Finite-Sample Convergence Bounds for Trust Region Policy Optimization in Mean-Field Games
Abstract
We introduce Mean-Field Trust Region Policy Optimization (MF-TRPO), a novel algorithm designed to compute approximate Nash equilibria for ergodic Mean-Field Games (MFG) in finite state-action spaces. Building on the well-established performance of TRPO in the reinforcement learning (RL) setting, we extend its methodology to the MFG framework, leveraging its stability and robustness in policy optimization. Under standard assumptions in the MFG literature, we provide a rigorous analysis of MF-TRPO, establishing theoretical guarantees on its convergence. Our results cover both the exact formulation of the algorithm and its sample-based counterpart, where we derive high-probability guarantees and finite sample complexity. This work advances MFG optimization by bridging RL techniques with mean-field decision-making, offering a theoretically grounded approach to solving complex multi-agent problems.
1 Introduction
In an increasingly interconnected world, autonomous systems capable of adaptive decision-making have become indispensable. Their range of applications is vast and continuously expanding, spanning from autonomous driving (see, e.g., Shalev-Shwartz et al., 2016) to energy market control (Samvelyan et al., 2019), and from traffic light optimization (Wiering et al., 2000) to advanced robotic systems (Matignon et al., 2007; Kober et al., 2013).
A powerful framework to model these adaptive decision agents is Multi-Agent Reinforcement Learning (MARL), which enables agents to learn optimal strategies through interaction with both the environment and other agents (see, e.g., Zhang et al., 2021; Gronauer & Diepold, 2022). However, MARL faces two major challenges: scalability and non-stationarity. As the number of agents increases, the joint state-action space grows exponentially, making learning computationally prohibitive. Additionally, because all agents are simultaneously updating their policies, the environment becomes non-stationary from the perspective of each individual agent, severely hindering convergence and stability. To address these issues, many techniques have been considered, like Centralized Training with Decentralized Execution (CTDE) (Foerster et al., 2018) and attention approaches (Iqbal & Sha, 2019). While these methods improve stability and learning efficiency, it comes at a significant computational cost, limiting its scalability in large-scale systems.
Under the assumptions of homogeneity and anonymity, complex multi-agent games can be effectively approximated using Mean-Field Games (MFG). MFG provide an asymptotic approximation of exchangeable particle systems as the number of agents grows. Originally introduced by Lasry & Lions (2006a, b, 2007) and Huang et al. (2003, 2005, 2006a, 2006b), MFG replace direct agent-to-agent interactions with a representative agent interacting with the statistical distribution of the population. Due to their analytical tractability and broad applicability, MFG have been widely adopted across various domains, including economic modeling (Bassière et al., 2024), finance (Lavigne & Tankov, 2023; Carmona et al., 2013), public health dynamics (Doncel et al., 2022), and energy storage (Alasseur et al., 2020).
In particular, Mean-Field Reinforcement Learning (MFRL) arises as the scaling limit of many MARL problems, positioning itself at the intersection of MFG and Reinforcement Learning (RL). In this setting, RL techniques are employed to solve equilibrium problems in large-scale multi-agent systems. At the core of this framework, each agent optimizes its objective while treating the mean-field distribution as fixed. This structure closely resembles CTDE in the MARL setting, but its computational cost is significantly reduced due to the mean-field approximation. In turn, the distribution evolves dynamically based on the collective behavior of all agents. This formulation extends the classical notion of Nash equilibrium to the Mean-Field Nash Equilibrium (MFNE), where equilibrium emerges from the interaction between individual decision-making and population-wide updates.
MFNE provide an adequate approximation for large-scale multi-agent interactions, achieving an -approximate Nash equilibrium in the corresponding -player game (Cardaliaguet et al., 2019; Fischer & Silva, 2021; Flandoli et al., 2022). This approximation drastically reduces the complexity of analyzing strategic interactions in large populations, establishing MFG as a scalable and computationally efficient framework for real-world applications.
Related works.
Recent work in MFG has explored the use of proximal methods due to their stability and empirical performance. Notably, Pérolat et al. (2022); Perrin et al. (2022) analyze Online Mirror Descent (OMD) from a model-specific perspective, while Yardim et al. (2023) extend this direction with a model-free approach. However, their analysis does not address population updates, operating in a restricted no-manipulation regime. These works have demonstrated the effectiveness of proximal methods in MFG, which is consistent with our research direction.
We extend this line of research by establishing finite-sample complexity guarantees providing a rigorous theoretical framework that ensures provable efficiency in solving MFG. Moreover, we explicitly incorporates the role of monotonicity in stabilizing population dynamics, a key aspect of the ergodic structure in MFG, and relax overly restrictive assumptions—such as uniformly bounded-away-from-zero policies or absolute continuity with respect to the uniform distribution. With a more flexible framework, we get refined learning guarantees under weaker assumptions, achieving an convergence rate in the optimization problem with improved sample efficiency, thus broadening the applicability of these methods. For an additional discussion on related work, we refer to Appendix A.
Contributions.
We propose Exact MF-TRPO and Sample-Based MF-TRPO, trust-region-based algorithms for computing approximate MFNE in the MFRL setting. Our method combines the structure of ergodic MFG with the stability of trust-region optimization to enable theoretically grounded learning in multi-agent environments. Our key contributions are:
-
1.
Theoretical analysis of Exact MF-TRPO with a bound on the exploitability of the learned policies, quantifying the achieved -MFNE after iterations.
-
2.
A sample-based variant, Sample-Based MF-TRPO, with finite-sample complexity guarantees under the -paradigm, requiring at most environment interactions to reach an -MFNE.
-
3.
Numerical experiments validating the efficiency and effectiveness of our approach in representative MFRL settings.
2 Framework
We use the discounted formulation as an approximation to the ergodic setting, a standard approach in the literature (see, e.g., Laurière et al., 2022). This methodology allows us to build on the well-established theoretical framework of discounted RL while capturing the long-term behavior of the ergodic formulation in a stepwise fashion. In particular, we adopt the Mean-Field Markov Decision Process (MF-MDP) framework, a natural extension of the infinite-horizon discounted MDP commonly studied in RL (Sutton & Barto, 2018a). This adaptation preserves the computational tractability and convergence properties of the discounted problem while approximating the stationary dynamics of the ergodic setting. The MF-MDP framework thus serves as a bridge between classical RL and MFG models and provides a structured approach to policy optimization that is valid for both finite horizons and stationary regimes. A detailed discussion of this approach and its theoretical foundations can be found in the Appendix B.
Notations.
For a finite set , let denote the set of probability distributions over . A finite MF-MDP is a tuple , where is a finite state space, is a finite action space, is the transition function, is the reward function, is the discount factor. Since we consider probability distributions over a finite state space of size , they can be identified as vectors in ; thus, we define the inner product and the Euclidean norm accordingly.
We assume that is continuous, thus bounded, and denote by its upper bound, i.e., , for Given a policy and a population profile , we define the transition operator as the transition matrix induced by the probability kernel, where actions are sampled as under the mean-field parameter , i.e.,
(1) |
Denote the stationary distribution of the Markov chain . Let to be set the policies, i.e., the set of functions from to . For , and , the Kullback–Leibler (KL) divergence between the two distributions and is defined as , if is absolutely continuous to , otherwise. We use the notation to hide polylogarithmic factors in the asymptotic complexity.
Problem formulation.
As in Laurière et al. (2022), the discounted stationary problem within a mean-field interaction setting is designed to approximate the -player game in the ergodic regime, as . In this setting, a representative agent in the mean-field approximation seeks to maximize the expected discounted sum of rewards while interacting with the population distribution under a policy . The objective function is given by
(2) |
Given an initial state , actions are sampled at each time step as , with state transitions governed by the kernel , i.e., . We consider an entropy-regularized variant of the classical MFG problem, where denotes the entropy regularization parameter.
As MFG are an -approximation of the -player game, it is natural to introduce an additional regularization term. If this extra bias remains within the order of the existing approximation error, model fidelity is thus preserved. Regularization enhances solution stability, which is particularly beneficial given the inherent nonlinearity of MFG optimization. This is especially advantageous in RL settings, where small perturbations in the value function or policy updates can otherwise lead to erratic behavior.
The value function is defined as
(3) |
With a slight abuse of notation, we use (resp. and ) to denote for (resp. and ).
Furthermore, let denote the regularized -function, defined as
(4) |
We denote the optimal policy of optimization problem , which is unique in the entropy-regularized RL (see, e.g. Haarnoja et al., 2017; Geist et al., 2019).
Define (resp. and ) the occupation measure of the process induced by this policy, under the population distribution and initial distribution (resp. and ), i.e.,
(5) |
with , , and , and its spatial marginal, i.e.,
(6) |
Interactions with the environment.
In the RL literature, various paradigms have been proposed to structure the interaction between an agent and its environment, influencing how data is collected and utilized for policy updates. In our setting, two fundamental actions can be performed: reset and step. The reset action initializes the environment by sampling a new state from the distribution , effectively allowing the agent to restart from a fresh initial condition. The step action, on the other hand, takes a chosen action and the mean-field distribution profile , and progresses the environment based on the current state . Specifically, it samples the next state from the transition kernel and updates the environment accordingly. This interaction model aligns with previous works (Kakade, 2003; Shani et al., 2020) and provides an intermediate assumption in the spectrum of data access models in RL. It is therefore weaker than assuming full access to the true model or a generative model, where arbitrary state-action pairs can be queried, but stronger than the setting where no restarts are allowed, restricting exploration to trajectories induced by the current policy. This ensures reliable state exploration and promotes stable convergence.
Nash Equilibrium.
Our work aims to compute a MFNE. A Nash equilibrium (NE), a fundamental concept in game theory, represents a stable state in which no player can improve her payoff by unilaterally changing her strategy, i.e., each player reacts optimally to the strategies of the others.
In the MFG framework, as the number of agents approaches infinity, this concept extends to a situation in which each agent optimally adapts its strategy to the behavior of the collective population. This leads to a mean-field fixed-point condition in which individual decisions shape and are shaped by the evolving population distribution. This formulation makes MFG a powerful tool for analyzing large-scale strategic interactions.
Definition 2.1 (MFNE).
A pair is said to be a MFNE if it satisfies the following two conditions:
-
1.
(Optimality under the population dynamics) The policy is an optimal solution to (3), given the population evolution , i.e.,
-
2.
(Consistency of the population evolution) The population distribution is a fixed point of the mean-field evolution equation
In this definition, the first condition ensures that no individual agent can improve their long-term objective by unilaterally deviating from the equilibrium policy . Given the equilibrium population distribution , this optimality condition guarantees that the policy remains the best possible strategy for each agent, thereby ensuring individual rationality.
The second condition enforces consistency in population dynamics: when all agents follow the equilibrium policy , the resulting population distribution remains stable over time. This fixed-point property ensures the system’s long-term evolution does not deviate from the equilibrium state.
This notion of equilibrium, introduced by Nash (1950) and extended to the mean-field setting by Lasry & Lions (2006a, b, 2007); Huang et al. (2003, 2005, 2006a), is fundamental in game theory. In these works, the authors generalize the results of Nash (1951), proving the existence of at least one NE using the Brouwer Fixed-Point Theorem. This non-constructive result holds under mild assumptions, such as the continuity of payoff functions and the compactness of strategy spaces, forming a cornerstone for analyzing strategic interactions. However, explicitly computing such equilibria remains challenging, particularly in the multi-player regime (Austrin et al., 2011).
Exploitability.
A key metric for evaluating the deviation from a NE is exploitability (Laurière et al., 2022). Formally, the mean-field setting, it is defined as:
(7) |
This quantity measures the potential improvement an individual agent could achieve by deviating unilaterally from the learned policy , given as mean-field parameter the stationary distribution .
Definition 2.2.
A pair is said to be a -MFNE, if its exploitability is bounded by , i.e., .
3 Assumptions
We outline the key assumptions that guarantee the well-posedness and stability of the MFG problem, providing a foundation for deriving finite-sample complexity bounds for the proposed algorithm.
A- 1.
Let (resp. ) be Lipschitz continuous with respect to , with Lipschitz constant (resp. ), i.e., for , , and , we have
Lipschitz continuity for the MDP parameters ensures that small perturbations in the state or action lead to proportionally small changes in the transition dynamics. This assumption is well established in the RL literature (Asadi et al., 2018; Le Lan et al., 2021) and it facilitates the derivation of meaningful error bounds and convergence rates in policy optimization. From this, we establish Lipschitz continuity of the optimal policies w.r.t. mean-field parameter as follows (cf. Corollary E.4).
Proposition 3.1.
Suppose Assumption 1 holds. Then, there exists a constant such that, for ,
where is the optimal policy associated with the mean-field distribution .
This result connects the structural properties of the MFG framework with the behavior of the associated optimal policies, forming a key theoretical foundation for the analysis of the proposed algorithms.
A- 2.
There exist an integer and a real number such that, for , we have
This monotonicity condition of the mean-field update is employed in various forms throughout the MFG literature (Angiuli et al., 2021, 2023; Yardim et al., 2023). This condition, originally introduced in the foundational works by Huang et al. (2003, 2005, 2006a, 2006b), ensures that iterative updates to the population distribution progressively aligns the system with the NE.
This condition extends the classical contractivity condition, which corresponds to (see, e.g., Guo et al., 2019). The condition might fail for but hold for some , reflecting the combined effect of the Lipschitz continuity of the regularized best response and the ergodicity of the Markov reward process . A full in-depth discussion of this condition and its implications is provided in Appendix E.4. As the regularized best response admits a unique optimizer , the previous condition ensures the uniqueness of the fixed point of the associated operator, i.e.,
(8) |
with the unique fixed point of this operator. This implies that, in this setting, we obtain the uniqueness of the MFNE, corresponding to the pair .
A- 3.
The MF-MDP is unichain: for every fixed , every stationary policy induces a unique stationary distribution over the state space, satisfying, for all , the following equation:
Moreover, there exist constants and such that the following uniform mixing property holds for all and :
(9) |
The unichain property eliminates ambiguities in the initial population distribution caused by multiple recurrent classes, which could otherwise complicate value function evaluation and policy improvement steps. As highlighted in Shani et al. (2020), optimal policies are defined only within the recurrent states of the Markov reward process at equilibrium. Moreover, the ergodicity property naturally aligns with the regularization framework, as it promotes exploration within the Markov chain, leading to a more robust and stable optimization landscape.
A- 4.
We have that
The structured interaction with the environment in the considered RL paradigm requires this concentration property of the occupation measure with respect to the reset distribution . This property enhances sample efficiency and aids in the accurate estimation of key quantities, such as the value function.
4 Exact algorithm: Exact MF-TRPO
We now introduce Exact MF-TRPO to solve the optimization problem (3). Note that the entropic regularization used in Exact MF-TRPO is particularly well-suited for TRPO (see, e.g., Shani et al., 2020), as the proximal policy update fully exploits the entropy term, enabling a soft-max closed-form updates in terms of the -functions. This property enhances convergence by ensuring smoother and more stable policy iterations.
Although (3) is not a convex optimization problem, the adaptive nature of the regularization term allows us to use mirror descent techniques from convex analysis to establish strong convergence guarantees. In particular, we derive finite-sample complexity bounds showing that show that Exact MF-TRPO converges at a rate of .
TRPO.
For a fixed mean-field population distribution , Exact TRPO provides a reliable approximation of the value function. The convergence of the algorithm is explicitly influenced by the regularization parameter , which determines the optimal step size .
The proposed algorithm relies on the subroutine PolicyUpdate, a key step in refining the policy at each iteration. This update mechanism is inherently tied to the regularization scheme employed in the mirror ascent formulation of TRPO. Specifically, with entropic regularization, as shown in Beck (2017), the policy update admits a closed-form solution in the form of a softmax function. This structure inherently ensures that the updated policy remains within the probabilistic simplex without requiring additional projections. This fosters smooth and stable learning while preventing overly aggressive updates that could destabilize the optimization process.
This policy update leverages a function and a policy at iteration , leading to an explicit update rule given by
(10) |
with learning rate .
Shani et al. (2020) establishes error bounds for the approximation of value functions that can be directly generalized to our setting. Building on these results, we quantify the gap between the value function of the computed policy and the optimal value function under the given mean-field distribution. This allows to control the policy itself, and derive policy guarantees.
Proposition 4.1.
Suppose that Assumptions 1 and 3 hold. Let be the sequence generated by the Exact TRPO algorithm. Then, there exists a constant such that, for all , we have
MF-TRPO.
We now present Exact MF-TRPO, which iteratively solves the MFG problem (3) by updating the population distribution using the output of Exact TRPO. This approach assumes direct access to the transition kernel and cost function, eliminating the need for stochastic approximation in policy updates.
The analysis of Exact MF-TRPO is instrumental to understand the performances of its sample-based counterpart. We do this providing precise theoretical guarantees on convergence rates without the additional complexity of sampling-induced errors. Building on this deterministic setup, we then focus on the convergence behavior of the broader Sample-Based MF-TRPO framework. We use the label “informal” to avoid overloading the main text with technical assumptions (e.g., on learning rates), rigorously stated in Appendix C.
Proposition 4.2 (informal).
Suppose that Assumptions 1, 2, and 3 hold. Then, there exists a constant such that the sequence generated by Exact MF-TRPO satisfies
for , with .
All constants appearing in our results are explicitly defined and detailed in Appendix C, where we also provide complete proofs supporting our theoretical guarantees.
This convergence result proves its effectiveness in tackling the challenges inherent in the non-linear and non-gradient structure of MFNE. Below, we summarize the key insights derived from this result:
-
•
Convergence. We establish an exponential rate of convergence in the first term of the bound to the equilibrium population distribution , while the second one captures the finite-sample bias in the best-response computation.
-
•
Learning Rates constraints. The theorem provides explicit constraints on the step size (cf. condition (17)), ensuring stability and preventing oscillations or divergence in the optimization process.
-
•
Explicit Dependence on Model Parameters. All constants in the convergence bound are fully characterized in terms of the structural parameters of the model (cf. Appendix C).
-
•
Controlled Policy Learning Bias. The bias introduced by policy updates in Exact TRPO, bounded by (Proposition 4.2), remains controlled throughout the iterative process. This ensures algorithmic stability, even with approximations in policy optimization.
With these results, we can now explicitly quantify the parameter corresponding to the proximity of our obtained solution with respect to the MFNE.
Corollary 4.3.
Suppose that Assumptions 1, 2, and 3 hold. Let (resp. ) the output of Exact MF-TRPO (resp. Exact TRPO). Then, there exists a constant , such that is -MFNE, with
5 Stochastic approximation: Sample-Based MF-TRPO
We provide Sample-Based MF-TRPO, a model-free variant of the previous algorithm designed to operate without explicit knowledge of the environment’s dynamics, nor the reward function. Sample-Based MF-TRPO utilizes sampled trajectories to estimate these updates, making it more applicable to real-world scenarios.
TRPO.
First, we adapt Exact TRPO to estimate policy updates in a data-driven manner. This approach leverages sampled trajectories to approximate the policy gradient, providing quantitative bounds on the proximity of the best response. Additionally, this TRPO formulation is particularly well-suited to the considered oracle-based framework, incorporating the -restart modeling. This structure ensures robust exploration while maintaining stability in policy updates, aligning naturally with the trust-region optimization paradigm.
The inherent stochasticity in the updates prevents us from establishing sample complexity bounds on the last iterate of the algorithm. However, by leveraging an averaging scheme, implemented through a dedicated subroutine (cf. Remark D.1), we mitigate this variability and provide clear quantitative bounds on the desired gap. Specifically, the policy we focus on is the uniform mixture over the first policies. This averaging scheme is standard in the RL literature and, in the unregularized case, satisfies the identity
(11) |
While we do not have a statistically feasible expression for this mixture policy, it is straightforward to sample from , using Uniform-Mixture, as discussed in Remark D.1.
Proposition 5.1.
Suppose Assumptions 1, 2, 3, and 4 hold. Fix . Let be the output of Sample-Based TRPO, over iterations.
Then, there exists such that the following holds with probability greater than
All the constants appearing in our results are explicitly defined and detailed in Appendix D, where we also provide the complete proofs supporting our theoretical guarantees.
MF-TRPO.
We present a version of Sample-Based MF-TRPO, with the full algorithm and detailed implementation provided in Appendix D.3.
One key aspect of the algorithm is the initialization step. Unlike in a generative model paradigm, access to a state is only available through a subroutine initialized at the restart distribution . For further details on this procedure, we refer to Section D.2.
The theoretical foundation of the convergence guarantees of this algorithm relies on deriving high-probability estimates
(cf. Proposition D.3). This is a crucial step in sample complexity bounds and is achieved using a martingale-based argument (Harvey et al., 2019). By leveraging concentration inequalities for martingales, the analysis ensures that the error in estimating key quantities, such as the policy value and state distributions, remains controlled with high probability throughout the learning process.
Proposition 5.2 (informal).
Suppose Assumptions 1, 2, 3, and 4 hold. Fix . Then, there exists a constant such that the sequence generated by Sample-Based MF-TRPO satisfies, for ,
All the constants appearing in our results are explicitly defined and detailed in Appendix D, where we also provide the complete proofs supporting our theoretical guarantees.
The sample-based convergence analysis of Sample-Based MF-TRPO in the model-free setting demonstrates that the algorithm preserves the same desirata described in Section 4 driving the convergence of its exact counterpart Exact MF-TRPO—proximal updates, entropic regularization, and trust-region optimization—remain intact. While having a model-agnostic nature, its theoretical power of estimation of the MFNE is preserved, as explicit knowledge of the environment’s transition dynamics or reward structure is required. It builds on trajectory sampling to iteratively refine policy updates while maintaining stability and efficiency.
This result entails that the stochastic error in the estimation of the mean-field distribution at each iteration does not compound throughout the iterative process. By leveraging concentration inequalities and high-probability guarantees, the cumulative impact of these estimation errors remains controlled, preventing divergence or instability. As a result, Sample-Based MF-TRPO retains strong convergence guarantees, making it a practical approach for solving MFG in a data-driven manner.
Moreover, we can also bound the exploitability as follows.
Corollary 5.3.
Suppose that Assumptions 1, 2, 3, and 4 hold. Fix . Let (resp. ) the output of Sample-Based MF-TRPO (resp. the iterates of Sample-Based TRPO). Then, there exists a constant such that
The complexity analysis demonstrates that the proposed Sample-Based MF-TRPO algorithm achieves a computational cost scaling as to reach an -MFNE (cf. Remark D.7). This scaling emerges naturally from two principal contributions: the inner loop performing the policy optimization via Sample-Based TRPO, and the outer population distribution update. These results align well with established convergence bounds in Shani et al. (2020), emphasizing a balanced trade-off between computational efficiency and the precision required to approximate the MFNE.
6 Numerical Experiments
We present here the results of the numerical experiments obtained with the Sample-Based MF-TRPO algorithm. The environment considered is a Grid-Based Crowd Modeling game where, from a given initial distribution, agents are tasked with moving through a grid, avoiding both static obstacles and potential overcrowding. A representative player’s state corresponds to her position within the grid, and at every time step, she can choose to move in any direction or stay in place. The reward structure imposes a small penalty for movement, offers a slight incentive for staying, and discourages agents from entering overcrowded areas. In addition, agents are encouraged to move toward a designated target, that is,
where is a -distance between the corresponding states, and the crowd reward is defined as
where , with being the indicator function and being a crowd-aversion parameter. We refer to Appendix F for additional details and experimental results related to the Exact MF-TRPO algorithm. The environment used here is a grid featuring three walls located at coordinates , , and . The point of interest is located in the bottom-right corner of the grid, and all the players start clustered in the top-left corner. In Figure 1 it is possible to observe that the exploitability behavior of the Sample-Based MF-TRPO algorithm converges after a few iterations, matching the theoretical predictions, whereas Figure 2 illustrates the progression of the mean field distribution across three different time steps during the learning phase, thus demonstrating that the players progressively learn to distribute themselves around the point of interest, preserving spread over the whole state space.





7 Conclusion
In this work, we introduced Exact MF-TRPO, a novel algorithm for computing MFNE in ergodic MFG. By leveraging the trust-region policy optimization framework, we established explicit non-asymptotic convergence guarantees, demonstrating that Exact MF-TRPO inherits the rate from TRPO, ensuring efficient learning in structured multi-agent systems.
To bridge the gap between theoretical guarantees and practical applicability, we further developed Sample-Based MF-TRPO, a model-free variant that estimates policy updates solely from sampled trajectories, under the -restart RL paradigm. Using concentration inequalities, we provided finite-sample complexity bounds for this algorithm, proving convergence under more relaxed assumptions compared to recent literature. Moreover, we show that a total number of calls to the environment that scales as , consistent with standard RL results (see, e.g., Shani et al., 2020). This result highlights the potential of RL techniques for scalable and data-driven MFG solutions.
Overall, our work contributes to the growing intersection of MFG and RL, providing both theoretical insights and algorithmic advancements. Future directions include extending these methods to more general MFG settings, such as those with continuous state spaces, and exploring adaptive sampling techniques to further improve efficiency in real-world applications.
Acknowledgements
The work of A.O. and E.M. was funded by the European Union (ERC-2022-SYG-OCEAN-101071601). Views and opinions expressed are however those of the author only and do not necessarily reflect those of the European Union or the European Research Council Executive Agency. Neither the European Union nor the granting authority can be held responsible for them. The work of L.M. and E.M. has been supported by Technology Innovation Institute (TII), project Fed2Learn. The work of D.T. has been supported by the Paris Île-de-France Région in the framework of DIM AI4IDF.
Impact Statement
This paper presents work whose goal is to advance the field of Machine Learning. There are many potential societal consequences of our work, none which we feel must be specifically highlighted here.
References
- Achdou & Capuzzo-Dolcetta (2010) Achdou, Y. and Capuzzo-Dolcetta, I. Mean field games: numerical methods. SIAM Journal on Numerical Analysis, 48(3):1136–1162, 2010.
- Achdou & Porretta (2016) Achdou, Y. and Porretta, A. Convergence of a finite difference scheme to weak solutions of the system of partial differential equations arising in mean field games. SIAM Journal on Numerical Analysis, 54(1):161–186, 2016.
- Achdou et al. (2012) Achdou, Y., Camilli, F., and Capuzzo-Dolcetta, I. Mean field games: numerical methods for the planning problem. SIAM Journal on Control and Optimization, 50(1):77–109, 2012.
- Achdou et al. (2020) Achdou, Y., Cardaliaguet, P., Delarue, F., Porretta, A., Santambrogio, F., Achdou, Y., and Laurière, M. Mean field games and applications: Numerical aspects. Mean Field Games: Cetraro, Italy 2019, pp. 249–307, 2020.
- Agarwal et al. (2020) Agarwal, A., Kakade, S., and Yang, L. F. Model-based reinforcement learning with a generative model is minimax optimal. In Conference on Learning Theory, pp. 67–83. PMLR, 2020.
- Alasseur et al. (2020) Alasseur, C., Ben Taher, I., and Matoussi, A. An extended mean field game for storage in smart grids. Journal of Optimization Theory and Applications, 184:644–670, 2020.
- Algumaei et al. (2023) Algumaei, T., Solozabal, R., Alami, R., Hacid, H., Debbah, M., and Takáč, M. Regularization of the policy updates for stabilizing mean field games. In Pacific-Asia Conference on Knowledge Discovery and Data Mining, pp. 361–372. Springer, 2023.
- Anahtarci et al. (2023a) Anahtarci, B., Kariksiz, C. D., and Saldi, N. Learning mean-field games with discounted and average costs. Journal of Machine Learning Research, 24(17):1–59, 2023a.
- Anahtarci et al. (2023b) Anahtarci, B., Kariksiz, C. D., and Saldi, N. Q-learning in regularized mean-field games. Dynamic Games and Applications, 13(1):89–117, 2023b.
- Anand et al. (2024) Anand, E., Karmarkar, I., and Qu, G. Mean-field sampling for cooperative multi-agent reinforcement learning. arXiv preprint arXiv:2412.00661, 2024.
- Angiuli et al. (2021) Angiuli, A., Fouque, J.-P., and Laurière, M. Unified reinforcement q-learning for mean field game and control problems. arXiv preprint arXiv:2006.13912, 2021.
- Angiuli et al. (2023) Angiuli, A., Fouque, J.-P., Laurière, M., and Zhang, M. Convergence of multi-scale reinforcement Qlearning algorithms for mean field game and control problems. arXiv preprint arXiv:2312.06659, 2023.
- Arapostathis et al. (2017) Arapostathis, A., Biswas, A., and Carroll, J. On solutions of mean field games with ergodic cost. Journal de Mathématiques Pures et Appliquées, 107(2):205–251, 2017.
- Asadi et al. (2018) Asadi, K., Misra, D., and Littman, M. Lipschitz continuity in model-based reinforcement learning. In International Conference on Machine Learning, pp. 264–273. PMLR, 2018.
- Austrin et al. (2011) Austrin, P., Braverman, M., and Chlamtáč, E. Inapproximability of np-complete variants of nash equilibrium. In International Workshop on Approximation Algorithms for Combinatorial Optimization, pp. 13–25. Springer, 2011.
- Azar et al. (2013) Azar, G. M., Munos, R., and Kappen, H. J. Minimax pac bounds on the sample complexity of reinforcement learning with a generative model. Machine learning, 91:325–349, 2013.
- Bardi & Priuli (2014) Bardi, M. and Priuli, F. S. Linear-quadratic n-person and mean-field games with ergodic cost. SIAM Journal on Control and Optimization, 52(5):3022–3052, 2014.
- Bassière et al. (2024) Bassière, A., Dumitrescu, R., and Tankov, P. A mean-field game model of electricity market dynamics. In Quantitative Energy Finance: Recent Trends and Developments, pp. 181–219. Springer, 2024.
- Becherer & Hesse (2024) Becherer, D. and Hesse, S. Common noise by random measures: Mean-field equilibria for competitive investment and hedging. arXiv preprint arXiv:2408.01175, 2024.
- Beck (2017) Beck, A. First-order methods in optimization. SIAM, 2017.
- Bhandari & Russo (2024) Bhandari, J. and Russo, D. Global optimality guarantees for policy gradient methods. Operations Research, 2024.
- Braun et al. (2011) Braun, D. A., Ortega, P. A., Theodorou, E., and Schaal, S. Path integral control and bounded rationality. In 2011 IEEE symposium on adaptive dynamic programming and reinforcement learning (ADPRL), pp. 202–209. IEEE, 2011.
- Cardaliaguet et al. (2019) Cardaliaguet, P., Delarue, F., Lasry, J.-M., and Lions, P.-L. The master equation and the convergence problem in mean field games:(ams-201). Princeton University Press, 2019.
- Carmona & Laurière (2021) Carmona, R. and Laurière, M. Convergence analysis of machine learning algorithms for the numerical solution of mean field control and games i: The ergodic case. SIAM Journal on Numerical Analysis, 59(3):1455–1485, 2021.
- Carmona et al. (2013) Carmona, R., Fouque, J.-P., and Sun, L.-H. Mean field games and systemic risk. arXiv preprint arXiv:1308.2172, 2013.
- Carmona et al. (2018) Carmona, R., Delarue, F., et al. Probabilistic theory of mean field games with applications I-II. Springer, 2018.
- Chassagneux et al. (2019) Chassagneux, J.-F., Crisan, D., and Delarue, F. Numerical method for fbsdes of mckean–vlasov type. The Annals of Applied Probability, 29(3):1640–1684, 2019.
- Cover (1999) Cover, T. M. Elements of information theory. John Wiley & Sons, 1999.
- Cui & Koeppl (2021) Cui, K. and Koeppl, H. Approximately solving mean field games via entropy-regularized deep reinforcement learning. In International Conference on Artificial Intelligence and Statistics, pp. 1909–1917. PMLR, 2021.
- Cui & Koeppl (2022) Cui, K. and Koeppl, H. Learning graphon mean field games and approximate nash equilibria. In International Conference on Learning Representations, 2022.
- Doncel et al. (2022) Doncel, J., Gast, N., and Gaujal, B. A mean field game analysis of sir dynamics with vaccination. Probability in the Engineering and Informational Sciences, 36(2):482–499, 2022.
- Espinosa & Touzi (2015) Espinosa, G.-E. and Touzi, N. Optimal investment under relative performance concerns. Mathematical Finance, 25(2):221–257, 2015.
- Fischer & Silva (2021) Fischer, M. and Silva, F. J. On the asymptotic nature of first order mean field games. Applied Mathematics & Optimization, 84:2327–2357, 2021.
- Flandoli et al. (2022) Flandoli, F., Ghio, M., and Livieri, G. N-player games and mean field games of moderate interactions. Applied Mathematics & Optimization, 85(3):38, 2022.
- Foerster et al. (2018) Foerster, J., Farquhar, G., Afouras, T., Nardelli, N., and Whiteson, S. Counterfactual multi-agent policy gradients. Proceedings of the AAAI conference on artificial intelligence, 32(1), 2018.
- Fort et al. (2011) Fort, G., Moulines, E., and Priouret, P. Convergence of adaptive and interacting markov chain monte carlo algorithms. The Annals of Statistics, 39(6):3262–3289, 2011.
- Fox et al. (2016) Fox, R., Pakman, A., and Tishby, N. Taming the noise in reinforcement learning via soft updates. Proceedings of the Thirty-Second Conference on Uncertainty in Artificial Intelligence, 2016.
- Geist et al. (2019) Geist, M., Scherrer, B., and Pietquin, O. A theory of regularized markov decision processes. In International Conference on Machine Learning, pp. 2160–2169. PMLR, 2019.
- Geist et al. (2022) Geist, M., Pérolat, J., Laurière, M., Elie, R., Perrin, S., Bachem, O., Munos, R., and Pietquin, O. Concave utility reinforcement learning: the mean-field game viewpoint, 2022. URL https://arxiv.org/abs/2106.03787.
- Germain et al. (2022) Germain, M., Mikael, J., and Warin, X. Numerical resolution of mckean-vlasov fbsdes using neural networks. Methodology and Computing in Applied Probability, 24(4):2557–2586, 2022.
- Gronauer & Diepold (2022) Gronauer, S. and Diepold, K. Multi-agent deep reinforcement learning: a survey. Artificial Intelligence Review, 55(2):895–943, 2022.
- Guo et al. (2019) Guo, X., Hu, A., Xu, R., and Zhang, J. Learning mean-field games. Advances in neural information processing systems, 32, 2019.
- Guo et al. (2023) Guo, X., Hu, A., Xu, R., and Zhang, J. A general framework for learning mean-field games. Mathematics of Operations Research, 48(2):656–686, 2023.
- Haarnoja et al. (2017) Haarnoja, T., Tang, H., Abbeel, P., and Levine, S. Reinforcement learning with deep energy-based policies. In International conference on machine learning, pp. 1352–1361. PMLR, 2017.
- Harvey et al. (2019) Harvey, N. J., Liaw, C., Plan, Y., and Randhawa, S. Tight analyses for non-smooth stochastic gradient descent. In Conference on Learning Theory, pp. 1579–1613. PMLR, 2019.
- Howard (1960) Howard, R. A. Dynamic programming and markov processes. John Wiley, 1960.
- Howard & Matheson (1972) Howard, R. A. and Matheson, J. E. Risk-sensitive markov decision processes. Management science, 18(7):356–369, 1972.
- Huang et al. (2003) Huang, M., Caines, P. E., and Malhamé, R. P. Individual and mass behaviour in large population stochastic wireless power control problems: centralized and nash equilibrium solutions. In 42nd IEEE international conference on decision and control (IEEE cat. No. 03CH37475), volume 1, pp. 98–103. IEEE, 2003.
- Huang et al. (2005) Huang, M., Malhamé, R. P., and Caines, P. E. Nash equilibria for large-population linear stochastic systems of weakly coupled agents. In Analysis, control and optimization of complex dynamic systems, pp. 215–252. Springer, 2005.
- Huang et al. (2006a) Huang, M., Malhamé, R. P., and Caines, P. E. Large population stochastic dynamic games: closed-loop mckean-vlasov systems and the nash certainty equivalence principle. Commun. Inf. Syst., 6(1):221–252, 2006a.
- Huang et al. (2006b) Huang, M., Malhamé, R. P., and Caines, P. E. Nash certainty equivalence in large population stochastic dynamic games: Connections with the physics of interacting particle systems. In Proceedings of the 45th IEEE Conference on Decision and Control, pp. 4921–4926. IEEE, 2006b.
- Iqbal & Sha (2019) Iqbal, S. and Sha, F. Actor-attention-critic for multi-agent reinforcement learning. In International conference on machine learning, pp. 2961–2970. PMLR, 2019.
- Kakade & Langford (2002) Kakade, S. and Langford, J. Approximately optimal approximate reinforcement learning. In Proceedings of the Nineteenth International Conference on Machine Learning, pp. 267–274, 2002.
- Kakade (2003) Kakade, S. M. On the sample complexity of reinforcement learning. University of London, University College London (United Kingdom), 2003.
- Kearns & Singh (1998) Kearns, M. and Singh, S. Finite-sample convergence rates for q-learning and indirect algorithms. Advances in neural information processing systems, 11, 1998.
- Kober et al. (2013) Kober, J., Bagnell, J. A., and Peters, J. Reinforcement learning in robotics: A survey. The International Journal of Robotics Research, 32(11):1238–1274, 2013.
- Lacker & Zariphopoulou (2019) Lacker, D. and Zariphopoulou, T. Mean field and n-agent games for optimal investment under relative performance criteria. Mathematical Finance, 29(4):1003–1038, 2019.
- Lasry & Lions (2006a) Lasry, J.-M. and Lions, P.-L. Jeux à champ moyen. i–le cas stationnaire. Comptes Rendus Mathématique, 343(9):619–625, 2006a.
- Lasry & Lions (2006b) Lasry, J.-M. and Lions, P.-L. Jeux à champ moyen. ii–horizon fini et contrôle optimal. Comptes Rendus. Mathématique, 343(10):679–684, 2006b.
- Lasry & Lions (2007) Lasry, J.-M. and Lions, P.-L. Mean field games. Japanese journal of mathematics, 2(1):229–260, 2007.
- Laurière et al. (2022) Laurière, M., Perrin, S., Geist, M., and Pietquin, O. Learning mean field games: A survey. arXiv preprint arXiv:2205.12944, pp. 19–49, 2022.
- Lavigne & Tankov (2023) Lavigne, P. and Tankov, P. Decarbonization of financial markets: a mean-field game approach. arXiv preprint arXiv:2301.09163, 2023.
- Le Lan et al. (2021) Le Lan, C., Bellemare, M. G., and Castro, P. S. Metrics and continuity in reinforcement learning. Proceedings of the AAAI Conference on Artificial Intelligence, 35(9):8261–8269, 2021.
- Mao et al. (2022) Mao, W., Qiu, H., Wang, C., Franke, H., Kalbarczyk, Z., Iyer, R., and Basar, T. A mean-field game approach to cloud resource management with function approximation. Advances in Neural Information Processing Systems, 35:36243–36258, 2022.
- Marcus et al. (1997) Marcus, S. I., Fernández-Gaucherand, E., Hernández-Hernandez, D., Coraluppi, S., and Fard, P. Risk sensitive markov decision processes. In Systems and control in the twenty-first century, pp. 263–279. Springer, 1997.
- Matignon et al. (2007) Matignon, L., Laurent, G. J., and Le Fort-Piat, N. Hysteretic q-learning: an algorithm for decentralized reinforcement learning in cooperative multi-agent teams. In 2007 IEEE/RSJ International Conference on Intelligent Robots and Systems, pp. 64–69. IEEE, 2007.
- Mei et al. (2020) Mei, J., Xiao, C., Szepesvari, C., and Schuurmans, D. On the global convergence rates of softmax policy gradient methods. In International conference on machine learning, pp. 6820–6829. PMLR, 2020.
- Mnih (2016) Mnih, V. Asynchronous methods for deep reinforcement learning. International Conference on Machine Learning, pp. 1928–1937, 2016.
- Nash (1950) Nash, J. Equilibrium points in n-person games. Proceedings of the national academy of sciences, 36(1):48–49, 1950.
- Nash (1951) Nash, J. Non-cooperative games. Annals of Mathematics, 54(2):286–295, 1951.
- Neu et al. (2017) Neu, G., Jonsson, A., and Gómez, V. A unified view of entropy-regularized markov decision processes. arXiv preprint arXiv:1705.07798, 2017.
- O’Donoghue et al. (2017) O’Donoghue, B., Munos, R., Kavukcuoglu, K., and Mnih, V. Combining policy gradient and q-learning. 5th International Conference on Learning Representations, 2017.
- Pérolat et al. (2022) Pérolat, J., Perrin, S., Elie, R., Laurière, M., Piliouras, G., Geist, M., Tuyls, K., and Pietquin, O. Scaling mean field games by online mirror descent. In Proceedings of the 21st International Conference on Autonomous Agents and Multiagent Systems, pp. 1028–1037, 2022.
- Perrin et al. (2020) Perrin, S., Pérolat, J., Laurière, M., Geist, M., Elie, R., and Pietquin, O. Fictitious play for mean field games: Continuous time analysis and applications. Advances in neural information processing systems, 33:13199–13213, 2020.
- Perrin et al. (2021) Perrin, S., Laurière, M., Pérolat, J., Geist, M., Élie, R., and Pietquin, O. Mean field games flock! the reinforcement learning way. arXiv preprint arXiv:2105.07933, 2021.
- Perrin et al. (2022) Perrin, S., Laurière, M., Pérolat, J., Élie, R., Geist, M., and Pietquin, O. Generalization in mean field games by learning master policies. Proceedings of the AAAI Conference on Artificial Intelligence, 36(9):9413–9421, 2022.
- Peters et al. (2010) Peters, J., Mulling, K., and Altun, Y. Relative entropy policy search. Proceedings of the AAAI Conference on Artificial Intelligence, 24(1):1607–1612, 2010.
- Puterman & Shin (1978) Puterman, M. L. and Shin, M. C. Modified policy iteration algorithms for discounted markov decision problems. Management Science, 24(11):1127–1137, 1978.
- Ruszczyński (2010) Ruszczyński, A. Risk-averse dynamic programming for markov decision processes. Mathematical programming, 125:235–261, 2010.
- Saldi (2020) Saldi, N. Discrete-time average-cost mean-field games on polish spaces. Turkish Journal of Mathematics, 44(2):463–480, 2020.
- Samvelyan et al. (2019) Samvelyan, M., Rashid, T., De Witt, C. S., Farquhar, G., Nardelli, N., Rudner, T. G., Hung, C.-M., Torr, P. H., Foerster, J., and Whiteson, S. The starcraft multi-agent challenge. arXiv preprint arXiv:1902.04043, 2019.
- Scherrer & Geist (2014) Scherrer, B. and Geist, M. Local policy search in a convex space and conservative policy iteration as boosted policy search. In Machine Learning and Knowledge Discovery in Databases: European Conference, ECML PKDD 2014, Nancy, France, September 15-19, 2014. Proceedings, Part III 14, pp. 35–50. Springer, 2014.
- Schulman et al. (2015) Schulman, J., Levine, S., Moritz, P., Jordan, M., and Abbeel, P. Trust region policy optimization. In Proceedings of the 32nd International Conference on International Conference on Machine Learning-Volume 37, pp. 1889–1897, 2015.
- Shalev-Shwartz et al. (2016) Shalev-Shwartz, S., Shammah, S., and Shashua, A. Safe, multi-agent, reinforcement learning for autonomous driving. arXiv preprint arXiv:1610.03295, 2016.
- Shani et al. (2020) Shani, L., Efroni, Y., and Mannor, S. Adaptive trust region policy optimization: Global convergence and faster rates for regularized mdps. Proceedings of the AAAI Conference on Artificial Intelligence, 34(04):5668–5675, 2020.
- Sidford et al. (2018) Sidford, A., Wang, M., Wu, X., Yang, L., and Ye, Y. Near-optimal time and sample complexities for solving markov decision processes with a generative model. Advances in Neural Information Processing Systems, 31, 2018.
- Sutton & Barto (2018a) Sutton, R. and Barto, A. G. Reinforcement learning: An introduction. SIAM Rev, 6(2):423, 2018a.
- Sutton & Barto (2018b) Sutton, R. S. and Barto, A. G. Reinforcement learning: An introduction. MIT press, 2018b.
- Szepesvári (2022) Szepesvári, C. Algorithms for reinforcement learning. Springer nature, 2022.
- Tangpi & Zhou (2024) Tangpi, L. and Zhou, X. Optimal investment in a large population of competitive and heterogeneous agents. Finance and Stochastics, 28(2):497–551, 2024.
- Weinan et al. (2017) Weinan, E., Han, J., and Jentzen, A. Deep learning-based numerical methods for high-dimensional parabolic partial differential equations and backward stochastic differential equations. Communications in Mathematics and Statistics, 4(5):349–380, 2017.
- Wiering et al. (2000) Wiering, M. A. et al. Multi-agent reinforcement learning for traffic light control. In Machine Learning: Proceedings of the Seventeenth International Conference (ICML’2000), pp. 1151–1158, 2000.
- Williams & Peng (1991) Williams, R. J. and Peng, J. Function optimization using connectionist reinforcement learning algorithms. Connection Science, 3(3):241–268, 1991.
- Yardim et al. (2023) Yardim, B., Cayci, S., Geist, M., and He, N. Policy mirror ascent for efficient and independent learning in mean field games. In International Conference on Machine Learning, pp. 39722–39754. PMLR, 2023.
- Zaman et al. (2023) Zaman, M. A. U., Koppel, A., Bhatt, S., and Basar, T. Oracle-free reinforcement learning in mean-field games along a single sample path. In International Conference on Artificial Intelligence and Statistics, pp. 10178–10206. PMLR, 2023.
- Zhang et al. (2021) Zhang, K., Yang, Z., and Başar, T. Multi-agent reinforcement learning: A selective overview of theories and algorithms. Handbook of reinforcement learning and control, pp. 321–384, 2021.
- Ziebart (2010) Ziebart, B. D. Modeling purposeful adaptive behavior with the principle of maximum causal entropy. Carnegie Mellon University, 2010.
- Ziebart et al. (2010) Ziebart, B. D., Bagnell, J. A., and Dey, A. K. Modeling interaction via the principle of maximum causal entropy. International Conference on Machine Learning (ICML), pp. 1255––1262, 2010.
Appendix
In the appendix, we provide a detailed exposition of the theoretical foundations and technical results supporting our main contributions. In Appendix A, we give a broad description of related work in MFG and MFRL, and in Appendix B, we lay out the fundamental framework of the MFG problem we aim to study. This section introduces the ergodic MFG formulation and its connection to the discounted setting. We also discuss the role of entropic regularization in RL and MFG, emphasizing its impact on stability and the approximation of Nash equilibria. In Section B.4, we present key assumptions we use in the proofs. In Appendix C, we present the exact TRPO-based algorithm for solving the MFG problem. This section provides a precise formulation of the exact methods and establishes its theoretical convergence guarantees. The adaptation of TRPO to the MFG setting is examined in detail, leveraging the structure of entropy-regularized RL. We present formal results on convergence rates and error bounds, ensuring the effectiveness and reliability of these methods in computing approximate MFNE. In Appendix D, we extend our analysis to the sample-based version of the algorithm. Here, we derive global sample complexity results and analyze the statistical error introduced by sampling. We establish high-probability bounds on the approximation error at each iterations. Next, in Appendix E, we provide additional proofs and auxiliary results that support the theoretical analysis conducted in the previous sections. These supplementary results play a crucial role in rigorously validating the convergence and stability properties of our proposed algorithms. Finally, in Appendix F we provide additional experimental details as well as experiments on exact versions of the algorithm.
Appendix A Related works
Except for the well-known Linear-Quadratic (LQ) case, where explicit solutions can be derived analytically or through simple ordinary differential equations, computing MFNE numerically remains a challenging and active research area. A vast body of literature has focused on addressing the computational complexity of these models, leading to three major methodological approaches.
The first class of methods relies on PDE approximations, leveraging the classical formulation of MFG through the Hamilton–Jacobi–Bellman equation coupled with the Fokker–Planck–Kolmogorov equation (Achdou & Capuzzo-Dolcetta, 2010; Achdou et al., 2012; Achdou & Porretta, 2016; Achdou et al., 2020). While mathematically elegant, these methods suffer from the well-documented curse of dimensionality, as solving PDEs numerically becomes intractable in high-dimensional state spaces.
The second approach leverages deep learning techniques with neural networks to approximate equilibrium solutions. These methods exploit function approximation to bypass explicit PDE resolution, making them a promising alternative in high-dimensional settings (Weinan et al., 2017; Chassagneux et al., 2019; Germain et al., 2022). However, they lack rigorous convergence guarantees. Additionally, these methods struggle to capture model specifications in a purely data-driven manner directly, limiting their adaptability in real-world applications.
The third category of numerical methods integrates RL techniques into the MFG framework, leading to two primary subcategories. The first subcategory employs RL as a solver for a given MFG model, using value-based or policy-based RL techniques to approximate Nash equilibria (Pérolat et al., 2022; Perrin et al., 2022). These approaches have achieved state-of-the-art performance in various settings and have been successfully deployed in large-scale simulations.
The second subcategory focuses on developing model-free RL algorithms for solving MFG, often incorporating regularization techniques to enhance stability (Cui & Koeppl, 2021; Perrin et al., 2021). While these methods show promising empirical performance, theoretical guarantees on finite-sample complexity remain limited, particularly for model-based RL approaches.
RL for MFG.
Among model-free RL approaches, we identify three main categories: value function-based methods, actor-critic methods—which combine value-based and policy-based approaches—and policy-oriented methods. As in classical RL, proximal methods have emerged as state-of-the-art techniques across a wide range of tasks, both in model-specific and model-agnostic settings.
Several studies have explored -learning-based algorithms for MFG, establishing theoretical convergence guarantees (Angiuli et al., 2021, 2023; Guo et al., 2019; Anahtarci et al., 2023b). However, these approaches rely on stringent assumptions that are often difficult to verify in practice. Moreover, -learning operates within the generative model oracle framework of RL, which assumes full query access to state-action transitions. In contrast, the restart oracle assumption provides a significantly weaker paradigm, offering a more practical and adaptable alternative for real-world learning settings.
Two-timescale updates have been employed in, e.g., Zaman et al. (2023) and Mao et al. (2022), where policy updates operate on a faster timescale, while the population distribution evolves more slowly based on model-based estimates of state dynamics. However, these approaches introduce significant challenges in both theoretical analysis and practical implementation. Their convergence often relies on strong assumptions that are difficult to verify, particularly in non-stationary environments.
In the MFG setting, proximal methods have been adopted for their stability properties and strong empirical performance. Among these approaches, we highlight the work of Pérolat et al. (2022), where the authors investigate Online Mirror Descent (OMD) from a model-specific perspective. Yardim et al. (2023) build up in this direction, developing the model-free analysis. Their approach, however, does not consider the crucial aspect of population updates, placing itself in the no-manipulation regime. In contrast, our work explicitly discusses how the monotonicity assumption can stabilize the population evolution up to a certain threshold, a perspective that aligns naturally with the ergodic nature of the Markov reward process once a policy has been selected. This additional consideration allows us to provide a more refined analysis of the learning dynamics in MFG, ensuring a more structured and well-posed approach to policy optimization.
Furthermore, Yardim et al. (2023) imposes stringent assumptions by requiring that policies remain uniformly bounded away from zero by a fixed constant, effectively enforcing an overly rigid form of regularization. This constraint exceeds the controlled bias typically introduced by entropic regularization. Notably, a similar assumption appears in Angiuli et al. (2021, 2023), where the Markov reward process is required to be absolutely continuous with respect to the uniform distribution over the state-action space—a significantly stronger condition than our unichain assumption.
By adopting a more flexible framework, we relax these restrictive conditions while improving sample complexity, particularly in terms of the number of required trajectories. Our analysis still achieves an error bound in trajectory estimates but under significantly milder assumptions.
An interesting direction for future work would be to connect our framework with that of Anand et al. (2024), which focuses on mean-field control. While their setting assumes cooperative agents optimizing a common objective, our work addresses non-cooperative mean-field games. Despite this key difference, it would be valuable to explore how our analytical tools—particularly for handling weak structural assumptions—could generalize their approach to broader settings.
Appendix B Framework
B.1 Ergodic MFG problem
The ergodic problem focuses on optimizing the long-term average performance of a stochastic system over an infinite time horizon. In contrast to finite time horizon problems, the emphasis is on stability and efficiency over time, making it essential for operations such as energy systems, financial markets and resource management. The goal is to find stationary policies that balance short-term costs with long-term gains. This approach is robust to uncertainties and ensures consistent performance despite stochastic disturbances.
Ergodic MFG have been first studied in continuous time and space problems (see, e.g., Bardi & Priuli, 2014; Arapostathis et al., 2017; Carmona & Laurière, 2021). In the context of discrete-time MFG, the ergodic setting has been studied by Saldi (2020) under the terminology of average-cost MFG. Anahtarci et al. (2023a) proposed a learning algorithm based on -learning and analyzed its convergence and sample complexity using a strict contraction argument.
Similarly, in Guo et al. (2023), the authors address the problem of evolving mean-field parameters, focusing on dynamically adjusting the population distribution over time. This differs from our approach, where we aim to learn policies without requiring explicit control or manipulation of the mean-field distribution at each step.
B.2 From ergodic MFG to discounted formulation
In this article, we focus on an ergodic equilibrium problem within the framework of mean-field games. This problem is traditionally defined as the unique solution resulting from the optimization of a long-term average cost function
for . We seek Nash equilibria with respect to this cost function, i.e., a tuple such that
-
•
;
-
•
.
Such a problem is independent of the initial condition and is stationary. Therefore, we can consider only constant vectors for , for any , a.
Moreover, given the stationarity of both the policy and the problem, we can further restrict ourselves to time-invariant policies and reformulate the problem as follows.
B.3 Regularization
Entropy regularization in RL.
Entropy regularization has been a prominent concept across various fields, including RL (Sutton & Barto, 2018b; Szepesvári, 2022). In dynamic programming and RL contexts, entropy-regularized Bellman equations and corresponding algorithms have been extensively studied to address key challenges. These include inducing safe exploration (Fox et al., 2016) and designing risk-sensitive policies (Howard & Matheson, 1972; Marcus et al., 1997; Ruszczyński, 2010). Additionally, these methods have been employed to model behaviors of imperfect decision-makers, as demonstrated by Ziebart et al. (2010); Ziebart (2010); Braun et al. (2011).
Beyond dynamic programming approaches, direct policy search methods have emerged as a powerful alternative for optimizing entropy-regularized objectives. These methods, which aim to drive safe online exploration in unknown Markov decision processes, have been explored in works such as Williams & Peng (1991); Peters et al. (2010); Schulman et al. (2015); Mnih (2016); O’Donoghue et al. (2017). Notably, state-of-the-art RL methods, including those by Mnih (2016); Schulman et al. (2015), leverage entropy-regularized policy search to balance exploration and exploitation effectively, highlighting the central role of regularization in achieving robust and safe learning.
Regularization, particularly the entropic one, has been extensively studied in the theoretical literature. In Neu et al. (2017), the authors provide a comprehensive analysis of mirror descent methods for RL, highlighting how regularization influences policy optimization and convergence properties. Similarly, Geist et al. (2019) formalize the theoretical impact of entropy regularization, demonstrating its role in stabilizing policy updates and improving exploration.
Regularization in RL for MFG.
In the inherently non-linear MFG setting, stabilizing policy updates is essential to ensuring convergence and preventing oscillatory behavior. In this setting, the underlying dynamics involve the interplay of numerous agents and necessitate a precise balance between individual and collective objectives. Regularization plays a key role by smoothing the cost landscape, mitigating instability, and creating well-conditioned optimization problems, ultimately leading to more reliable and efficient learning dynamics.
Moreover, MFG serve as approximations of the -player game problem in MARL, leveraging assumptions like anonymity and homogeneity to simplify the otherwise intractable dynamics of joint policy updates in large-scale systems. Introducing regularization into the MFG framework not only enhances stability but is theoretically justified, as the additional approximation error introduced by regularization is comparable to the inherent error of the MFG approximation itself.
Moreover, in the context of RL for MFG, regularize policies has been used to facilitates convergence of learning algorithms. Cui & Koeppl (2021) show that strict contraction property used by several other works fail to holds in general. The authors then studied a modified MFG with an entropy-regularized reward and showed that, for a sufficiently large degree of regularization, policy-iteration type RL algorithms can be shown to converge using contraction techniques. A similar approach has been used by Anahtarci et al. (2023b) to prove convergence of a -learning algorithm, and by Yardim et al. (2023) to prove the convergence of policies learned by independent learners in a regularized MFG. On the empirical side, policy regularization has also been used by Algumaei et al. (2023) through an algorithm relying on proximal policy optimization (PPO).
B.4 Discussion on the assumptions
Lipschitz property of the MDP parameters.
Assumption 1 of Lipschitz continuity on the parameters of the MDP, reward function and transition probability matrix , implies that the MDP does not change abruptly with respect to the state or action. This ensures that small perturbations in these variables lead to correspondingly small changes in the MDP’s dynamics.
This assumption is well-established in the RL literature (see, e.g., Asadi et al., 2018; Le Lan et al., 2021) and serves several critical purposes. First, it ensures the smoothness of value functions, essential for the stability of iterative optimization methods Second, it enables the derivation of meaningful error bounds and convergence rates, as shown in foundational works on policy optimization.
Moreover, as the reward function is defined on a compact domain, since and are finite, it is guaranteed that . A common practice in RL literature to normalize , i.e., , without loss of generality (Mei et al., 2020). This normalization simplifies expressions and makes algorithms scale-independent. However, in this work, we retain explicitly in our analysis to emphasize the clear dependence of all constants and bounds on the magnitude of the reward function. This approach ensures transparency in how the properties of influence the theoretical results and practical performance.
Unique recurrence property of the Markov reward process.
To ensure the well-posedness of the TRPO algorithm and derive meaningful performance bounds, we impose the unichain assumption 3. This assumption guarantees that the Markov chain induced by any admissible policy has a single recurrent class, potentially accompanied by a set of transient states. Such a property ensures that the long-term behavior of the Markov chain is well-defined, with a unique stationary distribution for each policy.
The unichain property plays a pivotal role in stabilizing the analysis of RL algorithms, particularly TRPO, as highlighted in Neu et al. (2017). It eliminates ambiguities on the initial population distribution arising from multiple recurrent classes, which could otherwise complicate the evaluation of value functions and policy improvement steps. As shown in in Puterman & Shin (1978), this condition is satisfied if all policies induce an irreducible and aperiodic Markov chain. Moreover, in the regularized setting, the regularization term helps in its satisfaction. Therefore, for any mean-field population profile , the Markov chain is irreducible and aperiodic, implying the existence of a unique stationary distribution and establishes the foundation for the mixing property (9).
The ergodicity assumption has been explored in various forms by different authors in the MFG literature. Notably, Angiuli et al. (2021, 2023) impose the condition that the induced Markov chain is aperiodic and absolutely continuous with respect to the uniform distribution over the state space. While this guarantees strong mixing properties, it is a highly restrictive assumption, as it effectively enforces immediate communication between all states, which is often unrealistic in practical applications. Instead, we generalize this assumption by adopting a standard ergodicity condition widely used in RL literature (Mei et al., 2020). This approach maintains the necessary stability properties while allowing for more realistic transition dynamics, ensuring broader applicability in complex multi-agent systems.
Finite concentration of the occupation measure.
In this paper, we adopt a RL paradigm where access to the environment is structured through a -restart, ensuring that each learning episode begins from a well-defined initial state distribution. The concentration of the occupation measure assumption is crucial for the convergence of the Sample-Based TRPO algorithm, ensuring that the estimation of the policy update remains stable and accurate over successive iterations. To compute this task, the algorithm operates in an episodic setting, where each episode begins by drawing an initial state from the restart distribution , followed by collecting a trajectory under the current policy . This episodic structure allows the algorithm to interact with the MDP in a controlled manner, facilitating the estimation of quantities like the value function .
This approach builds on the seminal work of Kakade (2003), which introduced the notion of a -restart model as an intermediary assumption in RL. The -restart model is weaker than having direct access to the true model or a generative model (Kearns & Singh, 1998; Azar et al., 2013; Sidford et al., 2018; Agarwal et al., 2020), as it does not require full knowledge of the transition kernel or the reward function. At the same time, it is stronger than the unrestricted case where no restarts are allowed, ensuring that the algorithm can sample states from a well-defined initial distribution at the start of each episode. This controlled interaction with the environment is crucial for accurately estimating value functions and gradients in the Sample-Based TRPO algorithm, ultimately enabling convergence guarantees.
The supremum in Assumption 4, often referred to as the concentrability coefficient, plays a critical role in the theoretical analysis of policy search algorithms. This concept was initially highlighted in the foundational work of Kakade & Langford (2002) and has since been extensively studied in the RL (RL) literature.
One of the reasons the concentrability coefficient has garnered attention is its frequent appearance in the analysis of approximate policy iteration schemes. Research by Scherrer & Geist (2014) and Bhandari & Russo (2024) shows that the concentrability coefficient often governs error propagation during learning. In essence, it provides bounds on how errors in approximating value functions or policies propagate through successive updates.
Appendix C Exact algorithms
Various algorithms have been proposed in the literature to address the exact MFG problem in scenarios where the MDP kernel and the reward function are fully accessible. In this case, the value function and a best response can be computed using dynamic programming and backward induction. This approach has been used, e.g., by Perrin et al. (2020) and Pérolat et al. (2022) to implement Fictitious Play (FP) and Online Mirror Descent (OMD) respectively. Cui & Koeppl (2022) presented a exact fixed point algorithm for graphon games. Angiuli et al. (2023) analyzed the convergence of a model-specific multi-scale algorithm for MFG.
This line of research often stems from the classical control theory and optimization frameworks, tailored to solve specific MFG problems with high precision. These methods focus on the exact representation of the MFG dynamics, providing critical insights into the equilibrium behavior of large-agent systems. In this work, we propose a novel adaptation of the TRPO algorithm, building on the robust framework of Shani et al. (2020). Our adaptation incorporates key elements of the MFG structure, leveraging entropic regularization and mean-field population dynamics. Moreover, we establish finite sample complexity results for this algorithm, demonstrating its theoretical convergence properties and its practical applicability in solving the ergodic MFG problem under a finite state-action setting.
C.1 TRPO - exact formulation
TRPO, inherently structured as a mirror descent method, proves particularly well-suited for entropy-regularized settings. This framework benefits from a significant simplification: the policy update admits a closed-form solution (Beck, 2017), expressed in terms of the -function associated with the current policy. By recasting the optimization problem in terms of -function computation, the algorithm focuses on the essential dynamics of the system, effectively bypassing the need for direct policy optimization over a high-dimensional space.
This closed-form update leverages the softmax form of the policy, a direct consequence of entropy regularization. The softmax structure ensures that the updated policies remain strictly in the interior of the probability simplex , avoiding deterministic solutions. This property not only facilitates numerical stability but also aligns with the theoretical foundations of the regularized problem. The use of first-order conditions becomes feasible and efficient, as the regularization term enforces a smooth, convex optimization landscape.
Moreover, the reward function’s linear dependence on the policy pairs seamlessly with the coercive nature of the entropy-regularized optimization problem. The coercivity guarantees that the optimal policies minimize the objective within the confines of the simplex, effectively balancing exploration and exploitation. This alignment between the problem structure and the algorithm’s mechanics underscores the power of TRPO in achieving convergence while maintaining theoretical guarantees in entropy-regularized MFG settings. By translating the original optimization problem into -function evaluations, the algorithm provides a practical yet robust pathway to finding approximate Nash equilibria in complex systems.
In the case where the mean-field population distribution parameter is fixed, the algorithm Exact TRPO provides a robust approximation to the value function. By iteratively updating the policy using trust region optimization techniques, the algorithm ensures convergence rates that explicitly depend on the regularization parameter . Given a fixed , the learning rate is optimally chosen to balance stability and efficiency in the policy updates. The following result establishes the error bounds for the value function approximation, highlighting the role of entropic regularization in convergence guarantees. Specifically, the bounds quantify the discrepancy between the value function induced by the computed policy and the optimal value function for the given mean-field population profile.
Theorem C.1 (Theorem 16 in Shani et al. (2020)).
Fix the initial distribution. Let be the sequence generated by the Exact TRPO algorithm. Then, there exists a constant such that
(12) |
Corollary C.2.
Proof.
Using Proposition E.2, we use the relationship between the total variation distance of a policy to the optimal policy and the corresponding difference in their value functions obtainine
C.2 Exact algorithm
We now analyze Algorithm 5, where no approximation is made. This algorithm is exact and does not involve any approximation. We provide a convergence result for this algorithm in the tabular setting.
Proposition C.3.
Proof.
We focus on the convergence of the sequence toward . Recall that is the fixed point (8). From this condition, we then have that
We then have that
Applying Assumptions 2, and Corollary E.4, together with Lemma E.1, the previous equality implies that
Since satisfy (14), we then obtain 15. We see that 15 is a contraction inequality. Combining this with (16), it implies that the sequence converges to exponentially fast, i.e.,
The rate of convergence is determined by the step-size . This concludes the proof. ∎
Remark C.4.
The exact algorithm ExactAlgo is a simplified version of the algorithm we consider in this paper. The exact algorithm does not involve any approximation, thus is deterministic. This convergence is in line with deterinistic optimization. In fact, to get a precision of , we need to be of order
We now analyze the convergence of the algorithm with approximation, i.e., Algorithm 2. We consider the following algorithm.
Theorem C.5.
Suppose that Assumptions 1 and 2 hold. Assume that, for any ,
(17) |
Then, the exact algorithm Exact MF-TRPO converges to the optimal policy in the tabular setting. In particular, we have that
(18) |
with
Proof.
We focus on the convergence of the sequence , with as in (8). Denote the output of Exact TRPO at each step. We then have that
Applying Assumptions 2 and Corollary E.4 together with Lemma E.1, and following the same lines as in the proof of Proposition C.3, the previous equality implies that
We now proceed in studying the terms , , and . Using Young’s inequality, we get that
and, using Lemma E.1 and Corollary E.4,
Since from Assumption 2 and satisfies (17), a straightforward computation shows that
-MFNE.
With the theoretical foundations established in Theorem C.1 and Theorem C.5, we can now derive a result on the closeness of the proposed algorithm to the MFNE. Specifically, we show that Exact MF-TRPO achieves an -MFNE, where the approximation error is explicitly quantified as follows.
Corollary C.6.
Suppose that Assumptions 1, 2, and 3 hold. Assume that, for any , the learning rate satisfies (17). Let (resp. ) the output of Exact MF-TRPO (resp. Exact TRPO). Then, is -MFNE, with
Proof.
From Proposition E.5, we have that the exploitability of a policy and a mean-field parameter can be bound by the gap of optimality of the w.r.t. the value function and the distance between and the stationary distribution .
From Theorem C.1, we have that
(20) |
On the other hand, using the fact that is a MFNE, we have
Then, applying Lemma E.1, together with Corollary E.4, we obtain
Moreover, from Theorem C.1 on the performances of Exact TRPO, together with Lemma E.1 and Proposition E.2, we have
Using the triangle inequality, together with Theorem C.5, we can bound as
Using the last inequality and (20), together with Proposition E.5, we have that
Therefore, is -MFNE as defined in Definition (2.1). ∎
Remark C.7.
From this corollary, it follows directly that to achieve an -MFNE, the required number of inner policy updates and outer population updates must satisfy the following scaling conditions:
This implies that the sample complexity of the proposed algorithm scales polynomially in with respect to the inner optimization steps and logarithmically with respect to the outer mean-field updates. This confirms the efficiency of our approach, ensuring that even for small values of , convergence to an approximate MFNE remains computationally feasible.
Appendix D Model free algorithms
Model-free approaches play a fundamental role in developing model-agnostic algorithms capable of autonomously adapting to diverse and evolving environments. In the MFG context, various models have been proposed across different domains (Perrin et al., 2021; Yardim et al., 2023), and recent efforts have explored data-driven methodologies to enhance their applicability. Following this line of research, we introduce Sample-Based MF-TRPO, a model-free approach tailored for MFG problems. By leveraging RL techniques with scalable sample-based updates, our method contributes to the growing body of work on data-driven MFG solutions, providing finite-sample complexity guarantees in this setting. This framework further aligns MFG with model-free learning paradigms, broadening their potential for real-world deployment in complex decision-making environments.
D.1 TRPO - sample-based formulation
In the framework established by Shani et al. (2020), it is important to note that the sample-based algorithm does not provide a last-iterate sample complexity guarantee. This contrasts with the exact algorithm, where the policy improvement lemma (Lemma 15, Shani et al., 2020) serves as a foundation for analyzing the convergence properties of the last iterate. In the exact update setting, this guarantee is analogous to Howard’s lemma (Howard, 1960). However, the presence of sampling errors in the sample-based setting hinders the attainment of such guarantees, necessitating a more refined approach when designing and analyzing RL algorithms for MFG.
To address this limitation, it becomes essential to consider alternative strategies than Theorem 5 in Shani et al. (2020). This theorem, however, still provides a framework for analyzing the uniform mixture of the policies generated during the iterative procedure, rather than relying solely on the last iterate. By shifting focus to such policy, we can generalize the theoretical guarantees of the algorithm—a property inherent to the MFG setting.
Additionally, the connection between the value function and the policy space plays a crucial role. Proposition E.2 ensures that the gap in value functions directly bounds the differences between policies. This property provides a pathway to refine the policy improvement process and derive meaningful finite-sample complexity guarantees. By combining these insights, we can propose a robust methodology where the sample-based algorithm achieves convergence with high probability, utilizing uniform mixture policies to overcome the challenges posed by the lack of last-iterate guarantees.
Overall, this refinement introduces a smarter utilization of the sample-based algorithm, emphasizing the role of averaging in mitigating the variability and uncertainty inherent in sample-based methods. This approach not only aligns with the theoretical underpinnings of convex optimization but also strengthens the practical applicability of RL algorithms in MFG, delivering finite-sample complexity results with rigorous probabilistic guarantees.
Remark D.1.
For a fixed , the output of Sample-Based TRPO is the uniform mixture policy . This policy is such that, in the unregularized case, we have (11). It consists on a mixture of , for . The following dedicated subroutine achieve the sampling process in a computationally efficient manner, without performing a direct mixture at every decision step.
This policy is defined is to sample from this policy efficiently without explicitly computing an arithmetic average at the sampling level, particularly in its use within the inner loop of Sample-Based TRPO. Moroever, given that the number of iterations is fixed beforehand, the procedure begins by drawing a random variable uniformly from the set . Once is selected, the sampling step follows the policy . This approach ensures that the selected action is drawn a policy without incurring unnecessary computational overhead during execution.
In particular, due to the regularization term, we have that the following inequality holds:
(21) |
In the absence of regularization, the objective is linear in the occupancy measure, which allows for exact equalities when considering mixtures of policies. However, once the entropic regularization term is introduced, the objective becomes concave in the occupancy measure (see, e.g., Neu et al. (2017) for a proof). As a result, we only obtain the previous inequality rather than (11) when averaging over iterates, as in the relation involving the mixture policy.
Theorem D.2 (Based on Theorem 5 in Shani et al. (2020)).
Suppose Assumption 4 holds. Fix . Let be the sequence generated by Sample-Based TRPO, using
trajectories in each iteration and a rollout up to time with
Then, there exists such that for all , the following holds with probability greater than
(22) |
Proof.
The proof of this result is based on the proof of Shani et al. (Theorem 5, 2020). The main difference is that we are considering the uniform mixture of the policies generated during the iterative procedure.
Applying Shani et al. (Lemma 19, 2020), we get
with
using that is a non-decreasing function. Multiplying both sides by , summing from to , and using the linearity of expectation, we get
with the occupancy measure defined as (6), where the second relation holds by the positivity of the Bregman distance, and the third relation by Shani et al. (Lemma 28, 2020) for uniformly initialized .
Dividing by , we obtain
Plugging in Shani et al. (Lemma 22 and Lemma 23, 2020), we get that for any , if the number of trajectories in the -th iteration satisfies
and the rollout is performed up to time with
then with probability at least ,
where we used Assumption 4 to bound the last term. Thus, combining this with (21), we obtain that
∎
D.2 Initialization step in the sample-based algorithm
The initialization step of the Sample-Based MF-TRPO algorithm presents particular challenges due to the limited operations allowed, specifically the reset and action operations, as described in Section 2. During each iteration of the algorithm, the initial state must be sampled from the distribution .
As detailed in Section 5, the distribution update in the algorithm follows the iterative rule:
where is the output of a single iteration of the Sample-Based MF-TRPO algorithm.
Consequently, at iteration , the distribution is an estimator of
since the update at iteration of Sample-Based MF-TRPO, is an unbiased estimator of the product distribution .
To correctly initialize the environment to a state s such that , the following subroutine is applied:
-
1.
Sampling a Level. Define the categorical random variable that takes value in the discrete space with probabilities given by
i.e.,
-
2.
Selecting a Level. Draw a sample .
-
3.
Rollout Procedure. Starting from an initial state sampled as , execute a rollout of the Markov transition kernels up to level , having to be the particle approximation of
D.3 Sample based algorithm and High Probability Estimates
Transitioning from exact computations to a sample-based setting, we introduce estimators for the key quantities involved in the learning process. These estimators leverage sampled trajectories to approximate the necessary expectations while maintaining computational efficiency.
To ensure the reliability of these approximations, we establish high-probability error bounds by leveraging concentration inequalities. This allows us to rigorously assess the performance of the algorithm, providing quantitative guarantees on the estimation error and its impact on the overall convergence rate. Through this probabilistic framework, we ensure that the sample-based algorithm retains stability and efficiency despite the inherent stochasticity.
Examining the Sample-Based MF-TRPO algorithm, we observe that two key approximations are introduced in the learning process. First, the policy update is performed through Sample-Based TRPO, whose finite-sample analysis in high probability is established in Theorem D.2. This result ensures that the policy iterates remain well-controlled throughout the optimization process in high probability. Secondly, in order to analyze the evolution of the mean-field population distribution, we need to establish a similar high-probability bound on the estimation of the term , which represents the transition dynamics under the estimated policy.
The unbiased estimator of this term uses the trajectories and is given by the empirical sum of . Note that each component of this vector is distributed according to a Bernoulli distribution and is centered in . Therefore, define the following martingale difference term
with defined as in Sample-Based MF-TRPO.
To address this, we first derive Proposition D.3, which is a preliminary concentration result that quantifies the approximation error in the estimation of this key quantity. The first one is Proposition D.3, which provides guarantees on the deviation of the error incurred in a single iteration of the algorithm, with high probability. Specifically, it establishes that the error made while estimating the error at each iteration, which is a bounded increment of a martingale. This result is pivotal, as it ensures that the errors introduced in each iteration of the algorithm are controlled and do not diverge as the algorithm progresses, and lays the foundation for a rigorous convergence analysis of Sample-Based MF-TRPO.
Proposition D.3.
For any and , if the number of trajectories in the -th iteration satisfies:
then, with probability at least , the following holds:
Proof.
From previous consideration, we have that is unbiased and is a martingale difference. Moreover, note that is a bounded vectore, i.e.
Moreover, using Jenses’s inequality, we have that
Therefore, is a bounded martingale difference. To show that the increment is bounded with high probability, we use Hoeffding’s inequality. Let , where . Then, for any , we have
This consideration is a special case of the Generalized Freedman inequality as presented in Harvey et al. (2019). The inequality provides sharp high-probability bounds for the sum of bounded, dependent random variables. In our case, the formulation is simplified due to the presence of a uniform bound on the variables we aim to control.
Therefore, in order to guarantee that
we need the number of trajectories to be at least
∎
D.4 Convergence of Sample-Based MF-TRPO
We now extend the exact analysis of Exact MF-TRPO to its sample-based counterpart, establishing global sample complexity bounds. While the exact algorithm benefits from having full knowledge of the transition kernel and reward function, the sample-based version introduces additional approximation errors due to finite sampling. We quantify these errors and derive high-probability guarantees on the convergence of the algorithm. This requires adapting the theoretical tools developed in the exact setting to account for trajectory-based estimations and ensuring that the resulting policy updates remain stable despite stochastic approximations.
Theorem D.4.
Suppose that Assumptions 1, 2, 3, and 4 hold. Assume that the following holds
(23) |
For any and , if the number of trajectories in each iteration for Sample-Based MF-TRPO satisfies
(24) |
and the number of iteration in each epoch of Sample-Based TRPO satisfies
(25) |
and the rollout is performed up to time with
(26) |
Then, with probability at least , we have that
(27) |
with
with the constant coming from Theorem D.2.
Proof.
We focus on the convergence of the sequence , with as in (8). Denote (resp. ) the output of Sample-Based TRPO at each step (resp. the estimator used in the update of in Sample-Based MF-TRPO). We then have that
Applying Assumptions 2 and Corollary E.4 together with Lemma E.1, and following the same lines as in the proof of Proposition C.3, the previous equality implies that
We now proceed in studying the terms , , , , , , and . Using Young’s inequality, we get that
where we used Lemma E.1 and Corollary E.4 in the last inequality. Using Young’s inequality, we get that
where we used Lemma E.1 and Corollary E.4 in the last inequality. Since from Assumption 2 and satisfies (23), a straightforward computation shows that
Since we have that (25) holds, we can apply Theorem D.2, together with Lemma E.1, to get that
Moreover, since (24) holds, using Proposition D.3, we have that, with probability at least ,
Therefore, combining the previous inequalities, we have that
(28) |
Developping the recursion (28), we obtain
Note that the second term of the r.h.s. of the previous inequality is a telescopic sum, as the central term can be rewritten as
Therefore, we get (18). Moreover, since satisfy (14), this concludes the proof. ∎
D.5 -MFNE
We aim to characterize the proximity of an approximate Nash equilibrium, specifically an -Nash equilibrium. In this context, we address two key questions:
-
1.
Given a fixed budget of sampled trajectories, how close the value function to the unique Nash equilibrium?
-
2.
Given a target approximation level , how many trajectories are required to achieve an -Nash equilibrium?
These questions are crucial for understanding the sample complexity of learning equilibria in mean-field settings and provide insights into the efficiency of our algorithmic approach.
Corollary D.5.
Suppose that Assumptions 1, 2, 3, and 4 hold. Fix . Assume that, for any , the learning rate satisfies (23), and let be the number of trajectories in each iteration for Sample-Based MF-TRPO satisfying (24). Let (resp. ) the output of Sample-Based MF-TRPO (resp. of Sample-Based TRPO). Then, we have the following bound on the exploitability
with
Proof.
From Proposition E.5, we have that the exploitability of a policy and a mean-field parameter can be bound by the gap of optimality of the w.r.t. the value function and the distance between and the stationary distribution .
From Theorem D.2, we have that
(29) |
On the other hand, using the fact that is a MFNE, we have
for . As in proof of Corollary C.6, we have
Moreover,
Using the triangle inequality, together with Theorem D.2, we can bound as
Using the last inequality and (20), together with Proposition E.5, we have that .
∎
Remark D.6.
It is important to note that our analysis does not directly bound the exploitability of the last iterate but on the uniform mixture of policies over the learning process. This distinction arises due to the absence of an exact counterpart to Howard’s theorem (Howard, 1960) in Sample-Based TRPO, as noted in Section D.1. Unlike Exact TRPO, where policy improvement guarantees can be established step by step, sampling errors introduce additional variability that prevents such guarantees in the sample-based setting.
Despite this limitation, our results demonstrate that the learned policies perform well on average and that we approximate the MFNE accordingly. The bounded average exploitability ensures that, over time, the algorithm remains close to an equilibrium, reinforcing the practical effectiveness of the proposed approach in large-scale multi-agent learning.
Remark D.7.
From the obtained sample complexity result, it follows directly that to achieve an -MFNE, the required number of inner policy updates and outer population updates must satisfy the following scaling conditions:
In addition to these requirements, we also establish that the number of episodes and the number of iterations per policy update must satisfy
These additional conditions ensure that the variance introduced by the sampling procedure remains controlled, allowing for a sufficiently accurate estimation of the value function and policy updates. This highlights the tradeoff between computational efficiency and precision in approximating the MFNE, showing that our algorithm achieves a well-balanced complexity while ensuring convergence guarantees.
At each iteration of Sample-Based MF-TRPO, the total number of calls to the environment consists of those required by the Sample-Based TRPO procedure plus the additional subroutine for updating the mean-field parameter. Sample-Based TRPO requires environment calls, scaling proportionally to the product . This aligns with Shani et al. (2020); however, we highlight a distinction stemming from the chosen metrics: the metric they use corresponds to the square root of our exploitability measure, introducing a cubic dependency in terms of . Additionally, the total complexity includes a multiplicative factor , whose contribution is negligible in practice due to its logarithmic scaling, preserving overall algorithmic efficiency.
On the other hand, in each iteration of Sample-Based MF-TRPO, the update step for the mean-field distribution scales as , which means calls to the MF-MDP. This complexity arises naturally from the oracle assumption adopted, which involves an initialization step at each iteration potentially requiring up to steps to accurately initialize the mean-field distribution. While introducing additional complexity, this initialization procedure is crucial for maintaining consistency across iterative population updates, thereby ensuring the stability and convergence accuracy of the algorithm towards the mean-field Nash equilibrium.
Combining these two contributions, we obtain an overall complexity that scales as , consistent with established convergence rates in the RL literature.
Appendix E Technical Lemmata
E.1 Lipschitzness of the Markov reward process iterates
We show in this section that Assumption 1 implies the existence of a Lipschitz constant for the operator and , for any and .
Lemma E.1.
Proof.
This proof is adapted from Fort et al. (Lemma 4.2, 2011) on parametrized Markov chains.
Step 1. Consider first . By employing a telescoping sum, we obtain
Consider a function with . Therefore, since is a difference of probabilities, we have that
(31) |
This is due to the fact that whenever we evaluate the previous difference of probabilities matrices on , the term is perceived a constant by the transition kernels and , summing this part to zero.
Define as
From Assumption 3, we have that
(32) |
Therefore, we have
and, applying Assumption 1, we get
for any . Combining this with (31), we get, from the characterization of the total variation norm of the integral with respect to the positive functions bounded in sup-norm by , that
where we have used the fact that a stochastic matrix, thus its biggest eigenvalue is and is a vector of just positive components.
E.2 From bound on Value function to bounds on Policy
In this section, we demonstrate how a bound on the value function naturally leads to a corresponding bound on the policies. In the seminal work by Shani et al. (2020), an bound was established for the cost functions. This result can be extended to derive a bound on the distance between policies by leveraging the properties of regularization. The connection between the value function and policies highlights the role of regularization in maintaining both theoretical guarantees and practical performance stability.
Indeed, from the entropic regularization, the optimization problem (3) with respect to the profile admits a unique solution . These considerations form the foundation of the following proposition.
Proposition E.2.
We have that
(33) |
for any .
Proof.
Denote the entropic regularization tem in the reward function (2) as a function of the occupation measure, i.e.,
with as in 5. Therefore, we can express as
(34) |
Taking the disentegration on the spatial component, we have the following relationship between the occupation measure and its marginal
(35) |
with as in 6. This also implies that
(36) |
From (34), as the optimal value does not depend on the initial condition, we have that the optimal policy of the previous optimization problem satisfies
for any . Combining this with (35), we obtain that the previous condition equivalent to
for any , . We recall that the Bregman divergence with respect to the regularization is defined as follows
Therefore,
(37) |
However, we have that with the Bregman divergence corresponding to the entropy regularization has the following expression (see, e.g., Neu et al., 2017)
Moreover, from the definition of , extracting the first term of the series, we obtain
Using the decomposition of as
we get
Therefore, for a function , we have
since is a positive function and is a positive measure. Applying this to the positive function , together with Pinsker’s inequality (see, e.g., Cover, 1999), we have
Proposition E.3.
Proof.
Step 1. Let us state the optimal Bellman equations
We notice that a function is -Lipschitz in -norm since the -norm of the gradient of this function always lies on a probability simplex (see, e.g., Geist et al., 2019). Thus, we have
Then, we study the Lipschitzness of optimal -values for arbitrary action
By Assumption 1, we have
thus
Overall, we have a recursive bound on difference between value functions
therefore
By a bound , we conclude the statement (38).
Step 2. Applying directly the Bellman equation, we have
Following the same lines as in the Step 1, we can then obtain (39).
∎
Corollary E.4.
Proof.
E.3 Bound on the Exploitability
To analyze the exploitability of a given policy and a given mean-field parameter , we decompose it into two key contributions. The first term captures the suboptimality of the best response against the mean-field distribution, quantifying how much an agent can improve its reward by deviating optimally. The second term accounts for the discrepancy between the current population distribution and the stationary distribution of the Markov reward process induced by . This decomposition allows us to explicitly bound the exploitability by controlling both the policy’s optimality and the convergence of the population dynamics to equilibrium.
Proposition E.5.
Fix a policy and two mean-field parameter . Then, we have that the exploitability as defined in (7) is bounded by
with
(41) |
Proof.
Fix a policy and two mean-field parameter . Then, we have that
On the one hand, from Proposition E.3, we have that
On the other hand, we have that
and a similar decomposition applies for . This means that
where we have applied Cauchy-Schwarz inequality in the last bound. Therefore, we can bound the exploitability as defined in (7) as
∎
E.4 Discussion on the monotonicity of the optimal Markov Kernel
Define the operator as . In this section, we outline sufficient conditions under which this operator, responsible for updating the population distribution in the MFG framework, exhibits monotonicity. Monotonicity of is a crucial property that ensures stability and convergence of the iterative updates toward the Nash equilibrium.
This operator represents a generalization of the standard contractivity condition, which is traditionally formulated with . This generalization is motivated by the fact that, as we aim at studying the regularized ergodic MFG problem (2)-(3), the condition can hold for some .
This contractivity condition reflects the combined effect of the Lipschitz continuity of the regularized best response and the ergodicity of the Markov reward process , ensuring the stability and convergence of the mean-field population updates in the ergodic setting.
Lemma E.6 (Strong monotonicity of ).
Proof.
Consider the following decomposition
The condition is satisfied when the Lipschitz constant associated with the transition kernel is sufficiently small, and the exponent is large enough.
Intuitively, a smaller indicates that the transition dynamics of the MDP are less sensitive to changes in the population distribution, reducing the potential for instability. A regularity condition on is a standard assumption in the literature to ensure the uniqueness of the MFNE. Similar assumptions have been employed in various works, including Becherer & Hesse (2024), Espinosa & Touzi (2015), Lacker & Zariphopoulou (2019), and Tangpi & Zhou (2024), among many others. These studies leverage regularity constraints to prevent degeneracies in equilibrium selection and to guarantee well-posedness in the associated fixed-point problems.
Meanwhile, a larger amplifies the effect of the contraction over multiple iterations of the operator, ensuring convergence even in cases where individual updates are not strongly contractive. This interplay between and highlights the importance of balancing the model’s inherent dynamics with the structural assumptions to guarantee monotonicity and stability in the population updates.
Appendix F Additional Experiments
We present results for the Exact MF-TRPO algorithm on two extensions of the Crowd Modeling game and we benchmark our results against Ficticious Play (FP) (Perrin et al., 2020) and Online Mirror Descent (OMD) (Pérolat et al., 2022). Our findings demonstrate that the exact algorithm matches the performance of state-of-the-art methods, highlighting its effectiveness in these settings. In the following, we provide a detailed overview of the games employed.
Grid-based Crowd Modeling Game. This environment, inspired by the Four Rooms example from Geist et al. (2022), is based on a two-dimensional grid with obstacles. Each agent’s state is defined by her position on the grid, and she can choose from five possible actions: moving left, right, up, down, or staying in place. The reward function is designed to discourage overcrowding by penalizing agents based on the population density at their next position. Specifically, agents receive a negative reward proportional to the logarithm of the density at their destination, encouraging a more even distribution across the state space. Additionally, a small bonus is given for staying in place, while moving in any direction results in a penalty. Formally, the reward function is defined as
where , with being the indicator function and being a crowd-aversion parameter.
In this environment, the transition matrix does not depend on the mean-field distribution ; however, some stochasticity is introduced through a slipperiness parameter: when an agent selects an action, she is most likely to follow it, but there remains a smaller probability of performing a different valid move. In particular, for each action, a total slipperiness probability of is evenly distributed among the alternative actions. Furthermore, this game can be extended by introducing a designated point of interest, denoted as , which guides the behavior of the players. The modified reward function is defined as
where denotes the previously defined reward function, and is the distance between state and the target state , computed as the norm of their coordinate difference.



Two-Islands-Graph Crowd Modeling. The Two Islands Crowd Modeling Game replaces the grid structure with two interconnected graphs, referred to as islands, connected by a single narrow bridge. The main challenge in this setting arises from the limited connectivity between the two sub-populations. The transition matrix is generated randomly, assigning to each node a probability distribution over its neighboring nodes, including itself. The reward function penalizes the logarithm of the mean-field distribution while encouraging movement toward the second island ,
The Exact MF-TRPO algorithm is evaluated on the two proposed variants of the Grid-Based environment and on the Two-Islands-Graph Crowd Modeling game. The former is modeled as an grid with walls delineating four symmetric and interconnected rooms, as in Geist et al. (2022), with all the players starting clustered in the top-left corner. For the latter, we consider a state space of size and an action space of size , with a branching factor of 2, that is, each state is connected to exactly two neighbors. Here, initially, all players are positioned at location 2 on the first island (see Figure 5).
F.1 Experimental setting
Results are presented for two different values of the regularization parameter: and and, throughout all experiments, the discount factor is set to . A key feature of both the exact and sample-based methods is the use of a warm start for the policy in the RL component. Rather than resetting the policy to a uniform distribution over actions at each iteration, it is initialized with the policy learned from the previous iteration. Moreover, the step size used for updating the distribution remains constant throughout the learning phase, i.e., . The key parameters for the two algorithms are summarized in Table 1.
Algorithm/Parameter | ||||||||
---|---|---|---|---|---|---|---|---|
Exact MF-TRPO | 0.9 | 10 | 0.01 | N/A | N/A | N/A | ||
Sample-Based MF-TRPO | 0.2 | 0.9 | 100 | 0.1 |
F.2 Results
The plots presented show the exploitability, defined in Equation 7, to evaluate the effectiveness of our approach, along with the evolution of the mean-field distribution over time. Compared to FP and OMD, Exact MF-TRPO performs competitively across all evaluated environments, demonstrating superior long-term performance. As training progresses, the model continually improves its policy and ultimately outperforms the other algorithms (see Figure 3). Moreover, players in grid-based games tend to move toward less crowded areas, gradually achieving a more uniform distribution (see Figure 4). Moreover, when a point of interest is introduced, the players manage to cluster around it (see Figure 6). Finally, as shown in Figure 5, the players progressively concentrate on the second island, attracted by the higher reward present in that region.








