multiple return instructions

Coalescing returns reduces code size by 1%.

Other optimization improvements

We improved a lot of our existing optimization passes. For this blog post, we will group them up in the same section, but they are independent of each other. All the optimizations in the following section contribute to a 5.7% code size reduction.

Code Sinking

Code sinking is an optimization pass that pushes down instructions to uncommon branches like paths that end with a throw. This is done to reduce wasted cycles on instructions that are likely not going to be used.

We improved code sinking in graphs with try catches: we now allow sinking code as long as we don't sink it inside of a different try than the one it started in (or inside of any try if it wasn't in one to begin with).

Code sinking optimizations in the presence of a try catch

In the first example, we can sink the Object creation since it will only be used in the if(flag) path and not the other and it is within the same try. With this change, at runtime it will only be run if flag is true. Without getting into too much technical detail, what we can sink is the actual object creation, but loading the Object class still remains before the if. This is hard to show with Kotlin code, as the same Kotlin line turns into several instructions at the ART compiler level.

In the second example, we cannot sink the code as we would be moving an instance creation (which may throw) inside of another try.

Code Sinking is mostly a runtime performance optimization, but it can help reduce the register pressure. By moving instructions closer to their uses, we can use fewer registers in some cases. Using fewer registers means fewer move instructions, which ends up helping code size.

Loop optimization

Loop optimization helps eliminate loops at compile time. In the following example, the loop in foo will multiply a by 10, 10 times. This is the same as multiplying by 100. We enabled loop optimization to work in graphs with try catches.

Loop optimization in the presence of a try catch

In foo, we can optimize the loop since the try catch is unrelated.

In bar or baz, however, we don't optimize it. It is not trivial to know the path the loop will take if we have a try in the loop, or if the loop as a whole is inside of a try.

Dead code elimination – Remove unneeded try blocks

We improved our dead code elimination phase by implementing an optimization that removes try blocks that don't contain throwing instructions. We are also able to remove some catch blocks, as long as no live try blocks point to it.

In the following example, we inline bar into foo. After that, we know that the division cannot throw. Later optimization passes can leverage this and improve the code.

We can remove tries which contain no throwing instructions

Just removing the dead code from the try catch is good enough, but even better is the fact that in some cases we allow other optimizations to take place. If you recall, we don't do loop optimization when the loop has a try, or it's inside of one. By eliminating this redundant try/catch, we can loop optimize producing smaller and faster code.

Another example of removing tries which contain no throwing instructions

Dead code elimination – SimplifyAlwaysThrows

During the dead code elimination phase, we have an optimization we call SimplifyAlwaysThrows. If we detect that an invoke will always throw, we can safely discard whatever code we have after that method call since it will never be executed.

We also updated SimplifyAlwaysThrows to work in graphs with try catches, as long as the invoke itself is not inside of a try. If it is inside of a try, we might jump to a catch block, and it gets harder to figure out the exact path that will be executed.

We can use the SimplifyAlwaysThrows optimization as long as the invoke itself is not inside of a try

We also improved:

  • Detection when an invoke will always throw by looking at their parameters. On the left, we will mark divide(1, 0) as always throwing even when the generic method doesn't always throw.
  • SimplifyAlwaysThrows to work on all invokes. Previously we had restrictions for example don't do it for invokes leading to an if, but we could remove all of the restrictions.

We improved detection, and removed some of the restrictions from this optimization

Load Store Elimination – Working with try catch blocks

Load store elimination (LSE) is an optimization pass that removes redundant loads and stores.

We improved this pass to work with try catches in the graph. In foo, we can see that we can do LSE normally if the stores/loads are not directly interacting with the try. In bar, we can see an example where we either go through the normal path and don't throw, in which case we return 1; or we throw, catch it and return 2. Since the value is known for every path, we can remove the redundant load.

Examples showing we can perform Load Store Elimination in graphs with try catches, as long as the instructions are not inside of a try

Load Store Elimination – Working with release/acquire operations

We improved our load store elimination pass to work in graphs with release/acquire operations. These are volatile loads, stores, and monitor operations. To clarify, this means that we allow LSE to work in graphs that have those operations, but we don't remove said operations.

In the example, i and j are regular ints, and vi is a volatile int. In foo, we can skip loading the values since there's not a release/acquire operation between the sets and the loads. In bar, the volatile operation happens between them so we can't eliminate the normal loads. Note that it doesn't matter that the volatile load result is not used—we cannot eliminate the acquire operation.

Examples showing that we can perform LSE in graphs with release/acquire operations. Note that the release/acquire operations themselves are not removed since they are needed for synchronization.

This optimization works similarly with volatile stores and monitor operations (which are synchronized blocks in Kotlin).

New inliner heuristic

Our inliner pass has a wide range of heuristics. Sometimes we decide not to inline a method because it is too big, or sometimes to force inlining of a method because it is too small (for example, empty methods like Object initialization).

We implemented a new inliner heuristic: Don't inline invokes leading to a throw. If we know we are going to throw we will skip inlining those methods, as throwing itself is costly enough that inlining that code path is not worth it.

We had three families of methods that we are skipping to inline:

  • Calculating and printing debug information before a throw.
  • Inlining the error constructor itself.
  • Finally blocks are duplicated in our optimizing compiler. We have one for the normal case (i.e. the try doesn't throw), and one for the exceptional case. We do this because in the exceptional case we have to: catch, execute the finally block, and rethrow. The methods in the exceptional case will now not be inlined, but the ones in the normal case will.

Examples showing:
* calculating and printing debug information before a throw
* inlining the error constructor itself
* methods in finally blocks

Constant folding

Constant folding is the optimization pass that changes operations into constants if possible. We implemented an optimization that propagates variables known to be constant when used in if guards. Having more constants in the graph allows us to perform more optimizations later.

In foo, we know that a has the value 2 in the if guard. We can propagate that information and deduce that b must be 4. In a similar vein, in bar we know that cond must be true in the if case and false in the else case (simplifying the graphs).

Example showing that if we know that a variable has a constant value within the scope of an 
 `if`, we will propagate that example within said scope

Putting it all together

If we take into account all code size optimizations in this blog post we achieved a 9.3% code size reduction!

To put things into perspective, an average phone can have 500MB-1GB in optimized code (The actual number can be higher or lower, depending on how many apps you have installed, and which particular apps you installed), so these optimizations save about 50-100MB per device. Since these optimizations reach 1B+ devices, we are saving 47-95 petabytes globally!

Further reading

If you are interested in the code changes themselves, feel free to take a look. All the improvements mentioned in this blog post are open source. If you want to help Android users worldwide, consider contributing to the Android Open Source Project!

  • Write barrier elimination: 1
  • Implicit suspend checks: 1
  • Coalescing returns: 1
  • Code sinking: 1, 2
  • Loop optimization: 1
  • Dead code elimination: 1, 2, 3, 4
  • Load store elimination: 1, 2, 3, 4
  • New inlining heuristic: 1
  • Constant folding: 1

Java is a trademark or registered trademark of Oracle and/or its affiliates

Re-weighted gradient descent via distributionally robust optimization

Deep neural networks (DNNs) have become essential for solving a wide range of tasks, from standard supervised learning (image classification using ViT) to meta-learning. The most commonly-used paradigm for learning DNNs is empirical risk minimization (ERM), which aims to identify a network that minimizes the average loss on training data points. Several algorithms, including stochastic gradient descent (SGD), Adam, and Adagrad, have been proposed for solving ERM. However, a drawback of ERM is that it weights all the samples equally, often ignoring the rare and more difficult samples, and focusing on the easier and abundant samples. This leads to suboptimal performance on unseen data, especially when the training data is scarce.

To overcome this challenge, recent works have developed data re-weighting techniques for improving ERM performance. However, these approaches focus on specific learning tasks (such as classification) and/or require learning an additional meta model that predicts the weights of each data point. The presence of an additional model significantly increases the complexity of training and makes them unwieldy in practice.

In “Stochastic Re-weighted Gradient Descent via Distributionally Robust Optimization” we introduce a variant of the classical SGD algorithm that re-weights data points during each optimization step based on their difficulty. Stochastic Re-weighted Gradient Descent (RGD) is a lightweight algorithm that comes with a simple closed-form expression, and can be applied to solve any learning task using just two lines of code. At any stage of the learning process, RGD simply reweights a data point as the exponential of its loss. We empirically demonstrate that the RGD reweighting algorithm improves the performance of numerous learning algorithms across various tasks, ranging from supervised learning to meta learning. Notably, we show improvements over state-of-the-art methods on DomainBed and Tabular classification. Moreover, the RGD algorithm also boosts performance for BERT using the GLUE benchmarks and ViT on ImageNet-1K.


Distributionally robust optimization

Distributionally robust optimization (DRO) is an approach that assumes a “worst-case” data distribution shift may occur, which can harm a model's performance. If a model has focussed on identifying few spurious features for prediction, these “worst-case” data distribution shifts could lead to the misclassification of samples and, thus, a performance drop. DRO optimizes the loss for samples in that “worst-case” distribution, making the model robust to perturbations (e.g., removing a small fraction of points from a dataset, minor up/down weighting of data points, etc.) in the data distribution. In the context of classification, this forces the model to place less emphasis on noisy features and more emphasis on useful and predictive features. Consequently, models optimized using DRO tend to have better generalization guarantees and stronger performance on unseen samples.

Inspired by these results, we develop the RGD algorithm as a technique for solving the DRO objective. Specifically, we focus on Kullback–Leibler divergence-based DRO, where one adds perturbations to create distributions that are close to the original data distribution in the KL divergence metric, enabling a model to perform well over all possible perturbations.

Figure illustrating DRO. In contrast to ERM, which learns a model that minimizes expected loss over original data distribution, DRO learns a model that performs well on several perturbed versions of the original data distribution.


Stochastic re-weighted gradient descent

Consider a random subset of samples (called a mini-batch), where each data point has an associated loss Li. Traditional algorithms like SGD give equal importance to all the samples in the mini-batch, and update the parameters of the model by descending along the averaged gradients of the loss of those samples. With RGD, we reweight each sample in the mini-batch and give more importance to points that the model identifies as more difficult. To be precise, we use the loss as a proxy to calculate the difficulty of a point, and reweight it by the exponential of its loss. Finally, we update the model parameters by descending along the weighted average of the gradients of the samples.

Due to stability considerations, in our experiments we clip and scale the loss before computing its exponential. Specifically, we clip the loss at some threshold T, and multiply it with a scalar that is inversely proportional to the threshold. An important aspect of RGD is its simplicity as it doesn’t rely on a meta model to compute the weights of data points. Furthermore, it can be implemented with two lines of code, and combined with any popular optimizers (such as SGD, Adam, and Adagrad.

Figure illustrating the intuitive idea behind RGD in a binary classification setting. Feature 1 and Feature 2 are the features available to the model for predicting the label of a data point. RGD upweights the data points with high losses that have been misclassified by the model.


Results

We present empirical results comparing RGD with state-of-the-art techniques on standard supervised learning and domain adaptation (refer to the paper for results on meta learning). In all our experiments, we tune the clipping level and the learning rate of the optimizer using a held-out validation set.


Supervised learning

We evaluate RGD on several supervised learning tasks, including language, vision, and tabular classification. For the task of language classification, we apply RGD to the BERT model trained on the General Language Understanding Evaluation (GLUE) benchmark and show that RGD outperforms the BERT baseline by +1.94% with a standard deviation of 0.42%. To evaluate RGD’s performance on vision classification, we apply RGD to the ViT-S model trained on the ImageNet-1K dataset, and show that RGD outperforms the ViT-S baseline by +1.01% with a standard deviation of 0.23%. Moreover, we perform hypothesis tests to confirm that these results are statistically significant with a p-value that is less than 0.05.

RGD’s performance on language and vision classification using GLUE and Imagenet-1K benchmarks. Note that MNLI, QQP, QNLI, SST-2, MRPC, RTE and COLA are diverse datasets which comprise the GLUE benchmark.

For tabular classification, we use MET as our baseline, and consider various binary and multi-class datasets from UC Irvine's machine learning repository. We show that applying RGD to the MET framework improves its performance by 1.51% and 1.27% on binary and multi-class tabular classification, respectively, achieving state-of-the-art performance in this domain.


Performance of RGD for classification of various tabular datasets.


Domain generalization

To evaluate RGD’s generalization capabilities, we use the standard DomainBed benchmark, which is commonly used to study a model’s out-of-domain performance. We apply RGD to FRR, a recent approach that improved out-of-domain benchmarks, and show that RGD with FRR performs an average of 0.7% better than the FRR baseline. Furthermore, we confirm with hypothesis tests that most benchmark results (except for Office Home) are statistically significant with a p-value less than 0.05.

Performance of RGD on DomainBed benchmark for distributional shifts.


Class imbalance and fairness

To demonstrate that models learned using RGD perform well despite class imbalance, where certain classes in the dataset are underrepresented, we compare RGD’s performance with ERM on long-tailed CIFAR-10. We report that RGD improves the accuracy of baseline ERM by an average of 2.55% with a standard deviation of 0.23%. Furthermore, we perform hypothesis tests and confirm that these results are statistically significant with a p-value of less than 0.05.

Performance of RGD on the long-tailed Cifar-10 benchmark for class imbalance domain.


Limitations

The RGD algorithm was developed using popular research datasets, which were already curated to remove corruptions (e.g., noise and incorrect labels). Therefore, RGD may not provide performance improvements in scenarios where training data has a high volume of corruptions. A potential approach to handle such scenarios is to apply an outlier removal technique to the RGD algorithm. This outlier removal technique should be capable of filtering out outliers from the mini-batch and sending the remaining points to our algorithm.


Conclusion

RGD has been shown to be effective on a variety of tasks, including out-of-domain generalization, tabular representation learning, and class imbalance. It is simple to implement and can be seamlessly integrated into existing algorithms with just two lines of code change. Overall, RGD is a promising technique for boosting the performance of DNNs, and could help push the boundaries in various domains.


Acknowledgements

The paper described in this blog post was written by Ramnath Kumar, Arun Sai Suggala, Dheeraj Nagaraj and Kushal Majmundar. We extend our sincere gratitude to the anonymous reviewers, Prateek Jain, Pradeep Shenoy, Anshul Nasery, Lovish Madaan, and the numerous dedicated members of the machine learning and optimization team at Google Research India for their invaluable feedback and contributions to this work.

Source: Google AI Blog


Neural network pruning with combinatorial optimization

Modern neural networks have achieved impressive performance across a variety of applications, such as language, mathematical reasoning, and vision. However, these networks often use large architectures that require lots of computational resources. This can make it impractical to serve such models to users, especially in resource-constrained environments like wearables and smartphones. A widely used approach to mitigate the inference costs of pre-trained networks is to prune them by removing some of their weights, in a way that doesn’t significantly affect utility. In standard neural networks, each weight defines a connection between two neurons. So after weights are pruned, the input will propagate through a smaller set of connections and thus requires less computational resources.

Original network vs. a pruned network.

Pruning methods can be applied at different stages of the network’s training process: post, during, or before training (i.e., immediately after weight initialization). In this post, we focus on the post-training setting: given a pre-trained network, how can we determine which weights should be pruned? One popular method is magnitude pruning, which removes weights with the smallest magnitude. While efficient, this method doesn’t directly consider the effect of removing weights on the network’s performance. Another popular paradigm is optimization-based pruning, which removes weights based on how much their removal impacts the loss function. Although conceptually appealing, most existing optimization-based approaches seem to face a serious tradeoff between performance and computational requirements. Methods that make crude approximations (e.g., assuming a diagonal Hessian matrix) can scale well, but have relatively low performance. On the other hand, while methods that make fewer approximations tend to perform better, they appear to be much less scalable.

In “Fast as CHITA: Neural Network Pruning with Combinatorial Optimization”, presented at ICML 2023, we describe how we developed an optimization-based approach for pruning pre-trained neural networks at scale. CHITA (which stands for “Combinatorial Hessian-free Iterative Thresholding Algorithm”) outperforms existing pruning methods in terms of scalability and performance tradeoffs, and it does so by leveraging advances from several fields, including high-dimensional statistics, combinatorial optimization, and neural network pruning. For example, CHITA can be 20x to 1000x faster than state-of-the-art methods for pruning ResNet and improves accuracy by over 10% in many settings.


Overview of contributions

CHITA has two notable technical improvements over popular methods:

  • Efficient use of second-order information: Pruning methods that use second-order information (i.e., relating to second derivatives) achieve the state of the art in many settings. In the literature, this information is typically used by computing the Hessian matrix or its inverse, an operation that is very difficult to scale because the Hessian size is quadratic with respect to the number of weights. Through careful reformulation, CHITA uses second-order information without having to compute or store the Hessian matrix explicitly, thus allowing for more scalability.
  • Combinatorial optimization: Popular optimization-based methods use a simple optimization technique that prunes weights in isolation, i.e., when deciding to prune a certain weight they don’t take into account whether other weights have been pruned. This could lead to pruning important weights because weights deemed unimportant in isolation may become important when other weights are pruned. CHITA avoids this issue by using a more advanced, combinatorial optimization algorithm that takes into account how pruning one weight impacts others.

In the sections below, we discuss CHITA’s pruning formulation and algorithms.


A computation-friendly pruning formulation

There are many possible pruning candidates, which are obtained by retaining only a subset of the weights from the original network. Let k be a user-specified parameter that denotes the number of weights to retain. Pruning can be naturally formulated as a best-subset selection (BSS) problem: among all possible pruning candidates (i.e., subsets of weights) with only k weights retained, the candidate that has the smallest loss is selected.

Pruning as a BSS problem: among all possible pruning candidates with the same total number of weights, the best candidate is defined as the one with the least loss. This illustration shows four candidates, but this number is generally much larger.

Solving the pruning BSS problem on the original loss function is generally computationally intractable. Thus, similar to previous work, such as OBD and OBS, we approximate the loss with a quadratic function by using a second-order Taylor series, where the Hessian is estimated with the empirical Fisher information matrix. While gradients can be typically computed efficiently, computing and storing the Hessian matrix is prohibitively expensive due to its sheer size. In the literature, it is common to deal with this challenge by making restrictive assumptions on the Hessian (e.g., diagonal matrix) and also on the algorithm (e.g., pruning weights in isolation).

CHITA uses an efficient reformulation of the pruning problem (BSS using the quadratic loss) that avoids explicitly computing the Hessian matrix, while still using all the information from this matrix. This is made possible by exploiting the low-rank structure of the empirical Fisher information matrix. This reformulation can be viewed as a sparse linear regression problem, where each regression coefficient corresponds to a certain weight in the neural network. After obtaining a solution to this regression problem, coefficients set to zero will correspond to weights that should be pruned. Our regression data matrix is (n x p), where n is the batch (sub-sample) size and p is the number of weights in the original network. Typically n << p, so storing and operating with this data matrix is much more scalable than common pruning approaches that operate with the (p x p) Hessian.

CHITA reformulates the quadratic loss approximation, which requires an expensive Hessian matrix, as a linear regression (LR) problem. The LR’s data matrix is linear in p, which makes the reformulation more scalable than the original quadratic approximation.


Scalable optimization algorithms

CHITA reduces pruning to a linear regression problem under the following sparsity constraint: at most k regression coefficients can be nonzero. To obtain a solution to this problem, we consider a modification of the well-known iterative hard thresholding (IHT) algorithm. IHT performs gradient descent where after each update the following post-processing step is performed: all regression coefficients outside the Top-k (i.e., the k coefficients with the largest magnitude) are set to zero. IHT typically delivers a good solution to the problem, and it does so iteratively exploring different pruning candidates and jointly optimizing over the weights.

Due to the scale of the problem, standard IHT with constant learning rate can suffer from very slow convergence. For faster convergence, we developed a new line-search method that exploits the problem structure to find a suitable learning rate, i.e., one that leads to a sufficiently large decrease in the loss. We also employed several computational schemes to improve CHITA’s efficiency and the quality of the second-order approximation, leading to an improved version that we call CHITA++.


Experiments

We compare CHITA’s run time and accuracy with several state-of-the-art pruning methods using different architectures, including ResNet and MobileNet.

Run time: CHITA is much more scalable than comparable methods that perform joint optimization (as opposed to pruning weights in isolation). For example, CHITA’s speed-up can reach over 1000x when pruning ResNet.

Post-pruning accuracy: Below, we compare the performance of CHITA and CHITA++ with magnitude pruning (MP), Woodfisher (WF), and Combinatorial Brain Surgeon (CBS), for pruning 70% of the model weights. Overall, we see good improvements from CHITA and CHITA++.

Post-pruning accuracy of various methods on ResNet20. Results are reported for pruning 70% of the model weights.
Post-pruning accuracy of various methods on MobileNet. Results are reported for pruning 70% of the model weights.

Next, we report results for pruning a larger network: ResNet50 (on this network, some of the methods listed in the ResNet20 figure couldn’t scale). Here we compare with magnitude pruning and M-FAC. The figure below shows that CHITA achieves better test accuracy for a wide range of sparsity levels.

Test accuracy of pruned networks, obtained using different methods.


Conclusion, limitations, and future work

We presented CHITA, an optimization-based approach for pruning pre-trained neural networks. CHITA offers scalability and competitive performance by efficiently using second-order information and drawing on ideas from combinatorial optimization and high-dimensional statistics.

CHITA is designed for unstructured pruning in which any weight can be removed. In theory, unstructured pruning can significantly reduce computational requirements. However, realizing these reductions in practice requires special software (and possibly hardware) that support sparse computations. In contrast, structured pruning, which removes whole structures like neurons, may offer improvements that are easier to attain on general-purpose software and hardware. It would be interesting to extend CHITA to structured pruning.


Acknowledgements

This work is part of a research collaboration between Google and MIT. Thanks to Rahul Mazumder, Natalia Ponomareva, Wenyu Chen, Xiang Meng, Zhe Zhao, and Sergei Vassilvitskii for their help in preparing this post and the paper. Also thanks to John Guilyard for creating the graphics in this post.

Source: Google AI Blog


Recommendations in Google Ads API Roundup

Google Ads API now supports 25 different recommendation types including the most frequently used types. With the wide array of types available and documentation with examples to help you get started, there has never been a better time to get started retrieving and applying recommendations with Google Ads API!

Key suggested uses
Recommendations provide customized suggestions to help increase your campaigns' performance. Recommendations can introduce you to new, relevant features, help you get more out of your budget by improving your bidding, keywords, and ads, and can increase the overall performance and efficiency of your campaigns. Here are a few examples of how recommendations can help with the management of your account:
  • Avoid getting limited by budget this holiday season. With CAMPAIGN_BUDGET and FORECASTING_CAMPAIGN_BUDGET recommendation types you’ll be sure to keep your ads running, so potential customers can see them by preventing your campaign from under-performing due to a limited budget.
  • Expand the reach of your ads with USE_BROAD_MATCH_KEYWORD, which will update your keyword match types to show your ads to relevant potential customers.
  • Upgrade to Performance Max with UPGRADE_SMART_SHOPPING_CAMPAIGN_TO_PERFORMANCE_MAX and UPGRADE_LOCAL_CAMPAIGN_TO_PERFORMANCE_MAX, which will take care of migrating your existing Smart Shopping and Local Campaigns for you.
Implementation guide
To help you get started, check out our implementation guide and YouTube deep dive tutorial for tips.

Code examples
We’ve also put together these code examples to save you time getting up to speed with Recommendations in Google Ads API.

Enhancing Backpropagation via Local Loss Optimization

While model design and training data are key ingredients in a deep neural network’s (DNN’s) success, less-often discussed is the specific optimization method used for updating the model parameters (weights). Training DNNs involves minimizing a loss function that measures the discrepancy between the ground truth labels and the model’s predictions. Training is carried out by backpropagation, which adjusts the model weights via gradient descent steps. Gradient descent, in turn, updates the weights by using the gradient (i.e., derivative) of the loss with respect to the weights.

The simplest weight update corresponds to stochastic gradient descent, which, in every step, moves the weights in the negative direction with respect to the gradients (with an appropriate step size, a.k.a. the learning rate). More advanced optimization methods modify the direction of the negative gradient before updating the weights by using information from the past steps and/or the local properties (such as the curvature information) of the loss function around the current weights. For instance, a momentum optimizer encourages moving along the average direction of past updates, and the AdaGrad optimizer scales each coordinate based on the past gradients. These optimizers are commonly known as first-order methods since they generally modify the update direction using only information from the first-order derivative (i.e., gradient). More importantly, the components of the weight parameters are treated independently from each other.

More advanced optimization, such as Shampoo and K-FAC, capture the correlations between gradients of parameters and have been shown to improve convergence, reducing the number of iterations and improving the quality of the solution. These methods capture information about the local changes of the derivatives of the loss, i.e., changes in gradients. Using this additional information, higher-order optimizers can discover much more efficient update directions for training models by taking into account the correlations between different groups of parameters. On the downside, calculating higher-order update directions is computationally more expensive than first-order updates. The operation uses more memory for storing statistics and involves matrix inversion, thus hindering the applicability of higher-order optimizers in practice.

In “LocoProp: Enhancing BackProp via Local Loss Optimization”, we introduce a new framework for training DNN models. Our new framework, LocoProp, conceives neural networks as a modular composition of layers. Generally, each layer in a neural network applies a linear transformation on its inputs, followed by a non-linear activation function. In the new construction, each layer is allotted its own weight regularizer, output target, and loss function. The loss function of each layer is designed to match the activation function of the layer. Using this formulation, training minimizes the local losses for a given mini-batch of examples, iteratively and in parallel across layers. Our method performs multiple local updates per batch of examples using a first-order optimizer (like RMSProp), which avoids computationally expensive operations such as the matrix inversions required for higher-order optimizers. However, we show that the combined local updates look rather like a higher-order update. Empirically, we show that LocoProp outperforms first-order methods on a deep autoencoder benchmark and performs comparably to higher-order optimizers, such as Shampoo and K-FAC, without the high memory and computation requirements.

Method
Neural networks are generally viewed as composite functions that transform model inputs into output representations, layer by layer. LocoProp adopts this view while decomposing the network into layers. In particular, instead of updating the weights of the layer to minimize the loss function at the output, LocoProp applies pre-defined local loss functions specific to each layer. For a given layer, the loss function is selected to match the activation function, e.g., a tanh loss would be selected for a layer with a tanh activation. Each layerwise loss measures the discrepancy between the layer's output (for a given mini-batch of examples) and a notion of a target output for that layer. Additionally, a regularizer term ensures that the updated weights do not drift too far from the current values. The combined layerwise loss function (with a local target) plus regularizer is used as the new objective function for each layer.

Similar to backpropagation, LocoProp applies a forward pass to compute the activations. In the backward pass, LocoProp sets per neuron "targets" for each layer. Finally, LocoProp splits model training into independent problems across layers where several local updates can be applied to each layer's weights in parallel.

Perhaps the simplest loss function one can think of for a layer is the squared loss. While the squared loss is a valid choice of a loss function, LocoProp takes into account the possible non-linearity of the activation functions of the layers and applies layerwise losses tailored to the activation function of each layer. This enables the model to emphasize regions at the input that are more important for the model prediction while deemphasizing the regions that do not affect the output as much. Below we show examples of tailored losses for the tanh and ReLU activation functions.

Loss functions induced by the (left) tanh and (right) ReLU activation functions. Each loss is more sensitive to the regions affecting the output prediction. For instance, ReLU loss is zero as long as both the prediction (â) and the target (a) are negative. This is because the ReLU function applied to any negative number equals zero.

After forming the objective in each layer, LocoProp updates the layer weights by repeatedly applying gradient descent steps on its objective. The update typically uses a first-order optimizer (like RMSProp). However, we show that the overall behavior of the combined updates closely resembles higher-order updates (shown below). Thus, LocoProp provides training performance close to what higher-order optimizers achieve without the high memory or computation needed for higher-order methods, such as matrix inverse operations. We show that LocoProp is a flexible framework that allows the recovery of well-known algorithms and enables the construction of new algorithms via different choices of losses, targets, and regularizers. LocoProp’s layerwise view of neural networks also allows updating the weights in parallel across layers.

Experiments
In our paper, we describe experiments on the deep autoencoder model, which is a commonly used baseline for evaluating the performance of optimization algorithms. We perform extensive tuning on multiple commonly used first-order optimizers, including SGD, SGD with momentum, AdaGrad, RMSProp, and Adam, as well as the higher-order Shampoo and K-FAC optimizers, and compare the results with LocoProp. Our findings indicate that the LocoProp method performs significantly better than first-order optimizers and is comparable to those of higher-order, while being significantly faster when run on a single GPU.

Train loss vs. number of epochs (left) and wall-clock time, i.e., the real time that passes during training, (right) for RMSProp, Shampoo, K-FAC, and LocoProp on the deep autoencoder model.

Summary and Future Directions
We introduced a new framework, called LocoProp, for optimizing deep neural networks more efficiently. LocoProp decomposes neural networks into separate layers with their own regularizer, output target, and loss function and applies local updates in parallel to minimize the local objectives. While using first-order updates for the local optimization problems, the combined updates closely resemble higher-order update directions, both theoretically and empirically.

LocoProp provides flexibility to choose the layerwise regularizers, targets, and loss functions. Thus, it allows the development of new update rules based on these choices. Our code for LocoProp is available online on GitHub. We are currently working on scaling up ideas induced by LocoProp to much larger scale models; stay tuned!

Acknowledgments
We would like to thank our co-author, Manfred K. Warmuth, for his critical contributions and inspiring vision. We would like to thank Sameer Agarwal for discussions looking at this work from a composite functions perspective, Vineet Gupta for discussions and development of Shampoo, Zachary Nado on K-FAC, Tom Small for development of the animation used in this blogpost and finally, Yonghui Wu and Zoubin Ghahramani for providing us with a nurturing research environment in the Google Brain Team.

Source: Google AI Blog


MLGO: A Machine Learning Framework for Compiler Optimization

The question of how to compile faster and smaller code arose together with the birth of modem computers. Better code optimization can significantly reduce the operational cost of large datacenter applications. The size of compiled code matters the most to mobile and embedded systems or software deployed on secure boot partitions, where the compiled binary must fit in tight code size budgets. With advances in the field, the headroom has been heavily squeezed with increasingly complicated heuristics, impeding maintenance and further improvements.

Recent research has shown that machine learning (ML) can unlock more opportunities in compiler optimization by replacing complicated heuristics with ML policies. However, adopting ML in general-purpose, industry-strength compilers remains a challenge.

To address this, we introduce “MLGO: a Machine Learning Guided Compiler Optimizations Framework”, the first industrial-grade general framework for integrating ML techniques systematically in LLVM (an open-source industrial compiler infrastructure that is ubiquitous for building mission-critical, high-performance software). MLGO uses reinforcement learning (RL) to train neural networks to make decisions that can replace heuristics in LLVM. We describe two MLGO optimizations for LLVM: 1) reducing code size with inlining; and 2) improving code performance with register allocation (regalloc). Both optimizations are available in the LLVM repository, and have been deployed in production.

How Does MLGO Work? With Inlining-for-Size As a Case Study
Inlining helps reduce code size by making decisions that enable the removal of redundant code. In the example below, the caller function foo() calls the callee function bar(), which itself calls baz(). Inlining both callsites returns a simple foo() function that reduces the code size.

Inlining reduces code size by removing redundant code.

In real code, there are thousands of functions calling each other, and thus comprise a call graph. During the inlining phase, the compiler traverses over the call graph on all caller-callee pairs, and makes decisions on whether to inline a caller-callee pair or not. It is a sequential decision process as previous inlining decisions will alter the call graph, affecting later decisions and the final result. In the example above, the call graph foo()bar()baz() needs a “yes” decision on both edges to make the code size reduction happen.

Before MLGO, the inline / no-inline decision was made by a heuristic that, over time, became increasingly difficult to improve. MLGO substitutes the heuristic with an ML model. During the call graph traversal, the compiler seeks advice from a neural network on whether to inline a particular caller-callee pair by feeding in relevant features (i.e., inputs) from the graph, and executes the decisions sequentially until the whole call graph is traversed.

Illustration of MLGO during inlining. “#bbs”, “#users”, and “callsite height” are example caller-callee pair features.

MLGO trains the decision network (policy) with RL using policy gradient and evolution strategies algorithms. While there is no ground truth about best decisions, online RL iterates between training and running compilation with the trained policy to collect data and improve the policy. In particular, given the current model under training, the compiler consults the model for inline / no-inline decision making during the inlining stage. After the compilation finishes, it produces a log of the sequential decision process (state, action, reward). The log is then passed to the trainer to update the model. This process repeats until we obtain a satisfactory model.

Compiler behavior during training. The compiler compiles the source code foo.cpp to an object file foo.o with a sequence of optimization passes, one of which is the inline pass.

The trained policy is then embedded into the compiler to provide inline / no-inline decisions during compilation. Unlike the training scenario, the policy does not produce a log. The TensorFlow model is embedded with XLA AOT, which converts the model into executable code. This avoids TensorFlow runtime dependency and overhead, minimizing the extra time and memory cost introduced by ML model inference at compilation time.

Compiler behavior in production.

We trained the inlining-for-size policy on a large internal software package containing 30k modules. The trained policy is generalizable when applied to compile other software and achieves a 3% ~ 7% size reduction. In addition to the generalizability across software, generalizability across time is also important — both the software and compiler are under active development so the trained policy needs to retain good performance for a reasonable time. We evaluated the model’s performance on the same set of software three months later and found only slight degradation.

Inlining-for-size policy size reduction percentages. The x-axis presents different software and the y-axis represents the percentage size reduction. “Training” is the software on which the model was trained and “Infra[1|2|3]” are different internal software packages.

The MLGO inlining-for-size training has been deployed on Fuchsia — a general purpose open source operating system designed to power a diverse ecosystem of hardware and software, where binary size is critical. Here, MLGO showed a 6.3% size reduction for C++ translation units.

Register-Allocation (for performance)
As a general framework, we used MLGO to improve the register allocation pass, which improves the code performance in LLVM. Register Allocation solves the problem of assigning physical registers to live ranges (i.e., variables).

As the code executes, different live ranges are completed at different times, freeing up registers for use by subsequent processing stages. In the example below, each “add” and “multiply” instruction requires all operands and the result to be in physical registers. The live range x is allocated to the green register and is completed before either live ranges in the blue or yellow registers. After x is completed, the green register becomes available and is assigned to live range t.

Register allocation example.

When it's time to allocate live range q, there are no available registers, so the register allocation pass must decide which (if any) live range can be "evicted" from its register to make room for q. This is referred to as the “live range eviction” problem, and is the decision for which we train the model to replace original heuristics. In this particular example, it evicts z from the yellow register, and assigns it to q and the first half of z.

We now consider the unassigned second half of live range z. We have a conflict again, and this time the live range t is evicted and split, and the first half of t and the final part of z end up using the green register. The middle part of z corresponds to the instruction q = t * y, where z is not being used, so it is not assigned to any register and its value is stored in the stack from the yellow register, which later gets reloaded to the green register. The same happens to t. This adds extra load/store instructions to the code and degrades performance. The goal of the register allocation algorithm is to reduce such inefficiencies as much as possible. This is used as the reward to guide RL policy training.

Similar to the inlining-for-size policy, the register allocation (regalloc-for-performance) policy is trained on a large Google internal software package, and is generalizable across different software, with 0.3% ~1.5% improvements in queries per second (QPS) on a set of internal large-scale datacenter applications. The QPS improvement has persisted for months after its deployment, showing the model’s generalizability across the time horizon.

Conclusion and Future Work
We propose MLGO, a framework for integrating ML techniques systematically in an industrial compiler, LLVM. MLGO is a general framework that can be expanded to be: 1) deeper, e.g., adding more features, and applying better RL algorithms; and 2) broader, by applying it to more optimization heuristics beyond inlining and regalloc. We are enthusiastic about the possibilities MLGO can bring to the compiler optimization domain and look forward to its further adoption and to future contributions from the research community.

Try it Yourself
Check out the open-sourced end-to-end data collection and training solution on github and a demo that uses policy gradient to train an inlining-for-size policy.

Acknowledgements
We’d like to thank MLGO’s contributors and collaborators Eugene Brevdo, Jacob Hegna, Gaurav Jain, David Li, Zinan Lin, Kshiteej Mahajan, Jack Morris, Girish Mururu, Jin Xin Ng, Robert Ormandi, Easwaran Raman, Ondrej Sykora, Maruf Zaber, Weiye Zhao. We would also like to thank Petr Hosek, Yuqian Li, Roland McGrath, Haowei Wu for trusting us and deploying MLGO in Fuchsia as MLGO’s very first customer; thank David Blaikie, Eric Christopher, Brooks Moses, Jordan Rupprecht for helping to deploy MLGO in Google internal large-scale datacenter applications; and thank Ed Chi, Tipp Moseley for their leadership support.

Source: Google AI Blog


Offline Optimization for Architecting Hardware Accelerators

Advances in machine learning (ML) often come with advances in hardware and computing systems. For example, the growth of ML-based approaches in solving various problems in vision and language has led to the development of application-specific hardware accelerators (e.g., Google TPUs and Edge TPUs). While promising, standard procedures for designing accelerators customized towards a target application require manual effort to devise a reasonably accurate simulator of hardware, followed by performing many time-intensive simulations to optimize the desired objective (e.g., optimizing for low power usage or latency when running a particular application). This involves identifying the right balance between total amount of compute and memory resources and communication bandwidth under various design constraints, such as the requirement to meet an upper bound on chip area usage and peak power. However, designing accelerators that meet these design constraints is often result in infeasible designs. To address these challenges, we ask: “Is it possible to train an expressive deep neural network model on large amounts of existing accelerator data and then use the learned model to architect future generations of specialized accelerators, eliminating the need for computationally expensive hardware simulations?

In “Data-Driven Offline Optimization for Architecting Hardware Accelerators”, accepted at ICLR 2022, we introduce PRIME, an approach focused on architecting accelerators based on data-driven optimization that only utilizes existing logged data (e.g., data leftover from traditional accelerator design efforts), consisting of accelerator designs and their corresponding performance metrics (e.g., latency, power, etc) to architect hardware accelerators without any further hardware simulation. This alleviates the need to run time-consuming simulations and enables reuse of data from past experiments, even when the set of target applications changes (e.g., an ML model for vision, language, or other objective), and even for unseen but related applications to the training set, in a zero-shot fashion. PRIME can be trained on data from prior simulations, a database of actually fabricated accelerators, and also a database of infeasible or failed accelerator designs1. This approach for architecting accelerators — tailored towards both single- and multi-applications — improves performance upon state-of-the-art simulation-driven methods by about 1.2x-1.5x, while considerably reducing the required total simulation time by 93% and 99%, respectively. PRIME also architects effective accelerators for unseen applications in a zero-shot setting, outperforming simulation-based methods by 1.26x.

PRIME uses logged accelerator data, consisting of both feasible and infeasible accelerators, to train a conservative model, which is used to design accelerators while meeting design constraints. PRIME architects accelerators with up to 1.5x smaller latency, while reducing the required hardware simulation time by up to 99%.

The PRIME Approach for Architecting Accelerators
Perhaps the simplest possible way to use a database of previously designed accelerators for hardware design is to use supervised machine learning to train a prediction model that can predict the performance objective for a given accelerator as input. Then, one could potentially design new accelerators by optimizing the performance output of this learned model with respect to the input accelerator design. Such an approach is known as model-based optimization. However, this simple approach has a key limitation: it assumes that the prediction model can accurately predict the cost for every accelerator that we might encounter during optimization! It is well established that most prediction models trained via supervised learning misclassify adversarial examples that “fool” the learned model into predicting incorrect values. Similarly, it has been shown that even optimizing the output of a supervised model finds adversarial examples that look promising under the learned model2, but perform terribly under the ground truth objective.

To address this limitation, PRIME learns a robust prediction model that is not prone to being fooled by adversarial examples (that we will describe shortly), which would be otherwise found during optimization. One can then simply optimize this model using any standard optimizer to architect simulators. More importantly, unlike prior methods, PRIME can also utilize existing databases of infeasible accelerators to learn what not to design. This is done by augmenting the supervised training of the learned model with additional loss terms that specifically penalize the value of the learned model on the infeasible accelerator designs and adversarial examples during training. This approach resembles a form of adversarial training.

In principle, one of the central benefits of a data-driven approach is that it should enable learning highly expressive and generalist models of the optimization objective that generalize over target applications, while also potentially being effective for new unseen applications for which a designer has never attempted to optimize accelerators. To train PRIME so that it generalizes to unseen applications, we modify the learned model to be conditioned on a context vector that identifies a given neural net application we wish to accelerate (as we discuss in our experiments below, we choose to use high-level features of the target application: such as number of feed-forward layers, number of convolutional layers, total parameters, etc. to serve as the context), and train a single, large model on accelerator data for all applications designers have seen so far. As we will discuss below in our results, this contextual modification of PRIME enables it to optimize accelerators both for multiple, simultaneous applications and new unseen applications in a zero-shot fashion.

Does PRIME Outperform Custom-Engineered Accelerators?
We evaluate PRIME on a variety of actual accelerator design tasks. We start by comparing the optimized accelerator design architected by PRIME targeted towards nine applications to the manually optimized EdgeTPU design. EdgeTPU accelerators are primarily optimized towards running applications in image classification, particularly MobileNetV2, MobileNetV3 and MobileNetEdge. Our goal is to check if PRIME can design an accelerator that attains a lower latency than a baseline EdgeTPU accelerator3, while also constraining the chip area to be under 27 mm2 (the default for the EdgeTPU accelerator). Shown below, we find that PRIME improves latency over EdgeTPU by 2.69x (up to 11.84x in t-RNN Enc), while also reducing the chip area usage by 1.50x (up to 2.28x in MobileNetV3), even though it was never trained to reduce chip area! Even on the MobileNet image-classification models, for which the custom-engineered EdgeTPU accelerator was optimized, PRIME improves latency by 1.85x.

Comparing latencies (lower is better) of accelerator designs suggested by PRIME and EdgeTPU for single-model specialization.
The chip area (lower is better) reduction compared to a baseline EdgeTPU design for single-model specialization.

Designing Accelerators for New and Multiple Applications, Zero-Shot
We now study how PRIME can use logged accelerator data to design accelerators for (1) multiple applications, where we optimize PRIME to design a single accelerator that works well across multiple applications simultaneously, and in a (2) zero-shot setting, where PRIME must generate an accelerator for new unseen application(s) without training on any data from such applications. In both settings, we train the contextual version of PRIME, conditioned on context vectors identifying the target applications and then optimize the learned model to obtain the final accelerator. We find that PRIME outperforms the best simulator-driven approach in both settings, even when very limited data is provided for training for a given application but many applications are available. Specifically in the zero-shot setting, PRIME outperforms the best simulator-driven method we compared to, attaining a reduction of 1.26x in latency. Further, the difference in performance increases as the number of training applications increases.

The average latency (lower is better) of test applications under zero-shot setting compared to a state-of-the-art simulator-driven approach. The text on top of each bar shows the set of training applications.

Closely Analyzing an Accelerator Designed by PRIME
To provide more insight to hardware architecture, we examine the best accelerator designed by PRIME and compare it to the best accelerator found by the simulator-driven approach. We consider the setting where we need to jointly optimize the accelerator for all nine applications, MobileNetEdge, MobileNetV2, MobileNetV3, M4, M5, M64, t-RNN Dec, and t-RNN Enc, and U-Net, under a chip area constraint of 100 mm2. We find that PRIME improves latency by 1.35x over the simulator-driven approach.

Per application latency (lower is better) for the best accelerator design suggested by PRIME and state-of-the-art simulator-driven approach for a multi-task accelerator design. PRIME reduces the average latency across all nine applications by 1.35x over the simulator-driven method.

As shown above, while the latency of the accelerator designed by PRIME for MobileNetEdge, MobileNetV2, MobileNetV3, M4, t-RNN Dec, and t-RNN Enc are better, the accelerator found by the simulation-driven approach yields a lower latency in M5, M6, and U-Net. By closely inspecting the accelerator configurations, we find that PRIME trades compute (64 cores for PRIME vs. 128 cores for the simulator-driven approach) for larger Processing Element (PE) memory size (2,097,152 bytes vs. 1,048,576 bytes). These results show that PRIME favors PE memory size to accommodate the larger memory requirements in t-RNN Dec and t-RNN Enc, where large reductions in latency were possible. Under a fixed area budget, favoring larger on-chip memory comes at the expense of lower compute power in the accelerator. This reduction in the accelerator's compute power leads to higher latency for the models with large numbers of compute operations, namely M5, M6, and U-Net.

Conclusion
The efficacy of PRIME highlights the potential for utilizing the logged offline data in an accelerator design pipeline. A likely avenue for future work is to scale this approach across an array of applications, where we expect to see larger gains because simulator-driven approaches would need to solve a complex optimization problem, akin to searching for needle in a haystack, whereas PRIME can benefit from generalization of the surrogate model. On the other hand, we would also note that PRIME outperforms prior simulator-driven methods we utilize and this makes it a promising candidate to be used within a simulator-driven method. More generally, training a strong offline optimization algorithm on offline datasets of low-performing designs can be a highly effective ingredient in at the very least, kickstarting hardware design, versus throwing out prior data. Finally, given the generality of PRIME, we hope to use it for hardware-software co-design, which exhibits a large search space but plenty of opportunity for generalization. We have also released both the code for training PRIME and the dataset of accelerators.

Acknowledgments
We thank our co-authors Sergey Levine, Kevin Swersky, and Milad Hashemi for their advice, thoughts and suggestions. We thank James Laudon, Cliff Young, Ravi Narayanaswami, Berkin Akin, Sheng-Chun Kao, Samira Khan, Suvinay Subramanian, Stella Aslibekyan, Christof Angermueller, and Olga Wichrowskafor for their help and support, and Sergey Levine for feedback on this blog post. In addition, we would like to extend our gratitude to the members of “Learn to Design Accelerators”, “EdgeTPU”, and the Vizier team for providing invaluable feedback and suggestions. We would also like to thank Tom Small for the animated figure used in this post.


1The infeasible accelerator designs stem from build errors in silicon or compilation/mapping failures. 
2This is akin to adversarial examples in supervised learning – these examples are close to the data points observed in the training dataset, but are misclassified by the classifier. 
3The performance metrics for the baseline EdgeTPU accelerator are extracted from an industry-based hardware simulator tuned to match the performance of the actual hardware. 
4These are proprietary object-detection models, and we refer to them as M4 (indicating Model 4), M5, and M6 in the paper. 

Source: Google AI Blog


A New Library for Network Optimization

Networks are all around us from the electrical circuits inside our computers to the multitude of internet servers that route packets of data around the globe. Even the web itself is a network of pages connected to each other by a myriad of blue links.

A network of rubber bands on a wooden pegboard


A network's structure is referred to as its topology. Network topologies can be physical or logical, centralized or decentralized, and fully or partially connected. Given a network with n nodes, the number of possible topologies grows exponentially with n; even just a dozen nodes admit nearly a trillion trillion possible configurations!

The network-opt logo


We are pleased to announce the open source release of network-opt, a C++ library that supports the optimization of network topologies. Using sophisticated techniques for combinatorial search, this algorithm can efficiently construct instances from a family of so-called series-parallel networks that commonly arise in electrical and telecommunications applications. For example, given 15 1-Ω resistors and a target resistance of π Ω, our tool can produce a circuit that achieves six digits of precision:

A 15-element circuit composed of 1-ohm resistors


For more details, refer to our paper: "Search Strategies for Topological Network Optimization," appearing this month at the Thirty-Sixth AAAI Conference on Artificial Intelligence (AAAI-22).


By Michael D. Moffitt, Core Enterprise Machine Learning