Externally Valid Selection of Experimental Sites via the k-Median Problem111We would like to thank Isaiah Andrews, Tim Armstrong, Michèle Belot, Michael Gechter, Jesse Goodman, Kei Hirano, Yiqi Liu, Francesca Molinari, Charles Manski, Guillaume Pouliot, David Shmoys, and Davide Viviano for helpful feedback, comments, and suggestions. We would also like to thank Eddie Ramirez Saquic for excellent research assistance. We gratefully acknowledge financial support from the NSF under grant SES-2315600.

   Brenda Prallon22footnotemark: 2    Chen Qiu22footnotemark: 2    Yiwei Sun22footnotemark: 2
(August 2024)
Abstract

We present a decision-theoretic justification for viewing the question of how to best choose where to experiment in order to optimize external validity as a k𝑘kitalic_k-median (clustering) problem, a popular problem in computer science and operations research. We present conditions under which minimizing the worst-case, welfare-based regret among all nonrandom schemes that select k𝑘kitalic_k sites to experiment is approximately equal—and sometimes exactly equal—to finding the k𝑘kitalic_k most central vectors of baseline site-level covariates. The k𝑘kitalic_k-median problem can be formulated as a linear integer program. Two empirical applications illustrate the theoretical and computational benefits of the suggested procedure.

1 Introduction

A common concern in randomized evaluations of new policies is their external validity; that is, whether the estimated effects of a policy intervention carry over to new samples or populations (Vivalt, 2020; Gechter, 2024; Duflo et al., 2007, Chapter 8). There is a large body of literature arguing that the external validity of randomized evaluations can be improved by explicitly incorporating this goal into the experimental design by, for example, carefully deciding where to experiment; see Degtiar and Rose (2023) for an overview and references, and also the recent work of Chassang and Kapon (2022). For instance, if a researcher has access to multiple sites to experimentally evaluate a new policy, it is possible to use the information available before the evaluation (such as site-level characteristics) to nonrandomly select one site—or more—that are “representative” of the populations of interest. Recently, both Egami and Lee (2024) and Gechter et al. (2024) have shown that a research design that nonrandomly selects experimental sites—referred to broadly in the literature as purposive sampling (Cook et al., 2002, p. 511)—is useful to improve the external validity of randomized evaluations.

This paper presents a theoretical justification for viewing the question of how to best design a purposive sampling scheme—or, equivalently, how to best nonrandomly select k𝑘kitalic_k sites to experiment—with the goal of optimizing external validity as a k𝑘kitalic_k-median problem. This is a classical problem in computer science and operations research (Williamson and Shmoys, 2011; Cohen-Addad et al., 2022). Broadly speaking, in a k𝑘kitalic_k-median problem, there is a set of facilities and a set of clients; the goal is to open at most k𝑘kitalic_k facilities and connect each client to at least one facility at minimal total connection cost. We next provide more specific details on how these problems relate.

Statistical Decision Theory: Under conditions that will be explained clearly in Sections 2 and 3, Theorem 1 in this paper shows that the worst-case (welfare-based) regret of any purposive sampling scheme that selects at most k𝑘kitalic_k sites to experiment is approximately equal—and sometimes exactly equal—to the objective function of a k𝑘kitalic_k-median problem with the following features. First, the sites available for experimentation are treated as facilities. Second, the sites where the policy maker would like to implement the new policy are treated as clients. Third, the connection cost between clients and facilities is proportional to the Euclidean distance between the available site-level covariates. Importantly, we show that the solution of the k𝑘kitalic_k-median problem is exactly “minimax-regret” optimal (among all purposive sampling schemes) when i) the candidate sites for experimentation and the policy-relevant sites are disjoint (Remark 4) and ii) the treatment effect heterogeneity across sites accommodated in the parameter space is substantial (in a sense Theorem 1 makes precise). When ii) holds but i) does not, we show that the optimized value of the k𝑘kitalic_k-median problem only approximates the minimax-regret value, and that the approximation error improves as either the number of policy-relevant sites increases or the experiments conducted become more precise (Remark 5). The link with the k𝑘kitalic_k-median problem established in this paper thus shows that selecting the k𝑘kitalic_k sites that have the most central vector of covariates tends to optimize external validity (in a minimax-regret sense).

In order to formalize this connection, we leverage recent developments in the literature of treatment choice problems with partial identification; for example Yata (2021); Ishihara and Kitagawa (2021); Montiel Olea et al. (2023). Note first that in the k𝑘kitalic_k-median problem, each client is typically connected to only one facility (otherwise, the connection cost would not be minimized). In the context of the site selection problem, a connection between a site i𝑖iitalic_i (where an experiment was conducted) and a site j𝑗jitalic_j (where no experimentation occurred) means that the estimated effects obtained in site i𝑖iitalic_i are used to inform policy decisions in site j𝑗jitalic_j. This means that even when experimental outcomes in k>1𝑘1k>1italic_k > 1 sites are available, the solution to the k𝑘kitalic_k-median problem would typically prescribe the policy maker to only use the information from the site with the smallest connection cost: the nearest neighbor. But when is it decision-theoretically optimal for a policy maker to behave in this way? After k𝑘kitalic_k sites have been selected for experimentation, the policy maker faces the “evidence aggregation” problem introduced in Ishihara and Kitagawa (2021) and recently discussed in Yata (2021), Christensen et al. (2022) and Montiel Olea et al. (2023). That is, the policy maker must decide how to use the estimated treatment effects in different—and potentially heterogeneous—sites to make a treatment choice (implement the new policy or preserve the status quo) in the site of interest. It is known that there are conditions under which it is minimax-regret optimal for the policy maker to base decisions on the nearest neighbor (the site with the most similar site-level characteristics), provided the true treatment effects are allowed to vary substantially as a function of site-level covariates; see, for example, Proposition 1 in Montiel Olea et al. (2023). Moreover, in this case, the optimized worst-case regret is proportional to the distance between the baseline covariates of the site of interest and those of its nearest neighbor. Thus, these results on treatment choice problems with partial identification are fundamental to provide a decision-theoretic justification for the use of the k𝑘kitalic_k-median (clustering) problem to optimize external validity.

Algorithms: The connection with the k𝑘kitalic_k-median problem clarifies the problem’s difficulty but also suggests efficient algorithms. To see the need for those, recall that any purposive sampling scheme optimizes over “n𝑛nitalic_n choose k𝑘kitalic_k” potential site combinations, where n𝑛nitalic_n is the total number of sites available for experimentation. Optimization over purposive sampling schemes also requires the evaluation of some measure of performance that depends on the dimension of the site-level covariates, d𝑑ditalic_d. Thus, optimally choosing a purposive sampling scheme by simply evaluating the performance of each combination is costly when “n𝑛nitalic_n choose k𝑘kitalic_k” or d𝑑ditalic_d is large.

Conceptually, the connection with the k𝑘kitalic_k-median problem allows us to understand the computational complexity of finding a minimax-regret optimal purposive sampling scheme under the conditions of Theorem 1. Since the k𝑘kitalic_k-median problem is known to be NP-hard (Kariv and Hakimi, 1979; Megiddo and Supowit, 1984; Cohen-Addad et al., 2018), there is no algorithm whose computational time scales polynomially in the problem’s inputs (d,k,n)𝑑𝑘𝑛(d,k,n)( italic_d , italic_k , italic_n ).

However, from a practical perspective, it is also known that the k𝑘kitalic_k-median problem admits a linear integer program formulation (Williamson and Shmoys, 2011, Chapter 7.7, p. 185), and is routinely solved with off-the-shelf algorithms such as different versions of the branch-and-bound method (Bertsimas and Weismantel, 2005, Chapter 11.1). In addition, different branch-and-bound algorithms either find a solution with provable optimality or, if stopped early, generate a report on the suboptimality of the solution found (Bertsimas et al., 2016). As we explain later, in our application with n=41𝑛41n=41italic_n = 41, d=13𝑑13d=13italic_d = 13, any problem for k{1,,10}𝑘110k\in\{1,\ldots,10\}italic_k ∈ { 1 , … , 10 } can be solved to provable optimality in just a few seconds using a personal laptop (see Figure 5 in Section 5).

Related Literature: Our results build on two recent papers that presented novel purposive sampling strategies to select experimental sites so as to optimize external validity. Gechter et al. (2024) present an elegant decision-theoretic approach that frames external validity as a policy problem and—under the assumption that the policy maker has a priori information about the effects of the new policy across sites—recommend a Bayesian approach for choosing where to experiment. Egami and Lee (2024) use the principle behind synthetic control (Abadie et al., 2010) to recommend what they call the synthetic purposive sampling of sites; specifically, they select good “donors,” defined as sites whose weighted average of observed baseline characteristics are close to the characteristics of the sites of interest. It is important to note that our decision-theoretic framework leans heavily on Gechter et al. (2024), but also that the sites selected through the k𝑘kitalic_k-median problem can be interpreted as a degenerate synthetic purposive sampling strategy, whereby each unit’s associated synthetic unit is just its nearest neighbor.

This paper also contributes to the literature arguing that a “modern, decision-theoretic framework can help clarify important practical questions of experimental design” (Banerjee et al., 2017). Although decision-theoretic approaches to external validity are recent, a large body of work used statistical decision theory to analyze other aspects of experimental design, such as sample size determination (Raiffa and Schlaifer, 1961; Manski and Tetenov, 2016, 2019; Azevedo et al., 2020, 2023; Hu et al., 2024). Finally, our notion of external validity is conceptually related to several areas of research in econometrics, machine learning, and statistics such as domain adaptation (Mansour et al., 2009; Ben-David et al., 2010), distributional shifts (Duchi and Namkoong, 2021; Sugiyama et al., 2007; Adjaho and Christensen, 2022), learning under biased sampling (Sahoo et al., 2022), and cross-domain transfer estimation and performance (Andrews et al., 2022; Menzel, 2023). To the best of our knowledge, none of these papers contain decision-theoretic analyses of where to experiment.

Outline: This paper is organized as follows. Section 2 introduces the formal framework. Section 3 presents the main result of this paper (Theorem 1) and explains how the k𝑘kitalic_k-median problem arises naturally when trying to incorporate external validity concerns into the decision problem of where to experiment. Section 4 presents the linear integer program formulation of the k𝑘kitalic_k-median problem. Section 5 presents two illustrative empirical applications. Section 6 considers extensions of the baseline model. Section 7 concludes. Proofs of the main results can be found in Appendix A. Additional results are collected in the Supplementary Appendix.

2 Setting up the Decision Problem

2.1 Notation

A policy maker considers a set of S𝑆S\in\mathbb{N}italic_S ∈ blackboard_N candidate sites to evaluate and, eventually, implement a new policy of interest. The sites are indexed by s𝒮{1,,S}𝑠𝒮1𝑆s\in\mathcal{S}\equiv\{1,\ldots,S\}italic_s ∈ caligraphic_S ≡ { 1 , … , italic_S }. In order to accommodate situations in which the policy maker is not necessarily able to experiment in all the candidate sites, we assume there is a nonempty subset 𝒮E𝒮subscript𝒮𝐸𝒮\mathcal{S}_{E}\subseteq\mathcal{S}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ⊆ caligraphic_S of what we term experimental sites.222As discussed in Allcott (2015), there are often systematic reasons determining the eligibility of certain sites for experimentation. For example, in microfinance RCT studies, experiments often require large sample sizes and well-managed microfinance institutions (MFIs), characteristics more commonly found in older and larger institutions. To qualify for clinical trials involving a new surgical procedure, hospitals and surgeons need to have both experience in the procedure and a history of low mortality rates. Throughout the paper, and to avoid a trivial instance of the site selection problem, we assume that there are at least two experimental sites (i.e., card(𝒮E)2cardsubscript𝒮𝐸2\textrm{card}(\mathcal{S}_{E})\geq 2card ( caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ) ≥ 2). It is also possible that institutional restrictions preclude the eventual implementation of the policy of interest in all of the candidate sites. Thus, it will be convenient to denote by 𝒮P𝒮subscript𝒮𝑃𝒮\mathcal{S}_{P}\subseteq\mathcal{S}caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ⊆ caligraphic_S the nonempty set of policy or policy-relevant sites.

In principle—and in order to accommodate different scenarios that could arise in empirical work—we allow for the case in which there is overlap between experimental and policy sites (i.e., 𝒮E𝒮Psubscript𝒮𝐸subscript𝒮𝑃\mathcal{S}_{E}\cap\mathcal{S}_{P}\neq\emptysetcaligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ≠ ∅). We note, however, that the case in which these two sets are disjoint (i.e., 𝒮E𝒮P=subscript𝒮𝐸subscript𝒮𝑃\mathcal{S}_{E}\cap\mathcal{S}_{P}=\emptysetcaligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = ∅) will be of particular interest. This is because when there is no overlap between experimental and policy sites, the extrapolation problem faced by the policy maker becomes evident: a policy decision needs to be made in sites where there has never been a previous experimental evaluation of the policy of interest.

For each site, the policy maker observes a vector of site characteristics Xsdsubscript𝑋𝑠superscript𝑑X_{s}\in\mathbb{R}^{d}italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT that may affect the treatment effect. Thus, we allow for treatment effect heterogeneity across sites, but we restrict this heterogeneity by assuming that it depends on observable characteristics. Specifically, let the function τ:d:𝜏superscript𝑑\tau:\mathbb{R}^{d}\rightarrow\mathbb{R}italic_τ : blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT → blackboard_R define conditional (on X𝑋Xitalic_X) average treatment effects. We posit that any pair of sites with “similar” observed characteristics also have “similar” treatment effects, formally by assuming that τ𝜏\tauitalic_τ is a Lipschitz function (with respect to the Euclidean norm) with known constant C𝐶Citalic_C. We formally state this assumption in Section 2.4. It is not innocuous, but we will argue that it can be replaced by other continuity-like conditions (such as Hölder continuity) while maintaining the main message of the paper.333In Section 6.3, we discuss how our results extend to other functional classes that could be used to restrict treatment effect heterogeneity, such as collections of functions that belong to a convex and centrosymmetric space. Such restrictions were recently used in the econometrics literature to analyze estimation, inference, and other decision problems that arise in a nonparametric regression setup (Yata, 2021; Armstrong and Kolesár, 2018). We let LipC(d)subscriptLip𝐶superscript𝑑\textrm{Lip}_{C}(\mathbb{R}^{d})Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) denote the space of all Lipschitz functions from dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT to \mathbb{R}blackboard_R with constant C𝐶Citalic_C.

2.2 Statistical Model for the Site Selection Problem

As in Gechter et al. (2024), the policy maker must choose a strict subset of experimental sites 𝒮𝒮E𝒮subscript𝒮𝐸\mathscr{S}\subset\mathcal{S}_{E}script_S ⊂ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT.444We require 𝒮𝒮\mathscr{S}script_S to be a strict subset of 𝒮Esubscript𝒮𝐸\mathcal{S}_{E}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT because if we allow the policy maker to experiment in all sites, and there is no cost of experimentation that varies at the site level, then there is no site selection problem. We consider the case in which 𝒮𝒮\mathscr{S}script_S is allowed to equal 𝒮Esubscript𝒮𝐸\mathcal{S}_{E}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT in Section 6.1. In this paper, we focus on the case in which there is a restriction on the total number of experimental sites that the policy maker can select. That is, there is an integer k𝑘k\in\mathbb{N}italic_k ∈ blackboard_N, k<card(𝒮E)𝑘cardsubscript𝒮𝐸k<\textrm{card}(\mathcal{S}_{E})italic_k < card ( caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ), such that 𝒮𝒮\mathscr{S}script_S must belong to the set

𝒜(k):={𝒮𝒮E|card(𝒮)k}.assign𝒜𝑘conditional-set𝒮subscript𝒮𝐸card𝒮𝑘\mathcal{A}(k):=\{\mathscr{S}\subset\mathcal{S}_{E}\>|\>\textrm{card}(\mathscr% {S})\leq k\}.caligraphic_A ( italic_k ) := { script_S ⊂ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT | card ( script_S ) ≤ italic_k } . (1)

What we have in mind with this restriction is that the policy maker may not have enough administrative resources to run more than k𝑘kitalic_k experiments to evaluate the policy of interest. Our notation also allows for the possibility that the policy maker does not want to experiment at all.

If the policy maker decides to experiment in a nonempty set 𝒮𝒜(k)𝒮𝒜𝑘\mathscr{S}\in\mathcal{A}(k)script_S ∈ caligraphic_A ( italic_k ) of cardinality card(𝒮)kcard𝒮𝑘\textrm{card}(\mathscr{S})\leq kcard ( script_S ) ≤ italic_k, then she will observe card(𝒮)card𝒮\textrm{card}(\mathscr{S})card ( script_S ) treatment effect estimates. We collect these estimates in a vector of dimension card(𝒮)card𝒮\textrm{card}(\mathscr{S})card ( script_S ). In a slight abuse of notation, let 𝒮1<𝒮2<<𝒮card(𝒮)subscript𝒮1subscript𝒮2subscript𝒮card𝒮\mathscr{S}_{1}<\mathscr{S}_{2}<\ldots<\mathscr{S}_{\textrm{card}(\mathscr{S})}script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < script_S start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < … < script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT denote the indices of the card(𝒮)card𝒮\textrm{card}(\mathscr{S})card ( script_S ) experimental sites in 𝒮𝒮\mathscr{S}script_S. Letting τ^ssubscript^𝜏𝑠\widehat{\mathcal{\tau}}_{s}over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT denote the estimated treatment effect in site s𝑠sitalic_s, we can define the vector

τ^𝒮:=(τ^𝒮1,,τ^𝒮card(𝒮)).assignsubscript^𝜏𝒮superscriptsubscript^𝜏subscript𝒮1subscript^𝜏subscript𝒮card𝒮top\widehat{\tau}_{\mathscr{S}}:=(\hat{\tau}_{\mathscr{S}_{1}},\ldots,\hat{\tau}_% {\mathscr{S}_{\textrm{card}(\mathscr{S})}})^{\top}.over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT := ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , … , over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT . (2)

Analogously, we can denote the vector of true treatment effects for the experimental sites in 𝒮𝒮\mathscr{S}script_S as

τ𝒮:=(τ(X𝒮1),,τ(X𝒮card(𝒮))).assignsubscript𝜏𝒮superscript𝜏subscript𝑋subscript𝒮1𝜏subscript𝑋subscript𝒮card𝒮top\tau_{\mathscr{S}}:=(\tau(X_{\mathscr{S}_{1}}),\ldots,\tau(X_{\mathscr{S}_{% \textrm{card}(\mathscr{S})}}))^{\top}.italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT := ( italic_τ ( italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) , … , italic_τ ( italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT . (3)

We assume that the treatment effect estimators obtained in each site are normally (and independently) distributed around the vector of true effects:

τ^𝒮𝒩card(𝒮)(τ𝒮,Σ𝒮), where Σ𝒮:=diag(σ𝒮12,,σ𝒮card(𝒮)2).formulae-sequencesimilar-tosubscript^𝜏𝒮subscript𝒩card𝒮subscript𝜏𝒮subscriptΣ𝒮assign where subscriptΣ𝒮diagsuperscriptsubscript𝜎subscript𝒮12superscriptsubscript𝜎subscript𝒮card𝒮2\widehat{\tau}_{\mathscr{S}}\sim\mathcal{N}_{\textrm{card}(\mathscr{S})}\left(% \tau_{\mathscr{S}}\>,\>\Sigma_{\mathscr{S}}\right),\textrm{ where }\Sigma_{% \mathscr{S}}:=\operatorname{diag}(\sigma_{\mathscr{S}_{1}}^{2},\ldots,\sigma_{% \mathscr{S}_{\textrm{card}(\mathscr{S})}}^{2}).over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ∼ caligraphic_N start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT ( italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT , roman_Σ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) , where roman_Σ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT := roman_diag ( italic_σ start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , … , italic_σ start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) . (4)

Following Gechter et al. (2024), treat Σ𝒮subscriptΣ𝒮\Sigma_{\mathscr{S}}roman_Σ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT as known.

The normality assumption in (4) is unlikely to hold exactly; however, it is common to assume that treatment effect estimates from randomized controlled trials are asymptotically normal with asymptotic variances that can be estimated consistently. Treating the limiting normal model as an exact finite-sample statistical model then eases exposition and allows us to focus on the core features of the site selection problem. Indeed, working directly with such a limiting model is common in applications of statistical decision theory to econometrics; see Müller (2011) and the references therein for theoretical support and applications in the context of testing problems and Ishihara and Kitagawa (2021), Stoye (2012), or Tetenov (2012) for precedents in closely related work. Gechter et al. (2024) use the same statistical model, but our parameter space has a more specific form as treatment effects are controlled by the Lipschitz function τ𝜏\mathcal{\tau}italic_τ.

After observing τ^𝒮subscript^𝜏𝒮\widehat{\mathcal{\tau}}_{\mathscr{S}}over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT, the policy maker chooses an action as[0,1]subscript𝑎𝑠01a_{s}\in[0,1]italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∈ [ 0 , 1 ] at each policy-relevant site s𝒮P𝑠subscript𝒮𝑃s\in\mathcal{S}_{P}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT. We interpret this action as the proportion of a population in the site that will be randomly assigned to the new policy. Thus, as=1subscript𝑎𝑠1a_{s}=1italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 1 means that everyone in site s𝑠sitalic_s is exposed to the new policy, and as=0subscript𝑎𝑠0a_{s}=0italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 0 means that the status quo at the site is preserved. Under this interpretation, as=.5subscript𝑎𝑠.5a_{s}=.5italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = .5 means that 50% of the population at site s𝑠sitalic_s will be exposed at random to the new policy; however, the formal development equally applies to either individual or population-level randomization. Our interpretation abstracts from integer issues arising with small populations.

Thus, we can define a treatment rule T𝑇Titalic_T as a (measurable) function T:card(𝒮)[0,1]card(𝒮P):𝑇superscriptcard𝒮superscript01cardsubscript𝒮𝑃T:\mathbb{R}^{\textrm{card}(\mathscr{S})}\rightarrow[0,1]^{\textrm{card}(% \mathcal{S}_{P})}italic_T : blackboard_R start_POSTSUPERSCRIPT card ( script_S ) end_POSTSUPERSCRIPT → [ 0 , 1 ] start_POSTSUPERSCRIPT card ( caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT that maps experimental outcomes to (possibly) randomized policy actions in each of the policy-relevant sites. It will be sometimes convenient to use Ts()subscript𝑇𝑠T_{s}(\cdot)italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( ⋅ ) to denote the specific treatment rule for site s𝒮P𝑠subscript𝒮𝑃s\in\mathcal{S}_{P}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT implied by T𝑇Titalic_T and 𝒯𝒮subscript𝒯𝒮\mathcal{T}_{\mathscr{S}}caligraphic_T start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT to denote the set of all treatment rules. Note that we index the treatment rules by the selected experimental sites, 𝒮𝒮\mathscr{S}script_S, to be explicit about the fact that the data used to inform policy will vary depending on the choice of 𝒮𝒮\mathscr{S}script_S. We call T𝒯𝒮𝑇subscript𝒯𝒮T\in\mathcal{T}_{\mathscr{S}}italic_T ∈ caligraphic_T start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT nonrandomized if for every s𝒮P𝑠subscript𝒮𝑃s\in\mathcal{S}_{P}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT we have Ts(z){0,1}subscript𝑇𝑠𝑧01T_{s}(z)\in\{0,1\}italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_z ) ∈ { 0 , 1 } for (Lebesgue) almost every zcard(𝒮)𝑧superscriptcard𝒮z\in\mathbb{R}^{\textrm{card}(\mathscr{S})}italic_z ∈ blackboard_R start_POSTSUPERSCRIPT card ( script_S ) end_POSTSUPERSCRIPT. Otherwise, we say that the rule is randomized. We will focus on decision rules that belong to the set

𝒯𝒮1/2:={T𝒯𝒮| for any diagonal matrix Σ, and s𝒮P,𝔼[Ts(U)]=1/2,U𝒩card(𝒮)(𝟎,Σ)}.assignsuperscriptsubscript𝒯𝒮12conditional-set𝑇subscript𝒯𝒮formulae-sequence for any diagonal matrix Σ and for-all𝑠subscript𝒮𝑃formulae-sequence𝔼delimited-[]subscript𝑇𝑠𝑈12similar-to𝑈subscript𝒩card𝒮0Σ\mathcal{T}_{\mathscr{S}}^{1/2}:=\left\{T\in\mathcal{T}_{\mathscr{S}}\>|\>% \textrm{ for any diagonal matrix }\Sigma,\textrm{ and }\forall s\in\mathcal{S}% _{P},\mathbb{E}[T_{s}(U)]=1/2,U\sim\mathcal{N}_{\textrm{card}(\mathscr{S})}(% \mathbf{0},\Sigma)\right\}.caligraphic_T start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT := { italic_T ∈ caligraphic_T start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT | for any diagonal matrix roman_Σ , and ∀ italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT , blackboard_E [ italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_U ) ] = 1 / 2 , italic_U ∼ caligraphic_N start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT ( bold_0 , roman_Σ ) } .

These are treatment rules in which the ex-ante probability of implementing the policy is 50%percent5050\%50 % whenever the true treatment effects at the sites experimented on are zero. For the moment, and for the sake of exposition, we assume that there is no cost of experimentation. While this assumption is clearly unrealistic, we later show that the main conclusions of our analysis are robust to adding fixed costs to the objective function of the k𝑘kitalic_k-median problem.

2.3 Welfare and Regret

Let #𝒮P:=card(𝒮P)assign#subscript𝒮𝑃cardsubscript𝒮𝑃\#\mathcal{S}_{P}:=\textrm{card}(\mathcal{S}_{P})# caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT := card ( caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ). We assume that the welfare of a decision rule T𝑇Titalic_T, given that sites 𝒮𝒮\mathscr{S}script_S are selected for experimentation, corresponds to the average expected welfare across policy-relevant sites:

𝒲(T,𝒮,τ):=1#𝒮Ps𝒮Pτ(Xs)𝔼τ𝒮[Ts(τ^𝒮)],assign𝒲𝑇𝒮𝜏1#subscript𝒮𝑃subscript𝑠subscript𝒮𝑃𝜏subscript𝑋𝑠subscript𝔼subscript𝜏𝒮delimited-[]subscript𝑇𝑠subscript^𝜏𝒮\mathcal{W}(T,\mathscr{S},\tau):=\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal% {S}_{P}}\tau(X_{s})\mathbb{E}_{\tau_{\mathscr{S}}}[T_{s}(\widehat{\tau}_{% \mathscr{S}})],caligraphic_W ( italic_T , script_S , italic_τ ) := divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) blackboard_E start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] , (5)

where 𝔼τ𝒮[Ts(τ^𝒮)]subscript𝔼subscript𝜏𝒮delimited-[]subscript𝑇𝑠subscript^𝜏𝒮\mathbb{E}_{\tau_{\mathscr{S}}}[T_{s}(\widehat{\tau}_{\mathscr{S}})]blackboard_E start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] means that the expectation is taken assuming τ^𝒮𝒩card(𝒮)(τ𝒮,Σ𝒮)similar-tosubscript^𝜏𝒮subscript𝒩card𝒮subscript𝜏𝒮subscriptΣ𝒮\widehat{\tau}_{\mathscr{S}}\sim\mathcal{N}_{\textrm{card}(\mathscr{S})}(\tau_% {\mathscr{S}},\Sigma_{\mathscr{S}})over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ∼ caligraphic_N start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT ( italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT , roman_Σ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ), and τ(Xs)𝜏subscript𝑋𝑠\tau(X_{s})italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) is the true treatment effect at site s𝑠sitalic_s.

The regret of policy (T,𝒮)𝑇𝒮(T,\mathscr{S})( italic_T , script_S ) can then be shown to equal

(T,𝒮,τ):=1#𝒮Ps𝒮Pτ(Xs)(𝟏{τ(Xs)0}𝔼τ𝒮[Ts(τ^𝒮)]).assign𝑇𝒮𝜏1#subscript𝒮𝑃subscript𝑠subscript𝒮𝑃𝜏subscript𝑋𝑠1𝜏subscript𝑋𝑠0subscript𝔼subscript𝜏𝒮delimited-[]subscript𝑇𝑠subscript^𝜏𝒮\mathcal{R}(T,\mathscr{S},\tau):=\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal% {S}_{P}}\tau(X_{s})\left(\mathbf{1}\{\tau(X_{s})\geq 0\}-\mathbb{E}_{\tau_{% \mathscr{S}}}[T_{s}(\widehat{\tau}_{\mathscr{S}})]\right).caligraphic_R ( italic_T , script_S , italic_τ ) := divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ( bold_1 { italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ≥ 0 } - blackboard_E start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] ) . (6)

Our focus will be on finding the purposive sampling scheme that minimizes worst-case regret.

Definition 1 (MMR optimal purposive sampling scheme and treatment rule).

The pair (T,𝒮)𝒯𝒮1/2×𝒜(k)superscript𝑇superscript𝒮superscriptsubscript𝒯𝒮12𝒜𝑘(T^{*},\mathscr{S}^{*})\in\mathcal{T}_{\mathscr{S}}^{1/2}\times\mathcal{A}(k)( italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ∈ caligraphic_T start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT × caligraphic_A ( italic_k ) is minimax-regret (MMR) optimal among all purposive sampling schemes and treatment rules if

supτLipC(d)(T,𝒮,τ)=inf𝒮𝒜(k),T𝒯𝒮1/2supτLipC(d)(T,𝒮,τ).subscriptsupremum𝜏subscriptLip𝐶superscript𝑑superscript𝑇superscript𝒮𝜏subscriptinfimumformulae-sequence𝒮𝒜𝑘𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}(T^{*},\mathscr{S}^{*% },\tau)=\inf_{\mathscr{S}\in\mathcal{A}(k),T\in\mathcal{T}^{1/2}_{\mathscr{S}}% }\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},\tau).roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_τ ) = roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) , italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) . (7)
Remark 1.

In the standard definition of minimax-regret optimality, it is common to allow the decision maker to select randomized decision rules. Definition 1 implies an asymmetric treatment of randomization: While we allow the policy maker to randomize policy implementation choices, we are restricting her to pick the experimental sites in a deterministic fashion. The nonrandom selection of experimental sites—referred to broadly in the literature as purposive sampling (Cook et al., 2002, p. 511)—is common in practice; see the discussion about purposive sampling in Egami and Lee (2024). In Section 6, we discuss the challenges we encountered in trying to allow for the random selection of experimental sites. ∎

Remark 2.

It will sometimes be convenient to rewrite the right-hand side of (7) as

inf𝒮𝒜(k)(infT𝒯𝒮1/2supτLipC(d)(T,𝒮,τ)).subscriptinfimum𝒮𝒜𝑘subscriptinfimum𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏\inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\inf_{T\in\mathcal{T}^{1/2}_{\mathscr% {S}}}\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},% \tau)\right).roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ) .

This suggests that, conceptually, the solutions of the minimax-regret problem can be done in two steps. First, one can analyze the problem of policy implementation given the experimental outcomes at sites 𝒮𝒮\mathscr{S}script_S. Then, one can optimize over the sites where to experiment.

This distinction is helpful because the first step is related to the “evidence aggregation” problem analyzed in Ishihara and Kitagawa (2021). As mentioned in the Introduction, we can leverage recent results in Yata (2021) and Montiel Olea et al. (2023) to better understand the minimax-regret solutions to this problem. A key difference from Ishihara and Kitagawa (2021) is that we typically have more than one policy-relevant site. ∎

2.4 Main Assumptions

We make the following assumptions. To begin, treatment effects vary continuously with the site-level covariates:

Assumption 1.

τ𝜏\tauitalic_τ is a Lipschitz function (with respect to the Euclidean norm) with known constant C𝐶Citalic_C. That is, for any x,xd𝑥superscript𝑥superscript𝑑x,x^{\prime}\in\mathbb{R}^{d}italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, |τ(x)τ(x)|Cxx,𝜏𝑥𝜏superscript𝑥𝐶norm𝑥superscript𝑥|\tau(x)-\tau(x^{\prime})|\leq C\|x-x^{\prime}\|,| italic_τ ( italic_x ) - italic_τ ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | ≤ italic_C ∥ italic_x - italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∥ , where \|\cdot\|∥ ⋅ ∥ denotes the Euclidean norm.

We also impose a regularity condition on site-level covariates. That is, we assume that all observed covariates are different:

Assumption 2.

XsXss,s𝒮formulae-sequencesubscript𝑋𝑠subscript𝑋superscript𝑠for-all𝑠superscript𝑠𝒮X_{s}\neq X_{s^{\prime}}\ \forall\ s,s^{\prime}\in\mathcal{S}italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≠ italic_X start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ∀ italic_s , italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ caligraphic_S.

Even if this were not the case in raw data, one would presumably want to induce it by adding site fixed effects.555In section 6.3, we discuss how Assumption 1 can be modified to accommodate other general distance measures, as well as some forms of unobserved treatment heterogeneity (for which Assumption 2 may be dropped).

Finally, for each policy site s𝒮P𝑠subscript𝒮𝑃s\in\mathcal{S}_{P}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT and every 𝒮𝒜(k)𝒮𝒜𝑘\mathscr{S}\in\mathcal{A}(k)script_S ∈ caligraphic_A ( italic_k ), denote by N𝒮(s)𝒮subscript𝑁𝒮𝑠𝒮N_{\mathscr{S}}(s)\in\mathscr{S}italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) ∈ script_S its nearest neighbor in 𝒮𝒮\mathscr{S}script_S (or the nearest neighbor with the smallest index in case of multiplicity). That is, for every s𝒮P𝑠subscript𝒮𝑃s\in\mathcal{S}_{P}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT:

XsXN𝒮(s)XsXss𝒮.normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠normsubscript𝑋𝑠subscript𝑋superscript𝑠for-allsuperscript𝑠𝒮\|X_{s}-X_{N_{\mathscr{S}}(s)}\|\leq\|X_{s}-X_{s^{\prime}}\|\;\>\forall s^{% \prime}\in\mathscr{S}.∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ≤ ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ∥ ∀ italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ script_S .

3 Main Result

This section presents our main result: The problem of finding the purposive sampling scheme that minimizes the worst-case regret is approximately equal—and sometimes equal—to solving a k𝑘kitalic_k-median problem.

In order to derive this, first recall the definition of the k𝑘kitalic_k-median problem.

Definition 2 (k𝑘kitalic_k-median problem).

We say that a purposive sampling scheme, 𝒮𝒜(k)𝒮𝒜𝑘\mathscr{S}\in\mathcal{A}(k)script_S ∈ caligraphic_A ( italic_k ), solves the k𝑘kitalic_k-median problem if it solves

inf𝒮𝒜(k)s𝒮PXsXN𝒮(s).subscriptinfimum𝒮𝒜𝑘subscript𝑠subscript𝒮𝑃normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\inf_{\mathscr{S}\in\mathcal{A}(k)}\sum_{s\in\mathcal{S}_{P}}\|X_{s}-X_{N_{% \mathscr{S}}(s)}\|.roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ . (8)

In Definition 2, the policy-relevant sites (with indexes in 𝒮Psubscript𝒮𝑃\mathcal{S}_{P}caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT) are clients and the experimental sites (with indexes in 𝒮Esubscript𝒮𝐸\mathcal{S}_{E}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT) are facilities. The connection cost between a facility i𝑖iitalic_i and a client j𝑗jitalic_j is the Euclidean distance XjXinormsubscript𝑋𝑗subscript𝑋𝑖\|X_{j}-X_{i}\|∥ italic_X start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∥. Since the goal of the k𝑘kitalic_k-median problem is to choose the k𝑘kitalic_k facilities that minimize connection cost, each client s𝒮P𝑠subscript𝒮𝑃s\in\mathcal{S}_{P}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT gets connected to the facility that is closest among those in 𝒮𝒮\mathscr{S}script_S. Hence, the term XN𝒮(s)subscript𝑋subscript𝑁𝒮𝑠X_{N_{\mathscr{S}}(s)}italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT appears in Definition 2. Finally, Equation (8) can be also written as

inf𝒮𝒜(k)(s𝒮P𝒮XsXN𝒮(s)+s𝒮P\𝒮XsXN𝒮(s))=inf𝒮𝒜(k)s𝒮P\𝒮XsXN𝒮(s),subscriptinfimum𝒮𝒜𝑘subscript𝑠subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠subscriptinfimum𝒮𝒜𝑘subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\sum_{s\in\mathcal{S}_{P}\cap\mathscr% {S}}\|X_{s}-X_{N_{\mathscr{S}}(s)}\|+\sum_{s\in\mathcal{S}_{P}\backslash% \mathscr{S}}\|X_{s}-X_{N_{\mathscr{S}}(s)}\|\right)=\inf_{\mathscr{S}\in% \mathcal{A}(k)}\sum_{s\in\mathcal{S}_{P}\backslash\mathscr{S}}\|X_{s}-X_{N_{% \mathscr{S}}(s)}\|,roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ∩ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ + ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) = roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ , (9)

where the equality follows from the fact that for each facility s𝒮P𝑠subscript𝒮𝑃s\in\mathcal{S}_{P}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT that is also a client s𝒮𝑠𝒮s\in\mathscr{S}italic_s ∈ script_S, connection cost becomes zero.

There is no scope for interesting analysis, or indeed for justification of experiments, unless noise in the signal is bounded. Therefore, define C:=max𝒮𝒜(k),s𝒮P\𝒮{π2σN𝒮(s)XsXN𝒮(s)}<assignsuperscript𝐶formulae-sequence𝒮𝒜𝑘𝑠\subscript𝒮𝑃𝒮𝜋2subscript𝜎subscript𝑁𝒮𝑠normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠C^{*}:=\underset{\mathscr{S}\in\mathcal{A}(k),s\in\mathcal{S}_{P}\backslash% \mathscr{S}}{\max}\left\{\sqrt{\frac{\pi}{2}}\frac{\sigma_{N_{{\mathscr{S}}(s)% }}}{\left\|X_{s}-X_{N_{{\mathscr{S}}(s)}}\right\|}\right\}<\inftyitalic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT := start_UNDERACCENT script_S ∈ caligraphic_A ( italic_k ) , italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_UNDERACCENT start_ARG roman_max end_ARG { square-root start_ARG divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_ARG divide start_ARG italic_σ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S ( italic_s ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S ( italic_s ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ end_ARG } < ∞. Let σEsubscript𝜎𝐸\sigma_{E}italic_σ start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT denote the largest standard deviation among the potential experimental sites. Our main theorem is the following:

Theorem 1.

Suppose Assumptions 1-2 hold. If C>C𝐶superscript𝐶C>C^{*}italic_C > italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, then for any 𝒮𝒜(k)𝒮𝒜𝑘\mathscr{S}\in\mathcal{A}(k)script_S ∈ caligraphic_A ( italic_k ):

|(infT𝒯𝒮1/2supτLipC(d)(T,𝒮,τ))C21#𝒮P(s𝒮P\𝒮XsXN𝒮(s))|subscriptinfimum𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏𝐶21#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\Bigg{|}\left(\inf_{T\in\mathcal{T}^{1/2}_{\mathscr{S}}}\sup_{\tau\in\textrm{% Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},\tau)\right)-\frac{C}{2}% \frac{1}{\#\mathcal{S}_{P}}\cdot\left(\sum_{s\in\mathcal{S}_{P}\backslash% \mathscr{S}}\|X_{s}-X_{N_{\mathscr{S}}(s)}\|\right)\Bigg{|}| ( roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ) - divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ⋅ ( ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) | (10)

is at most

BσEmin{#(𝒮E𝒮P),k}#𝒮P,𝐵subscript𝜎𝐸#subscript𝒮𝐸subscript𝒮𝑃𝑘#subscript𝒮𝑃B\cdot\sigma_{E}\cdot\frac{\min\{\#\left(\mathcal{S}_{E}\cap\mathcal{S}_{P}% \right),k\}}{\#\mathcal{S}_{P}},italic_B ⋅ italic_σ start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ⋅ divide start_ARG roman_min { # ( caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) , italic_k } end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG , (11)

where Bargmaxz0zΦ(z)𝐵subscript𝑧0𝑧Φ𝑧B\equiv\arg\max_{z\geq 0}z\Phi(-z)italic_B ≡ roman_arg roman_max start_POSTSUBSCRIPT italic_z ≥ 0 end_POSTSUBSCRIPT italic_z roman_Φ ( - italic_z ).

Proof.

Follows directly from Lemma 1 and 2 below. ∎

Remark 3.

The left-hand expression in (10), i.e., infT𝒯𝒮1/2supτLipC(d)(T,𝒮,τ)subscriptinfimum𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏\inf_{T\in\mathcal{T}^{1/2}_{\mathscr{S}}}\sup_{\tau\in\textrm{Lip}_{C}(% \mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},\tau)roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ), is the worst-case regret of the purposive sampling scheme 𝒮𝒮\mathscr{S}script_S assuming optimality of subsequent treatment choice T𝒮subscript𝑇𝒮T_{\mathscr{S}}italic_T start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT. Theorem 1 thus shows that the (optimized) worst-case regret of any purposive sampling scheme can be uniformly approximated (after scaling by the factor C/(2#𝒮PC/(2\#\mathcal{S}_{P}italic_C / ( 2 # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT)) by the objective function of the k-median problem in Definition 2—which is equivalent to the objective function of the problem in the right-hand side of Equation (9). ∎

Remark 4.

When the candidate sites for experimentation (𝒮E)subscript𝒮𝐸(\mathcal{S}_{E})( caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ) and the policy-relevant sites (𝒮P)subscript𝒮𝑃(\mathcal{S}_{P})( caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) are disjoint, then for any 𝒮𝒜(k)𝒮𝒜𝑘\mathscr{S}\in\mathcal{A}(k)script_S ∈ caligraphic_A ( italic_k ):

(infT𝒯𝒮1/2supτLipC(d)(T,𝒮,τ))=C21#𝒮P(s𝒮P\𝒮XsXN𝒮(s)).subscriptinfimum𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏𝐶21#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\left(\inf_{T\in\mathcal{T}^{1/2}_{\mathscr{S}}}\sup_{\tau\in\textrm{Lip}_{C}(% \mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},\tau)\right)=\frac{C}{2}\frac{1}{\#% \mathcal{S}_{P}}\cdot\left(\sum_{s\in\mathcal{S}_{P}\backslash\mathscr{S}}\|X_% {s}-X_{N_{\mathscr{S}}(s)}\|\right).( roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ) = divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ⋅ ( ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) . (12)

Consequently, any solution of the k𝑘kitalic_k-median problem is a fortiori an exact minimax-regret solution for the site selection problem. This follows from the fact that if a purposive sampling scheme, 𝒮superscript𝒮\mathscr{S}^{*}script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, minimizes the right-hand side of Equation (12), then it will automatically minimize the left-hand side, which, by Remark 2, defines a minimax-regret optimal purposive sampling scheme. ∎

Remark 5.

Theorem 1 implies that even in the case where 𝒮E𝒮Psubscript𝒮𝐸subscript𝒮𝑃\mathcal{S}_{E}\cap\mathcal{S}_{P}\neq\emptysetcaligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ≠ ∅,

|inf𝒮𝒜(k)(infT𝒯𝒮1/2supτLipC(d)(T,𝒮,τ))C21#𝒮Pinf𝒮𝒜(k)(s𝒮P\𝒮XsXN𝒮(s))|subscriptinfimum𝒮𝒜𝑘subscriptinfimum𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏𝐶21#subscript𝒮𝑃subscriptinfimum𝒮𝒜𝑘subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\Bigg{|}\inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\inf_{T\in\mathcal{T}^{1/2}_{% \mathscr{S}}}\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}(T,% \mathscr{S},\tau)\right)-\frac{C}{2}\frac{1}{\#\mathcal{S}_{P}}\cdot\inf_{% \mathscr{S}\in\mathcal{A}(k)}\left(\sum_{s\in\mathcal{S}_{P}\backslash\mathscr% {S}}\|X_{s}-X_{N_{\mathscr{S}}(s)}\|\right)\Bigg{|}| roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ) - divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ⋅ roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) | (13)

is at most equal to (11). Thus, the quality of this approximation improves as either the number of policy-relevant sites increases or the potential experiments conducted become more precise (σEsubscript𝜎𝐸\sigma_{E}italic_σ start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT becomes smaller). The approximation deteriorates as k𝑘kitalic_k increases. This happens because in approximating the worst-case regret of a purposive sampling scheme, we ignored the regret associated with sites selected for experimentation; we further discuss this below when we explain Lemma 1. This means that, as k𝑘kitalic_k increases, more sites will be ignored in our approximation and the upper bound will become looser. ∎

Remark 6.

The tight connection to the k𝑘kitalic_k-median problem relies on the Lipschitz constant to be large enough. In other words, the k𝑘kitalic_k-median problem only has a decision-theoretic justification when the treatment effect heterogeneity across sites accommodated in the parameter space is substantial. ∎

The proof of Theorem 1 follows from the two Lemmas below, which bound the site selection problem’s minimax-regret value from above and below, respectively.

Lemma 1 (Upper bound).

Suppose Assumptions 1-2 hold. For every 𝒮𝒜(k)𝒮𝒜𝑘\mathscr{S}\in\mathcal{A}(k)script_S ∈ caligraphic_A ( italic_k ), there exists a constant C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S ) such that, if C>C(𝒮)𝐶𝐶𝒮C>C(\mathscr{S})italic_C > italic_C ( script_S ), then

infT𝒯𝒮1/2supτLipC(d)(T,𝒮,τ)(B#𝒮Ps𝒮𝒮Pσs)+(C21#𝒮Ps𝒮P\𝒮XsXN𝒮(s)),subscriptinfimum𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏𝐵#subscript𝒮𝑃subscript𝑠𝒮subscript𝒮𝑃subscript𝜎𝑠𝐶21#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\inf_{T\in\mathcal{T}^{1/2}_{\mathscr{S}}}\sup_{\tau\in\textrm{Lip}_{C}(% \mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},\tau)\leq\;\left(\frac{B}{\#\mathcal% {S}_{P}}\sum_{s\in\mathscr{S}\cap\mathcal{S}_{P}}\sigma_{s}\right)+\left(\frac% {C}{2}\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal{S}_{P}\backslash\mathscr{S% }}\|X_{s}-X_{N_{\mathscr{S}}(s)}\|\right),roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ≤ ( divide start_ARG italic_B end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ script_S ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) + ( divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) , (14)

where Bargmaxz0zΦ(z)𝐵subscript𝑧0𝑧Φ𝑧B\equiv\arg\max_{z\geq 0}z\Phi(-z)italic_B ≡ roman_arg roman_max start_POSTSUBSCRIPT italic_z ≥ 0 end_POSTSUBSCRIPT italic_z roman_Φ ( - italic_z ).

Proof.

See Appendix A.1. ∎

Remark 7.

Lemma 1 implies that the minimax-regret value of the site selection problem in Definition 1 can be upper bounded by the solution of the uncapacitated k𝑘kitalic_k-facility location problem.666Uncapacitated means that there is no capacity constraint on the number of clients that each facility can accommodate. See (Williamson and Shmoys, 2011, Chapter 4.5) and Zhang (2007). Just as in the k𝑘kitalic_k-median problem, there is a set of facilities \mathcal{F}caligraphic_F and a set of clients (or cities) 𝒞𝒞\mathcal{C}caligraphic_C. Now we assume that there is an opening cost ci+subscript𝑐𝑖subscriptc_{i}\in\mathbb{R}_{+}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∈ blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT associated with each facility i𝑖i\in\mathcal{F}italic_i ∈ caligraphic_F. The connection cost between facilities and clients is as before. The goal is to open at most k𝑘kitalic_k facilities and connect each client to one facility so that total cost is minimized. Thus, the problem

inf𝒮𝒜(k)((B#𝒮Ps𝒮𝒮Pσs)+(C21#𝒮Ps𝒮P\𝒮XsXN𝒮(s)))subscriptinfimum𝒮𝒜𝑘𝐵#subscript𝒮𝑃subscript𝑠𝒮subscript𝒮𝑃subscript𝜎𝑠𝐶21#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\left(\frac{B}{\#\mathcal{S}_{P}}\sum% _{s\in\mathscr{S}\cap\mathcal{S}_{P}}\sigma_{s}\right)+\left(\frac{C}{2}\frac{% 1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal{S}_{P}\backslash\mathscr{S}}\|X_{s}-X_% {N_{\mathscr{S}}(s)}\|\right)\right)roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( ( divide start_ARG italic_B end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ script_S ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) + ( divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) ) (15)

can be viewed as a k𝑘kitalic_k-facility location problem. Just as before, the set of facilities is 𝒮Esubscript𝒮𝐸\mathcal{S}_{E}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT and the set of clients is 𝒮Psubscript𝒮𝑃\mathcal{S}_{P}caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT. The connection cost between sites s𝒮E𝑠subscript𝒮𝐸s\in\mathcal{S}_{E}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT and s𝒮psuperscript𝑠subscript𝒮𝑝s^{\prime}\in\mathcal{S}_{p}italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ caligraphic_S start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is (C/(2#𝒮P))XsXs𝐶2#subscript𝒮𝑃normsubscript𝑋𝑠subscript𝑋superscript𝑠(C/(2\#\mathcal{S}_{P}))\|X_{s}-X_{s^{\prime}}\|( italic_C / ( 2 # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) ) ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ∥. The opening cost for any facility s𝒮E𝑠subscript𝒮𝐸s\in\mathcal{S}_{E}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT is (B/#𝒮P)σs𝐵#subscript𝒮𝑃subscript𝜎𝑠(B/\#\mathcal{S}_{P})\sigma_{s}( italic_B / # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) italic_σ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. Minimizing the upper bound in Lemma 1 is thus equivalent to solving the k𝑘kitalic_k-facility location problem. As shown in Appendix A.1, the upper bound arises by bounding the worst-case sum of regrets across sites by the corresponding sum of worst-case regrets. For selected sites that are both facilities and clients, the worst-case regret is obtained by solving a point-identified treatment choice problem. For policy-relevant sites where no experiment was conducted, it is obtained by solving the partially-identified treatment choice problem in Ishihara and Kitagawa (2021). When 𝒮Esubscript𝒮𝐸\mathcal{S}_{E}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT and 𝒮Psubscript𝒮𝑃\mathcal{S}_{P}caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT are disjoint, the first component of the upper bound vanishes, and the bound becomes proportional to the solution of the k𝑘kitalic_k-median problem. ∎

Lemma 2 (Lower bound).

Suppose Assumptions 1-2 hold. For every 𝒮𝒜(k)𝒮𝒜𝑘\mathscr{S}\in\mathcal{A}(k)script_S ∈ caligraphic_A ( italic_k ):

infT𝒯𝒮1/2supτLipC(d)(T,𝒮,τ)(C21#𝒮Ps𝒮P\𝒮XsXN𝒮(s)).subscriptinfimum𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏𝐶21#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\inf_{T\in\mathcal{T}^{1/2}_{\mathscr{S}}}\sup_{\tau\in\textrm{Lip}_{C}(% \mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},\tau)\geq\;\left(\frac{C}{2}\frac{1}% {\#\mathcal{S}_{P}}\sum_{s\in\mathcal{S}_{P}\backslash\mathscr{S}}\|X_{s}-X_{N% _{\mathscr{S}}(s)}\|\right).roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ≥ ( divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) . (16)
Proof.

See Appendix A.2. ∎

Remark 8.

Lemma 2 implies that the minimax-regret value of the site selection problem in Definition 1 can be lower bounded by the solution of the k-median problem:

inf𝒮𝒜(k)(C21#𝒮Ps𝒮P\𝒮XsXN𝒮(s)).subscriptinfimum𝒮𝒜𝑘𝐶21#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\frac{C}{2}\frac{1}{\#\mathcal{S}_{P}% }\sum_{s\in\mathcal{S}_{P}\backslash\mathscr{S}}\|X_{s}-X_{N_{\mathscr{S}}(s)}% \|\right).roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) . (17)

Theorem 1 thus follows by noting that the upper and lower bound match up to the opening costs of the facilities. As noted before, the lower and upper bound match when 𝒮E𝒮P=subscript𝒮𝐸subscript𝒮𝑃\mathcal{S}_{E}\cap\mathcal{S}_{P}=\emptysetcaligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = ∅. ∎

Remark 9.

While the computation of an exact minimax-regret sampling scheme (whether purposive or randomized) appears intractable, our ability to provide nontrivial upper bounds on worst-case regret does not hinge on C𝐶Citalic_C being large enough. Indeed, one could always compute such a bound by (i) forcing the treatment decision Tssubscript𝑇𝑠T_{s}italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT to only depend on site N𝒮(s)subscript𝑁𝒮𝑠N_{\mathscr{S}}(s)italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) and (ii) computing worst-case expected regret separately across sites, ignoring that the regret-maximizing parameter configurations may be mutually inconsistent across sites. Both manipulations increase regret and therefore define an upper bound. This bound is easy to compute for any given sampling scheme because the induced minimax-regret treatment choice problem was solved in Stoye (2012). For C𝐶Citalic_C large enough, it will coincide with the bound in Theorem 1. The caveat is that this bound is nonlinear; hence, while it is easy to compute once, it does not generate a tight connection to the k𝑘kitalic_k-median problem and will be hard to optimize for large instances of the site selection problem. ∎

4 Integer Programming and the k𝑘kitalic_k-median Problem

It is well known (Williamson and Shmoys, 2011, Chapter 7.7, p. 185) that the k𝑘kitalic_k-median problem in (17) can be formulated as the following linear integer program:

min{yi,xi,j}i𝒮E,j𝒮Psubscriptsubscriptsubscript𝑦𝑖subscript𝑥𝑖𝑗formulae-sequence𝑖subscript𝒮𝐸𝑗subscript𝒮𝑃\displaystyle\min_{\{y_{i},x_{i,j}\}_{i\in\mathcal{S}_{E},j\in\mathcal{S}_{P}}}roman_min start_POSTSUBSCRIPT { italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT , italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_POSTSUBSCRIPT i𝒮E,j𝒮Pxi,jc(j,i)subscriptformulae-sequence𝑖subscript𝒮𝐸𝑗subscript𝒮𝑃subscript𝑥𝑖𝑗𝑐𝑗𝑖\displaystyle\sum_{i\in\mathcal{S}_{E},j\in\mathcal{S}_{P}}x_{i,j}\cdot c(j,i)∑ start_POSTSUBSCRIPT italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT , italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ⋅ italic_c ( italic_j , italic_i ) (18)
such that i𝒮Exi,j=1,j𝒮P,formulae-sequencesubscript𝑖subscript𝒮𝐸subscript𝑥𝑖𝑗1for-all𝑗subscript𝒮𝑃\displaystyle\sum_{i\in\mathcal{S}_{E}}x_{i,j}=1,\;\;\;\;\;\forall j\in% \mathcal{S}_{P},∑ start_POSTSUBSCRIPT italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT = 1 , ∀ italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT , (19)
i𝒮Eyik,subscript𝑖subscript𝒮𝐸subscript𝑦𝑖𝑘\displaystyle\sum_{i\in\mathcal{S}_{E}}y_{i}\leq k,∑ start_POSTSUBSCRIPT italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≤ italic_k , (20)
0xi,jyi,i𝒮E,j𝒮P,formulae-sequence0subscript𝑥𝑖𝑗subscript𝑦𝑖formulae-sequence𝑖subscript𝒮𝐸𝑗subscript𝒮𝑃\displaystyle 0\leq x_{i,j}\leq y_{i},\;\;\;\;\;i\in\mathcal{S}_{E},j\in% \mathcal{S}_{P},0 ≤ italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ≤ italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT , italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT , (21)
yi{0,1},xi,j{0,1},i𝒮E,j𝒮P.formulae-sequencesubscript𝑦𝑖01formulae-sequencesubscript𝑥𝑖𝑗01formulae-sequence𝑖subscript𝒮𝐸𝑗subscript𝒮𝑃\displaystyle y_{i}\in\{0,1\},x_{i,j}\in\{0,1\},\;\;\;\;\;i\in\mathcal{S}_{E},% j\in\mathcal{S}_{P}.italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∈ { 0 , 1 } , italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ∈ { 0 , 1 } , italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT , italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT . (22)

In this formulation, the choice variables are the {0,1}01\{0,1\}{ 0 , 1 }-valued variables yisubscript𝑦𝑖y_{i}italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and xi,jsubscript𝑥𝑖𝑗x_{i,j}italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT, for i𝒮E,j𝒮Pformulae-sequence𝑖subscript𝒮𝐸𝑗subscript𝒮𝑃i\in\mathcal{S}_{E},j\in\mathcal{S}_{P}italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT , italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT. The variable yisubscript𝑦𝑖y_{i}italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT indicates whether site i𝑖iitalic_i is chosen for experimentation, and xi,jsubscript𝑥𝑖𝑗x_{i,j}italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT indicates whether experimental site i𝑖iitalic_i is the nearest neighbor of policy site j𝑗jitalic_j. The total number of sites chosen for experimentation is restricted to be no greater than k𝑘kitalic_k, and site i𝑖iitalic_i can only be the nearest neighbor of policy site j𝑗jitalic_j if site i𝑖iitalic_i is chosen for experimentation. The connection cost between a facility i𝑖iitalic_i and a client j𝑗jitalic_j—denoted c(j,i)𝑐𝑗𝑖c(j,i)italic_c ( italic_j , italic_i )—is given by XiXjnormsubscript𝑋𝑖subscript𝑋𝑗\|X_{i}-X_{j}\|∥ italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ∥.777The connection cost in the objective function of the integer program differs from the connection cost in the k𝑘kitalic_k-median problem in (17) by a constant factor C/(2#𝒮P)𝐶2#subscript𝒮𝑃C/(2\#\mathcal{S}_{P})italic_C / ( 2 # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ), which does not affect the solution of the k𝑘kitalic_k-median problem. Solving the linear integer program described above is equivalent to solving (17). It is worth highlighting that the program above is linear in the choice variables xi,jsubscript𝑥𝑖𝑗x_{i,j}italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT and yisubscript𝑦𝑖y_{i}italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT.

A major advantage of the integer programming formulation of the k𝑘kitalic_k-median problem is that most ecosystems for scientific computing offer different solvers for linear and nonlinear integer programs. For the applications in this paper, we use the MIP solver in Gurobi (Gurobi Optimization, LLC, 2023) through the Python package gurobipy. The Gurobi software is highly optimized, especially for large-scale problems, and it offers an academic license. Even though the scale of the applications presented in Section 5 is not large enough for the efficiency advantages of Gurobi to become salient over other solvers, we wanted to showcase its ease of use. It also integrates seamlessly with Google Colab, providing us a way to build self-contained and reproducible examples.888https://colab.research.google.com/

The MIP solver uses a version of the branch-and-bound algorithm; see Bertsimas and Weismantel (2005), Chapter 11.1, for a general description of this algorithm. Broadly speaking, this algorithm works by first finding the solution to the linear programming (LP) relaxation of the original integer problem. This is known as the relaxation step. Then, the problem is split into two sub-problems (branching), according to the integer constraints on the solution. This method then gives bounds on the integer solution, and the algorithm is applied recursively until the lower bound and the upper bound converge up to a tolerance parameter. The recursion creates nodes, and there are some strategies to determine which nodes should be explored further; for example, nodes that have integer solutions do not require any more branching.

Gurobi implements several additional steps that help the branch-and-bound algorithm be more efficient.999See https://www.gurobi.com/resources/mixed-integer-programming-mip-a-primer-on-the-basics/. The presolve step reduces the number of effective constraints of the problem by checking if the integer requirement can eliminate some of them. As the name suggests, it is performed before the start of the branch-and-bound algorithm. Cutting planes tightens the feasible region by adding linear inequalities to eliminate fractional solutions; it is performed during the branch-and-bound process, and for this reason the algorithm used by Gurobi can be referred to as a version of a “branch-and-cut” algorithm. Finally, the MIP solver implements several heuristics, for example by rounding the component of a solution that is closest to an integer, fixing it, and hoping the other components will converge to integers quicker.

In principle, one can always solve the k𝑘kitalic_k-median problem in (17) by enumerating all possible size k𝑘kitalic_k subsets of 𝒮Esubscript𝒮𝐸\mathcal{S}_{E}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT. Such an algorithm runs in time proportional to

(#𝒮Ek)#𝒮Pkdbinomial#subscript𝒮𝐸𝑘#subscript𝒮𝑃𝑘𝑑{{\#\mathcal{S}_{E}}\choose{k}}\cdot\#\mathcal{S}_{P}\cdot k\cdot d( binomial start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT end_ARG start_ARG italic_k end_ARG ) ⋅ # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ⋅ italic_k ⋅ italic_d

time and therefore scales poorly when the term “#𝒮E#subscript𝒮𝐸\#\mathcal{S}_{E}# caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT choose k𝑘kitalic_k” is large. However, we are able to evaluate the performance of our preferred solver by brute-force solving smaller but not trivial instances of the problem, notably the entire application in Section 5.2.

Figure 9 in Appendix B.1 shows an example of the output obtained after using the MIP solver in Gurobi to solve the linear integer program for the application in Section 5.2. In this example, the scale of the problem is given by #𝒮E=#𝒮P=15,k=6formulae-sequence#subscript𝒮𝐸#subscript𝒮𝑃15𝑘6\#\mathcal{S}_{E}=\#\mathcal{S}_{P}=15,k=6# caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = 15 , italic_k = 6, and d=8𝑑8d=8italic_d = 8. We defer the details of the application to Section 5.2.

While we do not use it in this paper, we finally note that there is a large literature studying efficient (polynomial-time) approximation algorithms for the k𝑘kitalic_k-median problem, going back to seminal work of Charikar et al. (2002). A basic idea in these algorithms is to consider a linear programming relaxation of the integer program associated with the k𝑘kitalic_k-median problem (Williamson and Shmoys, 2011, Chapter 7.7). Even though the scale of the problems analyzed in this paper does not require the implementation of any of these algorithms, there are several papers that present theoretical performance guarantees for them; see for example Cohen-Addad et al. (2022) and also the references in Cohen-Addad et al. (2018). Finally, it is worth mentioning that when k𝑘kitalic_k is fixed (and not viewed as part of the problem’s input), there is an approximate algorithm that runs in time proportional to nd𝑛𝑑n\cdot ditalic_n ⋅ italic_d; see Kumar et al. (2010). Such an algorithm could be useful when n𝑛nitalic_n and d𝑑ditalic_d are large and k𝑘kitalic_k is small.

5 Applications

5.1 Mobile Financial Services in Bangladesh

Lee et al. (2021) conducted a randomized controlled trial in Bangladesh to estimate the effects of encouraging rural households to receive money transfers from migrant family members. They specifically conducted an encouragement design where poor rural households with family members who had migrated to a larger urban destination receive a 30–45 minute training about how to register and use the mobile banking service “bKash” to send instant remittances back home.

The experiment was conducted in the Gaibandha district, one of Bangladesh’s poorest regions. It focused on households that had migrant workers in the Dhaka district, the administrative unit in which the capital of Bangladesh is located. Lee et al. (2021) measure several outcomes of both receiving households and sender migrants; see their Figures 3 and 4. To give a concrete example of the measured outcomes, one question of interest is whether families that adopt the mobile banking technology are more (or less) likely to declare that the monga—the seasonal period of hunger in September through November—was not a problem for their household. Table 9, Column 7, p. 61 in Lee et al. (2021) presents results for this specific variable showing that households that used a bKash account in the treatment group are 9.2 percentage points more likely to declare that monga was not a problem.

We ask: Do the findings in Lee et al. (2021) generalize to other migration corridors, i.e., combinations of origin and destination districts, in Bangladesh? Is the corridor selected by Lee et al. (2021) a good choice for a researcher who is concerned about external validity? Following Gechter et al. (2024), we name the corridors using a destination-origin format; for example, the migration corridor studied in Lee et al. (2021) is “Dhaka-Gaibandha”. Figure 1 displays this corridor along with other common ones. The 41414141 migration corridors analyzed in Gechter et al. (2024) are depicted as dotted lines connecting an origin and a destination.101010We thank Michael Gechter for gracefully sharing part of the dataset used in Gechter et al. (2024).

Refer to caption
Figure 1: Bangladesh Migration Corridors

Notes: The map of Bangladesh with the origins of the migration corridors marked as light blue dots and the destinations marked as dark blue stars. Following the terminology used in Gechter et al. (2024), origin refers to a worker’s home, and destination refers to where the worker migrates for work. The corridor where the experiment was originally implemented in Lee et al. (2021), Dhaka-Gaibandha, is highlighted in yellow.

In Lee et al.’s 2021 words, “[t]he particular nature of our sample potentially limits the external validity” of the analysis. In short, migration corridors differ in characteristics ranging from the distance between the origin and the destination to the cost of living and the average wages. Figure 2 presents a box plot of d=13𝑑13d=13italic_d = 13 (standardized) characteristics that Gechter et al. (2024) collected for each of the 41414141 migration corridors. We take these corridors to be our potential experimental and policy-relevant sites. That is, 𝒮E=𝒮Psubscript𝒮𝐸subscript𝒮𝑃\mathcal{S}_{E}=\mathcal{S}_{P}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT, and card(𝒮Esubscript𝒮𝐸\mathcal{S}_{E}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT) = card(𝒮Psubscript𝒮𝑃\mathcal{S}_{P}caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT) = 41. Below we present results for k{1,2}𝑘12k\in\{1,2\}italic_k ∈ { 1 , 2 }.

Refer to caption
(a) k𝑘kitalic_k-median
Refer to caption
(b) Egami and Lee (2024) and Gechter et al. (2024)
Figure 2: Solution of the k𝑘kitalic_k-median Problem (k=1𝑘1k=1italic_k = 1) and Distribution of Site-Level Covariates

Notes: For each covariate, the box represents the interquartile range (IQR), the vertical black line represents the median, and the horizontal line shows the “theoretical minimum” (defined as Q11.5subscript𝑄11.5Q_{1}-1.5italic_Q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - 1.5IQR) and “theoretical maximum” (defined as Q3+1.5subscript𝑄31.5Q_{3}+1.5italic_Q start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + 1.5IQR). Black dots are outliers, defined as observations that fall beyond the theoretical minimum and maximum. Each panel depicts the sites selected by the different approaches when k=1𝑘1k=1italic_k = 1.

Selected Site when k=1𝑘1k=1italic_k = 1: When k=1𝑘1k=1italic_k = 1, the selected site based on the k𝑘kitalic_k-median problem is Dhaka-Pabna. This is also the solution obtained by using the synthetic purposive sampling approach (henceforth, SPS) in Egami and Lee (2024).

Figure 2(a) presents a visual comparison of the covariates of Dhaka-Pabna (blue circles) and Dhaka-Gaibandha (pink diamonds), the original site selected by Lee et al. (2021). The covariate values for Dhaka-Gaibandha are slightly outside the interquartile range for at least three covariates: migrant density, mean remittance, and mean household size. In comparison, all but one covariate value for Dhaka-Pabna are within the interquartile range. The figure also shows that two key covariates of Dhaka-Gaibandha are right at the edges of the interquartile range: distance between sites in the corridor (3rd quartile) and mean household income (1st quartile). One might conjecture that the effects of adopting a mobile banking technology to transfer money particularly depend on distance and household incomes, suggesting that the Dhaka-Gaibandha corridor may not be the most representative.111111Gechter et al. (2024) suggest that these qualities could explain the large treatment effects found by Lee et al. (2021). Interestingly, Dhaka-Pabna has opposite features: the distance between destination and origin in this corridor is short (1st quartile) and households are relatively better off in terms of income (3rd quartile). The use of the minimax criterion might explain why a corridor with these characteristics may be a good choice for extrapolating experimental results.

Figure 2(b) also presents the covariates of Dhaka-Noakhali (yellow triangles), the migration corridor selected by the Bayesian approach of Gechter et al. (2024).121212By construction, the solution of Gechter et al. (2024) depends on the choice of prior. The results herein reported are based on their preferred prior specification; see Section 5.3 p.p.23 in Gechter et al. (2024). We note that for 10 out of the 13 variables that control treatment effect heterogeneity, Dhaka-Noakhali has covariates that are typically outside the interquartile range.

Selected Sites when k=2𝑘2k=2italic_k = 2: When k=2𝑘2k=2italic_k = 2, the k𝑘kitalic_k-median solution is to pick Dhaka-Pabna and Dhaka-Pirojpur. The former also solved the k𝑘kitalic_k-median problem for k=1𝑘1k=1italic_k = 1.

Figure 3(a) presents the covariates of Dhaka-Pabna (filled, blue circle) and Dhaka-Pirojpur (hollow, blue circle). While four covariate values of Dhaka-Pirojpur (standard deviation of male migrant wage, standard deviation of household income, price index, and migrant density) are outside the interquartile range, they still appear more central than Dhaka-Gaibandha and Dhaka-Noakhali. However, relative to Dhaka-Pabna, the solution for k=2𝑘2k=2italic_k = 2 adds a considerably less central site; Figure 4 illustrates that this site is a good nearest neighbor for some sites that would not otherwise be well matched.

Figure 3(b) presents the solutions of Egami and Lee (2024) (squares) and Gechter et al. (2024) (triangles).131313Gechter et al. (2024) impose an additional constraint on purposive sampling schemes: they require the two migration corridors selected for experimentation to have origins in different divisions. We note that both the solutions of Egami and Lee (2024) and the k𝑘kitalic_k-median problem satisfy this constraint as well. The solution of the k𝑘kitalic_k-median problem and synthetic purposive sampling are no longer the same. We also note that the two sites selected by synthetic purposive sampling differ from the site selected by this procedure when k=1𝑘1k=1italic_k = 1.

Refer to caption
(a) k𝑘kitalic_k-median
Refer to caption
(b) Egami and Lee (2024) and Gechter et al. (2024)
Figure 3: Solution of the k𝑘kitalic_k-median Problem (k=2𝑘2k=2italic_k = 2) and Distribution of Site-level Covariates

Notes: Box plots for the distribution of covariates among migration corridors, constructed as explained in Figure 2. Each panel depicts the site selected by the different approaches when k=2𝑘2k=2italic_k = 2.

Refer to caption
(a) k𝑘kitalic_k-median
Refer to caption
(b) Egami and Lee (2024)
Figure 4: Optimal Experimental Sites for k=2𝑘2k=2italic_k = 2 on a Two-dimensional Covariate Plane

Notes: Each point represents a destination-origin migration corridor (site) on the two-dimensional covariate plane, using two covariates: distance between the two ends of each corridor and average household income in the home district (where the average is taken over households with a reported migrant). The solutions of each method are indicated by the shapes of the solid dots. In Panel 4(a), blue circles denote policy sites with Dhaka-Pabna as the nearest neighbor, and orange crosses represent those with Dhaka-Pirojpur as the nearest neighbor. Similarly, in Panel 4(b), blue circles and orange crosses follow the same coding, indicating that these policy sites rely on a single experimental site for constructing their synthetic control. Additionally, gray cross-circles represent policy sites that use both experimental sites for their synthetic control.

Figure 4 presents a simple visualization of optimal connections between sites under different optimality criteria. Note that the underlying scatter plots in both panels are the same; they visualize the location of corridors in (distance, household income)-space. However, Panel 4(a) indicates which sites were selected through solving the k𝑘kitalic_k-median problem and which of these sites any other site was matched to, where blue circles represent matches with Dhaka-Pabna and orange crosses represent matches with Dhaka-Pirojpur. Figure 4(a) also presents the sites selected by Gechter et al. (2024), i.e. Dhaka-Magura and Dhaka-Noakhali, both marked with dark blue triangles. They appear close to the sites selected by the k𝑘kitalic_k-median problem, but in terms of that problem’s criterion function, they only rank 455455455455 among 820820820820 candidate solutions and miss the problem’s optimal value by 17%percent1717\%17 %. Panel 4(b) similarly visualizes the sites selected by Egami and Lee’s (2024) synthetic purposive sampling approach, i.e. Dhaka-Gopalganj (orange square) and Dhaka-Barguna (blue square).141414We generated this figure by using Egami and Lee’s (2024) code on Gechter et al.’s (2024) data. We also visualize how policy sites are matched with experimental sites: Orange crosses correspond to migration corridors that assign all of their weight to Dhaka-Gopalganj; blue circles correspond to migration corridors that assign all of their weight to Dhaka-Barguna; all other sites (gray crossed circles) assign strictly positive weights to both donor sites. This illustrates that these sampling schemes meaningfully differ. This would even be true if the synthetic purposive sampling approach were implemented but forcing degenerate (single donor) matches, because Egami and Lee’s (2024) approach would then reduce to a k𝑘kitalic_k-mean problem, i.e. using squared Euclidean distance as connection cost.

Computational Costs: The above examples are small enough so that the k𝑘kitalic_k-median problem could easily be solved by brute-force enumeration of 41414141 and 820820820820 candidate solutions, respectively. Needless to say, such an approach would not scale—for example, in this same application, k=10𝑘10k=10italic_k = 10 induces 1,121,099,40811210994081,121,099,4081 , 121 , 099 , 408 candidate solutions.

Indeed, Figure 5(a) compares time to solve the k𝑘kitalic_k-median problem for k{1,,10}𝑘110k\in\{1,\ldots,10\}italic_k ∈ { 1 , … , 10 } using a) the integer program formulation of the k𝑘kitalic_k-median problem in Section 4 and b) brute-force enumeration. The MIP solver in Gurobi solves all instances of the problem to provable optimality in less than one second each. In contrast, brute-force enumeration takes approximately 5555 hours for k=10𝑘10k=10italic_k = 10.151515This was run in a Windows XPS with 10101010 cores and 32323232GB of RAM using R. The code is parallelized, and to avoid memory issues when k>8𝑘8k>8italic_k > 8, it runs a C++ function in the background to evaluate one combination at a time. This ensures that we are giving brute-force enumeration the best chance of success.

One potential benefit of brute-force enumeration is that one can check for multiple solutions, which actually occurred at k=6𝑘6k=6italic_k = 6. In Gurobi, an ad hoc search could be conducted by modifying the random number seed or using the concurrent optimizer, but discovery would not be guaranteed.

Figure 5(b) reports the time needed to implement the synthetic purposive sampling approach of Egami and Lee (2024). This was done by using the spsR package with the option to use the Gurobi solver on the background. We consider it to be fast, taking less than 40404040 seconds for k=10𝑘10k=10italic_k = 10. However, the synthetic purposive sampling problem can be formulated as a quadratic mixed integer program, while k𝑘kitalic_k-median is linear. This is evident from the graph; the k𝑘kitalic_k-median problem is solved almost instantly for every k𝑘kitalic_k up to 10101010.

Refer to caption
(a) Integer Program vs. Enumeration
Refer to caption
(b) Integer Program vs. SPS
Figure 5: Time Needed to Solve the k𝑘kitalic_k-median Problem k{1,10}𝑘110k\in\{1,\ldots 10\}italic_k ∈ { 1 , … 10 }

Notes: Time comparison of different purposive sampling approaches. Time (vertical axis) is in minutes. The dark, blue line with circles represents the time it takes to solve the integer program in Section 4 using the MIP solver in Gurobi to provable optimality (Gurobi gives the solution in less than a second). The light, blue line with squares in Panel a) represents the time it takes to solve the k𝑘kitalic_k-median problem using brute-force enumeration. The red line with triangles in Panel b) represents the time needed to implement the synthetic purposive sampling approach using the spsR package.

5.2 Multi-Country Survey Experiments in Europe

Our second application revisits a multi-country survey experiment originally conducted and analyzed in Naumann et al. (2018) and discussed in Egami and Lee (2024). The question of interest is whether native-born inhabitants of a particular country are more supportive of immigration depending on whether the potential migrants are high-skilled or low-skilled. Naumann et al. (2018) carried out a survey experiment in 15151515 European countries: Austria, Belgium, Czech Republic, Denmark, Finland, France, Germany, Ireland, Netherlands, Norway, Slovenia, Spain, Sweden, Switzerland, and the United Kingdom. Respondents were native-born individuals and were randomly assigned to report their attitudes towards either high-skilled (“treatment”) or low-skilled (“control”) immigrants.

While the experiments have already been conducted and outcomes of each experiment are available for all countries, we consider the situation of a researcher that can only conduct k=6𝑘6k=6italic_k = 6 experiments; here the choice of k=6𝑘6k=6italic_k = 6 mirrors Egami and Lee (2024). We let all 15151515 countries be both potential experimental and policy-relevant sites. That is, 𝒮E=𝒮Psubscript𝒮𝐸subscript𝒮𝑃\mathcal{S}_{E}=\mathcal{S}_{P}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT, and card(𝒮Esubscript𝒮𝐸\mathcal{S}_{E}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT) = card(𝒮Psubscript𝒮𝑃\mathcal{S}_{P}caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT) = 15. 161616A potential situation we have in mind is one of a researcher who would like to give policy advice to the policy makers in these countries on whether to initiate an immigration reform that could favor either high-skilled or low-skilled immigrants. The researcher knows that policy makers are interested in voters’ attitudes towards these types of reforms. We assume that the researcher is only able to experiment in a subset of countries due to administrative or budget constraints, and that he/she needs to extrapolate voters’ attitudes of the other policy-relevant sites based on the experimental estimates. The researcher needs to decide whether to recommend the implementation of an immigration reform directed to either high-skilled or low-skilled immigrants.

The only data needed to solve the k𝑘kitalic_k-median problem are site-level covariates of both experimental and policy-relevant sites. We use the same covariates as Egami and Lee (2024).171717These are: Gross Domestic Product (GDP), size of migrant population, unemployment rate, proportion of females, mean age, mean education, baseline level of support for immigration by the general public, and a categorical variable that indicates the subregions in Europe (i.e., South, North, East, and West). The covariate data can be accessed in the open-source software R package, spsR. They are in different scales: For example, GDP is measured in 2015 U.S. dollars, while the unemployment rate is reported in percentage points. As is commonly recommended in clustering problems and also done in Egami and Lee (2024), we standardize all of them.

When k=6𝑘6k=6italic_k = 6, the k𝑘kitalic_k-median problem is solved by the Czech Republic, Denmark, France, Ireland, Spain, and Switzerland. Figure 6 visualizes the distribution of standardized covariates, along with the sites selected by both the k𝑘kitalic_k-median and the synthetic purposive sampling approach. Four of the six selected sites are common to both approaches (Czech Republic, Denmark, Spain, Switzerland), but synthetic purposive sampling chooses Germany and the Netherlands instead of France and Ireland.

Refer to caption
(a) k𝑘kitalic_k-median
Refer to caption
(b) Egami and Lee (2024)
Figure 6: Distribution of Site-level Covariates in The Multi-Country Survey Experiment (k=6)𝑘6(k=6)( italic_k = 6 )

Notes: Box plots showing the distribution of covariates among the fifteen countries. The box plots are constructed as explained in Figure 2. Each panel depicts the site selected by the different approaches when k=6𝑘6k=6italic_k = 6. Solid shapes indicate sites that are in both the solution of the k𝑘kitalic_k-median problem and the synthetic purposive sampling approach. Hollow shapes indicate solutions that differ across methods.

Figure 7 visualizes the connection networks induced by the different solutions. In both panels, red circles represent countries that are selected for experimentation. The gray lines in Figure 7 indicate the connections between the experimental sites and the policy-relevant sites that were not selected for experimentation (blue circles). For example, we can see that each policy-relevant site is connected to exactly one experimental site, i.e. its nearest neighbor; for example, the United Kingdom uses only the information of France. The connection network in Figure 7(b) is considerably more dense, with all policy-relevant sites connected to more than one experimental site. For instance, the synthetic experiment for United Kingdom assigns positive weights to the Czech Republic, Netherlands, and Germany. To further aid the visual interpretation of the connection network, we color each connection differently to capture its strength. For example, for the United Kingdom, the strongest connection is to the Netherlands (0.74)0.74(0.74)( 0.74 ), whereas the weakest connection of the United Kingdom is to the Czech Republic (0.01)0.01(0.01)( 0.01 ).

Figure 7(a) also shows that three selected countries in the k𝑘kitalic_k-median problem (Switzerland, Czech Republic, and Spain) are not connected to any of the other policy-relevant sites. This suggests that these countries are selected because no other country provides a close enough match for them. In contrast, in Figure 7(b), these countries (Switzerland, Czech Republic, and Spain) receive positive weights from at least five other countries.

Refer to caption
(a) k𝑘kitalic_k-median

Refer to caption

(b) Egami and Lee (2024)
Figure 7: Connection Network for k=6𝑘6k=6italic_k = 6

Notes: In Panel (7(a)), a connection between a blue dot and a red dot indicates that the corresponding experimental site is the nearest neighbor of the policy site. In Panel (7(b)), each blue dot may be connected to one or multiple red dots, indicating that the corresponding policy site uses the weighted average of one or more experimental sites to construct its synthetic control. The width and the transparency of the connection line indicate the weight that each policy site puts on the connecting experimental site.

Approximation Error in Theorem 1: Because experimental estimates and corresponding standard errors are available for all 15151515 countries, we can compute the approximation error in Theorem 1. Define the relative approximation error of the k𝑘kitalic_k-median approximation to minimax regret as

inf𝒮𝒜(k)(infT𝒯𝒮1/2supτLipC(d)(T,𝒮,τ))C21#𝒮Pinf𝒮𝒜(k)(s𝒮P\𝒮XsXN𝒮(s)).subscriptinfimum𝒮𝒜𝑘subscriptinfimum𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏𝐶21#subscript𝒮𝑃subscriptinfimum𝒮𝒜𝑘subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\frac{\inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\inf_{T\in\mathcal{T}^{1/2}_{% \mathscr{S}}}\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}(T,% \mathscr{S},\tau)\right)}{\frac{C}{2}\frac{1}{\#\mathcal{S}_{P}}\inf_{\mathscr% {S}\in\mathcal{A}(k)}\left(\sum_{s\in\mathcal{S}_{P}\backslash\mathscr{S}}\|X_% {s}-X_{N_{\mathscr{S}}(s)}\|\right)}.divide start_ARG roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ) end_ARG start_ARG divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) end_ARG . (23)

Algebraic manipulations shows that this expression is bounded from below by

max{0,1BσEmin{#(𝒮E𝒮P),k}#𝒮PC21#𝒮Pinf𝒮𝒜(k)(s𝒮P\𝒮XsXN𝒮(s))},01𝐵subscript𝜎𝐸#subscript𝒮𝐸subscript𝒮𝑃𝑘#subscript𝒮𝑃𝐶21#subscript𝒮𝑃subscriptinfimum𝒮𝒜𝑘subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\max\left\{0,1-\frac{B\cdot\sigma_{E}\cdot\frac{\min\{\#\left(\mathcal{S}_{E}% \cap\mathcal{S}_{P}\right),k\}}{\#\mathcal{S}_{P}}}{\frac{C}{2}\frac{1}{\#% \mathcal{S}_{P}}\inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\sum_{s\in\mathcal{S}% _{P}\backslash\mathscr{S}}\|X_{s}-X_{N_{\mathscr{S}}(s)}\|\right)}\right\},roman_max { 0 , 1 - divide start_ARG italic_B ⋅ italic_σ start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ⋅ divide start_ARG roman_min { # ( caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) , italic_k } end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG end_ARG start_ARG divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) end_ARG } , (24)

and from above by

1+BσEmin{#(𝒮E𝒮P),k}#𝒮PC21#𝒮Pinf𝒮𝒜(k)(s𝒮P\𝒮XsXN𝒮(s)),1𝐵subscript𝜎𝐸#subscript𝒮𝐸subscript𝒮𝑃𝑘#subscript𝒮𝑃𝐶21#subscript𝒮𝑃subscriptinfimum𝒮𝒜𝑘subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠1+\frac{B\cdot\sigma_{E}\cdot\frac{\min\{\#\left(\mathcal{S}_{E}\cap\mathcal{S% }_{P}\right),k\}}{\#\mathcal{S}_{P}}}{\frac{C}{2}\frac{1}{\#\mathcal{S}_{P}}% \inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\sum_{s\in\mathcal{S}_{P}\backslash% \mathscr{S}}\|X_{s}-X_{N_{\mathscr{S}}(s)}\|\right)},1 + divide start_ARG italic_B ⋅ italic_σ start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ⋅ divide start_ARG roman_min { # ( caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) , italic_k } end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG end_ARG start_ARG divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) end_ARG , (25)

where Bargmaxz0zΦ(z)𝐵subscript𝑧0𝑧Φ𝑧B\equiv\arg\max_{z\geq 0}z\Phi(-z)italic_B ≡ roman_arg roman_max start_POSTSUBSCRIPT italic_z ≥ 0 end_POSTSUBSCRIPT italic_z roman_Φ ( - italic_z ). The closer the lower and upper bounds are, the better the k𝑘kitalic_k-median approximation is.

Figure 8 displays these bounds for values of k{1,,10}𝑘110k\in\{1,\ldots,10\}italic_k ∈ { 1 , … , 10 }. As k𝑘kitalic_k increases, the approximation becomes worse; as mentioned before, this is driven by ignoring the experimental sites themselves in bounding regret. In this application, the k𝑘kitalic_k-median solution still works relatively well when k=6𝑘6k=6italic_k = 6, with a relative error of about ±13%plus-or-minuspercent13\pm 13\%± 13 %. As k𝑘kitalic_k increases to 10101010, the relative error is about ±50%plus-or-minuspercent50\pm 50\%± 50 %.181818For these computations, we picked C𝐶Citalic_C to be the smallest Lipschitz constant needed to capture the heterogeneity of estimated treatment effects in the data. That is, we pick C𝐶Citalic_C as maxs,s𝒮E𝒮P|τ^sτ^s|XsXs,subscript𝑠superscript𝑠subscript𝒮𝐸subscript𝒮𝑃subscript^𝜏𝑠subscript^𝜏superscript𝑠normsubscript𝑋𝑠subscript𝑋superscript𝑠\max_{s,s^{\prime}\in\mathcal{S}_{E}\cup\mathcal{S}_{P}}\frac{|\hat{\tau}_{s}-% \hat{\tau}_{s^{\prime}}|}{\|X_{s}-X_{s^{\prime}}\|},roman_max start_POSTSUBSCRIPT italic_s , italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ∪ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG | over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | end_ARG start_ARG ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ∥ end_ARG , (26) where τ^ssubscript^𝜏𝑠\hat{\tau}_{s}over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT denote the estimated treatment effects for site s𝑠sitalic_s. It turns out that this C𝐶Citalic_C is comfortably “large” in the sense of Theorem 1. We provide additional details in Appendix B.3.

Refer to caption
Figure 8: Approximation Error of The k𝑘kitalic_k-median Solutions

Notes: Approximation error of the k𝑘kitalic_k-median problem in terms of lower and upper bound on the fraction of the true minimax-regret solution over the k𝑘kitalic_k-median solution; cf. (23). The Lipschitz constant is chosen to be the smallest Lipschitz constant that is needed to explain the data; cf. (26).

6 Extensions

6.1 Fixed Costs of Experimentation

We next allow for the possibility that running an experiment in a site s𝒮E𝑠subscript𝒮𝐸s\in\mathcal{S}_{E}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT has a fixed cost cssubscript𝑐𝑠c_{s}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. This means that the welfare of a decision rule T𝑇Titalic_T, given that sites 𝒮𝒮\mathscr{S}script_S are selected for experimentation, corresponds to

𝒲c(T,𝒮,τ):=1#𝒮P(s𝒮Pτ(Xs)𝔼τ𝒮[Ts(τ^𝒮)]s𝒮cs).assignsubscript𝒲𝑐𝑇𝒮𝜏1#subscript𝒮𝑃subscript𝑠subscript𝒮𝑃𝜏subscript𝑋𝑠subscript𝔼subscript𝜏𝒮delimited-[]subscript𝑇𝑠subscript^𝜏𝒮subscript𝑠𝒮subscript𝑐𝑠\mathcal{W}_{c}(T,\mathscr{S},\tau):=\frac{1}{\#\mathcal{S}_{P}}\left(\sum_{s% \in\mathcal{S}_{P}}\tau(X_{s})\mathbb{E}_{\tau_{\mathscr{S}}}[T_{s}(\widehat{% \tau}_{\mathscr{S}})]-\sum_{s\in\mathscr{S}}c_{s}\right).caligraphic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_T , script_S , italic_τ ) := divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ( ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) blackboard_E start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] - ∑ start_POSTSUBSCRIPT italic_s ∈ script_S end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) . (27)

Based on this welfare function, the oracle action for the policy maker is to implement the policy in any policy-relevant site s𝑠sitalic_s for which τ(Xs)0𝜏subscript𝑋𝑠0\tau(X_{s})\geq 0italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ≥ 0. Expected regret of (T,𝒮)𝑇𝒮(T,\mathscr{S})( italic_T , script_S ) becomes

c(T,𝒮,τ):=1#𝒮Ps𝒮cs+1#𝒮Ps𝒮Pτ(Xs)(𝟏{τ(Xs)0}𝔼τ𝒮[Ts(τ^𝒮)]).assignsubscript𝑐𝑇𝒮𝜏1#subscript𝒮𝑃subscript𝑠𝒮subscript𝑐𝑠1#subscript𝒮𝑃subscript𝑠subscript𝒮𝑃𝜏subscript𝑋𝑠1𝜏subscript𝑋𝑠0subscript𝔼subscript𝜏𝒮delimited-[]subscript𝑇𝑠subscript^𝜏𝒮\mathcal{R}_{c}(T,\mathscr{S},\tau):=\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in% \mathscr{S}}c_{s}+\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal{S}_{P}}\tau(X_% {s})\left(\mathbf{1}\{\tau(X_{s})\geq 0\}-\mathbb{E}_{\tau_{\mathscr{S}}}[T_{s% }(\widehat{\tau}_{\mathscr{S}})]\right).caligraphic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_T , script_S , italic_τ ) := divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ script_S end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ( bold_1 { italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ≥ 0 } - blackboard_E start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] ) . (28)

Under the assumptions of Theorem 1, it is possible to show (by extending Lemma 1 and 2 to account for the fixed costs of experimentation) that

inf𝒮𝒜(k)(infT𝒯𝒮1/2supτLipC(d)c(T,𝒮,τ))subscriptinfimum𝒮𝒜𝑘subscriptinfimum𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑subscript𝑐𝑇𝒮𝜏\inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\inf_{T\in\mathcal{T}^{1/2}_{\mathscr% {S}}}\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}_{c}(T,\mathscr{% S},\tau)\right)roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_T , script_S , italic_τ ) )

can be approximated by (C/2#𝒮P)𝐶2#subscript𝒮𝑃(C/2\#\mathcal{S}_{P})( italic_C / 2 # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) times

inf𝒮𝒜(k)(s𝒮(2C)cs+s𝒮P\𝒮XsXN𝒮(s)),subscriptinfimum𝒮𝒜𝑘subscript𝑠𝒮2𝐶subscript𝑐𝑠subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\sum_{s\in\mathscr{S}}\left(\frac{2}{% C}\right)c_{s}+\sum_{s\in\mathcal{S}_{P}\backslash\mathscr{S}}\|X_{s}-X_{N_{% \mathscr{S}}(s)}\|\right),roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( ∑ start_POSTSUBSCRIPT italic_s ∈ script_S end_POSTSUBSCRIPT ( divide start_ARG 2 end_ARG start_ARG italic_C end_ARG ) italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) , (29)

and that the approximation error is the same as the one given in Theorem 1.

The problem in (29) is the metric uncapacitated k𝑘kitalic_k-facility location problem that was discussed in Remark 7. This is a common extension of the k𝑘kitalic_k-median problem where there is a fixed cost of opening each facility. The problem can also be formulated as a linear integer program, namely

min{yi,xi,j}i𝒮E,j𝒮Psubscriptsubscriptsubscript𝑦𝑖subscript𝑥𝑖𝑗formulae-sequence𝑖subscript𝒮𝐸𝑗subscript𝒮𝑃\displaystyle\min_{\{y_{i},x_{i,j}\}_{i\in\mathcal{S}_{E},j\in\mathcal{S}_{P}}}roman_min start_POSTSUBSCRIPT { italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT , italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_POSTSUBSCRIPT (i𝒮Eyici+i𝒮E,j𝒮Pxi,jc(j,i))subscript𝑖subscript𝒮𝐸subscript𝑦𝑖subscriptsuperscript𝑐𝑖subscriptformulae-sequence𝑖subscript𝒮𝐸𝑗subscript𝒮𝑃subscript𝑥𝑖𝑗𝑐𝑗𝑖\displaystyle\left(\sum_{i\in\mathcal{S}_{E}}y_{i}c^{*}_{i}+\sum_{i\in\mathcal% {S}_{E},j\in\mathcal{S}_{P}}x_{i,j}\cdot c(j,i)\right)( ∑ start_POSTSUBSCRIPT italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT , italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ⋅ italic_c ( italic_j , italic_i ) ) (30)
such that i𝒮Exi,j=1,j𝒮P,formulae-sequencesubscript𝑖subscript𝒮𝐸subscript𝑥𝑖𝑗1for-all𝑗subscript𝒮𝑃\displaystyle\sum_{i\in\mathcal{S}_{E}}x_{i,j}=1,\;\;\;\;\;\forall j\in% \mathcal{S}_{P},∑ start_POSTSUBSCRIPT italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT = 1 , ∀ italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT , (31)
i𝒮Eyik,subscript𝑖subscript𝒮𝐸subscript𝑦𝑖𝑘\displaystyle\sum_{i\in\mathcal{S}_{E}}y_{i}\leq k,∑ start_POSTSUBSCRIPT italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≤ italic_k , (32)
0xi,jyi,i𝒮E,j𝒮P,formulae-sequence0subscript𝑥𝑖𝑗subscript𝑦𝑖formulae-sequence𝑖subscript𝒮𝐸𝑗subscript𝒮𝑃\displaystyle 0\leq x_{i,j}\leq y_{i},\;\;\;\;\;i\in\mathcal{S}_{E},j\in% \mathcal{S}_{P},0 ≤ italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ≤ italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT , italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT , (33)
yi{0,1},xi,j{0,1},i𝒮E,j𝒮P.formulae-sequencesubscript𝑦𝑖01formulae-sequencesubscript𝑥𝑖𝑗01formulae-sequence𝑖subscript𝒮𝐸𝑗subscript𝒮𝑃\displaystyle y_{i}\in\{0,1\},x_{i,j}\in\{0,1\},\;\;\;\;\;i\in\mathcal{S}_{E},% j\in\mathcal{S}_{P}.italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∈ { 0 , 1 } , italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ∈ { 0 , 1 } , italic_i ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT , italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT . (34)

Just as before, the choice variable yisubscript𝑦𝑖y_{i}italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT indicates whether facility i𝑖iitalic_i is open, and xi,jsubscript𝑥𝑖𝑗x_{i,j}italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT indicates whether client j𝑗jitalic_j is assigned to facility i𝑖iitalic_i. Constraints are imposed to ensure that each client j𝑗jitalic_j is assigned to at least one facility, that there are no more than k𝑘kitalic_k facilities in total, and that a client can only be assigned to an open facility. The connection cost between a facility i𝑖iitalic_i and a client j𝑗jitalic_j is still given by XiXjnormsubscript𝑋𝑖subscript𝑋𝑗\|X_{i}-X_{j}\|∥ italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ∥, but now there is a fixed cost of opening a facility i𝑖iitalic_i given by ci2ci/Csubscriptsuperscript𝑐𝑖2subscript𝑐𝑖𝐶c^{*}_{i}\equiv 2c_{i}/Citalic_c start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≡ 2 italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_C. This means that—in contrast to the case in which there is no fixed cost—choosing the optimal sites to maximize external validity now requires knowledge of the Lipschitz constant C𝐶Citalic_C (as this constant appears explicitly on the fixed cost cisubscriptsuperscript𝑐𝑖c^{*}_{i}italic_c start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT). This is simply because, as we have shown, the scale of regret absent costs of experimentation depends on C𝐶Citalic_C. Consequently, every time a new site is considered for experimentation, there is a trade-off between its contribution to reduce regret and the fixed cost of experimentation. We note that when C𝐶Citalic_C is large, the fixed costs in the objective function become negligible, and the solution of the uncapacitated facility location problem can be approximated by the solution of the k𝑘kitalic_k-median problem.

6.2 Random Selection of Experimental Sites

So far, our analysis focused on minimizing the worst-case (welfare-based) regret among all purposive sampling schemes that select at most k𝑘kitalic_k sites. That is, we excluded randomized (including nonuniformly randomized) sampling schemes. This is an important limitation—a widespread view among experimenters is that “the external validity of randomized evaluations for a given population (say, the population of a country) would be maximized by randomly selecting sites and, within these sites, by randomly selecting treatment and comparison groups” (Duflo et al., 2007, p. 3953).

We now extend our baseline framework (which excludes fixed costs of experimentation) to allow for randomized site selection. First, we note that whether randomization is potentially desired delicately depends on the decision-theoretic setup. For example, the optimal sampling scheme in the Bayesian setting of Gechter et al. (2024) will typically be purposive. Whether randomization improves minimax regret depends on how exactly the decision problem is formulated, which can be related to what we refer to as the timing assumptions in an implicit game that the decision maker plays against a malicious “nature”. We will first clarify this observation and then provide a brief illustration of randomized solutions.

To formalize the discussion, let M𝑀Mitalic_M denote the cardinality of 𝒜(k)𝒜𝑘\mathcal{A}(k)caligraphic_A ( italic_k ). Let Δ(𝒜(k))Δ𝒜𝑘\Delta\left(\mathcal{A}(k)\right)roman_Δ ( caligraphic_A ( italic_k ) ) denote the set of all probability distributions over the M𝑀Mitalic_M elements of 𝒜(k)𝒜𝑘\mathcal{A}(k)caligraphic_A ( italic_k ). We define a randomized site selection as a probability distribution p:=(p1,,pM)Δ(𝒜(k))assign𝑝subscript𝑝1subscript𝑝𝑀Δ𝒜𝑘p:=(p_{1},\ldots,p_{M})\in\Delta\left(\mathcal{A}(k)\right)italic_p := ( italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_p start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT ) ∈ roman_Δ ( caligraphic_A ( italic_k ) ). The econometrician will pick a subset from the experimental sites by drawing one realization of a distribution pΔ(𝒜(k))superscript𝑝Δ𝒜𝑘p^{*}\in\Delta\left(\mathcal{A}(k)\right)italic_p start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ∈ roman_Δ ( caligraphic_A ( italic_k ) ) that she specified. Denote the randomly selected sites by 𝒮superscript𝒮\mathscr{S}^{*}script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT.

In our setting, whether the decision maker will want to randomize crucially depends on timing assumptions; that is, the moment in the game in which nature can move to harm the decision maker. Consider first the case in which an adversarial nature may harm the policy maker by choosing τ𝜏\tauitalic_τ only after seeing the realization of 𝒮superscript𝒮\mathscr{S^{*}}script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT (and knowing 𝒯𝒮subscript𝒯superscript𝒮\mathcal{T}_{\mathscr{S}^{*}}caligraphic_T start_POSTSUBSCRIPT script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT). Then, the risk of using treatment rule T𝒯𝒮𝑇subscript𝒯superscript𝒮T\in\mathcal{T}_{\mathscr{S}^{*}}italic_T ∈ caligraphic_T start_POSTSUBSCRIPT script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT is

(T,𝒮,τ),𝑇superscript𝒮𝜏\mathcal{R}(T,\mathscr{S}^{*},\tau),caligraphic_R ( italic_T , script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_τ ) ,

and the worst-case payoff becomes

supτLipC(d)(T,𝒮,τ).subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇superscript𝒮𝜏\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S}^{*},% \tau).roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_τ ) .

The minimax problem faced by the econometrician after 𝒮superscript𝒮\mathscr{S}^{*}script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT has been realized is:

V(𝒮):=infT𝒯𝒮supτLipC(d)(T,𝒮,τ).assign𝑉superscript𝒮subscriptinfimum𝑇subscript𝒯superscript𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇superscript𝒮𝜏V(\mathscr{S}^{*}):=\inf_{T\in\mathcal{T}_{\mathscr{S}^{*}}}\sup_{\tau\in% \textrm{Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S}^{*},\tau).italic_V ( script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) := roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUBSCRIPT script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_τ ) .

With slight abuse of notation, let p(𝒮)𝑝superscript𝒮p(\mathscr{S}^{*})italic_p ( script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) denote the probability of choosing 𝒮superscript𝒮\mathscr{S}^{*}script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT at random under pΔ(𝒜(k))𝑝Δ𝒜𝑘p\in\Delta\left(\mathcal{A}(k)\right)italic_p ∈ roman_Δ ( caligraphic_A ( italic_k ) ). The (ex ante) expected payoff of any randomized site selection is

𝒮𝒜(k)p(𝒮)V(𝒮),subscriptsuperscript𝒮𝒜𝑘𝑝superscript𝒮𝑉superscript𝒮\sum_{\mathscr{S}^{*}\in\mathcal{A}(k)}p(\mathscr{S}^{*})\cdot V(\mathscr{S}^{% *}),∑ start_POSTSUBSCRIPT script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT italic_p ( script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ⋅ italic_V ( script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ,

and the optimal randomized site selection solves

infpΔ(𝒜(k))(𝒮𝒜(k)p(𝒮)V(𝒮)).subscriptinfimum𝑝Δ𝒜𝑘subscriptsuperscript𝒮𝒜𝑘𝑝superscript𝒮𝑉superscript𝒮\inf_{p\in\Delta(\mathcal{A}(k))}\left(\sum_{\mathscr{S}^{*}\in\mathcal{A}(k)}% p(\mathscr{S}^{*})\cdot V(\mathscr{S}^{*})\right).roman_inf start_POSTSUBSCRIPT italic_p ∈ roman_Δ ( caligraphic_A ( italic_k ) ) end_POSTSUBSCRIPT ( ∑ start_POSTSUBSCRIPT script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT italic_p ( script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ⋅ italic_V ( script_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ) .

But this problem is solved by any p𝑝pitalic_p supported on argmin𝒮V(𝒮)subscript𝒮𝑉𝒮\arg\min_{\mathscr{S}}V(\mathscr{S})roman_arg roman_min start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT italic_V ( script_S ), the set of purposive sampling schemes that solve the site selection problem. If that problem’s solution is unique, the policy maker will never randomize under this timing of the game. Furthermore, this “sequential” timing may feel natural in applications that we have in mind.

That said, the timing that seems more in line with Wald’s (1950) original application of the minimax principle is likely one in which the policy maker commits to both a randomized sampling scheme and a set of contingent (on sampling scheme) decision rules and nature adversarially picks τ𝜏\tauitalic_τ before any randomization was realized. We next briefly discuss this possibility.

To see that randomization might strictly speaking be optimal, consider a stylized example where k=1𝑘1k=1italic_k = 1 and the covariates of each site are equal to its index: SE={1,4}subscript𝑆𝐸14S_{E}=\{1,4\}italic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = { 1 , 4 }, and SP={2,3}subscript𝑆𝑃23S_{P}=\{2,3\}italic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = { 2 , 3 }. For simplicity, suppose furthermore that τ^s=τssubscript^𝜏𝑠subscript𝜏𝑠\hat{\tau}_{s}=\tau_{s}over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, i.e. there is no sampling uncertainty in the treatment effects. This example can be solved for those combinations of sampling scheme and treatment assignment rule that achieve exact MMR. As we formally show in Appendix B.2, the exact MMR attainable by purposive sampling equals 3C/43𝐶43C/43 italic_C / 4, whereas the exact MMR with randomized sampling equals C/2𝐶2C/2italic_C / 2.191919The assumption of perfect signals is for simplicity. The example is rigged to resemble cases analyzed in Montiel Olea et al. (2023) and Stoye (2012) and we would accordingly be able to generalize it, but for our present purpose, solving for arbitrary sampling variances would only add tedium. Thus, in principle there can be a gain to randomized sampling.

To see that solving this problem can quickly become very hard, consider now the same example, except that the experimental sites coincide with the policy sites at {1,2}12\{1,2\}{ 1 , 2 }. Then we can find the MMR optimal combination of purposive sampling scheme and treatment assignment rule, and we can also verify that randomized site selection will strictly reduce worst-case regret. However, we are unable to characterize the exact solution for this, still extremely structured, case. We leave further exploration of randomized site selection, including the potentially fruitful investigation of “good enough” (in terms of regret) such schemes, for future research.

6.3 Other Restrictions on Treatment Heterogeneity

With Assumption 1, the effect of the policy of interest is a Lipschitz function of observed covariates with respect to the Euclidean distance. This means that the effect of the policy of interest in two sites with the same covariates is assumed to be the same (in other words, there is no unobserved treatment effect heterogeneity). In this section, we show that our main result can also accommodate other general measures of the distance between covariates (thus allowing for other functional classes), as well as some forms of unobserved treatment effect heterogeneity (provided 𝒮E𝒮P=subscript𝒮𝐸subscript𝒮𝑃\mathcal{S}_{E}\cap\mathcal{S}_{P}=\emptysetcaligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = ∅).

6.3.1 Other Distance Measure Based on Observed Covariates

To see how our results can allow for other types of distances, let m(x,x):d×d+:𝑚𝑥superscript𝑥superscript𝑑superscript𝑑subscriptm(x,x^{\prime}):\mathbb{R}^{d}\times\mathbb{R}^{d}\rightarrow\mathbb{R}_{+}italic_m ( italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) : blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT × blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT → blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT be a measure of distance (metric) between x,xd𝑥superscript𝑥superscript𝑑x,x^{\prime}\in\mathbb{R}^{d}italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT.202020That is, we assume that m(x,x)>0𝑚𝑥superscript𝑥0m(x,x^{\prime})>0italic_m ( italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) > 0 for any xx𝑥superscript𝑥x\neq x^{\prime}italic_x ≠ italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, and that for any x𝑥xitalic_x we have m(x,x)=0𝑚𝑥𝑥0m(x,x)=0italic_m ( italic_x , italic_x ) = 0. We also assume that the function is symmetric, in that m(x,x)=m(x,x)𝑚𝑥superscript𝑥𝑚superscript𝑥𝑥m(x,x^{\prime})=m(x^{\prime},x)italic_m ( italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = italic_m ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x ). And finally, we assume that m𝑚mitalic_m satisfies the triangle inequality m(x,x)m(x,x′′)+m(x,x′′)𝑚𝑥superscript𝑥𝑚𝑥superscript𝑥′′𝑚superscript𝑥superscript𝑥′′m(x,x^{\prime})\leq m(x,x^{\prime\prime})+m(x^{\prime},x^{\prime\prime})italic_m ( italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ≤ italic_m ( italic_x , italic_x start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ) + italic_m ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ) for any x,x,x′′𝑥superscript𝑥superscript𝑥′′x,x^{\prime},x^{\prime\prime}italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_x start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT. Also see Dudley (2002, p.20). Then, Assumption 1 may be modified as:

Assumption 3.

τ𝜏\tauitalic_τ is a Lipschitz function (with respect to metric m(,)𝑚m(\cdot,\cdot)italic_m ( ⋅ , ⋅ )) with known constant C𝐶Citalic_C. That is, for any x,xd𝑥superscript𝑥superscript𝑑x,x^{\prime}\in\mathbb{R}^{d}italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, |τ(x)τ(x)|Cm(x,x)𝜏𝑥𝜏superscript𝑥𝐶𝑚𝑥superscript𝑥|\tau(x)-\tau(x^{\prime})|\leq Cm(x,x^{\prime})| italic_τ ( italic_x ) - italic_τ ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | ≤ italic_C italic_m ( italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ).

For example, m(x,x)=((xx)(xx))1/2𝑚𝑥superscript𝑥superscriptsuperscript𝑥superscript𝑥top𝑥superscript𝑥12m(x,x^{\prime})=\left(\left(x-x^{\prime}\right)^{\top}\left(x-x^{\prime}\right% )\right)^{1/2}italic_m ( italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = ( ( italic_x - italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT ( italic_x - italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT is the Euclidean distance in Assumption 1 and m(x,x)=((xx)W(xx))1/2𝑚𝑥superscript𝑥superscriptsuperscript𝑥superscript𝑥top𝑊𝑥superscript𝑥12m(x,x^{\prime})=\left(\left(x-x^{\prime}\right)^{\top}W\left(x-x^{\prime}% \right)\right)^{1/2}italic_m ( italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = ( ( italic_x - italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT italic_W ( italic_x - italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT for some positive definite matrix W𝑊Witalic_W is a W𝑊Witalic_W-weighted Euclidean distance. Choosing m(x,x)=[m~(x,x)]α𝑚𝑥superscript𝑥superscriptdelimited-[]~𝑚𝑥superscript𝑥𝛼m(x,x^{\prime})=\left[\tilde{m}(x,x^{\prime})\right]^{\alpha}italic_m ( italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = [ over~ start_ARG italic_m end_ARG ( italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT for some α(0,1)𝛼01\alpha\in(0,1)italic_α ∈ ( 0 , 1 ) and some distance measure m~(,)~𝑚\tilde{m}(\cdot,\cdot)over~ start_ARG italic_m end_ARG ( ⋅ , ⋅ ) also effectively allows us to model τ𝜏\tauitalic_τ as a Hölder continuous function of order α𝛼\alphaitalic_α (Dudley, 2002, p. 56). With Assumptions 2 and 3, N𝒮(s)subscript𝑁𝒮𝑠N_{\mathscr{S}}(s)italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) is understood to be a nearest neighbor in 𝒮𝒮\mathscr{S}script_S measured in terms of metric m(,)𝑚m(\cdot,\cdot)italic_m ( ⋅ , ⋅ ) that appears in Assumption 3. Moreover, LipC(d)subscriptLip𝐶superscript𝑑\text{Lip}_{C}(\mathbb{R}^{d})Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) now stands for the space of all Lipschitz functions from dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT to \mathbb{R}blackboard_R with constant C𝐶Citalic_C, but in terms of metric m(,)𝑚m(\cdot,\cdot)italic_m ( ⋅ , ⋅ ). We note that we are slightly abusing notation, because we are not using m𝑚mitalic_m to index the nearest-neighbor function N𝒮(s)subscript𝑁𝒮𝑠N_{\mathscr{S}}(s)italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) or the Lipschitz functional class LipC(d)subscriptLip𝐶superscript𝑑\text{Lip}_{C}(\mathbb{R}^{d})Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ), although their respective definitions explicitly on the choice of metric m(,)𝑚m(\cdot,\cdot)italic_m ( ⋅ , ⋅ ). Under a general metric m(,)𝑚m(\cdot,\cdot)italic_m ( ⋅ , ⋅ ), the k𝑘kitalic_k-median problem is simply modified as:

Definition 3.

We say that a purposive sampling scheme, 𝒮𝒜(k)𝒮𝒜𝑘\mathscr{S}\in\mathcal{A}(k)script_S ∈ caligraphic_A ( italic_k ), solves the k𝑘kitalic_k-median problem with a metric cost function m(,)𝑚m(\cdot,\cdot)italic_m ( ⋅ , ⋅ ) if it solves

inf𝒮𝒜(k)s𝒮Pm(Xs,XN𝒮(s)).subscriptinfimum𝒮𝒜𝑘subscript𝑠subscript𝒮𝑃𝑚subscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\inf_{\mathscr{S}\in\mathcal{A}(k)}\sum_{s\in\mathcal{S}_{P}}m(X_{s},X_{N_{% \mathscr{S}}(s)}).roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_m ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ) . (35)

Since the proof of Theorem 1 does not rely on any specific property of the Euclidean distance, the results of Theorem 1 still hold under a general metric m(,)𝑚m(\cdot,\cdot)italic_m ( ⋅ , ⋅ ), by replacing the Euclidean distance \left\|\cdot-\cdot\right\|∥ ⋅ - ⋅ ∥ with m(,)𝑚m(\cdot,\cdot)italic_m ( ⋅ , ⋅ ).

6.3.2 Unobserved Treatment Heterogeneity

Now, we show how our assumptions can be modified to accommodate some forms of unobserved treatment heterogeneity. Instead of viewing policy effects as a function of observed covariates X𝑋Xitalic_X only, we now model the policy effects at the site level. Specifically, for each s𝒮𝑠𝒮s\in\mathcal{S}italic_s ∈ caligraphic_S, denote by τssubscript𝜏𝑠\tau_{s}\in\mathbb{R}italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∈ blackboard_R the policy effect in site s𝑠sitalic_s. Then, instead of viewing τ𝜏\tauitalic_τ as a function, we will let τ:=(τ1,τ2,,τS)Sassign𝜏superscriptsubscript𝜏1subscript𝜏2subscript𝜏𝑆topsuperscript𝑆\tau:=(\tau_{1},\tau_{2},...,\tau_{S})^{\top}\in\mathbb{R}^{S}italic_τ := ( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_τ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT be a vector of dimension S𝑆Sitalic_S that represents the policy effects of all sites in 𝒮𝒮\mathcal{S}caligraphic_S. Then, Assumptions 1 and 2 can be replaced with the following:

Assumption 4.

For any s,s𝒮𝑠superscript𝑠𝒮s,s^{\prime}\in\mathcal{S}italic_s , italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ caligraphic_S, ss𝑠superscript𝑠s\neq s^{\prime}italic_s ≠ italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, τ𝜏\tauitalic_τ satisfies:

|τsτs|Cm(Xs,Xs)+c,subscript𝜏𝑠subscript𝜏superscript𝑠𝐶𝑚subscript𝑋𝑠subscript𝑋superscript𝑠𝑐|\tau_{s}-\tau_{s^{\prime}}|\leq Cm(X_{s},X_{s^{\prime}})+c,| italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | ≤ italic_C italic_m ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) + italic_c , (36)

where C>0𝐶0C>0italic_C > 0 and c>0𝑐0c>0italic_c > 0 are both known, and m(,)𝑚m(\cdot,\cdot)italic_m ( ⋅ , ⋅ ) is a metric.

Assumption 4 allows sites with the same covariates to have different policy effects. The difference, however, is assumed to be at most c𝑐citalic_c. We motivate Assumption 4 using a simple linear regression model in Appendix B.4. Denote by C,cSsubscript𝐶𝑐superscript𝑆\mathcal{F}_{C,c}\subset\mathbb{R}^{S}caligraphic_F start_POSTSUBSCRIPT italic_C , italic_c end_POSTSUBSCRIPT ⊂ blackboard_R start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT the collection of all vectors of dimension Ssuperscript𝑆\mathbb{R}^{S}blackboard_R start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT satisfying (36). For this parameter class, the minimax-regret optimal purposive sampling scheme and treatment rule are redefined to solve

inf𝒮𝒜(k),T𝒯𝒮1/2supτC,c(T,𝒮,τ).subscriptinfimumformulae-sequence𝒮𝒜𝑘𝑇superscriptsubscript𝒯𝒮12subscriptsupremum𝜏subscript𝐶𝑐𝑇𝒮𝜏\inf_{\mathscr{S}\in\mathcal{A}(k),T\in\mathcal{T}_{\mathscr{S}}^{1/2}}\sup_{% \tau\in\mathcal{F}_{C,c}}\mathcal{R}(T,\mathscr{S},\tau).roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) , italic_T ∈ caligraphic_T start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ caligraphic_F start_POSTSUBSCRIPT italic_C , italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) . (37)

It turns out that the purposive sampling scheme that solves (37) can still be approximated by the solution of the k𝑘kitalic_k-median problem in Definition 3 with a different cost function.

To see this, note that the functional class C,csubscript𝐶𝑐\mathcal{F}_{C,c}caligraphic_F start_POSTSUBSCRIPT italic_C , italic_c end_POSTSUBSCRIPT is still convex and centrosymmetric. Thus, we can show that, for the evidence aggregation problem studied in Montiel Olea et al. (2023), the conclusion of their Proposition 1(iii) extends to C,csubscript𝐶𝑐\mathcal{F}_{C,c}caligraphic_F start_POSTSUBSCRIPT italic_C , italic_c end_POSTSUBSCRIPT as well. Then, under Assumption 4, Lemmas 1 and 2 continue to hold with minor modifications. In particular, when C𝐶Citalic_C is large enough (with a threshold that can be exactly characterized), we have that

infT𝒯𝒮1/2supτC,c(T,𝒮,τ)(B#𝒮Ps𝒮𝒮Pσs)+(1#𝒮Ps𝒮P\𝒮Cm(Xs,XN𝒮(s))+c2).subscriptinfimum𝑇superscriptsubscript𝒯𝒮12subscriptsupremum𝜏subscript𝐶𝑐𝑇𝒮𝜏𝐵#subscript𝒮𝑃subscript𝑠𝒮subscript𝒮𝑃subscript𝜎𝑠1#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮𝐶𝑚subscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠𝑐2\inf_{T\in\mathcal{T}_{\mathscr{S}}^{1/2}}\sup_{\tau\in\mathcal{F}_{C,c}}% \mathcal{R}(T,\mathscr{S},\tau)\leq\;\left(\frac{B}{\#\mathcal{S}_{P}}\sum_{s% \in\mathscr{S}\cap\mathcal{S}_{P}}\sigma_{s}\right)+\left(\frac{1}{\#\mathcal{% S}_{P}}\sum_{s\in\mathcal{S}_{P}\backslash\mathscr{S}}\frac{Cm(X_{s},X_{N_{% \mathscr{S}}(s)})+c}{2}\right).roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ caligraphic_F start_POSTSUBSCRIPT italic_C , italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ≤ ( divide start_ARG italic_B end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ script_S ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) + ( divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT divide start_ARG italic_C italic_m ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ) + italic_c end_ARG start_ARG 2 end_ARG ) .

In addition, the following lower bound can also be verified to hold

infT𝒯𝒮1/2supτC,c(T,𝒮,τ)(1#𝒮Ps𝒮P\𝒮Cm(Xs,XN𝒮(s))+c2).subscriptinfimum𝑇superscriptsubscript𝒯𝒮12subscriptsupremum𝜏subscript𝐶𝑐𝑇𝒮𝜏1#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮𝐶𝑚subscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠𝑐2\inf_{T\in\mathcal{T}_{\mathscr{S}}^{1/2}}\sup_{\tau\in\mathcal{F}_{C,c}}% \mathcal{R}(T,\mathscr{S},\tau)\geq\;\left(\frac{1}{\#\mathcal{S}_{P}}\sum_{s% \in\mathcal{S}_{P}\backslash\mathscr{S}}\frac{Cm(X_{s},X_{N_{\mathscr{S}}(s)})% +c}{2}\right).roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ caligraphic_F start_POSTSUBSCRIPT italic_C , italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ≥ ( divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT divide start_ARG italic_C italic_m ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ) + italic_c end_ARG start_ARG 2 end_ARG ) .

The two bounds above imply that, even when treatment effect heterogeneity is characterized by Assumption 4, the optimized value of the minimax problem (37) can still be uniformly approximated by the following objective function

inf𝒮𝒜(k)(1#𝒮Ps𝒮P\𝒮(Cm(Xs,XN𝒮(s))+c2)).subscriptinfimum𝒮𝒜𝑘1#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮𝐶𝑚subscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠𝑐2\inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in% \mathcal{S}_{P}\backslash\mathscr{S}}\left(\frac{Cm(X_{s},X_{N_{\mathscr{S}}(s% )})+c}{2}\right)\right).roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ( divide start_ARG italic_C italic_m ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ) + italic_c end_ARG start_ARG 2 end_ARG ) ) .

When 𝒮E𝒮P=subscript𝒮𝐸subscript𝒮𝑃\mathcal{S}_{E}\cap\mathcal{S}_{P}=\emptysetcaligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = ∅, the solution to the problem above is the same as

inf𝒮𝒜(k)(1#𝒮Ps𝒮P(Cm(Xs,XN𝒮(s))+c2)).subscriptinfimum𝒮𝒜𝑘1#subscript𝒮𝑃subscript𝑠subscript𝒮𝑃𝐶𝑚subscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠𝑐2\inf_{\mathscr{S}\in\mathcal{A}(k)}\left(\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in% \mathcal{S}_{P}}\left(\frac{Cm(X_{s},X_{N_{\mathscr{S}}(s)})+c}{2}\right)% \right).roman_inf start_POSTSUBSCRIPT script_S ∈ caligraphic_A ( italic_k ) end_POSTSUBSCRIPT ( divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( divide start_ARG italic_C italic_m ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ) + italic_c end_ARG start_ARG 2 end_ARG ) ) .

This problem is equivalent to the k𝑘kitalic_k-median problem in Definition 3, where the connection cost between sites s𝑠sitalic_s and ssuperscript𝑠s^{\prime}italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT is given by m(Xs,Xs)𝑚subscript𝑋𝑠subscript𝑋superscript𝑠m(X_{s},X_{s^{\prime}})italic_m ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , italic_X start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) (neither C𝐶Citalic_C nor c𝑐citalic_c are required as inputs).

7 Conclusion

This paper presented a decision-theoretic justification for viewing the question of how to best choose where to experiment in order to optimize external validity as a k𝑘kitalic_k-median (clustering) problem. More concretely, we presented conditions under which minimizing the worst-case, welfare-based regret among all purposive (nonrandomized) schemes that select k𝑘kitalic_k sites is approximately equal, and can be exactly equal, to finding the k𝑘kitalic_k most central vectors of baseline site-level covariates.

We believe there are many interesting directions for future work. For example, while we focused on purposive sampling schemes, it would be interesting to better understand the value of randomized sampling schemes and whether site-level covariates can be used to design such randomized selection with an eye on external validity. We also think that discussions around the relation between the k𝑘kitalic_k-median problem and synthetic purposive sampling of Egami and Lee (2024) open interesting research directions to provide a decision-theoretic justification for the use the synthetic control of Abadie et al. (2010). Finally, it would be interesting to explore the use of (mixed) integer programming techniques to solve for the purposive sampling schemes suggested in Egami and Lee (2024) and Gechter et al. (2024).

Appendix A Proofs of Main Results

A.1 Proof of Lemma 1

Fix the selected sites 𝒮𝒮\mathscr{S}script_S, and denote the cardinality of 𝒮𝒮\mathscr{S}script_S as card(𝒮)card𝒮\textrm{card}(\mathscr{S})card ( script_S ). Let 𝒮1<𝒮2<<𝒮card(𝒮)subscript𝒮1subscript𝒮2subscript𝒮card𝒮\mathscr{S}_{1}<\mathscr{S}_{2}<\ldots<\mathscr{S}_{\textrm{card}(\mathscr{S})}script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < script_S start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < … < script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT denote the indices of the card(𝒮)card𝒮\textrm{card}(\mathscr{S})card ( script_S ) experimental sites in 𝒮𝒮\mathscr{S}script_S. For a given experimental site s𝒮𝑠𝒮s\in\mathscr{S}italic_s ∈ script_S, let τ^ssubscript^𝜏𝑠\widehat{\mathcal{\tau}}_{s}over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT denote its corresponding treatment effect estimate. Let

τ^𝒮:=(τ^𝒮1,,τ^𝒮card(𝒮))assignsubscript^𝜏𝒮superscriptsubscript^𝜏subscript𝒮1subscript^𝜏subscript𝒮card𝒮top\widehat{\tau}_{\mathscr{S}}:=(\hat{\tau}_{\mathscr{S}_{1}},\ldots,\hat{\tau}_% {\mathscr{S}_{\textrm{card}(\mathscr{S})}})^{\top}over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT := ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , … , over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT

denote the vector containing the estimates for each experimental site.

For each policy-relevant site, s𝒮P𝑠subscript𝒮𝑃s\in\mathcal{S}_{P}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT, recall that N𝒮(s)subscript𝑁𝒮𝑠N_{\mathscr{S}}(s)italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) denotes its nearest neighbor among the sites 𝒮𝒮\mathscr{S}script_S (or the nearest neighbor with the smallest index in case of multiplicity). Partition the policy-relevant sites as

𝒮P=(𝒮𝒮P)(𝒮P\𝒮).subscript𝒮𝑃𝒮subscript𝒮𝑃\subscript𝒮𝑃𝒮\mathcal{S}_{P}=\left(\mathscr{S}\cap\mathcal{S}_{P}\right)\cup\left(\mathcal{% S}_{P}\backslash\mathscr{S}\right).caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = ( script_S ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) ∪ ( caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S ) .

Consider the decision rule, T𝒯𝒮1/2superscript𝑇subscriptsuperscript𝒯12𝒮T^{*}\in\mathcal{T}^{1/2}_{\mathscr{S}}italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT, that recommends, for each policy-relevant site s𝒮P𝑠subscript𝒮𝑃s\in\mathcal{S}_{P}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT and given data τ^𝒮subscript^𝜏𝒮\widehat{\tau}_{\mathscr{S}}over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT, the following action:

  1. i)

    For s𝒮P𝒮𝑠subscript𝒮𝑃𝒮s\in\mathcal{S}_{P}\cap\mathscr{S}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ∩ script_S,

    Ts(τ^𝒮):=𝟏{τ^s0}.assignsubscriptsuperscript𝑇𝑠subscript^𝜏𝒮1subscript^𝜏𝑠0T^{*}_{s}\left(\widehat{\tau}_{\mathscr{S}}\right):=\mathbf{1}\{\widehat{% \mathcal{\tau}}_{s}\geq 0\}.italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) := bold_1 { over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≥ 0 } .
  2. ii)

    For s𝒮P\𝒮𝑠\subscript𝒮𝑃𝒮s\in\mathcal{S}_{P}\backslash\mathscr{S}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S, set

    Ts(τ^𝒮):=Φ(τ^N𝒮(s)σ~s),assignsubscriptsuperscript𝑇𝑠subscript^𝜏𝒮Φsubscript^𝜏subscript𝑁𝒮𝑠subscript~𝜎𝑠T^{*}_{s}\left(\widehat{\tau}_{\mathscr{S}}\right):=\Phi\left(\frac{\widehat{% \tau}_{N_{\mathscr{S}}(s)}}{\tilde{\sigma}_{s}}\right),italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) := roman_Φ ( divide start_ARG over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT end_ARG start_ARG over~ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG ) ,

    where

    σ~s:=(CXsXN𝒮(s)π/2)2σs2.assignsubscript~𝜎𝑠superscript𝐶normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠𝜋22subscriptsuperscript𝜎2𝑠\tilde{\sigma}_{s}:=\sqrt{\left(\frac{C\|X_{s}-X_{N_{\mathscr{S}}(s)}\|}{\sqrt% {\pi/2}}\right)^{2}-\sigma^{2}_{s}}.over~ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT := square-root start_ARG ( divide start_ARG italic_C ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ end_ARG start_ARG square-root start_ARG italic_π / 2 end_ARG end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG .

Note that the expression in ii) above is well-defined for every s𝒮P𝑠subscript𝒮𝑃s\in\mathcal{S}_{P}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT when C𝐶Citalic_C is large enough. This decision rule is the minimax-regret optimal rule (provided C𝐶Citalic_C is large enough) for the evidence aggregation framework discussed in Yata (2021) and Montiel Olea et al. (2023), both of which build upon Stoye (2012), and is indeed in 𝒯𝒮1/2subscriptsuperscript𝒯12𝒮\mathcal{T}^{1/2}_{\mathscr{S}}caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT.

Define C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S ) to be any value of C𝐶Citalic_C for which σ~s>0subscript~𝜎𝑠0\tilde{\sigma}_{s}>0over~ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT > 0 for every s𝒮P𝑠subscript𝒮𝑃s\in\mathcal{S}_{P}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT. Note that by definition of infimum,

infT𝒯𝒮supτLipC(d)(T,𝒮,τ)subscriptinfimum𝑇subscript𝒯𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏\displaystyle\inf_{T\in\mathcal{T}_{\mathscr{S}}}\sup_{\tau\in\textrm{Lip}_{C}% (\mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},\tau)roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) \displaystyle\leq supτLipC(d)(T,𝒮,τ),subscriptsupremum𝜏subscriptLip𝐶superscript𝑑superscript𝑇𝒮𝜏\displaystyle\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}(T^{*},% \mathscr{S},\tau),roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , script_S , italic_τ ) ,
\displaystyle\leq 1#𝒮Ps𝒮PsupτLipC(d)(τ(Xs)(𝟏{τ(Xs)0}𝔼τ𝒮[Ts(τ^𝒮)])),1#subscript𝒮𝑃subscript𝑠subscript𝒮𝑃subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝜏subscript𝑋𝑠1𝜏subscript𝑋𝑠0subscript𝔼subscript𝜏𝒮delimited-[]subscriptsuperscript𝑇𝑠subscript^𝜏𝒮\displaystyle\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal{S}_{P}}\sup_{\tau% \in\textrm{Lip}_{C}(\mathbb{R}^{d})}\left(\tau(X_{s})\left(\mathbf{1}\{\tau(X_% {s})\geq 0\}-\mathbb{E}_{\tau_{\mathscr{S}}}[T^{*}_{s}(\widehat{\tau}_{% \mathscr{S}})]\right)\right),divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT ( italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ( bold_1 { italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ≥ 0 } - blackboard_E start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] ) ) ,

where the second inequality follows from the fact that

(T,𝒮,τ):=1#𝒮Ps𝒮Pτ(Xs)(𝟏{τ(Xs)0}𝔼τ𝒮[Ts(τ^𝒮)]).assign𝑇𝒮𝜏1#subscript𝒮𝑃subscript𝑠subscript𝒮𝑃𝜏subscript𝑋𝑠1𝜏subscript𝑋𝑠0subscript𝔼subscript𝜏𝒮delimited-[]subscript𝑇𝑠subscript^𝜏𝒮\mathcal{R}(T,\mathscr{S},\tau):=\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal% {S}_{P}}\tau(X_{s})\left(\mathbf{1}\{\tau(X_{s})\geq 0\}-\mathbb{E}_{\tau_{% \mathscr{S}}}[T_{s}(\widehat{\tau}_{\mathscr{S}})]\right).caligraphic_R ( italic_T , script_S , italic_τ ) := divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ( bold_1 { italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ≥ 0 } - blackboard_E start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] ) .

It is a well-known result (and can be verified by algebra) that for any s𝒮𝒮P𝑠𝒮subscript𝒮𝑃s\in\mathscr{S}\cap\mathcal{S}_{P}italic_s ∈ script_S ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT,

supτLipC(d)(τ(Xs)(𝟏{τ(Xs)0}𝔼τ𝒮[Ts(τ^𝒮)]))=Bσs,subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝜏subscript𝑋𝑠1𝜏subscript𝑋𝑠0subscript𝔼subscript𝜏𝒮delimited-[]subscriptsuperscript𝑇𝑠subscript^𝜏𝒮𝐵subscript𝜎𝑠\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\left(\tau(X_{s})\left(\mathbf{1% }\{\tau(X_{s})\geq 0\}-\mathbb{E}_{\tau_{\mathscr{S}}}[T^{*}_{s}(\widehat{\tau% }_{\mathscr{S}})]\right)\right)=B\sigma_{s},roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT ( italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ( bold_1 { italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ≥ 0 } - blackboard_E start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] ) ) = italic_B italic_σ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , (38)

where Bargmaxz0zΦ(z)𝐵subscript𝑧0𝑧Φ𝑧B\equiv\arg\max_{z\geq 0}z\Phi(-z)italic_B ≡ roman_arg roman_max start_POSTSUBSCRIPT italic_z ≥ 0 end_POSTSUBSCRIPT italic_z roman_Φ ( - italic_z ).

Moreover, it follows from (Montiel Olea et al., 2023, Proposition 1 (iii) and its proof) that if C>maxs𝒮P\𝒮{π2σN𝒮(s)XsXN𝒮(s)}:=C(𝒮)𝐶𝑠\subscript𝒮𝑃𝒮𝜋2subscript𝜎subscript𝑁𝒮𝑠normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠assign𝐶𝒮C>\underset{s\in\mathcal{S}_{P}\backslash\mathscr{S}}{\max}\left\{\sqrt{\frac{% \pi}{2}}\frac{\sigma_{N_{{\mathscr{S}}(s)}}}{\left\|X_{s}-X_{N_{{\mathscr{S}}(% s)}}\right\|}\right\}:=C(\mathscr{S})italic_C > start_UNDERACCENT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_UNDERACCENT start_ARG roman_max end_ARG { square-root start_ARG divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_ARG divide start_ARG italic_σ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S ( italic_s ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S ( italic_s ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ end_ARG } := italic_C ( script_S ), where σN𝒮(s)subscript𝜎subscript𝑁𝒮𝑠\sigma_{N_{{\mathscr{S}}(s)}}italic_σ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S ( italic_s ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT denotes the standard deviation of the nearest neighbor estimate in 𝒮𝒮\mathscr{S}script_S for site s𝒮P\𝒮𝑠\subscript𝒮𝑃𝒮s\in\mathcal{S}_{P}\backslash\mathscr{S}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S, it holds that for any s𝒮P\𝒮𝑠\subscript𝒮𝑃𝒮s\in\mathcal{S}_{P}\backslash\mathscr{S}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S:

supτLipC(d)(τ(Xs)(𝟏{τ(Xs)0}𝔼τ𝒮[Ts(τ^𝒮)]))=C2XsXN𝒮(s).subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝜏subscript𝑋𝑠1𝜏subscript𝑋𝑠0subscript𝔼subscript𝜏𝒮delimited-[]subscriptsuperscript𝑇𝑠subscript^𝜏𝒮𝐶2normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\left(\tau(X_{s})\left(\mathbf{1% }\{\tau(X_{s})\geq 0\}-\mathbb{E}_{\tau_{\mathscr{S}}}[T^{*}_{s}(\widehat{\tau% }_{\mathscr{S}})]\right)\right)=\frac{C}{2}\|X_{s}-X_{N_{\mathscr{S}}(s)}\|.roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT ( italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ( bold_1 { italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ≥ 0 } - blackboard_E start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_T start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] ) ) = divide start_ARG italic_C end_ARG start_ARG 2 end_ARG ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ . (39)

Equations (38)-(39) imply:

infT𝒯𝒮1/2supτLipC(d)(T,𝒮,τ)(B#𝒮Ps𝒮𝒮Pσs)+(C21#𝒮Ps𝒮P\𝒮XsXN𝒮(s)).subscriptinfimum𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏𝐵#subscript𝒮𝑃subscript𝑠𝒮subscript𝒮𝑃subscript𝜎𝑠𝐶21#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\inf_{T\in\mathcal{T}^{1/2}_{\mathscr{S}}}\sup_{\tau\in\textrm{Lip}_{C}(% \mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},\tau)\leq\;\left(\frac{B}{\#\mathcal% {S}_{P}}\sum_{s\in\mathscr{S}\cap\mathcal{S}_{P}}\sigma_{s}\right)+\left(\frac% {C}{2}\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal{S}_{P}\backslash\mathscr{S% }}\|X_{s}-X_{N_{\mathscr{S}}(s)}\|\right).roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ≤ ( divide start_ARG italic_B end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ script_S ∩ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) + ( divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) . (40)

A.2 Proof of Lemma 2

Fix the selected sites 𝒮𝒮\mathscr{S}script_S, and denote the cardinality of 𝒮𝒮\mathscr{S}script_S as card(𝒮)card𝒮\textrm{card}(\mathscr{S})card ( script_S ). Let 𝒮1<𝒮2<<𝒮card(𝒮)subscript𝒮1subscript𝒮2subscript𝒮card𝒮\mathscr{S}_{1}<\mathscr{S}_{2}<\ldots<\mathscr{S}_{\textrm{card}(\mathscr{S})}script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < script_S start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < … < script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT denote the indices of the card(𝒮)card𝒮\textrm{card}(\mathscr{S})card ( script_S ) experimental sites in 𝒮𝒮\mathscr{S}script_S. For any Xd𝑋superscript𝑑X\in\mathbb{R}^{d}italic_X ∈ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, define X𝒩𝒮subscript𝑋subscript𝒩𝒮X_{\mathcal{N}_{\mathscr{S}}}italic_X start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT to be the element in {X𝒮1,,X𝒮card(𝒮)}subscript𝑋subscript𝒮1subscript𝑋subscript𝒮card𝒮\{X_{\mathscr{S}_{1}},\ldots,X_{\mathscr{S}_{\textrm{card}(\mathscr{S})}}\}{ italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , … , italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT } that is closest to X𝑋Xitalic_X in terms of \|\cdot\|∥ ⋅ ∥ (if there is more than one closest element, pick the X𝑋Xitalic_X associated to the smallest index). The proof has three parts.

Part I: Consider the function τ:d:superscript𝜏superscript𝑑\tau^{*}:\mathbb{R}^{d}\rightarrow\mathbb{R}italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT : blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT → blackboard_R such that:

τ(X)=CXX𝒩𝒮.superscript𝜏𝑋𝐶norm𝑋subscript𝑋subscript𝒩𝒮\tau^{*}(X)=C\|X-X_{\mathcal{N}_{\mathscr{S}}}\|.italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X ) = italic_C ∥ italic_X - italic_X start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ .

We start by showing that the function τsuperscript𝜏\tau^{*}italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT is Lipschitz with constant C𝐶Citalic_C. To see this, consider three cases:

Case 1: Suppose first that X,X{X𝒮1,,X𝒮card(𝒮)}𝑋superscript𝑋subscript𝑋subscript𝒮1subscript𝑋subscript𝒮card𝒮X,X^{\prime}\in\{X_{\mathscr{S}_{1}},\ldots,X_{\mathscr{S}_{\textrm{card}(% \mathscr{S})}}\}italic_X , italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ { italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , … , italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT }. In this case, we trivially have

|τ(X)τ(X)|=0CXX.superscript𝜏𝑋superscript𝜏superscript𝑋0𝐶norm𝑋superscript𝑋|\tau^{*}(X)-\tau^{*}(X^{\prime})|=0\leq C\|X-X^{\prime}\|.| italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X ) - italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | = 0 ≤ italic_C ∥ italic_X - italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∥ .

Case 2: Suppose now that X{X𝒮1,,X𝒮card(𝒮)}𝑋subscript𝑋subscript𝒮1subscript𝑋subscript𝒮card𝒮X\notin\{X_{\mathscr{S}_{1}},\ldots,X_{\mathscr{S}_{\textrm{card}(\mathscr{S})% }}\}italic_X ∉ { italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , … , italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT }, but X{X𝒮1,,X𝒮card(𝒮)}superscript𝑋subscript𝑋subscript𝒮1subscript𝑋subscript𝒮card𝒮X^{\prime}\in\{X_{\mathscr{S}_{1}},\ldots,X_{\mathscr{S}_{\textrm{card}(% \mathscr{S})}}\}italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ { italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , … , italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT }. In this case,

|τ(X)τ(X)|=CXX𝒩𝒮CXX,superscript𝜏𝑋superscript𝜏superscript𝑋𝐶norm𝑋subscript𝑋subscript𝒩𝒮𝐶norm𝑋superscript𝑋|\tau^{*}(X)-\tau^{*}(X^{\prime})|=C\|X-X_{\mathcal{N}_{\mathscr{S}}}\|\leq C% \|X-X^{\prime}\|,| italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X ) - italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | = italic_C ∥ italic_X - italic_X start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ ≤ italic_C ∥ italic_X - italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∥ ,

where the last inequality follows by the definition of X𝒩𝒮subscript𝑋subscript𝒩𝒮X_{\mathcal{N}_{\mathscr{S}}}italic_X start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT and the fact that X{X𝒮1,,X𝒮card(𝒮)}superscript𝑋subscript𝑋subscript𝒮1subscript𝑋subscript𝒮card𝒮X^{\prime}\in\{X_{\mathscr{S}_{1}},\ldots,X_{\mathscr{S}_{\textrm{card}(% \mathscr{S})}}\}italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ { italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , … , italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT }.

Case 3: Finally, take X,X{X𝒮1,,X𝒮card(𝒮)}𝑋superscript𝑋subscript𝑋subscript𝒮1subscript𝑋subscript𝒮card𝒮X,X^{\prime}\notin\{X_{\mathscr{S}_{1}},\ldots,X_{\mathscr{S}_{\textrm{card}(% \mathscr{S})}}\}italic_X , italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∉ { italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , … , italic_X start_POSTSUBSCRIPT script_S start_POSTSUBSCRIPT card ( script_S ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT }. In this case,

|τ(X)τ(X)|=C|XX𝒩𝒮XX𝒩𝒮|.superscript𝜏𝑋superscript𝜏superscript𝑋𝐶norm𝑋subscript𝑋subscript𝒩𝒮normsuperscript𝑋subscriptsuperscript𝑋subscript𝒩𝒮|\tau^{*}(X)-\tau^{*}(X^{\prime})|=C|\|X-X_{\mathcal{N}_{\mathscr{S}}}\|-\|X^{% \prime}-X^{\prime}_{\mathcal{N}_{\mathscr{S}}}\||.| italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X ) - italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | = italic_C | ∥ italic_X - italic_X start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ - ∥ italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ | .

Without loss of generality, assume that XX𝒩𝒮XX𝒩𝒮norm𝑋subscript𝑋subscript𝒩𝒮normsuperscript𝑋subscriptsuperscript𝑋subscript𝒩𝒮\|X-X_{\mathcal{N}_{\mathscr{S}}}\|\geq\|X^{\prime}-X^{\prime}_{\mathcal{N}_{% \mathscr{S}}}\|∥ italic_X - italic_X start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ ≥ ∥ italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥. Then,

|τ(X)τ(X)|superscript𝜏𝑋superscript𝜏superscript𝑋\displaystyle|\tau^{*}(X)-\tau^{*}(X^{\prime})|| italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X ) - italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | =\displaystyle== C(XX𝒩𝒮XX𝒩𝒮)𝐶norm𝑋subscript𝑋subscript𝒩𝒮normsuperscript𝑋subscriptsuperscript𝑋subscript𝒩𝒮\displaystyle C\left(\|X-X_{\mathcal{N}_{\mathscr{S}}}\|-\|X^{\prime}-X^{% \prime}_{\mathcal{N}_{\mathscr{S}}}\|\right)italic_C ( ∥ italic_X - italic_X start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ - ∥ italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ )
\displaystyle\leq C(XX𝒩𝒮XX𝒩𝒮)𝐶norm𝑋subscriptsuperscript𝑋subscript𝒩𝒮normsuperscript𝑋subscriptsuperscript𝑋subscript𝒩𝒮\displaystyle C\left(\|X-X^{\prime}_{\mathcal{N}_{\mathscr{S}}}\|-\|X^{\prime}% -X^{\prime}_{\mathcal{N}_{\mathscr{S}}}\|\right)italic_C ( ∥ italic_X - italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ - ∥ italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ )
=\displaystyle== C(XX+XX𝒩𝒮XX𝒩𝒮)𝐶norm𝑋superscript𝑋superscript𝑋subscriptsuperscript𝑋subscript𝒩𝒮normsuperscript𝑋subscriptsuperscript𝑋subscript𝒩𝒮\displaystyle C\left(\|X-X^{\prime}+X^{\prime}-X^{\prime}_{\mathcal{N}_{% \mathscr{S}}}\|-\|X^{\prime}-X^{\prime}_{\mathcal{N}_{\mathscr{S}}}\|\right)italic_C ( ∥ italic_X - italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT + italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ - ∥ italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ )
\displaystyle\leq CXX,𝐶norm𝑋superscript𝑋\displaystyle C\|X-X^{\prime}\|,italic_C ∥ italic_X - italic_X start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∥ ,

where the first inequality uses the definition of X𝒩𝒮subscript𝑋subscript𝒩𝒮X_{\mathcal{N}_{\mathscr{S}}}italic_X start_POSTSUBSCRIPT caligraphic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT, and the last display uses the triangle inequality. We conclude that τsuperscript𝜏\tau^{*}italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT is a Lipschitz function with constant C𝐶Citalic_C, which means it is included in our parameter space.

Part II: Since τLipC(d)superscript𝜏subscriptLip𝐶superscript𝑑\tau^{*}\in\textrm{Lip}_{C}(\mathbb{R}^{d})italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ), then for any treatment rule T𝑇Titalic_T:

supτLipC(d)(T,𝒮,τ)(T,𝒮,τ).subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏𝑇𝒮superscript𝜏\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},\tau)% \geq\mathcal{R}(T,\mathscr{S},\tau^{*}).roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ≥ caligraphic_R ( italic_T , script_S , italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) . (41)

Moreover, by definition,

(T,𝒮,τ):=1#𝒮Ps𝒮Pτ(Xs)(𝟏{τ(Xs)0}𝔼τ𝒮[Ts(τ^𝒮)]).assign𝑇𝒮𝜏1#subscript𝒮𝑃subscript𝑠subscript𝒮𝑃𝜏subscript𝑋𝑠1𝜏subscript𝑋𝑠0subscript𝔼subscript𝜏𝒮delimited-[]subscript𝑇𝑠subscript^𝜏𝒮\mathcal{R}(T,\mathscr{S},\tau):=\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal% {S}_{P}}\tau(X_{s})\left(\mathbf{1}\{\tau(X_{s})\geq 0\}-\mathbb{E}_{\tau_{% \mathscr{S}}}[T_{s}(\widehat{\tau}_{\mathscr{S}})]\right).caligraphic_R ( italic_T , script_S , italic_τ ) := divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ( bold_1 { italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ≥ 0 } - blackboard_E start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] ) .

Since τ(Xs)=0superscript𝜏subscript𝑋𝑠0\tau^{*}(X_{s})=0italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) = 0 for all s𝒮P𝒮𝑠subscript𝒮𝑃𝒮s\in\mathcal{S}_{P}\cap\mathscr{S}italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ∩ script_S, we have that

(T,𝒮,τ)𝑇𝒮superscript𝜏\displaystyle\mathcal{R}(T,\mathscr{S},\tau^{*})caligraphic_R ( italic_T , script_S , italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) =\displaystyle== 1#𝒮Ps𝒮P\𝒮τ(Xs)(𝟏{τ(Xs)0}𝔼τ𝒮[Ts(τ^𝒮)]),1#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮superscript𝜏subscript𝑋𝑠1superscript𝜏subscript𝑋𝑠0subscript𝔼subscriptsuperscript𝜏𝒮delimited-[]subscript𝑇𝑠subscript^𝜏𝒮\displaystyle\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal{S}_{P}\backslash% \mathscr{S}}\tau^{*}(X_{s})\left(\mathbf{1}\{\tau^{*}(X_{s})\geq 0\}-\mathbb{E% }_{\tau^{*}_{\mathscr{S}}}[T_{s}(\widehat{\tau}_{\mathscr{S}})]\right),divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ( bold_1 { italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ≥ 0 } - blackboard_E start_POSTSUBSCRIPT italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] ) , (42)
=\displaystyle== 1#𝒮Ps𝒮P\𝒮CXsXN𝒮(s)(1𝔼𝟎[Ts(τ^𝒮)]),1#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮𝐶normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠1subscript𝔼0delimited-[]subscript𝑇𝑠subscript^𝜏𝒮\displaystyle\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal{S}_{P}\backslash% \mathscr{S}}C\|X_{s}-X_{N_{\mathscr{S}}(s)}\|\left(1-\mathbb{E}_{\mathbf{0}}[T% _{s}(\widehat{\tau}_{\mathscr{S}})]\right),divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT italic_C ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ( 1 - blackboard_E start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT [ italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) ] ) ,

where the last line uses the definition of τsuperscript𝜏\tau^{*}italic_τ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. Equations (41) and (42) thus imply that for any T𝒯𝒮1/2𝑇subscriptsuperscript𝒯12𝒮T\in\mathcal{T}^{1/2}_{\mathscr{S}}italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT:

supτLipC(d)(T,𝒮,τ)(C21#𝒮Ps𝒮P\𝒮XsXN𝒮(s)).subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏𝐶21#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\sup_{\tau\in\textrm{Lip}_{C}(\mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},\tau)% \geq\;\left(\frac{C}{2}\frac{1}{\#\mathcal{S}_{P}}\sum_{s\in\mathcal{S}_{P}% \backslash\mathscr{S}}\|X_{s}-X_{N_{\mathscr{S}}(s)}\|\right).roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ≥ ( divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) . (43)

Part III: Equation (43) implies

infT𝒯𝒮1/2supτLipC(d)(T,𝒮,τ)(C21#𝒮Ps𝒮P\𝒮XsXN𝒮(s)).subscriptinfimum𝑇subscriptsuperscript𝒯12𝒮subscriptsupremum𝜏subscriptLip𝐶superscript𝑑𝑇𝒮𝜏𝐶21#subscript𝒮𝑃subscript𝑠\subscript𝒮𝑃𝒮normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠\inf_{T\in\mathcal{T}^{1/2}_{\mathscr{S}}}\sup_{\tau\in\textrm{Lip}_{C}(% \mathbb{R}^{d})}\mathcal{R}(T,\mathscr{S},\tau)\geq\;\left(\frac{C}{2}\frac{1}% {\#\mathcal{S}_{P}}\sum_{s\in\mathcal{S}_{P}\backslash\mathscr{S}}\|X_{s}-X_{N% _{\mathscr{S}}(s)}\|\right).roman_inf start_POSTSUBSCRIPT italic_T ∈ caligraphic_T start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_sup start_POSTSUBSCRIPT italic_τ ∈ Lip start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT caligraphic_R ( italic_T , script_S , italic_τ ) ≥ ( divide start_ARG italic_C end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG # caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_POSTSUBSCRIPT ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ( italic_s ) end_POSTSUBSCRIPT ∥ ) .

References

  • Abadie et al. (2010) Abadie, A., A. Diamond, and J. Hainmueller (2010): “Synthetic control methods for comparative case studies: Estimating the effect of California’s tobacco control program,” Journal of the American statistical Association, 105, 493–505.
  • Adjaho and Christensen (2022) Adjaho, C. and T. Christensen (2022): “Externally Valid Treatment Choice,” arXiv preprint arXiv:2205.05561.
  • Allcott (2015) Allcott, H. (2015): “Site selection bias in program evaluation,” The Quarterly Journal of Economics, 130, 1117–1165.
  • Andrews et al. (2022) Andrews, I., D. Fudenberg, A. Liang, and C. Wu (2022): “The Transfer Performance of Economic Models,” arXiv preprint arXiv:2202.04796.
  • Armstrong and Kolesár (2018) Armstrong, T. B. and M. Kolesár (2018): “Optimal inference in a class of regression models,” Econometrica, 86, 655–683.
  • Azevedo et al. (2020) Azevedo, E. M., A. Deng, J. L. Montiel Olea, J. Rao, and E. G. Weyl (2020): “A/b testing with fat tails,” Journal of Political Economy, 128, 4614–000.
  • Azevedo et al. (2023) Azevedo, E. M., D. Mao, J. L. Montiel Olea, and A. Velez (2023): “The A/B testing problem with Gaussian priors,” Journal of Economic Theory, 210, 105646.
  • Banerjee et al. (2017) Banerjee, A. V., S. Chassang, and E. Snowberg (2017): “Decision theoretic approaches to experiment design and external validity,” in Handbook of Economic Field Experiments, Elsevier, vol. 1, 141–174.
  • Ben-David et al. (2010) Ben-David, S., J. Blitzer, K. Crammer, A. Kulesza, F. Pereira, and J. W. Vaughan (2010): “A theory of learning from different domains,” Machine learning, 79, 151–175.
  • Bertsimas et al. (2016) Bertsimas, D., A. King, and R. Mazumder (2016): “Best subset selection via a modern optimization lens,” The Annals of Statistics, 44, 813 – 852.
  • Bertsimas and Weismantel (2005) Bertsimas, D. and R. Weismantel (2005): Optimization Over Integers, Dynamic Ideas.
  • Charikar et al. (2002) Charikar, M., S. Guha, É. Tardos, and D. B. Shmoys (2002): “A Constant-Factor Approximation Algorithm for the k-Median Problem,” Journal of Computer and System Sciences, 65, 129–149.
  • Chassang and Kapon (2022) Chassang, S. and S. Kapon (2022): “Designing randomized controlled trials with external validity in mind,” Tech. rep., National Bureau of Economic Research.
  • Christensen et al. (2022) Christensen, T., H. R. Moon, and F. Schorfheide (2022): “Optimal decision rules when payoffs are partially identified,” arXiv preprint arXiv:2204.11748.
  • Cohen-Addad et al. (2018) Cohen-Addad, V., A. De Mesmay, E. Rotenberg, and A. Roytman (2018): “The bane of low-dimensionality clustering,” in Proceedings of the Twenty-Ninth Annual ACM-SIAM Symposium on Discrete Algorithms, SIAM, 441–456.
  • Cohen-Addad et al. (2022) Cohen-Addad, V., H. Esfandiari, V. Mirrokni, and S. Narayanan (2022): “Improved approximations for Euclidean k-means and k-median, via nested quasi-independent sets,” in Proceedings of the 54th Annual ACM SIGACT Symposium on Theory of Computing, 1621–1628.
  • Cook et al. (2002) Cook, T. D., D. T. Campbell, and W. Shadish (2002): Experimental and quasi-experimental designs for generalized causal inference, vol. 1195, Houghton Mifflin Boston, MA.
  • Degtiar and Rose (2023) Degtiar, I. and S. Rose (2023): “A review of generalizability and transportability,” Annual Review of Statistics and Its Application, 10, 501–524.
  • Duchi and Namkoong (2021) Duchi, J. C. and H. Namkoong (2021): “Learning models with uniform performance via distributionally robust optimization,” The Annals of Statistics, 49, 1378–1406.
  • Dudley (2002) Dudley, R. (2002): Real Analysis and Probability, vol. 74, Cambridge University Press.
  • Duflo et al. (2007) Duflo, E., R. Glennerster, and M. Kremer (2007): “Using randomization in development economics research: A toolkit,” Handbook of development economics, 4, 3895–3962.
  • Egami and Lee (2024) Egami, N. and D. D. I. Lee (2024): “Designing Multi-Context Studies for External Validity: Site Selection via Synthetic Purposive Sampling,” .
  • Gechter (2024) Gechter, M. (2024): “Generalizing the Results from Social Experiments: Theory and Evidence from India,” Journal of Business & Economic Statistics, 42, 801–811.
  • Gechter et al. (2024) Gechter, M., K. Hirano, J. Lee, M. Mahmud, O. Mondal, J. Morduch, S. Ravindran, and A. S. Shonchoy (2024): “Selecting Experimental Sites for External Validity,” .
  • Gurobi Optimization, LLC (2023) Gurobi Optimization, LLC (2023): “Gurobi Optimizer Reference Manual,” .
  • Hu et al. (2024) Hu, Y., H. Zhu, E. Brunskill, and S. Wager (2024): “Minimax-Regret Sample Selection in Randomized Experiments,” arXiv preprint arXiv:2403.01386.
  • Ishihara and Kitagawa (2021) Ishihara, T. and T. Kitagawa (2021): “Evidence Aggregation for Treatment Choice,” ArXiv:2108.06473 [econ.EM], https://doi.org/10.48550/arXiv.2108.06473.
  • Kariv and Hakimi (1979) Kariv, O. and S. L. Hakimi (1979): “An algorithmic approach to network location problems. I: The p-centers,” SIAM journal on applied mathematics, 37, 513–538.
  • Kumar et al. (2010) Kumar, A., Y. Sabharwal, and S. Sen (2010): “Linear-time approximation schemes for clustering problems in any dimensions,” Journal of the ACM (JACM), 57, 1–32.
  • Lee et al. (2021) Lee, J. N., J. Morduch, S. Ravindran, A. Shonchoy, and H. Zaman (2021): “Poverty and migration in the digital age: Experimental evidence on mobile banking in Bangladesh,” American Economic Journal: Applied Economics, 13, 38–71.
  • Manski and Tetenov (2016) Manski, C. F. and A. Tetenov (2016): “Sufficient trial size to inform clinical practice,” Proceedings of the National Academy of Sciences, 113, 10518–10523.
  • Manski and Tetenov (2019) ——— (2019): “Trial Size for Near-Optimal Choice Between Surveillance and Aggressive Treatment: Reconsidering MSLT-II,” The American Statistician, 73, 305–311.
  • Mansour et al. (2009) Mansour, Y., M. Mohri, and A. Rostamizadeh (2009): “Domain Adaptation: Learning Bounds and Algorithms,” in Proceedings of The 22nd Annual Conference on Learning Theory (COLT 2009), Montréal, Canada.
  • Megiddo and Supowit (1984) Megiddo, N. and K. J. Supowit (1984): “On the complexity of some common geometric location problems,” SIAM journal on computing, 13, 182–196.
  • Menzel (2023) Menzel, K. (2023): “Transfer Estimates for Causal Effects across Heterogeneous Sites,” arXiv preprint arXiv:2305.01435.
  • Montiel Olea et al. (2023) Montiel Olea, J. L., C. Qiu, and J. Stoye (2023): “Decision Theory for Treatment Choice Problems with Partial Identification,” arXiv preprint arXiv:2312.17623.
  • Müller (2011) Müller, U. K. (2011): “Efficient tests under a weak convergence assumption,” Econometrica, 79, 395–435.
  • Naumann et al. (2018) Naumann, E., L. F. Stöetzer, and G. Pietrantuono (2018): “Attitudes towards highly skilled and low-skilled immigration in Europe: A survey experiment in 15 European countries,” European Journal of Political Research, 57, 1009–1030.
  • Raiffa and Schlaifer (1961) Raiffa, H. and R. Schlaifer (1961): “Applied statistical decision theory.” .
  • Sahoo et al. (2022) Sahoo, R., L. Lei, and S. Wager (2022): “Learning from a biased sample,” arXiv preprint arXiv:2209.01754.
  • Stoye (2012) Stoye, J. (2012): “Minimax regret treatment choice with covariates or with limited validity of experiments,” Journal of Econometrics, 166, 138–156.
  • Sugiyama et al. (2007) Sugiyama, M., M. Krauledat, and K.-R. Müller (2007): “Covariate shift adaptation by importance weighted cross validation.” Journal of Machine Learning Research, 8.
  • Tetenov (2012) Tetenov, A. (2012): “Statistical treatment choice based on asymmetric minimax regret criteria,” Journal of Econometrics, 166, 157–165.
  • Vivalt (2020) Vivalt, E. (2020): “How much can we generalize from impact evaluations?” Journal of the European Economic Association, 18, 3045–3089.
  • Wald (1950) Wald, A. (1950): Statistical Decision Functions, Wiley publications in statistics, New York; Chapman & Hall: London.
  • Williamson and Shmoys (2011) Williamson, D. P. and D. B. Shmoys (2011): The design of approximation algorithms, Cambridge university press.
  • Yata (2021) Yata, K. (2021): “Optimal Decision Rules Under Partial Identification,” ArXiv:2111.04926 [econ.EM], https://doi.org/10.48550/arXiv.2111.04926.
  • Zhang (2007) Zhang, P. (2007): “A new approximation algorithm for the k-facility location problem,” Theoretical Computer Science, 384, 126–135.

Appendix B Supplemental Appendix

B.1 Gurobi’s Output

In this section, we discuss some parameters in the Gurobi optimizer that can be tuned to improve its performance.212121For more specifics, see https://www.gurobi.com/documentation/current/refman/mip_models.html. First, MIPFocus specifies, on a high level, whether the solution should prioritize speed or optimality. The default value of this parameter is 00, which achieves a balance between searching for new solutions and proving optimality of the current solution. We set this parameter to 2222, prioritizing finding good-quality and optimal solutions. A second set of important parameters affect how the solver terminates. There are two termination choices for MIP models: 1) a restriction on the runtime of the algorithm, such as using TimeLimit to limit the wall-clock runtime, and 2) controlling the optimality gap, by setting a parameter MIPGap that stops the algorithm when the relative gap between the best-known solution and the best known bound on the solution objective is less than the specified value. In this application, we let the runtime to be the default value (infinity) and set the tolerance level to 106superscript10610^{-6}10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT. In Figure 9, the Gurobi solver outputs not only the solution and the optimal value of the problem but also the output gap. In this example, we can see that the gap between bounds is 00, demonstrating that the k𝑘kitalic_k-median problem has been solved to provable optimality. An interesting feature of using the integer programming formulation of the k𝑘kitalic_k-median problem is that even when we are not able to solve the problem to provable optimality, the optimality gap provides a suboptimality guarantee for the obtained solution. Finally, the output reports the algorithm’s runtime, which was only 0.050.050.050.05 seconds for this example. There are more parameters one can tune to control the root relaxation, the aggressiveness of the cutting plane strategies and the level of presolve, tolerance parameters for primal feasibility, the integer feasibility, and more. For the applications in this paper, we let the rest of the parameters be the default.

Refer to caption
Figure 9: Gurobi Output

Notes: An example output from the MIP solver in Gurobi using the multi-country experiment for k=6𝑘6k=6italic_k = 6. See details of the application in Section 5.2.

B.2 Exact Statement and Proof of Result Discussed in Section 6.2

Consider the following specialization of our setting: d=1𝑑1d=1italic_d = 1 and therefore X𝑋X\in\mathbb{R}italic_X ∈ blackboard_R, 𝒮E={1,4}subscript𝒮𝐸14\mathcal{S}_{E}=\{1,4\}caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = { 1 , 4 }, 𝒮P={2,3}subscript𝒮𝑃23\mathcal{S}_{P}=\{2,3\}caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = { 2 , 3 }, and for all sites we just have Xs=ssubscript𝑋𝑠𝑠X_{s}=sitalic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_s. (Intuitively, we consider 4 sites lined up equidistantly on a straight line, where the middle two sites are the policy sites.) Suppose furthermore that sampling distributions of signals are degenerate; formally, τ^s=τs:=τ(Xs)subscript^𝜏𝑠subscript𝜏𝑠assign𝜏subscript𝑋𝑠\hat{\tau}_{s}=\tau_{s}:=\tau(X_{s})over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT := italic_τ ( italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) for all s𝑠sitalic_s.

Lemma 3.

Under this section’s additional assumptions (see paragraph immediately above):

  1. 1.

    If sampling schemes may randomize, the lowest regret achievable by any combination of sampling scheme and treatment rule equals C/2𝐶2C/2italic_C / 2.

  2. 2.

    If sampling has to be purposive (nonrandomized), the lowest regret achievable in combination with any treatment rule equals 3C/43𝐶43C/43 italic_C / 4.

Proof.

To see claim 1, consider the distribution under which (τ1,τ2,τ3,τ4)subscript𝜏1subscript𝜏2subscript𝜏3subscript𝜏4(\tau_{1},\tau_{2},\tau_{3},\tau_{4})( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) is uniformly distributed over {(0,C,C,0),(0,C,C,0)}0𝐶𝐶00𝐶𝐶0\{(0,C,C,0),(0,-C,-C,0)\}{ ( 0 , italic_C , italic_C , 0 ) , ( 0 , - italic_C , - italic_C , 0 ) }. Then the experimental sites do not yield any information, and no sampling and decision rule can improve on tossing a coin, in which case expected regret equals C/2𝐶2C/2italic_C / 2. We conclude that the MMR value of this decision problem is at least C/2𝐶2C/2italic_C / 2.

Next, suppose the policy maker uniformly randomizes over 𝒮{1,4}𝒮14\mathscr{S}\subset\{1,4\}script_S ⊂ { 1 , 4 } for experimentation and then implements the new policy with probability a2=a3=[(C+τ^𝒮)/2C]01subscript𝑎2subscript𝑎3superscriptsubscriptdelimited-[]𝐶subscript^𝜏𝒮2𝐶01a_{2}=a_{3}=[(C+\hat{\tau}_{\mathscr{S}})/2C]_{0}^{1}italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = [ ( italic_C + over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT script_S end_POSTSUBSCRIPT ) / 2 italic_C ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT; here, the notation [X]01:=min{max{X,0},1}assignsuperscriptsubscriptdelimited-[]𝑋01𝑋01[X]_{0}^{1}:=\min\{\max\{X,0\},1\}[ italic_X ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT := roman_min { roman_max { italic_X , 0 } , 1 } indicates clamping of X𝑋Xitalic_X to [0,1]01[0,1][ 0 , 1 ]. We will show that the worst-case expected regret under this scheme is C/2𝐶2C/2italic_C / 2, which therefore is the problem’s MMR value and is attained by this rule.

Careful book-keeping reveals that the policy maker’s worst-case expected regret equals

max(τ1,τ2,τ3,τ4):|τsτt|C|st|12(τ2++τ3+)12([Cτ12C]01+[Cτ42C]01)+12(τ2+τ3)12([C+τ12C]01+[C+τ42C]01).subscript:subscript𝜏1subscript𝜏2subscript𝜏3subscript𝜏4absentsubscript𝜏𝑠subscript𝜏𝑡𝐶𝑠𝑡12superscriptsubscript𝜏2superscriptsubscript𝜏312superscriptsubscriptdelimited-[]𝐶subscript𝜏12𝐶01superscriptsubscriptdelimited-[]𝐶subscript𝜏42𝐶0112superscriptsubscript𝜏2superscriptsubscript𝜏312superscriptsubscriptdelimited-[]𝐶subscript𝜏12𝐶01superscriptsubscriptdelimited-[]𝐶subscript𝜏42𝐶01\max_{\begin{subarray}{c}(\tau_{1},\tau_{2},\tau_{3},\tau_{4}):\\ |\tau_{s}-\tau_{t}|\leq C|s-t|\end{subarray}}\frac{1}{2}(\tau_{2}^{+}+\tau_{3}% ^{+})\cdot\frac{1}{2}\left(\left[\frac{C-\tau_{1}}{2C}\right]_{0}^{1}+\left[% \frac{C-\tau_{4}}{2C}\right]_{0}^{1}\right)+\frac{1}{2}(\tau_{2}^{-}+\tau_{3}^% {-})\cdot\frac{1}{2}\left(\left[\frac{C+\tau_{1}}{2C}\right]_{0}^{1}+\left[% \frac{C+\tau_{4}}{2C}\right]_{0}^{1}\right).roman_max start_POSTSUBSCRIPT start_ARG start_ROW start_CELL ( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) : end_CELL end_ROW start_ROW start_CELL | italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT | ≤ italic_C | italic_s - italic_t | end_CELL end_ROW end_ARG end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ) ⋅ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( [ divide start_ARG italic_C - italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + [ divide start_ARG italic_C - italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) + divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT + italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ⋅ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( [ divide start_ARG italic_C + italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + [ divide start_ARG italic_C + italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) .

We will solve this by considering subcases. Suppose first that τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and τ3subscript𝜏3\tau_{3}italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT have different signs. Since both objective and constraints are invariant under multiplying (τ1,τ2,τ3,τ4)subscript𝜏1subscript𝜏2subscript𝜏3subscript𝜏4(\tau_{1},\tau_{2},\tau_{3},\tau_{4})( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) by 11-1- 1, suppose without further loss of generality that τ20τ3subscript𝜏20subscript𝜏3\tau_{2}\geq 0\geq\tau_{3}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≥ 0 ≥ italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. The optimization problem now simplifies to

max(τ1,τ2,τ3,τ4):|τsτt|C|st||τ2|4([Cτ12C]01+[Cτ42C]01)B[0,2]+|τ3|4([C+τ12C]01+[C+τ42C]01)=2Bmax{|τ2|,|τ3|}2C2,subscript:subscript𝜏1subscript𝜏2subscript𝜏3subscript𝜏4absentsubscript𝜏𝑠subscript𝜏𝑡𝐶𝑠𝑡subscript𝜏24absent𝐵02superscriptsubscriptdelimited-[]𝐶subscript𝜏12𝐶01superscriptsubscriptdelimited-[]𝐶subscript𝜏42𝐶01subscript𝜏34absent2𝐵superscriptsubscriptdelimited-[]𝐶subscript𝜏12𝐶01superscriptsubscriptdelimited-[]𝐶subscript𝜏42𝐶01subscript𝜏2subscript𝜏32𝐶2\max_{\begin{subarray}{c}(\tau_{1},\tau_{2},\tau_{3},\tau_{4}):\\ |\tau_{s}-\tau_{t}|\leq C|s-t|\end{subarray}}\frac{|\tau_{2}|}{4}\underset{% \equiv B\in[0,2]}{\underbrace{\left(\left[\frac{C-\tau_{1}}{2C}\right]_{0}^{1}% +\left[\frac{C-\tau_{4}}{2C}\right]_{0}^{1}\right)}}+\frac{|\tau_{3}|}{4}% \underset{=2-B}{\underbrace{\left(\left[\frac{C+\tau_{1}}{2C}\right]_{0}^{1}+% \left[\frac{C+\tau_{4}}{2C}\right]_{0}^{1}\right)}}\leq\frac{\max\{|\tau_{2}|,% |\tau_{3}|\}}{2}\leq\frac{C}{2},roman_max start_POSTSUBSCRIPT start_ARG start_ROW start_CELL ( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) : end_CELL end_ROW start_ROW start_CELL | italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT | ≤ italic_C | italic_s - italic_t | end_CELL end_ROW end_ARG end_POSTSUBSCRIPT divide start_ARG | italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | end_ARG start_ARG 4 end_ARG start_UNDERACCENT ≡ italic_B ∈ [ 0 , 2 ] end_UNDERACCENT start_ARG under⏟ start_ARG ( [ divide start_ARG italic_C - italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + [ divide start_ARG italic_C - italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) end_ARG end_ARG + divide start_ARG | italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT | end_ARG start_ARG 4 end_ARG start_UNDERACCENT = 2 - italic_B end_UNDERACCENT start_ARG under⏟ start_ARG ( [ divide start_ARG italic_C + italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + [ divide start_ARG italic_C + italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) end_ARG end_ARG ≤ divide start_ARG roman_max { | italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | , | italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT | } end_ARG start_ARG 2 end_ARG ≤ divide start_ARG italic_C end_ARG start_ARG 2 end_ARG ,

where the first inequality is justified in the display and the second one follows because τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and τ3subscript𝜏3\tau_{3}italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT have different signs but differ by at most C𝐶Citalic_C.

Now let τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and τ3subscript𝜏3\tau_{3}italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT have the same sign, which we take without further loss of generality to be positive. Then we initially observe simplification to

max(τ1,τ2,τ3,τ4):|τsτt|C|st|14(τ2+τ3)([Cτ12C]01+[Cτ42C]01),subscript:subscript𝜏1subscript𝜏2subscript𝜏3subscript𝜏4absentsubscript𝜏𝑠subscript𝜏𝑡𝐶𝑠𝑡14subscript𝜏2subscript𝜏3superscriptsubscriptdelimited-[]𝐶subscript𝜏12𝐶01superscriptsubscriptdelimited-[]𝐶subscript𝜏42𝐶01\max_{\begin{subarray}{c}(\tau_{1},\tau_{2},\tau_{3},\tau_{4}):\\ |\tau_{s}-\tau_{t}|\leq C|s-t|\end{subarray}}\frac{1}{4}(\tau_{2}+\tau_{3})% \left(\left[\frac{C-\tau_{1}}{2C}\right]_{0}^{1}+\left[\frac{C-\tau_{4}}{2C}% \right]_{0}^{1}\right),roman_max start_POSTSUBSCRIPT start_ARG start_ROW start_CELL ( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) : end_CELL end_ROW start_ROW start_CELL | italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT | ≤ italic_C | italic_s - italic_t | end_CELL end_ROW end_ARG end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 4 end_ARG ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) ( [ divide start_ARG italic_C - italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + [ divide start_ARG italic_C - italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) ,

and we can furthermore concentrate out τ1=τ2Csubscript𝜏1subscript𝜏2𝐶\tau_{1}=\tau_{2}-Citalic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_C, τ4=τ3Csubscript𝜏4subscript𝜏3𝐶\tau_{4}=\tau_{3}-Citalic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT - italic_C to get

max(τ2,τ3):|τ2τ3|C14(τ2+τ3)([2Cτ22C]01+[2Cτ32C]01).subscript:subscript𝜏2subscript𝜏3absentsubscript𝜏2subscript𝜏3𝐶14subscript𝜏2subscript𝜏3superscriptsubscriptdelimited-[]2𝐶subscript𝜏22𝐶01superscriptsubscriptdelimited-[]2𝐶subscript𝜏32𝐶01\max_{\begin{subarray}{c}(\tau_{2},\tau_{3}):\\ |\tau_{2}-\tau_{3}|\leq C\end{subarray}}\frac{1}{4}(\tau_{2}+\tau_{3})\left(% \left[\frac{2C-\tau_{2}}{2C}\right]_{0}^{1}+\left[\frac{2C-\tau_{3}}{2C}\right% ]_{0}^{1}\right).roman_max start_POSTSUBSCRIPT start_ARG start_ROW start_CELL ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) : end_CELL end_ROW start_ROW start_CELL | italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT | ≤ italic_C end_CELL end_ROW end_ARG end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 4 end_ARG ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) ( [ divide start_ARG 2 italic_C - italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + [ divide start_ARG 2 italic_C - italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) .

Clamping of expressions at 1111 cannot bind because τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and τ3subscript𝜏3\tau_{3}italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT are positive. If clamping at 0 binds for both fractions, then the objective equals 00. Suppose clamping at 00 binds for one expression, say (without further loss of generality) because τ2>2Csubscript𝜏22𝐶\tau_{2}>2Citalic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 2 italic_C, then we have simplification to

max(τ2,τ3):|τ2τ3|C14(τ2+τ3)2Cτ32C.subscript:subscript𝜏2subscript𝜏3absentsubscript𝜏2subscript𝜏3𝐶14subscript𝜏2subscript𝜏32𝐶subscript𝜏32𝐶\max_{\begin{subarray}{c}(\tau_{2},\tau_{3}):\\ |\tau_{2}-\tau_{3}|\leq C\end{subarray}}\frac{1}{4}(\tau_{2}+\tau_{3})\frac{2C% -\tau_{3}}{2C}.roman_max start_POSTSUBSCRIPT start_ARG start_ROW start_CELL ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) : end_CELL end_ROW start_ROW start_CELL | italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT | ≤ italic_C end_CELL end_ROW end_ARG end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 4 end_ARG ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) divide start_ARG 2 italic_C - italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG .

Keeping in mind that τ2>2Csubscript𝜏22𝐶\tau_{2}>2Citalic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 2 italic_C and therefore also τ3>Csubscript𝜏3𝐶\tau_{3}>Citalic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT > italic_C in this subcase, evaluation of derivatives shows that this expression decreases in τ3subscript𝜏3\tau_{3}italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT; hence, τ3=τ2Csubscript𝜏3subscript𝜏2𝐶\tau_{3}=\tau_{2}-Citalic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_C. Substituting this in, one can further verify the expression to be decreasing in τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT; therefore, the maximal value in this subcase is attained at a boundary point also covered by the next case (and, though not essential for the argument, this value can be verified to be 3C/83𝐶83C/83 italic_C / 8).

Finally, if no clamping binds, we can reduce the problem to

max(τ2,τ3):|τ2τ3|C14(τ2+τ3)4Cτ2τ32C=C2,subscript:subscript𝜏2subscript𝜏3absentsubscript𝜏2subscript𝜏3𝐶14subscript𝜏2subscript𝜏34𝐶subscript𝜏2subscript𝜏32𝐶𝐶2\max_{\begin{subarray}{c}(\tau_{2},\tau_{3}):\\ |\tau_{2}-\tau_{3}|\leq C\end{subarray}}\frac{1}{4}(\tau_{2}+\tau_{3})\frac{4C% -\tau_{2}-\tau_{3}}{2C}=\frac{C}{2},roman_max start_POSTSUBSCRIPT start_ARG start_ROW start_CELL ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) : end_CELL end_ROW start_ROW start_CELL | italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT | ≤ italic_C end_CELL end_ROW end_ARG end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 4 end_ARG ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) divide start_ARG 4 italic_C - italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_C end_ARG = divide start_ARG italic_C end_ARG start_ARG 2 end_ARG ,

where the maximum is attained by setting τ2+τ3=2Csubscript𝜏2subscript𝜏32𝐶\tau_{2}+\tau_{3}=2Citalic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 2 italic_C.

Regarding claim 2, by the decision problem’s symmetry, it is without further loss of generality to assume that site 1111 is being sampled. MMR is then at least 3C/43𝐶43C/43 italic_C / 4 because this value is achieved if the true parameter values (τ1,τ2,τ3,τ4)subscript𝜏1subscript𝜏2subscript𝜏3subscript𝜏4(\tau_{1},\tau_{2},\tau_{3},\tau_{4})( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) are equally likely to be (0,C,2C,3C)0𝐶2𝐶3𝐶(0,C,2C,3C)( 0 , italic_C , 2 italic_C , 3 italic_C ) or (0,C,2C,3C)0𝐶2𝐶3𝐶-(0,C,2C,3C)- ( 0 , italic_C , 2 italic_C , 3 italic_C ).

We next show that this value is attained by uniformly assigning treatment with probability a2=a3=[(3C2τ^1)/6C]01subscript𝑎2subscript𝑎3superscriptsubscriptdelimited-[]3𝐶2subscript^𝜏16𝐶01a_{2}=a_{3}=[(3C-2\hat{\tau}_{1})/6C]_{0}^{1}italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = [ ( 3 italic_C - 2 over^ start_ARG italic_τ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) / 6 italic_C ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT. Indeed, worst case regret of this decision rule equals

max(τ1,τ2,τ3,τ4):|τsτt|C|st|12(τ2++τ3+)[3C2τ16C]01+12(τ2+τ3)[3C+2τ16C]01.subscript:subscript𝜏1subscript𝜏2subscript𝜏3subscript𝜏4absentsubscript𝜏𝑠subscript𝜏𝑡𝐶𝑠𝑡12superscriptsubscript𝜏2superscriptsubscript𝜏3superscriptsubscriptdelimited-[]3𝐶2subscript𝜏16𝐶0112superscriptsubscript𝜏2superscriptsubscript𝜏3superscriptsubscriptdelimited-[]3𝐶2subscript𝜏16𝐶01\max_{\begin{subarray}{c}(\tau_{1},\tau_{2},\tau_{3},\tau_{4}):\\ |\tau_{s}-\tau_{t}|\leq C|s-t|\end{subarray}}\frac{1}{2}(\tau_{2}^{+}+\tau_{3}% ^{+})\left[\frac{3C-2\tau_{1}}{6C}\right]_{0}^{1}+\frac{1}{2}(\tau_{2}^{-}+% \tau_{3}^{-})\left[\frac{3C+2\tau_{1}}{6C}\right]_{0}^{1}.roman_max start_POSTSUBSCRIPT start_ARG start_ROW start_CELL ( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) : end_CELL end_ROW start_ROW start_CELL | italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT | ≤ italic_C | italic_s - italic_t | end_CELL end_ROW end_ARG end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ) [ divide start_ARG 3 italic_C - 2 italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 6 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT + italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) [ divide start_ARG 3 italic_C + 2 italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 6 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT .

If τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and τ3subscript𝜏3\tau_{3}italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT have different signs, we can bound this value by C/2𝐶2C/2italic_C / 2 just as before. Suppose they have the same sign, which we take to be positive without further loss of generality, then the problem simplifies to

max(τ1,τ2,τ3,τ4):|τsτt|C|st|12(τ2+τ3)[3C2τ16C]01=maxτ112(2τ1+3C)[3C2τ16C]01=3C4.subscript:subscript𝜏1subscript𝜏2subscript𝜏3subscript𝜏4absentsubscript𝜏𝑠subscript𝜏𝑡𝐶𝑠𝑡12subscript𝜏2subscript𝜏3superscriptsubscriptdelimited-[]3𝐶2subscript𝜏16𝐶01subscriptsubscript𝜏1122subscript𝜏13𝐶superscriptsubscriptdelimited-[]3𝐶2subscript𝜏16𝐶013𝐶4\max_{\begin{subarray}{c}(\tau_{1},\tau_{2},\tau_{3},\tau_{4}):\\ |\tau_{s}-\tau_{t}|\leq C|s-t|\end{subarray}}\frac{1}{2}(\tau_{2}+\tau_{3})% \left[\frac{3C-2\tau_{1}}{6C}\right]_{0}^{1}=\max_{\tau_{1}}\frac{1}{2}(2\tau_% {1}+3C)\left[\frac{3C-2\tau_{1}}{6C}\right]_{0}^{1}=\frac{3C}{4}.roman_max start_POSTSUBSCRIPT start_ARG start_ROW start_CELL ( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) : end_CELL end_ROW start_ROW start_CELL | italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT | ≤ italic_C | italic_s - italic_t | end_CELL end_ROW end_ARG end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) [ divide start_ARG 3 italic_C - 2 italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 6 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT = roman_max start_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 2 italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 3 italic_C ) [ divide start_ARG 3 italic_C - 2 italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 6 italic_C end_ARG ] start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT = divide start_ARG 3 italic_C end_ARG start_ARG 4 end_ARG .

Here, the first equality concentrates out τ2=τ1+Csubscript𝜏2subscript𝜏1𝐶\tau_{2}=\tau_{1}+Citalic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_C and τ3=τ2+Csubscript𝜏3subscript𝜏2𝐶\tau_{3}=\tau_{2}+Citalic_τ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_C; the second equality uses that clamping cannot bind (clamping at 00 would set the expression to 00, clamping at 1111 would imply that τ2<0subscript𝜏20\tau_{2}<0italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < 0), after which the problem is straightforwardly solved by τ1=0subscript𝜏10\tau_{1}=0italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0. ∎

B.3 Additional Analysis of the Survey Experiment

As mentioned in Section 5, we can obtain experimental estimates of all fifteen countries using the original experiments conducted in Naumann et al. (2018). Generally, experimental estimates of all policy sites are unknown and unattainable in most real-world applications; otherwise, there is no need to solve the site selection problem. However, as an illustrative example, we will leverage the information in these experiments to quantify the magnitude of the Lipschitz constant C𝐶Citalic_C needed to explain the treatment heterogeneity in the data and the constant C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S ) in the assumption of Lemma 1 that gives the result of this paper.

The outcome of interest from the survey is a categorical variable indicating survey respondents’ attitudes towards immigrants: 1111 for Allow None; 2222 for Allow A Few; 3333 for Allow Some; 4444 for Allow Many. For a more straightforward interpretation, we redefine the outcome variable to be binary: we let the outcome of the survey respondent be 1111 if she answers 3333 or 4444, indicating “support”; otherwise, we let her outcome be 00, indicating “oppose.” The treatment is also a binary variable, which equals 1111 (0)0(0)( 0 ) if the survey is about high-skilled (low-skilled) immigrants. We use a simple difference-in-means estimator to estimate the treatment effect of each country. The table below shows the point estimates and their standard errors. The point estimates speak to the difference between the percentage of people who support high-skilled immigrants and the percentage of people who support low-skilled immigrants. For example, 25.8% more survey respondents in Austria are more supportive of high-skilled immigrants, compared to low-skilled immigrants.

Table 1: Experimental Estimates of Each Country

Country Estimate Standard Error Austria 0.258906 0.026490 Belgium 0.232145 0.024740 Switzerland 0.285371 0.027139 Czechia 0.222865 0.019601 Germany 0.339650 0.016396 Denmark 0.293745 0.025156 Spain 0.265763 0.022948 Finland 0.403363 0.020222 France 0.275320 0.022045 United Kingdom 0.407362 0.022651 Ireland 0.238961 0.020889 Netherlands 0.301243 0.022853 Norway 0.262747 0.025613 Sweden 0.149249 0.019271 Slovenia 0.301862 0.025517

Notes: This table presents the experimental estimates of the treatment effect of the policy in each country. The original outcome from the survey is a categorical variable indicating survey respondents’ attitudes towards immigrants: 1111 for Allow None; 2222 for Allow A Few; 3333 for Allow Some; 4444 for Allow Many. We redefine the outcome variable to be binary: we let the outcome of the survey respondent be 1111 if she answers 3333 or 4444, indicating “support”; otherwise, we let her outcome be 00, indicating “oppose.” The treatment is also a binary variable, which equals 1111 (0)0(0)( 0 ) if the survey is about high-skilled (low-skilled) immigrants. We use a simple difference-in-means estimator to estimate the treatment effect of each country and their standard error.

In Figure 10(a), each point represents a pair of two countries, and the slope from the origin to each point represents the smallest Lipschitz constant needed to explain the data observed for these two countries. Hence, the Lipschitz constant C𝐶Citalic_C that is able to explain the treatment effect heterogeneity in the data for all countries is at least

max{|τ^(Xi)τ^(Xj)|XiXj},i,j𝒮E𝒮P,^𝜏subscript𝑋𝑖^𝜏subscript𝑋𝑗normsubscript𝑋𝑖subscript𝑋𝑗for-all𝑖𝑗subscript𝒮𝐸subscript𝒮𝑃\max\left\{\frac{|\hat{\tau}(X_{i})-\hat{\tau}(X_{j})|}{\|X_{i}-X_{j}\|}\right% \},\quad\forall i,j\in\mathcal{S}_{E}\cup\mathcal{S}_{P},roman_max { divide start_ARG | over^ start_ARG italic_τ end_ARG ( italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - over^ start_ARG italic_τ end_ARG ( italic_X start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) | end_ARG start_ARG ∥ italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ∥ end_ARG } , ∀ italic_i , italic_j ∈ caligraphic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT ∪ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT , (44)

which equals 0.0892 and corresponds to the slope of the red dashed line in Figure 10. The pair of countries that give the maximum of equation 44 is Finland and Sweden. Additionally, Montiel Olea et al. (2023) show that, for each possible set of experimental sites 𝒮𝒮\mathscr{S}script_S, the C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S ) that gives the solution in Lemma 1 is defined as

C(𝒮):=maxs𝒮P\𝒮{π2σN𝒮(s)XsXN𝒮(s)}.assign𝐶𝒮𝑠\subscript𝒮𝑃𝒮𝜋2subscript𝜎subscript𝑁𝒮𝑠normsubscript𝑋𝑠subscript𝑋subscript𝑁𝒮𝑠C(\mathscr{S}):=\underset{s\in\mathcal{S}_{P}\backslash\mathscr{S}}{\max}\left% \{\sqrt{\frac{\pi}{2}}\frac{\sigma_{N_{{\mathscr{S}}(s)}}}{\left\|X_{s}-X_{N_{% {\mathscr{S}}(s)}}\right\|}\right\}.italic_C ( script_S ) := start_UNDERACCENT italic_s ∈ caligraphic_S start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT \ script_S end_UNDERACCENT start_ARG roman_max end_ARG { square-root start_ARG divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_ARG divide start_ARG italic_σ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S ( italic_s ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG ∥ italic_X start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S ( italic_s ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ end_ARG } . (45)

Replacing σN𝒮(s)subscript𝜎subscript𝑁𝒮𝑠\sigma_{N_{{\mathscr{S}}(s)}}italic_σ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT script_S ( italic_s ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT with the corresponding estimated standard errors, we get C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S ) equals 0.02330.02330.02330.0233, which corresponds to the slope of the blue dashed line in figure 10(a). Both the numbers and the plot indicate the smallest Lipschitz constant C𝐶Citalic_C needed to explain the data is bigger than the largest lower bound of C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S ) that gives the nearest neighbor result. Additionally, Figure 10(b) shows a histogram of C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S ) for all possible 𝒮𝒮\mathscr{S}script_S, and the red dashed line is the smallest Lipschitz constant compatible with data. By visual inspection, we can conclude that, in this application, the assumption C>C(𝒮)𝐶𝐶𝒮C>C(\mathscr{S})italic_C > italic_C ( script_S ) is likely to hold. It is worth pointing out that this assumption is in general not testable because the experimental estimates of all policy sites are unknown and we cannot compute C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S ).

Refer to caption
(a) The C𝐶Citalic_C Compatible with Data and C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S )
Refer to caption
(b) Histogram of C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S )
Figure 10: Lipschitz Constant C𝐶Citalic_C and C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S )

Notes: The dots in Panel 10(a) represent all possible pairs among the 15151515 countries. The slope of each point connecting to the origin represents the value of the Lipschitz constant needed to explain the estimated treatment effects and site-level covariates observed for that pair of countries. The red dash line is the smallest C𝐶Citalic_C that is needed to explain all the data, computed using equation 44. The slope of the blue dashed line represents C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S ), computed using 45. Panel 10(b) presents a histogram of C(𝒮)𝐶𝒮C(\mathscr{S})italic_C ( script_S ) for all possible choices of experimental sites 𝒮𝒮\mathscr{S}script_S. The red dashed line is the smallest Lipschitz constant needed to explain the data, corresponding to the same red dashed line in panel 10(a).

B.4 A Simple Example that Motivates Assumption 4

In this section, we provide a simple linear regression example to motivate Assumption 4. Suppose the effect of the status-quo is known in all sites and normalized to zero. For each site s𝒮𝑠𝒮s\in\mathcal{S}italic_s ∈ caligraphic_S, we have a random sample of nssubscript𝑛𝑠n_{s}italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT units in the experiment. Let Yi,ssubscript𝑌𝑖𝑠Y_{i,s}italic_Y start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT be the outcome under the policy of interest for unit i{1,,ns}𝑖1subscript𝑛𝑠i\in\{1,...,n_{s}\}italic_i ∈ { 1 , … , italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT }. We assume that Yi,ssubscript𝑌𝑖𝑠Y_{i,s}italic_Y start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT is generated as follows:

Yi,s=βXi,s+γZi,s+εi,s,subscript𝑌𝑖𝑠𝛽subscript𝑋𝑖𝑠𝛾subscript𝑍𝑖𝑠subscript𝜀𝑖𝑠Y_{i,s}=\beta X_{i,s}+\gamma Z_{i,s}+\varepsilon_{i,s},italic_Y start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT = italic_β italic_X start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT + italic_γ italic_Z start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT + italic_ε start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT ,

where Xi,ssubscript𝑋𝑖𝑠X_{i,s}\in\mathbb{R}italic_X start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT ∈ blackboard_R is the observed unit-level covariate for individual i𝑖iitalic_i in site s𝑠sitalic_s, Zi,ssubscript𝑍𝑖𝑠Z_{i,s}\in\mathbb{R}italic_Z start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT ∈ blackboard_R is the unobserved unit-level covariate, εi,s𝒩(0,σε,s2)similar-tosubscript𝜀𝑖𝑠𝒩0superscriptsubscript𝜎𝜀𝑠2\varepsilon_{i,s}\sim\mathcal{N}(0,\sigma_{\varepsilon,s}^{2})italic_ε start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT ∼ caligraphic_N ( 0 , italic_σ start_POSTSUBSCRIPT italic_ε , italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) is an error term with σε,s2>0superscriptsubscript𝜎𝜀𝑠20\sigma_{\varepsilon,s}^{2}>0italic_σ start_POSTSUBSCRIPT italic_ε , italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT > 0 and is also independent of (Xi,s,Zi,s)subscript𝑋𝑖𝑠subscript𝑍𝑖𝑠\left(X_{i,s},Z_{i,s}\right)( italic_X start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT , italic_Z start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT ), and β,γ𝛽𝛾\beta,\gamma\in\mathbb{R}italic_β , italic_γ ∈ blackboard_R are the same across different sites. For simplicity, suppose that Xi,ssubscript𝑋𝑖𝑠X_{i,s}italic_X start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT and Zi,ssubscript𝑍𝑖𝑠Z_{i,s}italic_Z start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT are jointly normal: (Xi,s,Zi,s)𝒩(μs,Σs)similar-tosuperscriptsubscript𝑋𝑖𝑠subscript𝑍𝑖𝑠top𝒩subscript𝜇𝑠subscriptΣ𝑠\left(X_{i,s},Z_{i,s}\right)^{\top}\sim\mathcal{N}(\mu_{s},\varSigma_{s})( italic_X start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT , italic_Z start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT ∼ caligraphic_N ( italic_μ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , roman_Σ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ), where μs2subscript𝜇𝑠superscript2\mu_{s}\in\mathbb{R}^{2}italic_μ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and ΣssubscriptΣ𝑠\varSigma_{s}roman_Σ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is a 2×2222\times 22 × 2 positive-definite covariance matrix.

Let Y¯s:=1nsi=1nsYi,sassignsubscript¯𝑌𝑠1subscript𝑛𝑠superscriptsubscript𝑖1subscript𝑛𝑠subscript𝑌𝑖𝑠\bar{Y}_{s}:=\frac{1}{n_{s}}\sum_{i=1}^{n_{s}}Y_{i,s}over¯ start_ARG italic_Y end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT := divide start_ARG 1 end_ARG start_ARG italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_Y start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT be the sample average of the observed outcome at site s𝑠sitalic_s, and X¯s:=1nsi=1nsXi,s,Z¯s:=1nsi=1nsZi,sformulae-sequenceassignsubscript¯𝑋𝑠1subscript𝑛𝑠superscriptsubscript𝑖1subscript𝑛𝑠subscript𝑋𝑖𝑠assignsubscript¯𝑍𝑠1subscript𝑛𝑠superscriptsubscript𝑖1subscript𝑛𝑠subscript𝑍𝑖𝑠\bar{X}_{s}:=\frac{1}{n_{s}}\sum_{i=1}^{n_{s}}X_{i,s},\bar{Z}_{s}:=\frac{1}{n_% {s}}\sum_{i=1}^{n_{s}}Z_{i,s}over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT := divide start_ARG 1 end_ARG start_ARG italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_X start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT , over¯ start_ARG italic_Z end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT := divide start_ARG 1 end_ARG start_ARG italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_Z start_POSTSUBSCRIPT italic_i , italic_s end_POSTSUBSCRIPT be the observed and unobserved site-level aggregate covariates, respectively. Under the above assumptions and conditional on observed site-level aggregate covariate X¯ssubscript¯𝑋𝑠\bar{X}_{s}over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, we have

Y¯sX¯s𝒩(βX¯s+γ𝔼[Z¯s|X¯s],σs2),similar-toconditionalsubscript¯𝑌𝑠subscript¯𝑋𝑠𝒩𝛽subscript¯𝑋𝑠𝛾𝔼delimited-[]conditionalsubscript¯𝑍𝑠subscript¯𝑋𝑠superscriptsubscript𝜎𝑠2\bar{Y}_{s}\mid\bar{X}_{s}\sim\mathcal{N}(\beta\bar{X}_{s}+\gamma\mathbb{E}[% \bar{Z}_{s}|\bar{X}_{s}],\sigma_{s}^{2}),over¯ start_ARG italic_Y end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∣ over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∼ caligraphic_N ( italic_β over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + italic_γ blackboard_E [ over¯ start_ARG italic_Z end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT | over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ] , italic_σ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , (46)

where 𝔼[Z¯s|X¯s]𝔼delimited-[]conditionalsubscript¯𝑍𝑠subscript¯𝑋𝑠\mathbb{E}[\bar{Z}_{s}|\bar{X}_{s}]blackboard_E [ over¯ start_ARG italic_Z end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT | over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ] is the expectation of Z¯ssubscript¯𝑍𝑠\bar{Z}_{s}over¯ start_ARG italic_Z end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT conditional X¯ssubscript¯𝑋𝑠\bar{X}_{s}over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, and σs2>0superscriptsubscript𝜎𝑠20\sigma_{s}^{2}>0italic_σ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT > 0. Then, Assumption 4 models a case for which the policy effect of interest is

τs:=βX¯s+γ𝔼[Z¯s|X¯s],s𝒮.formulae-sequenceassignsubscript𝜏𝑠𝛽subscript¯𝑋𝑠𝛾𝔼delimited-[]conditionalsubscript¯𝑍𝑠subscript¯𝑋𝑠for-all𝑠𝒮\tau_{s}:=\beta\bar{X}_{s}+\gamma\mathbb{E}[\bar{Z}_{s}|\bar{X}_{s}],\forall s% \in\mathcal{S}.italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT := italic_β over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + italic_γ blackboard_E [ over¯ start_ARG italic_Z end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT | over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ] , ∀ italic_s ∈ caligraphic_S .

Furthermore, (46) implies that, conditional on X¯ssubscript¯𝑋𝑠\bar{X}_{s}over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, we have an unbiased and normal estimator for τssubscript𝜏𝑠\tau_{s}italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. We may also calculate that 𝔼[Z¯s|X¯s]=αs𝔼delimited-[]conditionalsubscript¯𝑍𝑠subscript¯𝑋𝑠subscript𝛼𝑠\mathbb{E}[\bar{Z}_{s}|\bar{X}_{s}]=\alpha_{s}blackboard_E [ over¯ start_ARG italic_Z end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT | over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ] = italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT for some αssubscript𝛼𝑠\alpha_{s}italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT that depends on each site s𝑠sitalic_s. Then, for s,s𝒮𝑠superscript𝑠𝒮s,s^{\prime}\in\mathcal{S}italic_s , italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ caligraphic_S, ss𝑠superscript𝑠s\neq s^{\prime}italic_s ≠ italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, we have:

τsτssubscript𝜏𝑠subscript𝜏superscript𝑠\displaystyle\tau_{s}-\tau_{s^{\prime}}italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT =β(X¯sX¯s)+γ(αsαs),absent𝛽subscript¯𝑋𝑠subscript¯𝑋superscript𝑠𝛾subscript𝛼𝑠subscript𝛼superscript𝑠\displaystyle=\beta\left(\bar{X}_{s}-\bar{X}_{s^{\prime}}\right)+\gamma\left(% \alpha_{s}-\alpha_{s^{\prime}}\right),= italic_β ( over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) + italic_γ ( italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_α start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) , (47)

implying that |τsτs|subscript𝜏𝑠subscript𝜏superscript𝑠\left|\tau_{s}-\tau_{s^{\prime}}\right|| italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | can be bounded as in Assumption 4 for some positive C𝐶Citalic_C and c𝑐citalic_c, as long as we are willing to assume that β,𝛽\beta,italic_β , γ𝛾\gammaitalic_γ are bounded, and that |αsαs|subscript𝛼𝑠subscript𝛼superscript𝑠\left|\alpha_{s}-\alpha_{s^{\prime}}\right|| italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_α start_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | are bounded uniformly among all s,s𝒮𝑠superscript𝑠𝒮s,s^{\prime}\in\mathcal{S}italic_s , italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ caligraphic_S.