Learning NP-Hard Multi-Agent Assignment Planning Using GNN: Inference On A Random Graph and Provable Auction-Fitted Q-Learning

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Learning NP-Hard Multi-Agent Assignment Planning

using GNN: Inference on a Random Graph and


Provable Auction-Fitted Q-learning

Hyunwook Kang∗ Taehwan Kwon


Department of Computer Science Kakao Brain
Texas A&M University [email protected]
[email protected]

Jinkyoo Park† James R. Morrison†


Industrial & Systems Engineering Electrical Engineering
KAIST Central Michigan University
[email protected] [email protected]

Abstract

This paper explores the possibility of near-optimally solving multi-agent, multi-task


NP-hard planning problems with time-dependent rewards using a learning-based
algorithm. In particular, we consider a class of robot/machine scheduling problems
called the multi-robot reward collection problem (MRRC). Such MRRC problems
well model ride-sharing, pickup-and-delivery, and a variety of related problems.
In representing the MRRC problem as a sequential decision-making problem, we
observe that each state can be represented as an extension of probabilistic graphical
models (PGMs), which we refer to as random PGMs. We then develop a mean-field
inference method for random PGMs. We then propose (1) an order-transferable
Q-function estimator and (2) an order-transferability-enabled auction to select a
joint assignment in polynomial-time. These result in a reinforcement learning
framework with at least 1 − 1/e optimality. Experimental results on solving MRRC
problems highlight the near-optimality and transferability of the proposed methods.
We also consider identical parallel machine scheduling problems (IPMS) and
minimax multiple traveling salesman problems (minimax-mTSP).

1 Introduction

Motivation Consider a set of identical robots seeking to serve a set of spatially distributed tasks.
Each task is given an initial age (which then increases linearly in time). Greater rewards are given to
younger tasks when service is complete according to a predetermined reward rule. Such problems
prevail in operations research, e.g., dispatching drivers to transport customers or scheduling machines
in a factory. Solving such highly structured NP-hard problems with the constraint of ‘no possibility
of two robots assigned to one task at once’ using mathematical optimization schemes is infeasible
or ineffective due to the expensive computational cost, especially when the problem size is large.
Applying decentralized approach using multi-agent modeling frameworkLong et al. (2020); Rashid
et al. (2018); Sunehag et al. (2017) is a possible way to solve a large scale problems. However, due
to the impossibility of inducing consensus among agents in achieving the global objective without

First author

correspondence to: Jinkyoo Park ([email protected]), James R. Morrison ([email protected])

36th Conference on Neural Information Processing Systems (NeurIPS 2022).


effective communication Fischer et al. (1985), such decentralized approaches are rarely used in
industries (e.g., factories). Thus, this study focuses on centralized methods for solving MRRC.
Research Questions. Many of such NP-hard scheduling problems have time-dependent rewards.
To the best of our knowledge, these problems have not yet been addressed by non-decentralized
learning-based methods. Even if one can, it must also be able to simultaneously address a fundamental
challenge: the number of possible robot-task pairs to be considered increases exponentially. For
instance, scheduling 8 robots and 50 tasks involves 1013 possible joint assignments at each time-step.
The main research question that the current study seeks to resolve is "how to design a computationally
effective (i.e., scalability in terms of learning and decision making) a learning based centralized
decision making scheme for solving a large-scale NP-hard scheduling problems?"
Proposed method and contributions. The present paper explores the possibility of near-optimally
solving multi-robot, multi-task NP-hard scheduling problems with time-dependent rewards using a
learning-based algorithm. The study formulates multi-robot, multi-task NP-hard scheduling problems
in a sequential decision-making framework and derives a joint scheduling policy with a theoretical
performance bound under reasonable assumptions. The novelties of the current study are as follows:

• The study first observes that a state-joint assignment pair can be represented as a random PGM.
After developing a theory of random PGM-based mean-field inference, we derive random struc-
ture2vec, a random PGM-based extension of structure2vec Dai et al. (2016).
• We estimate the Q-function Q(sk , ak ) using layers of random structure2vec, where (sk , ak ) is
the state-joint action pair. Using an interpretation of a layer of structure2vec as a Weisfeiler-
Lehman kernel (as in Dai et al. (2016)), we design the estimator to possess a property we call
order-transferability. This property enables transferability in problem size.
• We propose a joint assignment rule called order-transferability-enabled auction policy (OTAP)
to address exponential growth in joint assignment space. We propose auction-fitted Q-iteration
(AFQI) by substitution of the argmax operation of fitted Q-iteration with OTAP to train the
Q-function in a scalable manner. We prove that AFQI results in a policy with polynomial-time
computation that achieves at least 1 − 1/e performance compared with the optimal policy.

Results and Impacts. Using simulation experiments, we show that the proposed policy typically
achieves 97% optimality for the multi-robot reward collection (MRRC) problem in a deterministic
environment with linearly time-varying rewards. This performance is well extended to experiments
with stochastic traveling times. To the best of our knowledge, this result is the first to learn a
near-optimal NP-hard multi-robot/machine scheduling policy with time-dependent rewards.

2 Related studies.

Reinforcement Learning based Vehicle Routing Problems. Mazyavkina et al. (2020) have catego-
rized the RL approaches solving vehicle routing problems into two: (1) the improvement heuristics
that learn an operator can iteratively improve the entire routing plans until there is no improvement is
made (Wu et al., 2020; da Costa et al., 2020; Chen & Tian, 2019; Lu et al., 2020; Kim et al., 2021), (2)
the construction heuristics that learn a policy that sequentially make a single routing action given the
partial solution (state) (Bello et al., 2016; Nazari et al., 2018; Kool et al., 2018; Khalil et al., 2017),
and (3) the hybrid approaches that mix these two approaches (Joshi et al., 2020; Fu et al., 2021; Kool
et al., 2021; Ahn et al., 2020). These RL approaches have mainly considered a single-agent routing
problem. Although these methods solve CVRP with multiple vehicles, they solve this problem from
a single-agent perspective. In addition, most of these approaches consider the static reward function
setting. On the contrary, our study explicitly considers multi-vehicle interaction while considering
the time-varying reward, which is a more realistic routing problem setting.
Graph Inference Based Approach. Dai et al. (2017) showed that a graph neural network (GNN)
called structure2vec Dai et al. (2016) can construct a solution for the Traveling Salesman Problem
(TSP). structure2vec is a popular GNN derived from mean-field inference with a probabilistic
graphical model (PGM). Dai et al. (2017) formulates the TSP as a Markov decision process (MDP)
where a heuristically constructed PGM represents each state-next assignment pair. They employ
structure2vec derived from the heuristic PGM to infer the Q-function, which they use to select the
next assignment. While their choice of PGM was heuristic, their approach achieved near-optimality

2
Figure 1: Representation of the scheduling problem using a random PGM

and transferability of their trained single-robot scheduling algorithm to new single-robot scheduling
problems with an unseen number of tasks.

3 Multi-Robot Reward Collection Problem


In the main text, we model a general multi robot/machine scheduling problem as a discrete-time,
discrete-state (DTDS) sequential decision-making problem. In the DTDS model, time advances
in fixed increments ∆, i.e., tk = t0 + ∆ × k where tk is the actual time after k decision epochs
have passed. For simplicity, we use k as a time index representing the kth decision epoch. In
this framework, sk denotes a state, and action ak denotes a joint assignment of robots/machines to
unfinished tasks at the kth epoch. The objective of the problem is to learn the optimum scheduling
policy πθ : sk → ak that maximizes the reward collected or minimizes the total completion time
(a.k.a. makespan minimization). Below is the formulation of MRRC. We additionally propose a
continuous-time continuous-state (CTCS) problem to identical parallel machine scheduling problems
(IPMS) in Appendix A and minimax multiple traveling salesman problems (minimax-mTSP) in
Appendix B.

3.1 State

The state sk at epoch k is represented as (gk , Dk ) where a graph gk = ((R, Tk ), (EkT T , EkRT )) and
associated feature set Dk = (DkR , DkT , DkT T , DkRT ). The elements of graph gk are defined as (See
Figure 1):
• R = {1, ..., M } is the index set of all robots. The index i and j will be used to specifically denote
robots.
• Tk = {1, ..., N } is the index set of all remaining unserved tasks at decision epoch k. The index p
and q will be used to specifically denote tasks.
• EkT T = {ϵTpqT |p ∈ Tk , q ∈ Tk } is the set of all directed edges from a task in Tk to any task in
Tk . We consider each edge as a random variable. The task-to-task edge ϵTpqT = 1 indicates the
event that a robot that has just completed task p subsequently completes task q. We denote the
probability p(ϵTpqT = 1) ∈ [0, 1] the presence probability of the edge ϵTpqT .
• EkRT = {ϵRTip |i ∈ R, p ∈ Tk } is the set of all directed edges from a robot in R to a task in Tk .
We say the robot-to-task edge ϵRT ip = 1 indicates the event that robot i is assigned to the task
p. This edge is defined deterministically depending on the joint assignment action. If robot i is
assigned to task p, then p(ϵRT
ip ) = 1, otherwise 0.

The element of feature set Dk associated with the graph gk is defined as:
• DkR = {dR i |i ∈ R} is the set of node features for the robot nodes in R at epoch k. In MRRC, di
R

is defined as the location of robot i at epoch k (epoch index k is omitted).


• DkT = {dTp |p ∈ Tk } is the set of node features for the task nodes in Tk at epoch k. In MRRC, dTp
is defined as the age of task p at epoch k (epoch index k is omitted).

3
• DkT T = {dTpqT |p ∈ Tk , q ∈ Tk } is the set of task-task edge features at epoch k. dTpqT denotes the
duration for a robot that has just completed task p to subsequently compete task q. We call this
duration task completion time. In MRRC, a task completion time is given as a random variable
(in practice, our method only requires a set of samples of random variable).
• DkRT = {dRT RT
ip |i ∈ R, p ∈ Tk } is the set of robot-task edge features at epoch k. dip denotes the
traveling time for robot i to reach task p.

3.2 Action

An action ak , a joint assignment at epoch k, is defined as a maximal bipartite matching of the


complete bipartite graph (R, Tk , EkRT ) composed of the robot nodes R, the remaining task nodes Tk ,
and the fully connected edges between them EkRT . That is, given the current state sk = (gk , Dk ), ak
is a subset of EkRT satisfying (1) no two robots can be assigned to the same tasks, and (ii) a robot may
only remain without assignment when the number of robots exceeds the number of remaining tasks.
If ϵRT
ip ∈ ak , it means that robot i is assigned with task p at epoch k. For example, Figure 1 shows
the case where ak = (ϵRT RT RT RT
1,1 , ϵ2,3 ). (Note, we equivalently may say this as ϵ1,1 = ϵ2,3 = 1 and 0
otherwise.) In MRRC, all robots are allowed to change their assignments at each decision epoch. (In
IPMS and mTSP, only free machines/salesmen are newly assigned.)

3.3 State transition

As the joint assignment ak is executed given the current state sk = (gk , Dk ) , the next state
sk+1 = (gk+1 , Dk+1 ) is determined with the updated graph gk+1 and features Dk+1 .
Graph update. When the decision epoch corresponds to the point when task p is completed, the
corresponding task node will be removed in the updated task nodes as Tk+1 = Tk /{p}, and the
TT RT
task-task edges and robot-task edges, Ek+1 and Ek+1 , will be accordingly updated.
R T TT RT
Feature update. At decision epoch k + 1, Dk+1 = (Dk+1 , Dk+1 , Dk+1 , Dk+1 ) is determined. In
R R T T
MRRC, task locations Dk+1 = {di |i ∈ R} and task ages Dk+1 = {dp |p ∈ Tk+1 } are updated.
RT R
The robot-task edge features Dk+1 will be updated according to Dk+1 as well.

3.4 Reward and objective

At time 0, each task is given an initial age which increases linearly in time. A reward rk = r dTp is


given when a task p ∈ Tk , whose age is dTp , is served at epoch k. We consider linear and nonlinear
reward functions r for MRRC. The objective is to learn a stationary policy π, a function that maps
current
Pstate s into current action a, to maximize expected total collected rewards Qπ (s, a) =:

EP,π [ k=0 R(stk , atk , stk+1 )|st0 = s, at0 = a].

4 Random graph embedding: RandStructure2Vec


We observe that when task completion time is not deterministic, a scheduling problem can be
represented as an extension of probabilistic graphical models (PGMs), which we refer to as random
PGMs. This section proposes a mean-field inference method, random struct2vec, for random PGMs
to estimate the state-action value Q(sk , ak ) in solving MRRC.

4.1 Random PGM for representing a state of MRRC

Q variables X = {Xp }, suppose that we can factor the joint distribution p (X ) as


Given random
p (X )= Z1 i ϕi (Di ) where ϕi (Di ) denotes a marginal distribution or conditional distribution as-
sociated with a set of random variables Di ; Z is a normalizing constant. Then {Xp } is called a
probabilistic graphical model (PGM). In a PGM, Di is called a clique and ϕi (Di ) is called the clique
potential for Di , and Di is called the scope of ϕi . We often write simply ϕi , suppressing Di .
Starting from a state sk and an action ak , one can conduct a random experiment of “sequential
decision making using policy π”. In this random experiment, we can denote the events ‘How robots
serve all remaining tasks in which sequence’ as scenarios. For example, suppose that at time-step k we

4
are given robots {r1 , r2 }, tasks {t1 , t2 , t3 , t4 , t5 , t6 } and we follow the policy π onward. One possible
scenario is that robot r1 serves tasks {t1 → t4 } and robot r2 serves tasks {t3 → t2 → t5 → t6 }
(see Figure 1). Note that the time when t5 is served depends on the time when t2 is served (state
transition is Markovian); thus the reward from t5 depends on the reward from t2 . As shown in Figure
1, {{t1 → t4 }, {t3 → t2 → t5 → t6 }} can be represented as a single instance of a Bayesian Network.
Since scenario realization is random, we can construct the distribution over such scenarios using
’random’ Bayesian Network with random node Xk = (sk , ak ) and clique potential ϕ. For details, see
section 5.1.

4.2 Mean-field inference with random PGM

Let X = {Xp } be the set of all random variables in the inference problem. Let GX be the set of all
possible PGMs on X . Let P : GX 7→ [0, 1] be a probability measure on GX . Define a random PGM
on X as {GX , P}. Note that the inference of {GX , P} will be difficult; |GX | is too large for inferring
P even using Monte-Carlo sampling approach. To avoid this difficulty, we use the approximated
inference using semi-cliques. Suppose that we are given the set of all possible cliques on X as CX .
As a PGM will be realized according to P, only a few of the possible cliques in CX will be actually
realized as an element of the PGM and become real cliques. We call such potential clique elements
if we were given P, we could calculate the presence probability pm
of CX as semi-cliques. Note that P
of the semi-clique Dm as pm = G∈GX P(G)1Dm ∈G , where 1 denotes the indicator function.

Mean-field inference with random PGM. We start from a specific inference problem and state the
main theorem in more general way. Consider a random PGM on X = ({Hi }, {Xj }) where Hk is
the latent variable corresponding to the observed variable Xk . Our goal is to infer {Hi } given {Xj }
by finding p({Hi }|{xj }). In mean-field inference, we instead find a set of surrogate distributions
{q {xj } (Hi )} for which {Hi } are independent. Here q {xj } means that q is a function of {xj }).
We next state Theorem 1. in a very general manner. The statement is the same as that of mean-field
inference with PGM Koller & Friedman (2009) except that ours has the presence probability terms
{pm } of semi-cliques {Dm }. The implication is that inference of presence probability of each
semi-clique is enough to conduct mean-field inference, and the inference of {GX , P} is not needed.

Theorem 1. Random PGM based mean field inference. Suppose a random PGM on X = {Xp }
is given, and the presence probability {pm } for all semi-cliques Dm ∈ CX are known.
nPsurrogate distribution {qp (xp )} in mean-fieldoinference is optimal only if qp (xp ) =
Then, the
1
Zp exp m:Xp ∈Dm pm E(Dm −{Xp })∼q [ln ϕm (Dm , xp )] where Zp is a normalizer and ϕm clique
potential for Dm .
For the general background and proof, see Appendix C.

RandStructure2Vec. In Dai et al. (2016), structure2vec was derived as a vector space embedding of
mean-field inference with PGM. For detailed background on vector space embedding, see Appendix
D. From Theorem 1, we derive random structure2vec as a vector-space embedding of mean-field
inference with random PGM. Suppose Q that once aQPGM is realized the PGM has joint distribution
proportional to some factorization p ϕ (Hp |Ip ) p,q ϕ (Hp |Hq ) (as in Dai et al. (2016)). Under
this assumption, we can write {q {xj } (Hi )} as {q xi (Hi )}. In Dai et al. (2016),
P they suggest that
structure2vec is essentially a fixed point iteration µ̃p ← σ(W1 xp + W2 q̸=p µ̃q ) where µ̃p is a
latent vector for node p and xp is input for nodeR p. They show that, when we interpret µ̃i as a vector
space injective embedding expressed as µ̃i = H ϕ(hi )q xi (hi )dhi for some ϕ, structure2vec’s fixed
point iteration is the embedding of fixed point iteration of mean-field inference with the PGM. Lemma
1 states that we have a similar result for random PGM by including only the {pqp } information.

Lemma 1. Structure2vec for random PGM. Assume that the presence probabilities {pqp } for all
pairwise semi-cliques Dqp ∈ CX are given.Then embedding the fixedpoint equation in Theorem 1
P
generates the fixed point equation µ̃p ← σ W1 xp + W2 q̸=p pqp µ̃q . We refer to this fixed point
iteration as random structure2vec. (The proof of Lemma 1 can be found in Appendix E.)

5
Remarks. Note that inference of {GX , P} is in general a difficult task. One implication of Theorem
1 is that we transformed a difficult inference task into a simple inference task: inferring the presence
probability of each semi-clique. (See Appendix F for the algorithm that conducts this task.) In
addition, Lemma 1 provides a theoretical justification to ignore the inter-dependencies among edge
presences when embedding a random graph using GNN. When graph edges are not explicitly given or
known to be random, the simplest heuristic one can use is to separately infer the presence probabilities
of all edges and adjust the weights of GNN’s message propagation. According to Lemma 1, possible
inter-dependencies among edges would not affect the quality of such heuristic inference.

5 Solving MRRC with RandStructure2Vec

Figure 2: State representation and main inference procedure

This section describes how the proposed method solves the MRRC problem with the newly developed
random structure2vec method. The total solution (i.e., the sequential assignments of robots/machines
to tasks) is found by iteratively repeating a sequential decision making. This section specifically
describes how to choose a joint assignment ak given the current state sk = (gk , Dk ). The procedure
for determining a joint assignment action is composed of (1) represent the state using a random
Bayesian Network, (2) estimate the Q-value using random graph embedding, and (3) select a joint
assignment. Figure 2 depicts the overall procedure. This section focuses on MRRC; the appendices
will provide the formulation, solution procedure, and results for IPMS and mTSP problems as well.

5.1 Representing a state using a random PGM

In this section, we describe how the MRRC problem state sk and one possible joint assignment
ak can be represented as a Bayesian network. Assume Hp , the hidden random variable for task
p, carries information about the benefit of serving task p. Given a scenario (see section 3.1), Hp
depends on Hq only if q is served after p by the same robot, and Hp depends on Xp , the observed
features for task p. For an MRRC problem, Xp can be either a task’s assignment information dRT ip or
T
a task’s age dp . BasedQon these definitions,
Q a Bayesian Network for {Hp } and {Xp } is constructed
as p({Hp }|{Xp }) = p ϕ (Hp |Xp ) p,q ϕ (Hp |Hq ). This Bayesian Network corresponds to one
scenario; there are many possible scenarios that can be realized randomly. Since random PGM natu-
rally models this characteristic, according to Lemma 1, we are justified to use random structure2vec
with edge (semi-clique) presence probabilities {p(ϵTpqT )} from in section 2.1.

5.2 Estimating state-action value using Order trainability-enabled Q-function

We illustrate how we can estimate a Q-function for MRRC by applying random structure2vec
tothe random PGM that represents
 (sk , ak ). Lemma 1 provides the fixed point equation µ̃p =
P
σ W1 xp + W2 j̸=k pqp µ̃q to compute the embeddings µ̃p for node task p in a random PGM.
The embeddings µ̃p for p ∈ Tk are computed in an iterative manner using random structure2vec as:

6
 
(τ +1) P (τ )
µ̃p = σ W1 xp + W2 q̸=p pqp µ̃q . We propose a network architecture composed of two-step
sequential random structure2vec, which is composed of action embedding and value embedding.
Action Embedding. The first random structure2vec layer embeds the joint assignment into the task
nodes Tk = {1, ..., n}. The action embedding µ̃A p for task node p is defined as the fixed point for the
 
A A A A A RT
where xA
P
equation µ̃p = σ W1 xp + W2 q̸=p pqp µ̃q p = dip (distance from robot i to task p)
when task p is assigned to robot i, xAp = 0 when task p is not assigned. The action node embeddings
{µ̃A
p |p ∈ T k }, computed via iteration simultaneously by employing random structure2vec, provide
sufficient information about the relative locations between robots and their assigned tasks.
Value Embedding. The second random structure2vec layer embeds the task ages into the task
nodes. The value embedding µ̃Vp for task node p is defined as the fixed point for the equation
 
T
µ̃Vp = σ W1V xVp + W2V q̸=p pqp µ̃Vq where xVp = (µ̃A
P
p , dp ) is the concatenation of the action
T
embedding µ̃A
p computed by the first layer and the task age dp for node p. The resulting value node
V
embeddings {µ̃p |p ∈ Tk }, computed in an iterative manner, provide sufficient information about
how much value is likely in the local graph around each task by the specified joint assignment.
P Qθ (sk , ak ). To derive Qθ (sk , ak ), we aggregate the embedding vectors for all nodes by
Computing
µ̃V = p µ̃Vp to obtain one global vector µ̃V to embed the value affinity of the global graph. We
then use a neural network to map µ̃V into Qθ (sk , ak ). The overall pseudo-code for estimating steps
is provided in Appendix G.
It is essential for the proposed inference method can estimate the state action value Qθ (sk , ak )
for varying graph size gk . Let us provide the intuition related to scuh problem-size transferability.
For Action Embedding, transferability is trivial; the inference problem is a scale-free task locally
around each node. For Value Embedding, consider the ratio of robots to tasks. The overall value
affinity embedding will be underestimated if this ratio in the training environment is smaller than
this ratio in the testing environment; overestimated overall otherwise. The intuition is that this
over/under-estimation does not matter in Q-function based policies as discussed in van Hasselt et al.
(2015) as long as the order of Q-function value among actions are the same. That is, as long as
the best assignments chosen are the same, i.e., arg maxak Q(sk , ak ) = arg maxak Qθ (sk , ak ), the
magnitude of imprecision |Q(sk , ak ) − Qθ (sk , ak )| does not matter. We call this property of an
estimator order-transferability (with respect to the max operation).

5.3 Selecting a joint assignment using OTAP

With the previously introduced way to estimate Qθ (sk , ak ), we illustrate how to compute the joint
assignment (action) a∗k , a maximal bipartite matching in the bipartite graph (R, Tk , EkRT ), given
the state sk = (gk , Dk ). Specifically, we propose the order transferability-enabled auction policy
(OTAP) that constructs a joint assignment ak through N = max (|R|, |k|) iterations of Bidding and
Consensus phases. This auction follows the spirit of Sequential Single Item (SSI) auctioning (Koenig
et al. (2006)); each iteration adds one robot-task assignment to construct a full joint assignment.
(n−1)
Bidding-phase. In the nth iteration of the bidding phase, given Mθ , the ordered set of n − 1
robot-task edges in EkRT determined by the previous n − 1 iterations, all the unassigned robots bid
their the most preferable task to conduct. Respecting robot-task assignments determined in previous
n − 1 iterations and ignoring other unassigned robots, each unassigned robot i select the best task
(n−1)
assignment ϵRTil that maximizes Qnθ (sk , Mθ ∪ {ϵRT
ip }) among all unassigned tasks p ∈ Tk ,
n
where Qθ is the θ-parameterized network with superscript n indicating that the state action value is
(n−1)
estimated at nth iteration. Then, robot i bids {ϵRT n
iℓ , Qθ (sk , Mθ ∪ {ϵRTiℓ })} to the auctioneer.
This bidding occurs simultaneously by all the unassigned robots at the nth iteration. Since the number
of ignored robots varies at each iteration, transferability of Q-function inference is crucial.
Consensus-phase. In the consensus phase of nth iteration, the centralized auctioneer finds the bid
(n−1)
with the best bid value, say {ϵRT n
i∗ p∗ , Qθ (sk , Mθ ∪ {ϵRT ∗ ∗
i∗ p∗ })} (Here i and p denote the best robot
(n) (n) (n) (n−1) (n)
task pair.) Denote ϵRT
i∗ p∗ := mθ . The auctioneer sets everyone’s Mθ as Mθ = Mθ ∪ mθ
and initiate bidding phase for the remaining unassigned robots.

7
(N ) (1) (N ) (N )
These two phases iterate until reaching Mθ = {mθ , . . . , mθ }. This Mθ is chosen as the
joint assignment a∗k of N -robots at time step k. That is, πθ (sk ) = a∗k . The computational complexity
for computing πθ is O (|R| |T|) and is only polynomial; see Appendix L.

5.4 Training Q-function using AFQI

The fitted Q-iteration (FQI) finds θ that minimizes E(sk ,ak ,rk ,sk+1 )∼D [Qθ (sk , ak ) − [r (sk , ak ) +
γ maxa Qθ (sk+1 , a)]] where D denotes the distribution of training data. We propose a new rein-
forcement learning method, which we call Auction-fitted Q-iteration (AFQI), which replaces maxa
Qθ (sk , a) used in the conventional FQI with OTAP. That is, writing OTAP as πQθ , AFQI finds θ that
empirically minimizes E(sk ,ak ,rk ,sk+1 )∼D [Qθ (sk , ak ) − [r (sk , ak ) + γQθ (sk+1 , πQθ (sk+1 ))]] .
In learning the parameters θ for Qθ (sk , ak ), we use the exploration strategy that perturbs the
parameters θ randomly to actively explore the joint assignment space with OTAP. While this method
was originally developed for policy-gradient based methods Plappert et al. (2017), exploration in
parameter space is useful in our auction-fitted Q-iteration since it generates a reasonable combination
of assignments.

6 Theoretical analysis

We show the proposed AFQI obtains at least 1 − 1/e optimality and enables computation of the joint
assignment in polynomial time. This result is achieved by the order-transferability of the proposed
Q-function estimator and its use in selecting the joint assignment.

6.1 Performance bound of OTAP


(N )
Recall that Qn denotes the n-robot problem’s true Q-function. In the same way as we defined Mθ
above, denote the joint assignment chosen by OTAP as {Qn }N n=1 as M
(N )
= {m(1) , . . . , m(N ) }.
(N )
Lemma 2. If the Q-function approximator has order transferability, then M(N ) = Mθ .
For any decision epoch k, let M denote a set of robot-task pairs (a subset of EkRT ). For any robot-task
pair m ∈ EkRT , define ∆(m | M) := Q|M∪{m}| (sk , M ∪ {m}) − Q|M| (sk , M) as the the marginal
value (under the true Q-functions) of adding robot-task pair m ∈ EkRT . Lemma 2 enables us to use
the result discussed in Nemhauser et al. (1978) and achieve the result of Theorem 2.
Theorem 2. Suppose that the Q-function approximation with the parameter value θ exhibits order
(N ) N
transferability. Denote Mθ as the result of OTAP using {Qnθ }n=1 and let M∗ = argmaxak
Q|ak | (sk , ak ). If ∆(m | M) ≥ 0, ∀M ⊂ EkRT , ∀m ∈ EkRT , and the marginal value of adding one
robot diminishes as the number of robots increases, i.e., ∆(m | M) ≤ ∆(m | N ), ∀N ⊂ M ⊂ EkRT ,
∀m ∈ EkRT , then the result of OTAP is at least better than 1 − 1/e of an optimal assignment. That is,
(N ) |M∗ |
QNθ (sk , Mθ )≥Q (sk , M∗ )(1 − 1/e) . See Appendix H and K for the proofs.

6.2 Performance bound of AFQI

AFQI seeks to find θ that minimizes E(sk ,ak ,rk ,sk+1 )∼D [Qθ (sk , ak ) − [r (sk , ak ) +
γQθ (sk+1 , πQθ (sk+1 ))]]. Here, we use OTAP, denoted as πQθ , instead of maxa Qθ (sk , a) which is
used in general fitted-Q iteration. As we have seen in section 5.1, OTAP replaces the max operation
with the auction algorithm with a provable performance bound compared with the max operation.
Lemma 3 allows us to use this performance bound to obtain a performance assurance on AFQI
compared with FQI. We only write an abbreviated version of the statement for brevity. The formal
description of conditions and the proof is provided in Appendix I.
Lemma 3. Kang & Kumar (2021) Suppose that a 1 − 1/r approximation algorithm is substituted for
the max operation in FQI. Then, the corresponding new Fitted Q-iteration’s performance is at least
1 − 1/r optimal.
Corollary 1. AFQI achieves at least 1 − 1/e performance compared with the optimal policy.

8
Table 1: Performance test (50 trials of training for each case)
Testing size : Robot (R) / Task (T)
Reward Environment Baseline
2R/20T 3R/20T 3R/30T 5R/30T 5R/40T 8R/40T 8R/50T
98.31 97.50 97.80 95.35 96.99 96.11 96.85
Optimal
(±4.23) (±4.71) (±5.14) (±5.28) (±5.42) (±4.56) (±3.40)
99.86 97.50 118.33 110.42 105.14 104.63 120.16
Deterministic Ekisi et al.
(±3.24) (±2.65) (±2.84) (±2.97) (±3.78) (±2.50) (±3.94)
Linear
137.3 120.6 129.7 110.4 123.0 119.9 119.8
SGA
(±5.65) (±5.03) (±5.54) (±4.34) (±4.97) (±4.74) (±5.84)
130.9 115.7 122.8 115.6 122.3 113.3 115.9
Stochastic SGA
(±4.02) (±4.03) (±5.21) (±6.23) (±4.94) (±5.53) (±4.08)
111.5 118.1 118.0 110.9 118.7 111.2 112.6
Deterministic SGA
(±3.71) (±5.56) (±5.09) (±4.64) (±5.23) (±5.38) (±5.07)
Nonlinear
110.8 117.4 119.7 111.9 120.0 110.4 112.4
Stochastic SGA
(±5.17) (±6.22) (±4.48) (±4.70) (±6.38) (±5.14) (±5.30)

Table 2: Transferability test (linear & deterministic env, standard dev. provided in the appendix)

Training size Testing size : Robot (R) / Task (T)


#R/ #T 2R/20T 3R/20T 3R/30T 5R/30T 5R/40T 8R/40T 8R/50T

2R/20T 98.31 (±4.23) 93.61(±4.98) 97.31 (±4.25) 92.16 (±3.49) 92.83(±4.25) 90.94(±3.98) 93.44 (±4.02)
3R/20T 95.98(±4.75) 97.50(±3.71) 96.11(±3.63) 93.64(±4.54) 91.75(±5.71) 91.60(±5.03) 92.77(±4.74)
3R/30T 94.16(±4.97) 96.17(±4.22) 97.80(±5.14) 94.79(±3.53) 93.19(±3.78) 93.14(±4.50) 93.28(±3.99)
5R/30T 97.83(±3.11) 94.89(±4.43) 96.43(±4.23) 95.35±5.28) 93.28(±4.18) 92.63(±5.07) 92.40(±4.10)
5R/40T 97.39(±4.65) 94.69(±4.01) 95.22(±4.88) 93.15(±5.09) 96.99±4.42) 94.96(±3.94) 93.65 (±5.66)
8R/40T 95.44(±4.32) 94.43(±4.88) 93.48(±4.37) 93.93(±5.05) 96.41(±3.96) 96.11±4.56) 95.24(±4.44)
8R/50T 95.69(±3.18) 96.68(±2.81) 97.35(±4.20) 94.02(±2.69) 94.50(±4.44) 94.86(±3.26) 96.85±3.40)

Table 3: Training complexity (mean of 20 trials of training, linear & deterministic env.)

Testing size : Robot (R) / Task (T)


Linear & Deterministic
2R/20T 3R/20T 3R/30T 5R/30T 5R/40T 8R/40T 8R/50T
Performance with full training 98.31 97.50 97.80 95.35 96.99 96.11 96.85
Training time for 93% optimality 19261.2 61034.0 99032.7 48675.3 48217.5 45360.0 47244.2

6.3 Experiment settings

In the main text, we focus on discrete-time & discrete-state (DTDS) MRRC problems with deter-
ministic and stochastic task completion times. For CTCS deterministic problems with real-world
datasets, see IPMS (Appendix A) and mTSP (Appendix B).
Environment. Since there is no standard dataset for MRRC problems, we used the complex maze-like
environment generator of Neller et al. (2010) (code provided in Appendix 10). This complex maze
mimics the complex road layout of a city and random traffic, inducing nontrivial task completion
times. See the leftmost image of Figure 2 and the supplementary video. We randomly generated a
new maze for every training and testing experiment with randomly chosen initial task/robot locations.
To generate the task completion times, Dijkstra’s algorithm and dynamic programming were used for
deterministic and stochastic environments, respectively.
In the stochastic environment, a robot makes its intended move with a certain probability. (Cells
with a dot: success with 55%, every other direction with 15% each. Cells without a dot: 70% and
10%, respectively.) A task is considered served when a robot reaches it. We consider two reward
rules: linearly decaying rewards f (age) = max{200 − age, 0} and nonlinearly decaying rewards
f (age) = λage with λ = 0.99, where age is the task age when served. The initial age of tasks are
uniformly distributed in the interval [0, 100].
Baselines. For deterministic environments with linear rewards, where the corresponding MRRC can
be formulated as a mixed-integer linear program (MILP), we consider the following two baselines:
• Optimal: Gurobi Gurobi Optimization (2019), an off-the-self the optimization solver for MILP,
was used to solve the problems with 60-min time limit.
• Ekici et al.: Ekici & Retharekar (2013), the most up-to-date heuristic for solving MRRC in the
Operations Research community, was used the problems.

9
For stochastic environments or exponential rewards, to our knowledge, there is no literature addressing
MRRC with. Thus, we construct an indirect baseline:
• Sequential Greedy Algorithm (SGA): a general-purpose multi-robot task allocation algorithm
called SGA Han-Lim Choi et al. (2009).

The performance measure we used is ρ = Rewards collected by the proposed method


Reward collected by the baseline . Thus, the value of ρ
greater than 100% indicates the proposed method collects more reward than the corresponding
baseline algorithm. Note that ρ against Optimal is always lower than 100%.
Note that we cannot provide other reinforcement learning-based heuristics as additional baselines
since, to the best of our knowledge (for the class of NP-hard multi-robot/machine scheduling problems
with decaying rewards), this paper is the first to propose a reinforcement learning-based heuristic.

6.4 Performance test.

Performance was tested under four environments: deterministic/linear rewards, determinis-


tic/nonlinear rewards, stochastic/linear rewards, stochastic/nonlinear rewards. See Table 1. Our
method achieves near-optimality for linear/deterministic rewards with 3% fewer rewards than optimal
on average. The standard deviation for ρ is provided in parentheses. For other environments, we
see that the %SGA ratio for linear/deterministic is well maintained. Due to dataset generation’s
dynamic programming computation complexity, we only consider 8 robots/50 tasks at maximum. We
considered larger size problems in IPMS experiments discussed in Appendix A.

6.5 Transferability test.

Table 2 provides comprehensive transferability test results. The rows indicate training conditions,
while the columns indicate testing conditions. The results in the diagonal cells in red (cells with
the same training size and testing size) serve as baselines (direct testing). The results in the off-
diagonal show the results for the transferability testing and demonstrate how the algorithms trained
with different problem sizes perform well on test problems (zero-shot transfer). We can see that
lower-direction transfer tests (trained with larger size problems and tested with smaller size problems)
show only a small loss in performance. For upper-direction transfer tests (trained with smaller size
problems and tested with larger size problems), the loss was up to 4 percent.

6.6 Scalability analysis.

For training complexity, we measured the training time required to achieve 93% optimality considering
a deterministic environment with linear rewards. Table 4 shows that training time may not necessarily
increase as problem size gets larger, while the performance is fairly maintained.
MRRC can be formulated as a semi-MDP (SMDP) based multi-robot planning problem (e.g., Omid-
shafiei et al. (2017)). This problem’s complexity with R robots and T tasks and maximum H time
horizon is O((R!/T !(R − T )!)H ). In our proposed method, this complexity is addressed by a combi-
nation of two complexities: computational complexity and training complexity. For computational
complexity of joint assignment decision at each timestep is O(|R||T |3 ). See Appendix L for details.

7 Concluding Remarks
In this paper, we addressed the challenge of developing a near-optimal learning-based method
for solving NP-hard multi-robot/machine scheduling problems. We developed a theory of mean-
field inference for scheduling problems and a corresponding theoretically justified GNN method to
precisely infer the Q-function. We addressed the scalability issue of Fitted Q-Iteration methods for
multi-robot/machine scheduling problems by providing a polynomial-time algorithm with a provable
performance guarantee. Simulation results demonstrate the effectiveness of the our methods.

Acknowledgement

Jinkyoo Park was supported by Institute of Information & communications Technology Planning
Evaluation (IITP) grant funded by the Korea government(MSIT)(2022-0-01032, Development of
Collective Collaboration Intelligence Framework for Internet of Autonomous Things).

10
References
Agarwal, A., Jiang, N., and Kakade, S. M. Reinforcement learning: Theory and algorithms. Technical
report, 2019.
Ahn, S., Seo, Y., and Shin, J. Learning what to defer for maximum independent sets. In III, H. D. and
Singh, A. (eds.), Proceedings of the 37th International Conference on Machine Learning, volume
119 of Proceedings of Machine Learning Research, pp. 134–144. PMLR, 13–18 Jul 2020.
Bello, I., Pham, H., Le, Q. V., Norouzi, M., and Bengio, S. Neural combinatorial optimization with
reinforcement learning. arXiv preprint arXiv:1611.09940, 2016.
Chen, X. and Tian, Y. Learning to perform local rewriting for combinatorial optimization. In Wallach,
H., Larochelle, H., Beygelzimer, A., d'Alché-Buc, F., Fox, E., and Garnett, R. (eds.), Advances in
Neural Information Processing Systems, volume 32. Curran Associates, Inc., 2019.
da Costa, P. R. d. O., Rhuggenaath, J., Zhang, Y., and Akcay, A. Learning 2-opt heuristics for the
traveling salesman problem via deep reinforcement learning. In Pan, S. J. and Sugiyama, M. (eds.),
Proceedings of The 12th Asian Conference on Machine Learning, volume 129 of Proceedings of
Machine Learning Research, pp. 465–480, Bangkok, Thailand, 18–20 Nov 2020. PMLR.
Dai, H., Dai, B., and Song, L. Discriminative Embeddings of Latent Variable Models for Structured
Data. 48:1–23, 2016. doi: 1603.05629.
Dai, H., Khalil, E. B., Zhang, Y., Dilkina, B., and Song, L. Learning Combinatorial Optimization
Algorithms over Graphs. (Nips), 2017.
Ekici, A. and Retharekar, A. Multiple agents maximum collection problem with time dependent
rewards. Computers and Industrial Engineering, 64(4):1009–1018, 2013. ISSN 03608352. doi:
10.1016/j.cie.2013.01.010. URL http://dx.doi.org/10.1016/j.cie.2013.01.010.
Fischer, M. J., Lynch, N. A., and Paterson, M. S. Impossibility of distributed consensus with one
faulty process. Journal of the ACM (JACM), 32(2):374–382, 1985.
Fu, Z.-H., Qiu, K.-B., and Zha, H. Generalize a small pre-trained model to arbitrarily large tsp
instances, 2021.
Google. Google OR-Tools, 2012. URL https://developers.google.com/optimization/.
Gurobi Optimization, L. Gurobi optimizer reference manual, 2019. URL http://www.gurobi.com.
Han-Lim Choi, Brunet, L., and How, J. Consensus-Based Decentralized Auctions for Robust Task
Allocation. IEEE Transactions on Robotics, 25(4):912–926, aug 2009. ISSN 1552-3098. doi:
10.1109/TRO.2009.2022423.
Joshi, C. K., Cappart, Q., Rousseau, L.-M., Laurent, T., and Bresson, X. Learning tsp requires
rethinking generalization, 2020.
Kang, H. and Kumar, P. R. 1+r-approximate policy and fitted q-iteration
for problems with large action space. Unpublished working paper
("http://people.tamu.edu/~hwkang/KangKumar2021A.pdf"), 2021.
Khalil, E., Dai, H., Zhang, Y., Dilkina, B., and Song, L. Learning combinatorial optimization
algorithms over graphs. In Advances in Neural Information Processing Systems, pp. 6348–6358,
2017.
Kim, M., Park, J., et al. Learning collaborative policies to solve np-hard routing problems. Advances
in Neural Information Processing Systems, 34, 2021.
Koenig, S., Tovey, C., Lagoudakis, M., Markakis, V., Kempe, D., Keskinocak, P., Kleywegt, A.,
Meyerson, A., and Jain, S. The power of sequential single-item auctions for agent coordination. In
AAAI, volume 2006, pp. 1625–1629, 2006.
Koller, D. and Friedman, N. Probabilistic graphical models : principles and techniques, page 449-453.
The MIT Press, 1st edition, 2009. ISBN 9780262013192.

11
Kool, W., Van Hoof, H., and Welling, M. Attention, learn to solve routing problems! arXiv preprint
arXiv:1803.08475, 2018.
Kool, W., van Hoof, H., Gromicho, J., and Welling, M. Deep policy dynamic programming for
vehicle routing problems, 2021.
Kurz, M. E., Askin, R. G., Kurzy, M. E., and Askiny, R. G. Heuristic scheduling of parallel machines
with sequence-dependent set-up times. International Journal of Production Research, 39(16):
3747–3769, 2001. ISSN 0020-7543. doi: 10.1080/00207540110064938.
Long, Q., Zhou, Z., Gupta, A., Fang, F., Wu, Y., and Wang, X. Evolutionary population curriculum
for scaling multi-agent reinforcement learning, 2020.
Lu, H., Zhang, X., and Yang, S. A learning-based iterative method for solving vehicle routing
problems. In International Conference on Learning Representations, 2020.
Mazyavkina, N., Sviridov, S., Ivanov, S., and Burnaev, E. Reinforcement learning for combinatorial
optimization: A survey. arXiv preprint arXiv:2003.03600, 2020.
minmaxTSPlib. minmax multiple-tsp library. https://profs.info.uaic.ro/∼mtsplib/MinMaxMTSP/,
2021.
Nazari, M., Oroojlooy, A., Snyder, L., and Takác, M. Reinforcement learning for solving the vehicle
routing problem. In Advances in Neural Information Processing Systems, pp. 9839–9849, 2018.
Neller, T., DeNero, J., Klein, D., Koenig, S., Yeoh, W., Zheng, X., Daniel, K., Nash, A., Dodds, Z.,
Carenini, G., Poole, D., and Brooks, C. Model AI Assignments. Proceedings of the Twenty-Fourth
AAAI Conference on Artificial Intelligence (AAAI-10), pp. 1919–1921, 2010.
Nemhauser, G. L., Wolsey, L. A., and Fisher, M. L. An analysis of approximations for maximizing
submodular set functions—i. Mathematical programming, 14(1):265–294, 1978.
Omidshafiei, S., Agha–Mohammadi, A., Amato, C., Liu, S., How, J. P., and Vian, J. Decentralized
control of multi-robot partially observable Markov decision processes using belief space macro-
actions. The International Journal of Robotics Research, 36(2):231–258, 2017. doi: 10.1177/
0278364917692864.
Parisotto, E., Song, F., Rae, J., Pascanu, R., Gulcehre, C., Jayakumar, S., Jaderberg, M., Kauf-
man, R. L., Clark, A., Noury, S., et al. Stabilizing transformers for reinforcement learning. In
International conference on machine learning, pp. 7487–7498. PMLR, 2020.
Plappert, M., Houthooft, R., Dhariwal, P., Sidor, S., Chen, R. Y., Chen, X., Asfour, T., Abbeel, P., and
Andrychowicz, M. Parameter Space Noise for Exploration. pp. 1–18, 2017.
Rashid, T., Samvelyan, M., de Witt, C. S., Farquhar, G., Foerster, J., and Whiteson, S. Qmix:
Monotonic value function factorisation for deep multi-agent reinforcement learning, 2018.
Smola, A., Gretton, A., Song, L., and Schölkopf, B. A hilbert space embedding for distributions. In
International Conference on Algorithmic Learning Theory, pp. 13–31. Springer, 2007.
Sriperumbudur, B. K., Gretton, A., Fukumizu, K., Lanckriet, G., and Schölkopf, B. Injective hilbert
space embeddings of probability measures. In 21st Annual Conference on Learning Theory (COLT
2008), pp. 111–122. Omnipress, 2008.
Sunehag, P., Lever, G., Gruslys, A., Czarnecki, W. M., Zambaldi, V., Jaderberg, M., Lanctot, M.,
Sonnerat, N., Leibo, J. Z., Tuyls, K., and Graepel, T. Value-decomposition networks for cooperative
multi-agent learning, 2017.
van Hasselt, H., Guez, A., and Silver, D. Deep Reinforcement Learning with Double Q-learning.
2015.
Wu, Y., Song, W., Cao, Z., Zhang, J., and Lim, A. Learning improvement heuristics for solving
routing problems, 2020.

12

You might also like