Tag Archives: reinforcement learning

World scale inverse reinforcement learning in Google Maps

Routing in Google Maps remains one of our most helpful and frequently used features. Determining the best route from A to B requires making complex trade-offs between factors including the estimated time of arrival (ETA), tolls, directness, surface conditions (e.g., paved, unpaved roads), and user preferences, which vary across transportation mode and local geography. Often, the most natural visibility we have into travelers' preferences is by analyzing real-world travel patterns.

Learning preferences from observed sequential decision making behavior is a classic application of inverse reinforcement learning (IRL). Given a Markov decision process (MDP) — a formalization of the road network — and a set of demonstration trajectories (the traveled routes), the goal of IRL is to recover the users' latent reward function. Although past research has created increasingly general IRL solutions, these have not been successfully scaled to world-sized MDPs. Scaling IRL algorithms is challenging because they typically require solving an RL subroutine at every update step. At first glance, even attempting to fit a world-scale MDP into memory to compute a single gradient step appears infeasible due to the large number of road segments and limited high bandwidth memory. When applying IRL to routing, one needs to consider all reasonable routes between each demonstration's origin and destination. This implies that any attempt to break the world-scale MDP into smaller components cannot consider components smaller than a metropolitan area.

To this end, in "Massively Scalable Inverse Reinforcement Learning in Google Maps", we share the result of a multi-year collaboration among Google Research, Maps, and Google DeepMind to surpass this IRL scalability limitation. We revisit classic algorithms in this space, and introduce advances in graph compression and parallelization, along with a new IRL algorithm called Receding Horizon Inverse Planning (RHIP) that provides fine-grained control over performance trade-offs. The final RHIP policy achieves a 16–24% relative improvement in global route match rate, i.e., the percentage of de-identified traveled routes that exactly match the suggested route in Google Maps. To the best of our knowledge, this represents the largest instance of IRL in a real world setting to date.

Google Maps improvements in route match rate relative to the existing baseline, when using the RHIP inverse reinforcement learning policy.


The benefits of IRL

A subtle but crucial detail about the routing problem is that it is goal conditioned, meaning that every destination state induces a slightly different MDP (specifically, the destination is a terminal, zero-reward state). IRL approaches are well suited for these types of problems because the learned reward function transfers across MDPs, and only the destination state is modified. This is in contrast to approaches that directly learn a policy, which typically require an extra factor of S parameters, where S is the number of MDP states.

Once the reward function is learned via IRL, we take advantage of a powerful inference-time trick. First, we evaluate the entire graph's rewards once in an offline batch setting. This computation is performed entirely on servers without access to individual trips, and operates only over batches of road segments in the graph. Then, we save the results to an in-memory database and use a fast online graph search algorithm to find the highest reward path for routing requests between any origin and destination. This circumvents the need to perform online inference of a deeply parameterized model or policy, and vastly improves serving costs and latency.

Reward model deployment using batch inference and fast online planners.


Receding Horizon Inverse Planning

To scale IRL to the world MDP, we compress the graph and shard the global MDP using a sparse Mixture of Experts (MoE) based on geographic regions. We then apply classic IRL algorithms to solve the local MDPs, estimate the loss, and send gradients back to the MoE. The worldwide reward graph is computed by decompressing the final MoE reward model. To provide more control over performance characteristics, we introduce a new generalized IRL algorithm called Receding Horizon Inverse Planning (RHIP).

IRL reward model training using MoE parallelization, graph compression, and RHIP.

RHIP is inspired by people’s tendency to perform extensive local planning ("What am I doing for the next hour?") and approximate long-term planning ("What will my life look like in 5 years?"). To take advantage of this insight, RHIP uses robust yet expensive stochastic policies in the local region surrounding the demonstration path, and switches to cheaper deterministic planners beyond some horizon. Adjusting the horizon H allows controlling computational costs, and often allows the discovery of the performance sweet spot. Interestingly, RHIP generalizes many classic IRL algorithms and provides the novel insight that they can be viewed along a stochastic vs. deterministic spectrum (specifically, for H=∞ it reduces to MaxEnt, for H=1 it reduces to BIRL, and for H=0 it reduces to MMP).

Given a demonstration from so to sd, (1) RHIP follows a robust yet expensive stochastic policy in the local region surrounding the demonstration (blue region). (2) Beyond some horizon H, RHIP switches to following a cheaper deterministic planner (red lines). Adjusting the horizon enables fine-grained control over performance and computational costs.


Routing wins

The RHIP policy provides a 15.9% and 24.1% lift in global route match rate for driving and two-wheelers (e.g., scooters, motorcycles, mopeds) relative to the well-tuned Maps baseline, respectively. We're especially excited about the benefits to more sustainable transportation modes, where factors beyond journey time play a substantial role. By tuning RHIP's horizon H, we're able to achieve a policy that is both more accurate than all other IRL policies and 70% faster than MaxEnt.

Our 360M parameter reward model provides intuitive wins for Google Maps users in live A/B experiments. Examining road segments with a large absolute difference between the learned rewards and the baseline rewards can help improve certain Google Maps routes. For example:

Nottingham, UK. The preferred route (blue) was previously marked as private property due to the presence of a large gate, which indicated to our systems that the road may be closed at times and would not be ideal for drivers. As a result, Google Maps routed drivers through a longer, alternate detour instead (red). However, because real-world driving patterns showed that users regularly take the preferred route without an issue (as the gate is almost never closed), IRL now learns to route drivers along the preferred route by placing a large positive reward on this road segment.


Conclusion

Increasing performance via increased scale – both in terms of dataset size and model complexity – has proven to be a persistent trend in machine learning. Similar gains for inverse reinforcement learning problems have historically remained elusive, largely due to the challenges with handling practically sized MDPs. By introducing scalability advancements to classic IRL algorithms, we're now able to train reward models on problems with hundreds of millions of states, demonstration trajectories, and model parameters, respectively. To the best of our knowledge, this is the largest instance of IRL in a real-world setting to date. See the paper to learn more about this work.


Acknowledgements

This work is a collaboration across multiple teams at Google. Contributors to the project include Matthew Abueg, Oliver Lange, Matt Deeds, Jason Trader, Denali Molitor, Markus Wulfmeier, Shawn O'Banion, Ryan Epp, Renaud Hartert, Rui Song, Thomas Sharp, Rémi Robert, Zoltan Szego, Beth Luan, Brit Larabee and Agnieszka Madurska.

We’d also like to extend our thanks to Arno Eigenwillig, Jacob Moorman, Jonathan Spencer, Remi Munos, Michael Bloesch and Arun Ahuja for valuable discussions and suggestions.

Source: Google AI Blog


Preference learning with automated feedback for cache eviction

Caching is a ubiquitous idea in computer science that significantly improves the performance of storage and retrieval systems by storing a subset of popular items closer to the client based on request patterns. An important algorithmic piece of cache management is the decision policy used for dynamically updating the set of items being stored, which has been extensively optimized over several decades, resulting in several efficient and robust heuristics. While applying machine learning to cache policies has shown promising results in recent years (e.g., LRB, LHD, storage applications), it remains a challenge to outperform robust heuristics in a way that can generalize reliably beyond benchmarks to production settings, while maintaining competitive compute and memory overheads.

In “HALP: Heuristic Aided Learned Preference Eviction Policy for YouTube Content Delivery Network”, presented at NSDI 2023, we introduce a scalable state-of-the-art cache eviction framework that is based on learned rewards and uses preference learning with automated feedback. The Heuristic Aided Learned Preference (HALP) framework is a meta-algorithm that uses randomization to merge a lightweight heuristic baseline eviction rule with a learned reward model. The reward model is a lightweight neural network that is continuously trained with ongoing automated feedback on preference comparisons designed to mimic the offline oracle. We discuss how HALP has improved infrastructure efficiency and user video playback latency for YouTube’s content delivery network.


Learned preferences for cache eviction decisions

The HALP framework computes cache eviction decisions based on two components: (1) a neural reward model trained with automated feedback via preference learning, and (2) a meta-algorithm that combines a learned reward model with a fast heuristic. As the cache observes incoming requests, HALP continuously trains a small neural network that predicts a scalar reward for each item by formulating this as a preference learning method via pairwise preference feedback. This aspect of HALP is similar to reinforcement learning from human feedback (RLHF) systems, but with two important distinctions:

  • Feedback is automated and leverages well-known results about the structure of offline optimal cache eviction policies.
  • The model is learned continuously using a transient buffer of training examples constructed from the automated feedback process.

The eviction decisions rely on a filtering mechanism with two steps. First, a small subset of candidates is selected using a heuristic that is efficient, but suboptimal in terms of performance. Then, a re-ranking step optimizes from within the baseline candidates via the sparing use of a neural network scoring function to “boost” the quality of the final decision.

As a production ready cache policy implementation, HALP not only makes eviction decisions, but also subsumes the end-to-end process of sampling pairwise preference queries used to efficiently construct relevant feedback and update the model to power eviction decisions.


A neural reward model

HALP uses a light-weight two-layer multilayer perceptron (MLP) as its reward model to selectively score individual items in the cache. The features are constructed and managed as a metadata-only “ghost cache” (similar to classical policies like ARC). After any given lookup request, in addition to regular cache operations, HALP conducts the book-keeping (e.g., tracking and updating feature metadata in a capacity-constrained key-value store) needed to update the dynamic internal representation. This includes: (1) externally tagged features provided by the user as input, along with a cache lookup request, and (2) internally constructed dynamic features (e.g., time since last access, average time between accesses) constructed from lookup times observed on each item.

HALP learns its reward model fully online starting from a random weight initialization. This might seem like a bad idea, especially if the decisions are made exclusively for optimizing the reward model. However, the eviction decisions rely on both the learned reward model and a suboptimal but simple and robust heuristic like LRU. This allows for optimal performance when the reward model has fully generalized, while remaining robust to a temporarily uninformative reward model that is yet to generalize, or in the process of catching up to a changing environment.

Another advantage of online training is specialization. Each cache server runs in a potentially different environment (e.g., geographic location), which influences local network conditions and what content is locally popular, among other things. Online training automatically captures this information while reducing the burden of generalization, as opposed to a single offline training solution.


Scoring samples from a randomized priority queue

It can be impractical to optimize for the quality of eviction decisions with an exclusively learned objective for two reasons.

  1. Compute efficiency constraints: Inference with a learned network can be significantly more expensive than the computations performed in practical cache policies operating at scale. This limits not only the expressivity of the network and features, but also how often these are invoked during each eviction decision.
  2. Robustness for generalizing out-of-distribution: HALP is deployed in a setup that involves continual learning, where a quickly changing workload might generate request patterns that might be temporarily out-of-distribution with respect to previously seen data.

To address these issues, HALP first applies an inexpensive heuristic scoring rule that corresponds to an eviction priority to identify a small candidate sample. This process is based on efficient random sampling that approximates exact priority queues. The priority function for generating candidate samples is intended to be quick to compute using existing manually-tuned algorithms, e.g., LRU. However, this is configurable to approximate other cache replacement heuristics by editing a simple cost function. Unlike prior work, where the randomization was used to tradeoff approximation for efficiency, HALP also relies on the inherent randomization in the sampled candidates across time steps for providing the necessary exploratory diversity in the sampled candidates for both training and inference.

The final evicted item is chosen from among the supplied candidates, equivalent to the best-of-n reranked sample, corresponding to maximizing the predicted preference score according to the neural reward model. The same pool of candidates used for eviction decisions is also used to construct the pairwise preference queries for automated feedback, which helps minimize the training and inference skew between samples.

An overview of the two-stage process invoked for each eviction decision.


Online preference learning with automated feedback

The reward model is learned using online feedback, which is based on automatically assigned preference labels that indicate, wherever feasible, the ranked preference ordering for the time taken to receive future re-accesses, starting from a given snapshot in time among each queried sample of items. This is similar to the oracle optimal policy, which, at any given time, evicts an item with the farthest future access from all the items in the cache.

Generation of the automated feedback for learning the reward model.

To make this feedback process informative, HALP constructs pairwise preference queries that are most likely to be relevant for eviction decisions. In sync with the usual cache operations, HALP issues a small number of pairwise preference queries while making each eviction decision, and appends them to a set of pending comparisons. The labels for these pending comparisons can only be resolved at a random future time. To operate online, HALP also performs some additional book-keeping after each lookup request to process any pending comparisons that can be labeled incrementally after the current request. HALP indexes the pending comparison buffer with each element involved in the comparison, and recycles the memory consumed by stale comparisons (neither of which may ever get a re-access) to ensure that the memory overhead associated with feedback generation remains bounded over time.

Overview of all main components in HALP.


Results: Impact on the YouTube CDN

Through empirical analysis, we show that HALP compares favorably to state-of-the-art cache policies on public benchmark traces in terms of cache miss rates. However, while public benchmarks are a useful tool, they are rarely sufficient to capture all the usage patterns across the world over time, not to mention the diverse hardware configurations that we have already deployed.

Until recently, YouTube servers used an optimized LRU-variant for memory cache eviction. HALP increases YouTube’s memory egress/ingress — the ratio of the total bandwidth egress served by the CDN to that consumed for retrieval (ingress) due to cache misses — by roughly 12% and memory hit rate by 6%. This reduces latency for users, since memory reads are faster than disk reads, and also improves egressing capacity for disk-bounded machines by shielding the disks from traffic.

The figure below shows a visually compelling reduction in the byte miss ratio in the days following HALP’s final rollout on the YouTube CDN, which is now serving significantly more content from within the cache with lower latency to the end user, and without having to resort to more expensive retrieval that increases the operating costs.

Aggregate worldwide YouTube byte miss ratio before and after rollout (vertical dashed line).

An aggregated performance improvement could still hide important regressions. In addition to measuring overall impact, we also conduct an analysis in the paper to understand its impact on different racks using a machine level analysis, and find it to be overwhelmingly positive.


Conclusion

We introduced a scalable state-of-the-art cache eviction framework that is based on learned rewards and uses preference learning with automated feedback. Because of its design choices, HALP can be deployed in a manner similar to any other cache policy without the operational overhead of having to separately manage the labeled examples, training procedure and the model versions as additional offline pipelines common to most machine learning systems. Therefore, it incurs only a small extra overhead compared to other classical algorithms, but has the added benefit of being able to take advantage of additional features to make its eviction decisions and continuously adapt to changing access patterns.

This is the first large-scale deployment of a learned cache policy to a widely used and heavily trafficked CDN, and has significantly improved the CDN infrastructure efficiency while also delivering a better quality of experience to users.


Acknowledgements

Ramki Gummadi is now part of Google DeepMind. We would like to thank John Guilyard for help with the illustrations and Richard Schooler for feedback on this post.

Source: Google AI Blog


Using reinforcement learning for dynamic planning in open-ended conversations

As virtual assistants become ubiquitous, users increasingly interact with them to learn about new topics or obtain recommendations and expect them to deliver capabilities beyond narrow dialogues of one or two turns. Dynamic planning, namely the capability to look ahead and replan based on the flow of the conversation, is an essential ingredient for the making of engaging conversations with the deeper, open-ended interactions that users expect.

While large language models (LLMs) are now beating state-of-the-art approaches in many natural language processing benchmarks, they are typically trained to output the next best response, rather than planning ahead, which is required for multi-turn interactions. However, in the past few years, reinforcement learning (RL) has delivered incredible results addressing specific problems that involve dynamic planning, such as winning games and protein folding.

Today, we are sharing our recent advances in dynamic planning for human-to-assistant conversations, in which we enable an assistant to plan a multi-turn conversation towards a goal and adapt that plan in real-time by adopting an RL-based approach. Here we look at how to improve long interactions by applying RL to compose answers based on information extracted from reputable sources, rather than relying on content generated by a language model. We expect that future versions of this work could combine LLMs and RL in multi-turn dialogues. The deployment of RL “in the wild” in a large-scale dialogue system proved a formidable challenge due to the modeling complexity, tremendously large state and action spaces, and significant subtlety in designing reward functions.


What is dynamic planning?

Many types of conversations, from gathering information to offering recommendations, require a flexible approach and the ability to modify the original plan for the conversation based on its flow. This ability to shift gears in the middle of a conversation is known as dynamic planning, as opposed to static planning, which refers to a more fixed approach. In the conversation below, for example, the goal is to engage the user by sharing interesting facts about cool animals. To begin, the assistant steers the conversation to sharks via a sound quiz. Given the user's lack of interest in sharks, the assistant then develops an updated plan and pivots the conversation to sea lions, lions, and then cheetahs.

The assistant dynamically modifies its original plan to talk about sharks and shares facts about other animals.

Dynamic composition

To cope with the challenge of conversational exploration, we separate the generation of assistant responses into two parts: 1) content generation, which extracts relevant information from reputable sources, and 2) flexible composition of such content into assistant responses. We refer to this two-part approach as dynamic composition. Unlike LLM methods, this approach gives the assistant the ability to fully control the source, correctness, and quality of the content that it may offer. At the same time, it can achieve flexibility via a learned dialogue manager that selects and combines the most appropriate content.

In an earlier paper, “Dynamic Composition for Conversational Domain Exploration”, we describe a novel approach which consists of: (1) a collection of content providers, which offer candidates from different sources, such as news snippets, knowledge graph facts, and questions; (2) a dialogue manager; and (3) a sentence fusion module. Each assistant response is incrementally constructed by the dialogue manager, which selects candidates proposed by the content providers. The selected sequence of utterances is then fused into a cohesive response.


Dynamic planning using RL

At the core of the assistant response composition loop is a dialogue manager trained using off-policy RL, namely an algorithm that evaluates and improves a policy that is different from the policy used by the agent (in our case, the latter is based on a supervised model). Applying RL to dialogue management presents several challenges, including a large state space (as the state represents the conversation state, which needs to account for the whole conversation history) and an effectively unbounded action space (that may include all existing words or sentences in natural language).

We address these challenges using a novel RL construction. First, we leverage powerful supervised models — specifically, recurrent neural networks (RNNs) and transformers — to provide a succinct and effective dialogue state representation. These state encoders are fed with the dialogue history, composed of a sequence of user and assistant turns, and output a representation of the dialogue state in the form of a latent vector.

Second, we use the fact that a relatively small set of reasonable candidate utterances or actions can be generated by content providers at each conversation turn, and limit the action space to these. Whereas the action space is typically fixed in RL settings, because all states share the same action space, ours is a non-standard space in which the candidate actions may differ with each state, since content providers generate different actions depending on the dialogue context. This puts us in the realm of stochastic action sets, a framework that formalizes cases where the set of actions available in each state is governed by an exogenous stochastic process, which we address using Stochastic Action Q-Learning, a variant of the Q-learning approach. Q-learning is a popular off-policy RL algorithm, which does not require a model of the environment to evaluate and improve the policy. We trained our model on a corpus of crowd-compute–rated conversations obtained using a supervised dialogue manager.

Given the current dialogue history and a new user query, content providers generate candidates from which the assistant selects one. This process runs in a loop, and at the end the selected utterances are fused into a cohesive response.

Reinforcement learning model evaluation

We compared our RL dialogue manager with a launched supervised transformer model in an experiment using Google Assistant, which conversed with users about animals. A conversation starts when a user triggers the experience by asking an animal-related query (e.g., “How does a lion sound?”). The experiment was conducted using an A/B testing protocol, in which a small percentage of Assistant users were randomly sampled to interact with our RL-based assistant while other users interacted with the standard assistant.

We found that the RL dialogue manager conducts longer, more engaging conversations. It increases conversation length by 30% while improving user engagement metrics. We see an increase of 8% in cooperative responses to the assistant’s questions — e.g., “Tell me about lions,” in response to “Which animal do you want to hear about next?” Although there is also a large increase in nominally “non-cooperative” responses (e.g., “No,” as a reply to a question proposing additional content, such as “Do you want to hear more?”), this is expected as the RL agent takes more risks by asking pivoting questions. While a user may not be interested in the conversational direction proposed by the assistant (e.g., pivoting to another animal), the user will often continue to engage in a dialogue about animals.

From the non-cooperative user response in the 3rd turn (“No.”) and the query “Make a dog sound,” in the 5th turn, the assistant recognizes that the user is mostly interested in animal sounds and modifies its plan, providing sounds and sound quizzes.

In addition, some user queries contain explicit positive (e.g., “Thank you, Google,” or “I’m happy.”) or negative (e.g., “Shut up,” or “Stop.”) feedback. While an order of magnitude fewer than other queries, they offer a direct measure of user (dis)satisfaction. The RL model increases explicit positive feedback by 32% and reduces negative feedback by 18%.


Learned dynamic planning characteristics and strategies

We observe several characteristics of the (unseen) RL plan to improve user engagement while conducting longer conversations. First, the RL-based assistant ends 20% more turns in questions, prompting the user to choose additional content. It also better harnesses content diversity, including facts, sounds, quizzes, yes/no questions, open questions, etc. On average, the RL assistant uses 26% more distinct content providers per conversation than the supervised model.

Two observed RL planning strategies are related to the existence of sub-dialogues with different characteristics. Sub-dialogues about animal sounds are poorer in content and exhibit entity pivoting at every turn (i.e., after playing the sound of a given animal, we can either suggest the sound of a different animal or quiz the user about other animal sounds). In contrast, sub-dialogues involving animal facts typically contain richer content and have greater conversation depth. We observe that RL favors the richer experience of the latter, selecting 31% more fact-related content. Lastly, when restricting analysis to fact-related dialogues, the RL assistant exhibits 60% more focus-pivoting turns, that is, conversational turns that change the focus of the dialogue.

Below, we show two example conversations, one conducted by the supervised model (left) and the second by the RL model (right), in which the first three user turns are identical. With a supervised dialogue manager, after the user declined to hear about “today’s animal”, the assistant pivots back to animal sounds to maximize the immediate user satisfaction. While the conversation conducted by the RL model begins identically, it exhibits a different planning strategy to optimize the overall user engagement, introducing more diverse content, such as fun facts.

In the left conversation, conducted by the supervised model, the assistant maximizes the immediate user satisfaction. The right conversation, conducted by the RL model, shows different planning strategies to optimize the overall user engagement.

Future research and challenges

In the past few years, LLMs trained for language understanding and generation have demonstrated impressive results across multiple tasks, including dialogue. We are now exploring the use of an RL framework to empower LLMs with the capability of dynamic planning so that they can dynamically plan ahead and delight users with a more engaging experience.


Acknowledgements

The work described is co-authored by: Moonkyung Ryu, Yinlam Chow, Orgad Keller, Ido Greenberg, Avinatan Hassidim, Michael Fink, Yossi Matias, Idan Szpektor and Gal Elidan. We would like to thank: Roee Aharoni, Moran Ambar, John Anderson, Ido Cohn, Mohammad Ghavamzadeh, Lotem Golany, Ziv Hodak, Adva Levin, Fernando Pereira, Shimi Salant, Shachar Shimoni, Ronit Slyper, Ariel Stolovich, Hagai Taitelbaum, Noam Velan, Avital Zipori and the CrowdCompute team led by Ashwin Kakarla. We thank Sophie Allweis for her feedback on this blogpost and Tom Small for the visualization.

Source: Google AI Blog


IndoorSim-to-OutdoorReal: Learning to navigate outdoors without any outdoor experience

Teaching mobile robots to navigate in complex outdoor environments is critical to real-world applications, such as delivery or search and rescue. However, this is also a challenging problem as the robot needs to perceive its surroundings, and then explore to identify feasible paths towards the goal. Another common challenge is that the robot needs to overcome uneven terrains, such as stairs, curbs, or rockbed on a trail, while avoiding obstacles and pedestrians. In our prior work, we investigated the second challenge by teaching a quadruped robot to tackle challenging uneven obstacles and various outdoor terrains.

In “IndoorSim-to-OutdoorReal: Learning to Navigate Outdoors without any Outdoor Experience”, we present our recent work to tackle the robotic challenge of reasoning about the perceived surroundings to identify a viable navigation path in outdoor environments. We introduce a learning-based indoor-to-outdoor transfer algorithm that uses deep reinforcement learning to train a navigation policy in simulated indoor environments, and successfully transfers that same policy to real outdoor environments. We also introduce Context-Maps (maps with environment observations created by a user), which are applied to our algorithm to enable efficient long-range navigation. We demonstrate that with this policy, robots can successfully navigate hundreds of meters in novel outdoor environments, around previously unseen outdoor obstacles (trees, bushes, buildings, pedestrians, etc.), and in different weather conditions (sunny, overcast, sunset).





PointGoal navigation

User inputs can tell a robot where to go with commands like “go to the Android statue”, pictures showing a target location, or by simply picking a point on a map. In this work, we specify the navigation goal (a selected point on a map) as a relative coordinate to the robot’s current position (i.e., “go to ∆x, ∆y”), this is also known as the PointGoal Visual Navigation (PointNav) task. PointNav is a general formulation for navigation tasks and is one of the standard choices for indoor navigation tasks. However, due to the diverse visuals, uneven terrains and long distance goals in outdoor environments, training PointNav policies for outdoor environments is a challenging task.


Indoor-to-outdoor transfer

Recent successes in training wheeled and legged robotic agents to navigate in indoor environments were enabled by the development of fast, scalable simulators and the availability of large-scale datasets of photorealistic 3D scans of indoor environments. To leverage these successes, we develop an indoor-to-outdoor transfer technique that enables our robots to learn from simulated indoor environments and to be deployed in real outdoor environments.

To overcome the differences between simulated indoor environments and real outdoor environments, we apply kinematic control and image augmentation techniques in our learning system. When using kinematic control, we assume the existence of a reliable low-level locomotion controller that can control the robot to precisely reach a new location. This assumption allows us to directly move the robot to the target location during simulation training through a forward Euler integration and relieves us from having to explicitly model the underlying robot dynamics in simulation, which drastically improves the throughput of simulation data generation. Prior work has shown that kinematic control can lead to better sim-to-real transfer compared to a dynamic control approach, where full robot dynamics are modeled and a low-level locomotion controller is required for moving the robot.



Left Kinematic control; Right: Dynamic control

We created an outdoor maze-like environment using objects found indoors for initial experiments, where we used Boston Dynamics' Spot robot for test navigation. We found that the robot could navigate around novel obstacles in the new outdoor environment.



The Spot robot successfully navigates around obstacles found in indoor environments, with a policy trained entirely in simulation.

However, when faced with unfamiliar outdoor obstacles not seen during training, such as a large slope, the robot was unable to navigate the slope.



The robot is unable to navigate up slopes, as slopes are rare in indoor environments and the robot was not trained to tackle it.

To enable the robot to walk up and down slopes, we apply an image augmentation technique during the simulation training. Specifically, we randomly tilt the simulated camera on the robot during training. It can be pointed up or down within 30 degrees. This augmentation effectively makes the robot perceive slopes even though the floor is level. Training on these perceived slopes enables the robot to navigate slopes in the real-world.



By randomly tilting the camera angle during training in simulation, the robot is now able to walk up and down slopes.

Since the robots were only trained in simulated indoor environments, in which they typically need to walk to a goal just a few meters away, we find that the learned network failed to process longer-range inputs — e.g., the policy failed to walk forward for 100 meters in an empty space. To enable the policy network to handle long-range inputs that are common for outdoor navigation, we normalize the goal vector by using the log of the goal distance.


Context-Maps for complex long-range navigation

Putting everything together, the robot can navigate outdoors towards the goal, while walking on uneven terrain, and avoiding trees, pedestrians and other outdoor obstacles. However, there is still one key component missing: the robot’s ability to plan an efficient long-range path. At this scale of navigation, taking a wrong turn and backtracking can be costly. For example, we find that the local exploration strategy learned by standard PointNav policies are insufficient in finding a long-range goal and usually leads to a dead end (shown below). This is because the robot is navigating without context of its environment, and the optimal path may not be visible to the robot from the start.



Navigation policies without context of the environment do not handle complex long-range navigation goals.

To enable the robot to take the context into consideration and purposefully plan an efficient path, we provide a Context-Map (a binary image that represents a top-down occupancy map of the region that the robot is within) as additional observations for the robot. An example Context-Map is given below, where the black region denotes areas occupied by obstacles and white region is walkable by the robot. The green and red circle denotes the start and goal location of the navigation task. Through the Context-Map, we can provide hints to the robot (e.g., the narrow opening in the route below) to help it plan an efficient navigation route. In our experiments, we create the Context-Map for each route guided by Google Maps satellite images. We denote this variant of PointNav with environmental context, as Context-Guided PointNav.

Example of the Context-Map (right) for a navigation task (left).

It is important to note that the Context-Map does not need to be accurate because it only serves as a rough outline for planning. During navigation, the robot still needs to rely on its onboard cameras to identify and adapt its path to pedestrians, which are absent on the map. In our experiments, a human operator quickly sketches the Context-Map from the satellite image, masking out the regions to be avoided. This Context-Map, together with other onboard sensory inputs, including depth images and relative position to the goal, are fed into a neural network with attention models (i.e., transformers), which are trained using DD-PPO, a distributed implementation of proximal policy optimization, in large-scale simulations.

The Context-Guided PointNav architecture consists of a 3-layer convolutional neural network (CNN) to process depth images from the robot's camera, and a multilayer perceptron (MLP) to process the goal vector. The features are passed into a gated recurrent unit (GRU). We use an additional CNN encoder to process the context-map (top-down map). We compute the scaled dot product attention between the map and the depth image, and use a second GRU to process the attended features (Context Attn., Depth Attn.). The output of the policy are linear and angular velocities for the Spot robot to follow.


Results

We evaluate our system across three long-range outdoor navigation tasks. The provided Context-Maps are rough, incomplete environment outlines that omit obstacles, such as cars, trees, or chairs.

With the proposed algorithm, our robot can successfully reach the distant goal location 100% of the time, without a single collision or human intervention. The robot was able to navigate around pedestrians and real-world clutter that are not present on the context-map, and navigate on various terrain including dirt slopes and grass.


Route 1


  


Route 2


  


Route 3


  


Conclusion

This work opens up robotic navigation research to the less explored domain of diverse outdoor environments. Our indoor-to-outdoor transfer algorithm uses zero real-world experience and does not require the simulator to model predominantly-outdoor phenomena (terrain, ditches, sidewalks, cars, etc). The success in the approach comes from a combination of a robust locomotion control, low sim-to-real gap in depth and map sensors, and large-scale training in simulation. We demonstrate that providing robots with approximate, high-level maps can enable long-range navigation in novel outdoor environments. Our results provide compelling evidence for challenging the (admittedly reasonable) hypothesis that a new simulator must be designed for every new scenario we wish to study. For more information, please see our project page.


Acknowledgements

We would like to thank Sonia Chernova, Tingnan Zhang, April Zitkovich, Dhruv Batra, and Jie Tan for advising and contributing to the project. We would also like to thank Naoki Yokoyama, Nubby Lee, Diego Reyes, Ben Jyenis, and Gus Kouretas for help with the robot experiment setup.

Source: Google AI Blog


Robotic deep RL at scale: Sorting waste and recyclables with a fleet of robots

Reinforcement learning (RL) can enable robots to learn complex behaviors through trial-and-error interaction, getting better and better over time. Several of our prior works explored how RL can enable intricate robotic skills, such as robotic grasping, multi-task learning, and even playing table tennis. Although robotic RL has come a long way, we still don't see RL-enabled robots in everyday settings. The real world is complex, diverse, and changes over time, presenting a major challenge for robotic systems. However, we believe that RL should offer us an excellent tool for tackling precisely these challenges: by continually practicing, getting better, and learning on the job, robots should be able to adapt to the world as it changes around them.

In “Deep RL at Scale: Sorting Waste in Office Buildings with a Fleet of Mobile Manipulators”, we discuss how we studied this problem through a recent large-scale experiment, where we deployed a fleet of 23 RL-enabled robots over two years in Google office buildings to sort waste and recycling. Our robotic system combines scalable deep RL from real-world data with bootstrapping from training in simulation and auxiliary object perception inputs to boost generalization, while retaining the benefits of end-to-end training, which we validate with 4,800 evaluation trials across 240 waste station configurations.


Problem setup

When people don’t sort their trash properly, batches of recyclables can become contaminated and compost can be improperly discarded into landfills. In our experiment, a robot roamed around an office building searching for “waste stations” (bins for recyclables, compost, and trash). The robot was tasked with approaching each waste station to sort it, moving items between the bins so that all recyclables (cans, bottles) were placed in the recyclable bin, all the compostable items (cardboard containers, paper cups) were placed in the compost bin, and everything else was placed in the landfill trash bin. Here is what that looks like:






This task is not as easy as it looks. Just being able to pick up the vast variety of objects that people deposit into waste bins presents a major learning challenge. Robots also have to identify the appropriate bin for each object and sort them as quickly and efficiently as possible. In the real world, the robots can encounter a variety of situations with unique objects, like the examples from real office buildings below:


Learning from diverse experience

Learning on the job helps, but before even getting to that point, we need to bootstrap the robots with a basic set of skills. To this end, we use four sources of experience: (1) a set of simple hand-designed policies that have a very low success rate, but serve to provide some initial experience, (2) a simulated training framework that uses sim-to-real transfer to provide some initial bin sorting strategies, (3) “robot classrooms” where the robots continually practice at a set of representative waste stations, and (4) the real deployment setting, where robots practice in real office buildings with real trash.

A diagram of RL at scale. We bootstrap policies from data generated with a script (top-left). We then train a sim-to-real model and generate additional data in simulation (top-right). At each deployment cycle, we add data collected in our classrooms (bottom-right). We further deploy and collect data in office buildings (bottom-left).

Our RL framework is based on QT-Opt, which we previously applied to learn bin grasping in laboratory settings, as well as a range of other skills. In simulation, we bootstrap from simple scripted policies and use RL, with a CycleGAN-based transfer method that uses RetinaGAN to make the simulated images appear more life-like.




From here, it’s off to the classroom. While real-world office buildings can provide the most representative experience, the throughput in terms of data collection is limited — some days there will be a lot of trash to sort, some days not so much. Our robots collect a large portion of their experience in “robot classrooms.” In the classroom shown below, 20 robots practice the waste sorting task:




While these robots are training in the classrooms, other robots are simultaneously learning on the job in 3 office buildings, with 30 waste stations:





Sorting performance

In the end, we gathered 540k trials in the classrooms and 32.5k trials from deployment. Overall system performance improved as more data was collected. We evaluated our final system in the classrooms to allow for controlled comparisons, setting up scenarios based on what the robots saw during deployment. The final system could accurately sort about 84% of the objects on average, with performance increasing steadily as more data was added. In the real world, we logged statistics from three real-world deployments between 2021 and 2022, and found that our system could reduce contamination in the waste bins by between 40% and 50% by weight. Our paper provides further insights on the technical design, ablations studying various design decisions, and more detailed statistics on the experiments.


Conclusion and future work

Our experiments showed that RL-based systems can enable robots to address real-world tasks in real office environments, with a combination of offline and online data enabling robots to adapt to the broad variability of real-world situations. At the same time, learning in more controlled “classroom” environments, both in simulation and in the real world, can provide a powerful bootstrapping mechanism to get the RL “flywheel” spinning to enable this adaptation. There is still a lot left to do: our final RL policies do not succeed every time, and larger and more powerful models will be needed to improve their performance and extend them to a broader range of tasks. Other sources of experience, including from other tasks, other robots, and even Internet videos may serve to further supplement the bootstrapping experience that we obtained from simulation and classrooms. These are exciting problems to tackle in the future. Please see the full paper here, and the supplementary video materials on the project webpage.


Acknowledgements

This research was conducted by multiple researchers at Robotics at Google and Everyday Robots, with contributions from Alexander Herzog, Kanishka Rao, Karol Hausman, Yao Lu, Paul Wohlhart, Mengyuan Yan, Jessica Lin, Montserrat Gonzalez Arenas, Ted Xiao, Daniel Kappler, Daniel Ho, Jarek Rettinghouse, Yevgen Chebotar, Kuang-Huei Lee, Keerthana Gopalakrishnan, Ryan Julian, Adrian Li, Chuyuan Kelly Fu, Bob Wei, Sangeetha Ramesh, Khem Holden, Kim Kleiven, David Rendleman, Sean Kirmani, Jeff Bingham, Jon Weisz, Ying Xu, Wenlong Lu, Matthew Bennice, Cody Fong, David Do, Jessica Lam, Yunfei Bai, Benjie Holson, Michael Quinlan, Noah Brown, Mrinal Kalakrishnan, Julian Ibarz, Peter Pastor, Sergey Levine and the entire Everyday Robots team.

Source: Google AI Blog


UniPi: Learning universal policies via text-guided video generation

Building models that solve a diverse set of tasks has become a dominant paradigm in the domains of vision and language. In natural language processing, large pre-trained models, such as PaLM, GPT-3 and Gopher, have demonstrated remarkable zero-shot learning of new language tasks. Similarly, in computer vision, models like CLIP and Flamingo have shown robust performance on zero-shot classification and object recognition. A natural next step is to use such tools to construct agents that can complete different decision-making tasks across many environments.

However, training such agents faces the inherent challenge of environmental diversity, since different environments operate with distinct state action spaces (e.g., the joint space and continuous controls in MuJoCo are fundamentally different from the image space and discrete actions in Atari). This environmental diversity hampers knowledge sharing, learning, and generalization across tasks and environments. Furthermore, it is difficult to construct reward functions across environments, as different tasks generally have different notions of success.

In “Learning Universal Policies via Text-Guided Video Generation”, we propose a Universal Policy (UniPi) that addresses environmental diversity and reward specification challenges. UniPi leverages text for expressing task descriptions and video (i.e., image sequences) as a universal interface for conveying action and observation behavior in different environments. Given an input image frame paired with text describing a current goal (i.e., the next high-level step), UniPi uses a novel video generator (trajectory planner) to generate video with snippets of what an agent’s trajectory should look like to achieve that goal. The generated video is fed into an inverse dynamics model that extracts underlying low-level control actions, which are then executed in simulation or by a real robot agent. We demonstrate that UniPi enables the use of language and video as a universal control interface for generalizing to novel goals and tasks across diverse environments.

Video policies generated by UniPi.
UniPi may be applied to downstream multi-task settings that require combinatorial language generalization, long-horizon planning, or internet-scale knowledge. In the bottom example, UniPi takes the image of the white robot arm from the internet and generates video snippets according to the text description of the goal.


UniPi implementation

To generate a valid and executable plan, a text-to-video model must synthesize a constrained video plan starting at the current observed image. We found it more effective to explicitly constrain a video synthesis model during training (as opposed to only constraining videos at sampling time) by providing the first frame of each video as explicit conditioning context.

At a high level, UniPi has four major components: 1) consistent video generation with first-frame tiling, 2) hierarchical planning through temporal super resolution, 3) flexible behavior synthesis, and 4) task-specific action adaptation. We explain the implementation and benefit of each component in detail below.


Video generation through tiling

Existing text-to-video models like Imagen typically generate videos where the underlying environment state changes significantly throughout the duration. To construct an accurate trajectory planner, it is important that the environment remains consistent across all time points. We enforce environment consistency in conditional video synthesis by providing the observed image as additional context when denoising each frame in the synthesized video. To achieve context conditioning, UniPi directly concatenates each intermediate frame sampled from noise with the conditioned observed image across sampling steps, which serves as a strong signal to maintain the underlying environment state across time.

Text-conditional video generation enables UniPi to train general purpose policies on a wide range of data sources (simulated, real robots and YouTube).


Hierarchical planning

When constructing plans in high-dimensional environments with long time horizons, directly generating a set of actions to reach a goal state quickly becomes intractable due to the exponential growth of the underlying search space as the plan gets longer. Planning methods often circumvent this issue by leveraging a natural hierarchy in planning. Specifically, planning methods first construct coarse plans (the intermediate key frames spread out across time) operating on low-dimensional states and actions, which are then refined into plans in the underlying state and action spaces.

Similar to planning, our conditional video generation procedure exhibits a natural temporal hierarchy. UniPi first generates videos at a coarse level by sparsely sampling videos (“abstractions”) of desired agent behavior along the time axis. UniPi then refines the videos to represent valid behavior in the environment by super-resolving videos across time. Meanwhile, coarse-to-fine super-resolution further improves consistency via interpolation between frames.

Given an input observation and text instruction, we plan a set of images representing agent behavior. Images are converted to actions using an inverse dynamics model.


Flexible behavioral modulation

When planning a sequence of actions for a given sub-goal, one can readily incorporate external constraints to modulate a generated plan. Such test-time adaptability can be implemented by composing a probabilistic prior incorporating properties of the desired plan to specify desired constraints across the synthesized action trajectory, which is also compatible with UniPi. In particular, the prior can be specified using a learned classifier on images to optimize a particular task, or as a Dirac delta distribution on a particular image to guide a plan towards a particular set of states. To train the text-conditioned video generation model, we utilize the video diffusion algorithm, where pre-trained language features from the Text-To-Text Transfer Transformer (T5) are encoded.


Task-specific action adaptation

Given a set of synthesized videos, we train a small task-specific inverse dynamics model to translate frames into a set of low-level control actions. This is independent from the planner and can be done on a separate, smaller and potentially suboptimal dataset generated by a simulator.

Given the input frame and text description of the current goal, the inverse dynamics model synthesizes image frames and generates a control action sequence that predicts the corresponding future actions. An agent then executes inferred low-level control actions via closed-loop control.


Capabilities and evaluation of UniPi

We measure the task success rate on novel language-based goals, and find that UniPi generalizes well to both seen and novel combinations of language prompts, compared to baselines such as Transformer BC, Trajectory Transformer (TT), and Diffuser.

UniPi generalizes well to both seen and novel combinations of language prompts in Place (e.g., “place X in Y”) and Relation (e.g., “place X to the left of Y”) tasks.

Below, we illustrate generated videos on unseen combinations of goals. UniPi is able to synthesize a diverse set of behaviors that satisfy unseen language subgoals:

Generated videos for unseen language goals at test time.


Multi-environment transfer

We measure the task success rate of UniPi and baselines on novel tasks not seen during training. UniPi again outperforms the baselines by a large margin:

UniPi generalizes well to new environments when trained on a set of different multi-task environments.

Below, we illustrate generated videos on unseen tasks. UniPi is further able to synthesize a diverse set of behaviors that satisfy unseen language tasks:

Generated video plans on different new test tasks in the multitask setting.


Real world transfer

Below, we further illustrate generated videos given language instructions on unseen real images. Our approach is able to synthesize a diverse set of different behaviors which satisfy language instructions:

Using internet pre-training enables UniPi to synthesize videos of tasks not seen during training. In contrast, a model trained from scratch incorrectly generates plans of different tasks:

To evaluate the quality of videos generated by UniPi when pre-trained on non-robot data, we use the Fréchet Inception Distance (FID) and Fréchet Video Distance (FVD) metrics. We used Contrastive Language-Image Pre-training scores (CLIPScores) to measure the language-image alignment. We demonstrate that pre-trained UniPi achieves significantly higher FID and FVD scores and a better CLIPScore compared to UniPi without pre-training, suggesting that pre-training on non-robot data helps with generating plans for robots. We report the CLIPScore, FID, and VID scores for UniPi trained on Bridge data, with and without pre-training:


Model (24x40)       CLIPScore ↑       FID ↓       FVD ↓      
No pre-training       24.43 ± 0.04       17.75 ± 0.56       288.02 ± 10.45      
Pre-trained       24.54 ± 0.03       14.54 ± 0.57       264.66 ± 13.64      

Using existing internet data improves video plan predictions under all metrics considered.


The future of large-scale generative models for decision making

The positive results of UniPi point to the broader direction of using generative models and the wealth of data on the internet as powerful tools to learn general-purpose decision making systems. UniPi is only one step towards what generative models can bring to decision making. Other examples include using generative foundation models to provide photorealistic or linguistic simulators of the world in which artificial agents can be trained indefinitely. Generative models as agents can also learn to interact with complex environments such as the internet, so that much broader and more complex tasks can eventually be automated. We look forward to future research in applying internet-scale foundation models to multi-environment and multi-embodiment settings.


Acknowledgements

We’d like to thank all remaining authors of the paper including Bo Dai, Hanjun Dai, Ofir Nachum, Joshua B. Tenenbaum, Dale Schuurmans, and Pieter Abbeel. We would like to thank George Tucker, Douglas Eck, and Vincent Vanhoucke for the feedback on this post and on the original paper.

Source: Google AI Blog


Towards ML-enabled cleaning robots

Over the past several years, the capabilities of robotic systems have improved dramatically. As the technology continues to improve and robotic agents are more routinely deployed in real-world environments, their capacity to assist in day-to-day activities will take on increasing importance. Repetitive tasks like wiping surfaces, folding clothes, and cleaning a room seem well-suited for robots, but remain challenging for robotic systems designed for structured environments like factories. Performing these types of tasks in more complex environments, like offices or homes, requires dealing with greater levels of environmental variability captured by high-dimensional sensory inputs, from images plus depth and force sensors.

For example, consider the task of wiping a table to clean a spill or brush away crumbs. While this task may seem simple, in practice, it encompasses many interesting challenges that are omnipresent in robotics. Indeed, at a high-level, deciding how to best wipe a spill from an image observation requires solving a challenging planning problem with stochastic dynamics: How should the robot wipe to avoid dispersing the spill perceived by a camera? But at a low-level, successfully executing a wiping motion also requires the robot to position itself to reach the problem area while avoiding nearby obstacles, such as chairs, and then to coordinate its motions to wipe clean the surface while maintaining contact with the table. Solving this table wiping problem would help researchers address a broader range of robotics tasks, such as cleaning windows and opening doors, which require both high-level planning from visual observations and precise contact-rich control.

   

Learning-based techniques such as reinforcement learning (RL) offer the promise of solving these complex visuo-motor tasks from high-dimensional observations. However, applying end-to-end learning methods to mobile manipulation tasks remains challenging due to the increased dimensionality and the need for precise low-level control. Additionally, on-robot deployment either requires collecting large amounts of data, using accurate but computationally expensive models, or on-hardware fine-tuning.

In “Robotic Table Wiping via Reinforcement Learning and Whole-body Trajectory Optimization”, we present a novel approach to enable a robot to reliably wipe tables. By carefully decomposing the task, our approach combines the strengths of RL — the capacity to plan in high-dimensional observation spaces with complex stochastic dynamics — and the ability to optimize trajectories, effectively finding whole-body robot commands that ensure the satisfaction of constraints, such as physical limits and collision avoidance. Given visual observations of a surface to be cleaned, the RL policy selects wiping actions that are then executed using trajectory optimization. By leveraging a new stochastic differential equation (SDE) simulator of the wiping task to train the RL policy for high-level planning, the proposed end-to-end approach avoids the need for task-specific training data and is able to transfer zero-shot to hardware.


Combining the strengths of RL and of optimal control

We propose an end-to-end approach for table wiping that consists of four components: (1) sensing the environment, (2) planning high-level wiping waypoints with RL, (3) computing trajectories for the whole-body system (i.e., for each joint) with optimal control methods, and (4) executing the planned wiping trajectories with a low-level controller.

System Architecture

The novel component of this approach is an RL policy that effectively plans high-level wiping waypoints given image observations of spills and crumbs. To train the RL policy, we completely bypass the problem of collecting large amounts of data on the robotic system and avoid using an accurate but computationally expensive physics simulator. Our proposed approach relies on a stochastic differential equation (SDE) to model latent dynamics of crumbs and spills, which yields an SDE simulator with four key features:

  • It can describe both dry objects pushed by the wiper and liquids absorbed during wiping.
  • It can simultaneously capture multiple isolated spills.
  • It models the uncertainty of the changes to the distribution of spills and crumbs as the robot interacts with them.
  • It is faster than real-time: simulating a wipe only takes a few milliseconds.
   
The SDE simulator allows simulating dry crumbs (left), which are pushed during each wipe, and spills (right), which are absorbed while wiping. The simulator allows modeling particles with different properties, such as with different absorption and adhesion coefficients and different uncertainty levels.

This SDE simulator is able to rapidly generate large amounts of data for RL training. We validate the SDE simulator using observations from the robot by predicting the evolution of perceived particles for a given wipe. By comparing the result with perceived particles after executing the wipe, we observe that the model correctly predicts the general trend of the particle dynamics. A policy trained with this SDE model should be able to perform well in the real world.


Using this SDE model, we formulate a high-level wiping planning problem and train a vision-based wiping policy using RL. We train entirely in simulation without collecting a dataset using the robot. We simply randomize the initial state of the SDE to cover a wide range of particle dynamics and spill shapes that we may see in the real world.

In deployment, we first convert the robot's image observations into black and white to better isolate the spills and crumb particles. We then use these “thresholded” images as the input to the RL policy. With this approach we do not require a visually-realistic simulator, which would be complex and potentially difficult to develop, and we are able to minimize the sim-to-real gap.

The RL policy’s inputs are thresholded image observations of the cleanliness state of the table. Its outputs are the desired wiping actions. The policy uses a ResNet50 neural network architecture followed by two fully-connected (FC) layers.

The desired wiping motions from the RL policy are executed with a whole-body trajectory optimizer that efficiently computes base and arm joint trajectories. This approach allows satisfying constraints, such as avoiding collisions, and enables zero-shot sim-to-real deployment.

   

Experimental results

We extensively validate our approach in simulation and on hardware. In simulation, our RL policies outperform heuristics-based baselines, requiring significantly fewer wipes to clean spills and crumbs. We also test our policies on problems that were not observed at training time, such as multiple isolated spill areas on the table, and find that the RL policies generalize well to these novel problems.

     
Example of wiping actions selected by the RL policy (left) and wiping performance compared with a baseline (middle, right). The baseline wipes to the center of the table, rotating after each wipe. We report the total dirty surface of the table (middle) and the spread of crumbs particles (right) after each additional wipe.

Our approach enables the robot to reliably wipe spills and crumbs (without accidentally pushing debris from the table) while avoiding collisions with obstacles like chairs.


For further results, please check out the video below:



Conclusion

The results from this work demonstrate that complex visuo-motor tasks such as table wiping can be reliably accomplished without expensive end-to-end training and on-robot data collection. The key consists of decomposing the task and combining the strengths of RL, trained using an SDE model of spill and crumb dynamics, with the strengths of trajectory optimization. We see this work as an important step towards general-purpose home-assistive robots. For more details, please check out the original paper.


Acknowledgements

We'd like to thank our coauthors Sumeet Singh, Mario Prats, Jeffrey Bingham, Jonathan Weisz, Benjie Holson, Xiaohan Zhang, Vikas Sindhwani, Yao Lu, Fei Xia, Peng Xu, Tingnan Zhang, and Jie Tan. We'd also like to thank Benjie Holson, Jake Lee, April Zitkovich, and Linda Luu for their help and support in various aspects of the project. We’re particularly grateful to the entire team at Everyday Robots for their partnership on this work, and for developing the platform on which these experiments were conducted.



Source: Google AI Blog


Pre-training generalist agents using offline reinforcement learning

Reinforcement learning (RL) algorithms can learn skills to solve decision-making tasks like playing games, enabling robots to pick up objects, or even optimizing microchip designs. However, running RL algorithms in the real world requires expensive active data collection. Pre-training on diverse datasets has proven to enable data-efficient fine-tuning for individual downstream tasks in natural language processing (NLP) and vision problems. In the same way that BERT or GPT-3 models provide general-purpose initialization for NLP, large RL–pre-trained models could provide general-purpose initialization for decision-making. So, we ask the question: Can we enable similar pre-training to accelerate RL methods and create a general-purpose “backbone” for efficient RL across various tasks?

In “Offline Q-learning on Diverse Multi-Task Data Both Scales and Generalizes”, to be published at ICLR 2023, we discuss how we scaled offline RL, which can be used to train value functions on previously collected static datasets, to provide such a general pre-training method. We demonstrate that Scaled Q-Learning using a diverse dataset is sufficient to learn representations that facilitate rapid transfer to novel tasks and fast online learning on new variations of a task, improving significantly over existing representation learning approaches and even Transformer-based methods that use much larger models.



Scaled Q-learning: Multi-task pre-training with conservative Q-learning

To provide a general-purpose pre-training approach, offline RL needs to be scalable, allowing us to pre-train on data across different tasks and utilize expressive neural network models to acquire powerful pre-trained backbones, specialized to individual downstream tasks. We based our offline RL pre-training method on conservative Q-learning (CQL), a simple offline RL method that combines standard Q-learning updates with an additional regularizer that minimizes the value of unseen actions. With discrete actions, the CQL regularizer is equivalent to a standard cross-entropy loss, which is a simple, one-line modification on standard deep Q-learning. A few crucial design decisions made this possible:

  • Neural network size: We found that multi-game Q-learning required large neural network architectures. While prior methods often used relatively shallow convolutional networks, we found that models as large as a ResNet 101 led to significant improvements over smaller models.
  • Neural network architecture: To learn pre-trained backbones that are useful for new games, our final architecture uses a shared neural network backbone, with separate 1-layer heads outputting Q-values of each game. This design avoids interference between the games during pre-training, while still providing enough data sharing to learn a single shared representation. Our shared vision backbone also utilized a learned position embedding (akin to Transformer models) to keep track of spatial information in the game.
  • Representational regularization: Recent work has observed that Q-learning tends to suffer from representational collapse issues, where even large neural networks can fail to learn effective representations. To counteract this issue, we leverage our prior work to normalize the last layer features of the shared part of the Q-network. Additionally, we utilized a categorical distributional RL loss for Q-learning, which is known to provide richer representations that improve downstream task performance.

The multi-task Atari benchmark

We evaluate our approach for scalable offline RL on a suite of Atari games, where the goal is to train a single RL agent to play a collection of games using heterogeneous data from low-quality (i.e., suboptimal) players, and then use the resulting network backbone to quickly learn new variations in pre-training games or completely new games. Training a single policy that can play many different Atari games is difficult enough even with standard online deep RL methods, as each game requires a different strategy and different representations. In the offline setting, some prior works, such as multi-game decision transformers, proposed to dispense with RL entirely, and instead utilize conditional imitation learning in an attempt to scale with large neural network architectures, such as transformers. However, in this work, we show that this kind of multi-game pre-training can be done effectively via RL by employing CQL in combination with a few careful design decisions, which we describe below.


Scalability on training games

We evaluate the Scaled Q-Learning method’s performance and scalability using two data compositions: (1) near optimal data, consisting of all the training data appearing in replay buffers of previous RL runs, and (2) low quality data, consisting of data from the first 20% of the trials in the replay buffer (i.e., only data from highly suboptimal policies). In our results below, we compare Scaled Q-Learning with an 80-million parameter model to multi-game decision transformers (DT) with either 40-million or 80-million parameter models, and a behavioral cloning (imitation learning) baseline (BC). We observe that Scaled Q-Learning is the only approach that improves over the offline data, attaining about 80% of human normalized performance.

Further, as shown below, Scaled Q-Learning improves in terms of performance, but it also enjoys favorable scaling properties: just as how the performance of pre-trained language and vision models improves as network sizes get bigger, enjoying what is typically referred as “power-law scaling”, we show that the performance of Scaled Q-learning enjoys similar scaling properties. While this may be unsurprising, this kind of scaling has been elusive in RL, with performance often deteriorating with larger model sizes. This suggests that Scaled Q-Learning in combination with the above design choices better unlocks the ability of offline RL to utilize large models.



Fine-tuning to new games and variations

To evaluate fine-tuning from this offline initialization, we consider two settings: (1) fine-tuning to a new, entirely unseen game with a small amount of offline data from that game, corresponding to 2M transitions of gameplay, and (2) fine-tuning to a new variant of the games with online interaction. The fine-tuning from offline gameplay data is illustrated below. Note that this condition is generally more favorable to imitation-style methods, Decision Transformer and behavioral cloning, since the offline data for the new games is of relatively high-quality. Nonetheless, we see that in most cases Scaled Q-learning improves over alternative approaches (80% on average), as well as dedicated representation learning methods, such as MAE or CPC, which only use the offline data to learn visual representations rather than value functions.

In the online setting, we see even larger improvements from pre-training with Scaled Q-learning. In this case, representation learning methods like MAE yield minimal improvement during online RL, whereas Scaled Q-Learning can successfully integrate prior knowledge about the pre-training games to significantly improve the final score after 20k online interaction steps.

These results demonstrate that pre-training generalist value function backbones with multi-task offline RL can significantly boost performance of RL on downstream tasks, both in offline and online mode. Note that these fine-tuning tasks are quite difficult: the various Atari games, and even variants of the same game, differ significantly in appearance and dynamics. For example, the target blocks in Breakout disappear in the variation of the game as shown below, making control difficult. However, the success of Scaled Q-learning, particularly as compared to visual representation learning techniques, such as MAE and CPC, suggests that the model is in fact learning some representation of the game dynamics, rather than merely providing better visual features.

Fine-tuning with online RL for variants of the game Freeway, Hero, and Breakout. The new variant used in fine-tuning is shown in the bottom row of each figure, the original game seen in pre-training is in the top row. Fine-tuning from Scaled Q-Learning significantly outperforms MAE (a visual representation learning method) and learning from scratch with single-game DQN.

Conclusion and takeaways

We presented Scaled Q-Learning, a pre-training method for scaled offline RL that builds on the CQL algorithm, and demonstrated how it enables efficient offline RL for multi-task training. This work made initial progress towards enabling more practical real-world training of RL agents as an alternative to costly and complex simulation-based pipelines or large-scale experiments. Perhaps in the long run, similar work will lead to generally capable pre-trained RL agents that develop broadly applicable exploration and interaction skills from large-scale offline pre-training. Validating these results on a broader range of more realistic tasks, in domains such as robotics (see some initial results) and NLP, is an important direction for future research. Offline RL pre-training has a lot of potential, and we expect that we will see many advances in this area in future work.


Acknowledgements

This work was done by Aviral Kumar, Rishabh Agarwal, Xinyang Geng, George Tucker, and Sergey Levine. Special thanks to Sherry Yang, Ofir Nachum, and Kuang-Huei Lee for help with the multi-game decision transformer codebase for evaluation and the multi-game Atari benchmark, and Tom Small for illustrations and animation.

Source: Google AI Blog


Beyond Tabula Rasa: Reincarnating Reinforcement Learning

Reinforcement learning (RL) is an area of machine learning that focuses on training intelligent agents using related experiences so they can learn to solve decision making tasks, such as playing video games, flying stratospheric balloons, and designing hardware chips. Due to the generality of RL, the prevalent trend in RL research is to develop agents that can efficiently learn tabula rasa, that is, from scratch without using previously learned knowledge about the problem. However, in practice, tabula rasa RL systems are typically the exception rather than the norm for solving large-scale RL problems. Large-scale RL systems, such as OpenAI Five, which achieves human-level performance on Dota 2, undergo multiple design changes (e.g., algorithmic or architectural changes) during their developmental cycle. This modification process can last months and necessitates incorporating such changes without re-training from scratch, which would be prohibitively expensive. 

Furthermore, the inefficiency of tabula rasa RL research can exclude many researchers from tackling computationally-demanding problems. For example, the quintessential benchmark of training a deep RL agent on 50+ Atari 2600 games in ALE for 200M frames (the standard protocol) requires 1,000+ GPU days. As deep RL moves towards more complex and challenging problems, the computational barrier to entry in RL research will likely become even higher.

To address the inefficiencies of tabula rasa RL, we present “Reincarnating Reinforcement Learning: Reusing Prior Computation To Accelerate Progress” at NeurIPS 2022. Here, we propose an alternative approach to RL research, where prior computational work, such as learned models, policies, logged data, etc., is reused or transferred between design iterations of an RL agent or from one agent to another. While some sub-areas of RL leverage prior computation, most RL agents are still largely trained from scratch. Until now, there has been no broader effort to leverage prior computational work for the training workflow in RL research. We have also released our code and trained agents to enable researchers to build on this work.

Tabula rasa RL vs. Reincarnating RL (RRL). While tabula rasa RL focuses on learning from scratch, RRL is based on the premise of reusing prior computational work (e.g., prior learned agents) when training new agents or improving existing agents, even in the same environment. In RRL, new agents need not be trained from scratch, except for initial forays into new problems.

Why Reincarnating RL?

Reincarnating RL (RRL) is a more compute and sample-efficient workflow than training from scratch. RRL can democratize research by allowing the broader community to tackle complex RL problems without requiring excessive computational resources. Furthermore, RRL can enable a benchmarking paradigm where researchers continually improve and update existing trained agents, especially on problems where improving performance has real-world impact, such as balloon navigation or chip design. Finally, real-world RL use cases will likely be in scenarios where prior computational work is available (e.g., existing deployed RL policies).

RRL as an alternative research workflow. Imagine a researcher who has trained an agent A1 for some time, but now wants to experiment with better architectures or algorithms. While the tabula rasa workflow requires retraining another agent from scratch, RRL provides the more viable option of transferring the existing agent A1 to another agent and training this agent further, or simply fine-tuning A1.

While there have been some ad hoc large-scale reincarnation efforts with limited applicability, e.g., model surgery in Dota2, policy distillation in Rubik’s cube, PBT in AlphaStar, RL fine-tuning a behavior-cloned policy in AlphaGo / Minecraft, RRL has not been studied as a research problem in its own right. To this end, we argue for developing general-purpose RRL approaches as opposed to prior ad-hoc solutions.


Case Study: Policy to Value Reincarnating RL

Different RRL problems can be instantiated depending on the kind of prior computational work provided. As a step towards developing broadly applicable RRL approaches, we present a case study on the setting of Policy to Value reincarnating RL (PVRL) for efficiently transferring an existing sub-optimal policy (teacher) to a standalone value-based RL agent (student). While a policy directly maps a given environment state (e.g., a game screen in Atari) to an action, value-based agents estimate the effectiveness of an action at a given state in terms of achievable future rewards, which allows them to learn from previously collected data.

For a PVRL algorithm to be broadly useful, it should satisfy the following requirements:

  • Teacher Agnostic: The student shouldn’t be constrained by the existing teacher policy’s architecture or training algorithm.
  • Weaning off the teacher: It is undesirable to maintain dependency on past suboptimal teachers for successive reincarnations.
  • Compute / Sample Efficient: Reincarnation is only useful if it is cheaper than training from scratch.

Given the PVRL algorithm requirements, we evaluate whether existing approaches, designed with closely related goals, will suffice. We find that such approaches either result in small improvements over tabula rasa RL or degrade in performance when weaning off the teacher.

To address these limitations, we introduce a simple method, QDagger, in which the agent distills knowledge from the suboptimal teacher via an imitation algorithm while simultaneously using its environment interactions for RL. We start with a deep Q-network (DQN) agent trained for 400M environment frames (a week of single-GPU training) and use it as the teacher for reincarnating student agents trained on only 10M frames (a few hours of training), where the teacher is weaned off over the first 6M frames. For benchmark evaluation, we report the interquartile mean (IQM) metric from the RLiable library. As shown below for the PVRL setting on Atari games, we find that the QDagger RRL method outperforms prior approaches.

Benchmarking PVRL algorithms on Atari, with teacher-normalized scores aggregated across 10 games. Tabula rasa DQN (–·–) obtains a normalized score of 0.4. Standard baseline approaches include kickstarting, JSRL, rehearsal, offline RL pre-training and DQfD. Among all methods, only QDagger surpasses teacher performance within 10 million frames and outperforms the teacher in 75% of the games.

Reincarnating RL in Practice

We further examine the RRL approach on the Arcade Learning Environment, a widely used deep RL benchmark. First, we take a Nature DQN agent that uses the RMSProp optimizer and fine-tune it with the Adam optimizer to create a DQN (Adam) agent. While it is possible to train a DQN (Adam) agent from scratch, we demonstrate that fine-tuning Nature DQN with the Adam optimizer matches the from-scratch performance using 40x less data and compute.

Reincarnating DQN (Adam) via Fine-Tuning. The vertical separator corresponds to loading network weights and replay data for fine-tuning. Left: Tabula rasa Nature DQN nearly converges in performance after 200M environment frames. Right: Fine-tuning this Nature DQN agent using a reduced learning rate with the Adam optimizer for 20 million frames obtains similar results to DQN (Adam) trained from scratch for 400M frames.

Given the DQN (Adam) agent as a starting point, fine-tuning is restricted to the 3-layer convolutional architecture. So, we consider a more general reincarnation approach that leverages recent architectural and algorithmic advances without training from scratch. Specifically, we use QDagger to reincarnate another RL agent that uses a more advanced RL algorithm (Rainbow) and a better neural network architecture (Impala-CNN ResNet) from the fine-tuned DQN (Adam) agent.

Reincarnating a different architecture / algorithm via QDagger. The vertical separator is the point at which we apply offline pre-training using QDagger for reincarnation. Left: Fine-tuning DQN with Adam. Right: Comparison of a tabula rasa Impala-CNN Rainbow agent (sky blue) to an Impala-CNN Rainbow agent (pink) trained using QDagger RRL from the fine-tuned DQN (Adam). The reincarnated Impala-CNN Rainbow agent consistently outperforms its scratch counterpart. Note that further fine-tuning DQN (Adam) results in diminishing returns (yellow).

Overall, these results indicate that past research could have been accelerated by incorporating a RRL approach to designing agents, instead of re-training agents from scratch. Our paper also contains results on the Balloon Learning Environment, where we demonstrate that RRL allows us to make progress on the problem of navigating stratospheric balloons using only a few hours of TPU-compute by reusing a distributed RL agent trained on TPUs for more than a month.


Discussion

Fairly comparing reincarnation approaches involves using the exact same computational work and workflow. Furthermore, the research findings in RRL that broadly generalize would be about how effective an algorithm is given access to existing computational work, e.g., we successfully applied QDagger developed using Atari for reincarnation on Balloon Learning Environment. As such, we speculate that research in reincarnating RL can branch out in two directions:

  • Standardized benchmarks with open-sourced computational work: Akin to NLP and vision, where typically a small set of pre-trained models are common, research in RRL may also converge to a small set of open-sourced computational work (e.g., pre-trained teacher policies) on a given benchmark.
  • Real-world domains: Since obtaining higher performance has real-world impact in some domains, it incentivizes the community to reuse state-of-the-art agents and try to improve their performance.

See our paper for a broader discussion on scientific comparisons, generalizability and reproducibility in RRL. Overall, we hope that this work motivates researchers to release computational work (e.g., model checkpoints) on which others could directly build. In this regard, we have open-sourced our code and trained agents with their final replay buffers. We believe that reincarnating RL can substantially accelerate research progress by building on prior computational work, as opposed to always starting from scratch.


Acknowledgements

This work was done in collaboration with Pablo Samuel Castro, Aaron Courville and Marc Bellemare. We’d like to thank Tom Small for the animated figure used in this post. We are also grateful for feedback by the anonymous NeurIPS reviewers and several members of the Google Research team, DeepMind and Mila.

Source: Google AI Blog


PI-ARS: Accelerating Evolution-Learned Visual-Locomotion with Predictive Information Representations

Evolution strategy (ES) is a family of optimization techniques inspired by the ideas of natural selection: a population of candidate solutions are usually evolved over generations to better adapt to an optimization objective. ES has been applied to a variety of challenging decision making problems, such as legged locomotion, quadcopter control, and even power system control.

Compared to gradient-based reinforcement learning (RL) methods like proximal policy optimization (PPO) and soft actor-critic (SAC), ES has several advantages. First, ES directly explores in the space of controller parameters, while gradient-based methods often explore within a limited action space, which indirectly influences the controller parameters. More direct exploration has been shown to boost learning performance and enable large scale data collection with parallel computation. Second, a major challenge in RL is long-horizon credit assignment, e.g., when a robot accomplishes a task in the end, determining which actions it performed in the past were the most critical and should be assigned a greater reward. Since ES directly considers the total reward, it relieves researchers from needing to explicitly handle credit assignment. In addition, because ES does not rely on gradient information, it can naturally handle highly non-smooth objectives or controller architectures where gradient computation is non-trivial, such as meta–reinforcement learning. However, a major weakness of ES-based algorithms is their difficulty in scaling to problems that require high-dimensional sensory inputs to encode the environment dynamics, such as training robots with complex vision inputs.

In this work, we propose “PI-ARS: Accelerating Evolution-Learned Visual-Locomotion with Predictive Information Representations”, a learning algorithm that combines representation learning and ES to effectively solve high dimensional problems in a scalable way. The core idea is to leverage predictive information, a representation learning objective, to obtain a compact representation of the high-dimensional environment dynamics, and then apply Augmented Random Search (ARS), a popular ES algorithm, to transform the learned compact representation into robot actions. We tested PI-ARS on the challenging problem of visual-locomotion for legged robots. PI-ARS enables fast training of performant vision-based locomotion controllers that can traverse a variety of difficult environments. Furthermore, the controllers trained in simulated environments successfully transfer to a real quadruped robot.

PI-ARS trains reliable visual-locomotion policies that are transferable to the real world.

Predictive Information
A good representation for policy learning should be both compressive, so that ES can focus on solving a much lower dimensional problem than learning from raw observations would entail, and task-critical, so the learned controller has all the necessary information needed to learn the optimal behavior. For robotic control problems with high-dimensional input space, it is critical for the policy to understand the environment, including the dynamic information of both the robot itself and its surrounding objects.

As such, we propose an observation encoder that preserves information from the raw input observations that allows the policy to predict the future states of the environment, thus the name predictive information (PI). More specifically, we optimize the encoder such that the encoded version of what the robot has seen and planned in the past can accurately predict what the robot might see and be rewarded in the future. One mathematical tool to describe such a property is that of mutual information, which measures the amount of information we obtain about one random variable X by observing another random variable Y. In our case, X and Y would be what the robot saw and planned in the past, and what the robot sees and is rewarded in the future. Directly optimizing the mutual information objective is a challenging problem because we usually only have access to samples of the random variables, but not their underlying distributions. In this work we follow a previous approach that uses InfoNCE, a contrastive variational bound on mutual information to optimize the objective.

Left: We use representation learning to encode PI of the environment. Right: We train the representation by replaying trajectories from the replay buffer and maximize the predictability between the observation and motion plan in the past and the observation and reward in the future of the trajectory.

Predictive Information with Augmented Random Search
Next, we combine PI with Augmented Random Search (ARS), an algorithm that has shown excellent optimization performance for challenging decision-making tasks. At each iteration of ARS, it samples a population of perturbed controller parameters, evaluates their performance in the testing environment, and then computes a gradient that moves the controller towards the ones that performed better.

We use the learned compact representation from PI to connect PI and ARS, which we call PI-ARS. More specifically, ARS optimizes a controller that takes as input the learned compact representation PI and predicts appropriate robot commands to achieve the task. By optimizing a controller with smaller input space, it allows ARS to find the optimal solution more efficiently. Meanwhile, we use the data collected during ARS optimization to further improve the learned representation, which is then fed into the ARS controller in the next iteration.

An overview of the PI-ARS data flow. Our algorithm interleaves between two steps: 1) optimizing the PI objective that updates the policy, which is the weights for the neural network that extracts the learned representation; and 2) sampling new trajectories and updating the controller parameters using ARS.

Visual-Locomotion for Legged Robots
We evaluate PI-ARS on the problem of visual-locomotion for legged robots. We chose this problem for two reasons: visual-locomotion is a key bottleneck for legged robots to be applied in real-world applications, and the high-dimensional vision-input to the policy and the complex dynamics in legged robots make it an ideal test-case to demonstrate the effectiveness of the PI-ARS algorithm. A demonstration of our task setup in simulation can be seen below. Policies are first trained in simulated environments, and then transferred to hardware.

An illustration of the visual-locomotion task setup. The robot is equipped with two cameras to observe the environment (illustrated by the transparent pyramids). The observations and robot state are sent to the policy to generate a high-level motion plan, such as feet landing location and desired moving speed. The high-level motion plan is then achieved by a low-level Motion Predictive Control (MPC) controller.

Experiment Results
We first evaluate the PI-ARS algorithm on four challenging simulated tasks:

  • Uneven stepping stones: The robot needs to walk over uneven terrain while avoiding gaps.
  • Quincuncial piles: The robot needs to avoid gaps both in front and sideways.
  • Moving platforms: The robot needs to walk over stepping stones that are randomly moving horizontally or vertically. This task illustrates the flexibility of learning a vision-based policy in comparison to explicitly reconstructing the environment.
  • Indoor navigation: The robot needs to navigate to a random location while avoiding obstacles in an indoor environment.

As shown below, PI-ARS is able to significantly outperform ARS in all four tasks in terms of the total task reward it can obtain (by 30-50%).

Left: Visualization of PI-ARS policy performance in simulation. Right: Total task reward (i.e., episode return) for PI-ARS (green line) and ARS (red line). The PI-ARS algorithm significantly outperforms ARS on four challenging visual-locomotion tasks.

We further deploy the trained policies to a real Laikago robot on two tasks: random stepping stone and indoor navigation. We demonstrate that our trained policies can successfully handle real-world tasks. Notably, the success rate of the random stepping stone task improved from 40% in the prior work to 100%.

PI-ARS trained policy enables a real Laikago robot to navigate around obstacles.

Conclusion
In this work, we present a new learning algorithm, PI-ARS, that combines gradient-based representation learning with gradient-free evolutionary strategy algorithms to leverage the advantages of both. PI-ARS enjoys the effectiveness, simplicity, and parallelizability of gradient-free algorithms, while relieving a key bottleneck of ES algorithms on handling high-dimensional problems by optimizing a low-dimensional representation. We apply PI-ARS to a set of challenging visual-locomotion tasks, among which PI-ARS significantly outperforms the state of the art. Furthermore, we validate the policy learned by PI-ARS on a real quadruped robot. It enables the robot to walk over randomly-placed stepping stones and navigate in an indoor space with obstacles. Our method opens the possibility of incorporating modern large neural network models and large-scale data into the field of evolutionary strategy for robotics control.

Acknowledgements
We would like to thank our paper co-authors: Ofir Nachum, Tingnan Zhang, Sergio Guadarrama, and Jie Tan. We would also like to thank Ian Fischer and John Canny for valuable feedback.

Source: Google AI Blog