<<

PRELIMINARY VERSION: DO NOT CITE The AAAI Digital Library will contain the published version some time after the conference

A Generative Adversarial Framework for Bounding Confounded Causal Effects

Yaowei Hu,1 Yongkai Wu, 2 Lu Zhang, 1 Xintao Wu 1 1 University of Arkansas 2 Clemson University [email protected], [email protected], [email protected], [email protected]

Abstract ACE may not be uniquely calculated from the observational data without further assumptions, known as the unidentifi- Causal inference from observational data is receiving wide able situation (Shpitser and Pearl 2008). In an unidentifiable applications in many fields. However, unidentifiable situa- tions, where causal effects cannot be uniquely computed from situation, any estimation of ACEs only based on the observa- observational data, pose critical barriers to applying causal in- tional distribution is not guaranteed to be correct, since there ference to complicated real applications. In this paper, we de- can be other underlying causal models that also agree with velop a bounding method for estimating the average causal ef- the observational distribution but result in different ACEs fect (ACE) under unidentifiable situations due to hidden con- (Avin, Shpitser, and Pearl 2005). founding based on Pearl’s structural causal model. We pro- In order to estimate the ACE in unidentifiable situations, pose to parameterize the unknown exogenous random vari- researchers seek for bounding approaches. In (Balke and ables and structural equations of a causal model using neural Pearl 1997), the authors develop a constrained optimization networks and implicit generative models. Then, using an ad- versarial learning framework, we search the parameter space problem for discovering bounds from the observational data to explicitly traverse causal models that agree with the given which are guaranteed to be tightest. The general idea is to observational distribution, and find those that minimize or shift the randomness of the causal model from the distribu- maximize the ACE to obtain its lower and upper bounds. The tions of U to the distributions of mappings so that it covers proposed method does not make assumption about the type all possible domains of U, either categorical, continuous, or of structural equations and variables. using both mixed. Then, the optimization is to search mapping distri- synthetic and real-world datasets are conducted. butions with the maximum/minimum ACEs subject to the constraint that the joint distribution must be consistent with Introduction the data. In a recent work (Wu et al. 2019), the authors ex- tend this idea to bound unidentifiable path-specific effects Inferring causal effects from observational data is an impor- and counterfactual effects in addition to the ACE. However, tant task in many fields, including economics (Zhao, Run- these methods are limited to categorical observed variables fola, and Kemper 2017), healthcare (Hu and Kerschberg since the number of distinct functional mappings between 2018; Wang and Wu 2019), fair machine learning (Zhang continuous variables are infinite, leading to an infinite num- and Bareinboim 2018; Zhang, Wu, and Wu 2018; Kusner ber of variables in the optimization problem. et al. 2017), etc. Pearl’s structural causal model (Pearl 2009) In this paper, we extend the method in (Balke and Pearl is a decent and widely adopted mathematical framework for 1997) for bounding ACEs to continuous and possibly high- conducting causal inference. In the structural causal model, dimensional variables, thanks to the significant progress of intervention is a technique that fixes some variable to certain generative models made by the machine learning commu- constant without changing other parts of the model. Facili- nity in past years. Following the spirit of (Balke and Pearl tated by intervention, the average causal effect (ACE) of one 1997), we also shift the randomness of the causal model variable A on another variable B can be formulated as the from the distribution of U to the distribution of all possi- average change in B due to interventions on A. The calcu- ble mappings from A to B. We define a distribution over lation of the ACE from the observational data depends on a the space of all mappings from A to B, which summarizes causal graph that specifies the parental relationship among all possible equations from (U, A) to B for all possible U. variables. Under the assumption of no hidden Since the mapping from A to B can be in arbitrary forms, (i.e., no unobserved common causes of A and B), the ACE we use the neural network as a universal estimator of map- can be calculated using the well-known truncated factor- pings so that the space of all possible mappings is estimated ization formula (Pearl 2009). However, the no-hidden con- by the parameter space Θ of the neural network. Then, given founding assumption is usually over-simplified and needs to any distribution over Θ, i.e., P (θ), both the joint distribution be relaxed in practice. When hidden confounders exist, the and the ACE can be expressed as functions of P (θ). In or- Copyright © 2021, Association for the Advancement of Artificial der for the induced joint distribution to fit the observational Intelligence (www.aaai.org). All rights reserved. data, we adopt the generative adversarial learning frame- work, where a generator is designed to produce P (θ), and of a given causal graph. A recent work (Li et al. 2020) de- a discriminator is designed to distinguish the generated dis- veloped a GAN based deconfouding algorithm assuming no tribution from the real distribution. Finally, we combine task hidden confounding. of finding the highest and lowest values of the ACE with the One critical issue in causal inference is the identifiabil- adversarial learning. The training procedure of the generator ity, which is not paid enough attention in many of these is to find a P (θ) that maximizes/minimizes the ACE, under work. Unidentifiable situations refer to that the unique es- the condition that the discriminator is unable to identify the timation of certain causal effect on a causal graph is theo- generated distribution. In the experiments we demonstrate retically infeasible (Shpitser and Pearl 2008). In this sense, that, our method can provide more accurate estimations to the ACE in unidentifiable situations cannot be point iden- the ACE than several widely used causal inference meth- tified from the data and one can only seek for estimating ods including instrumental variable estimation (Bowden and bounds. The of the ACE based on the Turkington 1984) and propensity score adjusted regression potential framework has been studied in (Kallus, (Abdia et al. 2017). Mao, and Zhou 2018), where a functional interval estima- Handling hidden confounders is one of the most important tor is derived from a weighted kernel regression. However, aspects of inferring ACEs from observational data. Some re- this work only applies to the case where A is a binary treat- cent works try to estimate the representation of hidden con- ment. In the structural causal model framework, the rea- founders using latent-variable models (e.g., (Louizos et al. sons of leading to unidentifiable situations have been stud- 2017; Chiappa 2019; Madras et al. 2019)). However, these ied in the causal inference literature, as summarized in (Wu works don’t consider the identifiability issue. Grounded on et al. 2019), where hidden confounding is one of the typical the method in (Balke and Pearl 1997), our work provides reasons. Different bounding techniques have been proposed a much more reliable estimation of the ACE than previous to address different unidentifiable situations. For example, works. Since we don’t make any assumption about the type (Miles et al. 2015) derived bounds for the natural direct ef- of structural equations and variables, the proposed method fect with a discrete mediator. In (Zhang, Wu, and Wu 2018), can serve as a general basis of applying the structural causal the authors proposed a method to bound unidentifiable path- model to a wider of applications. Last but not least, in specific effects due to the existence of the “kite graph”. In some practical situations it may be reasonable to assume a (Wu, Zhang, and Wu 2019), a method for bounding uniden- certain type of equations. Our method can be simply mod- tifiable counterfactual effects due to the existence of the “w ified to encode such assumption. We include a subsection graph” was developed. However, none of these methods can showing that encoding the linear equation assumption can apply to unidentifiable situations due to hidden confounders. make the bounds converge to a fixed value. In (Wu et al. 2019), the authors proposed a general frame- work for bounding all unidentifiable situations, which is ex- Related Work tended from (Balke and Pearl 1997) and also inherit its limi- To infer causal effects from observational data, many causal tation that all endogenous variables must be categorical. Our inference methods have been proposed, either based on work addresses the limitation of (Balke and Pearl 1997; Wu Pearl’s structural causal model framework (Pearl 2009) or et al. 2019) by a novel adaptation of the generative adver- Rubin’s potential outcome framework (Rubin 2005). Widely sarial learning framework. Following the idea of (Wu et al. adopted classic methods include propensity score based 2019), our method can also extend to all unidentifiable sit- methods (Abdia et al. 2017), causal graph based methods uations in addition to those caused by hidden confounders. (Pearl 2010), instrumental variable estimation (Bowden and This will be our future work. Turkington 1984), etc. In recent years, it has been proposed to use machine learning models to facilitate causal infer- Preliminaries ence. In (Shalit, Johansson, and Sontag 2017), the authors We develop our method based on Pearl’s structural causal estimated the causal effect under the no-hidden confound- model framework, which is formally defined as follows. ing assumption by using a neural network structure that learns a balanced representation for the treated and con- Definition 1 (Structural Causal Model (Pearl 2009)). A trol distributions. In (Louizos et al. 2017; Chiappa 2019; structural causal model M is represented by a quadriple Madras et al. 2019), a latent representation was learned hU, V, F,P (U)i where to summarize the exogenous variable space of the causal 1. U is a set of exogenous random variables that are deter- model by using deep latent-variable models such as the vari- mined by factors outside the model. ational auto-encoding (VAE). Then, causal effects can be es- 2. P (U) is a joint defined over U. timated based on the latent-variable models via approaches. In (Yoon, Jordon, and van der Schaar 2018), 3. V is a set of endogenous variables that are determined by the authors used generative adversarial nets (GAN) to gener- variables in U ∪ V. ate counterfactual outcomes such that the discriminator can- 4. F is a set of structural equations from U∪V to V. Specif- not distinguish the counterfactual outcomes from the factual ically, for each V ∈ V, there is a function fV ∈ F map- outcomes, and then estimate the ACE based on the poten- ping from U∪(V\V ) to V , i.e., v = fV (paV , uV ), where tial outcome framework. In (Kocaoglu et al. 2018; Xu et al. paV and uV are realization of a set of endogenous vari- 2019), the authors also used GAN but propose to arrange the ables PAV ∈ V \ V and a set of exogenous variables UV network structure of the generator to preserve the structure respectively. C U with the same causal graph shown in Figure1. Assume that the two models have the same equations for determining the values of C and A, but differ in the equation for determin- A B ing the value of B, as shown in Tables1 and2. Denoting the probability P (U = 1) as p, we can compute the joint Figure 1: Example: A, B, C are observed variables and U is distribution P (A = a, B = b, C = c) represented by each a hidden variable. model by summarizing the probabilities of all values of U that lead to A = a, B = b and C = c. For example, we can Table 1: Equation Table 2: Equations obtain that P (A = 0,B = 0,C = 0) = (1 − p)P (C = 0) 1 2 fA(c, u) for determin- fB(a, u) and fB(a, u) for for both models since in both models U must be 0 to get ing values of A. determining values of B. A = 0 and B = 0 when C = 0. In fact, one can verify that two models completely agree with the joint distribution 1 2 U C A = fA U A B = fB B = fB P (A, B, C). On the other hand, P (B = b|do(A = a)) = 0 0 0 0 0 0 0 P P (u) is computed by summarizing the prob- u:fB (a,u)=b 0 1 0 0 1 1 0 abilities of all values of U that lead to B = b when fix- 1 0 1 1 0 0 0 ing the value of A to a. One can verify that, for the first 1 1 1 1 1 1 1 1 model (with equation fB), P (B = 1|do(A = 1)) = 1 and P (B = 1|do(A = 0)) = 0; for the second model f 2 P (B = 1|do(A = 1)) = p If all exogenous variables in U are assumed to be mutu- (with equation B), and P (B = 1|do(A = 0)) = 0 ally independent, then the causal model is called a Marko- . Thus, the ACE is 1 for the first p vian model; otherwise, it is called a semi-Markovian model. model and for the second. Assume that the true model is For example, a causal model represented by Figure1 is a either one of the two. Since they represent the same joint distribution, there is no way to distinguish between them semi-Markovian model in which UA and UB are completely correlated and denoted by a single exogenous variable U. based on the observational data. If we randomly guess the In general, f (·) can be an equation of any type. In some true model from the two, the bias in estimating ACE can V vary from 0 to 1 depending on the value of p. cases people may assume fV (·) to be a certain type of equa- tion. For example, if for each node V , fV (·) is a linear equa- tion, then the causal model is called a linear causal model; if Bounding ACEs via Optimization f (·) is an additive equation, i.e., v = f (pa ) + u , then V V V V According to (Avin, Shpitser, and Pearl 2005), if an ACE is the causal model is called an additive causal model. unidentifiable, then there must exist multiple causal mod- Each causal model M is associated with a causal graph els that have the same observational distribution but pro- G = hV, Ei where V is a set of nodes and E is a set of edges. duce different values of the ACE. Since there is no way Each node of V corresponds to a variable of V in M. Each to distinguish these causal models using the observational edge in E, denoted by a directed arrow →, points from a data, we are unable to compute the accurate ACE with- node A ∈ U∪V to a different node B ∈ V if f uses values B out further knowledge or assumption about the underlying of A as input. Apparently, two different causal models are data generating mechanism. However, from all these pos- associated with the same causal graph if they have the same sible causal models if we can find the ones that produce inputs for all equations. the maximal/minimal ACEs, then the maximum/minimum An intervention on endogenous variable A is defined as in fact provide the tightest bounds of the unidentifiable ACE. the substitution of structural equation f (PA ,U ) with a A A A As such, the bounding exercise can be formulated as math- constant a, denoted as do(A = a) or do(a) for short. For ematical optimization problems for maximizing/minimizing another endogenous variable B which is affected by the the ACE, where we need to examine all possible causal mod- intervention, its distribution under the intervention, called els. The challenge is that, in general we have no knowledge the post-intervention distribution, is denoted as P (B|do(a)). about the equations and exogenous variables of the possible The ACE is defined as the difference of expected values of causal models. Thus, for optimization we need to intention- B under two different interventions (Pearl 2009). ally traverse all possible variable dimensions and domains, Definition 2. The average causal effect (ACE) of A on B is all kinds of variable distributions, and all types of equations, given by E[B|do(a1)] − E[B|do(a0)]. which is computationally impossible. Given a causal graph, if there exist two or more different In (Balke and Pearl 1997), the authors proposed to par- causal models associated with the causal graph that agree tition the domain of each exogenous variable into a lim- with an observational distribution but result in different ACE ited number of equivalent classes, each inducing a distinct values, then this ACE is said to be unidentifiable (Avin, Sh- functional mapping between endogenous variables. These pitser, and Pearl 2005). The complete graphical criterion of functional mappings are called the response functions. Con- ACE identifiability has been studied in (Shpitser and Pearl sider an endogenous variable V ∈ V, whose equation 2008) known as the “hedge criterion”. v = fV (paV , uV ) is a mapping from PAV ,UV to V . In gen- An unidentifiable example. Figure1 gives a toy exam- eral, UV can be with arbitrary domain size, and fV can be ple where the ACE of A and B cannot be uniquely identified any function. However, for a given value uV , no matter what due to a hidden confounder U. Consider two causal models dimension and domain UV has, fV is a deterministic map- ping from endogenous variables PAV to V , i.e., a response our problem, random noise zV is taken as input and trans- function from PAV to V . Thus, by denoting response func- formed into θV via a neural network GV (zV ). By conven- tions as values of a variable RV , called the response-function tion, we refer to this neural network as a generator. Since variable, the distribution of UV , i.e., P (uV ), can be equiva- θV represents the response function for computing v, with lently partitioned and translated into the distribution P (rV ). GV (zV ) we in fact define an implicit generative model for Then, as UV varies along its domain, the only effect it can generating v from zV , i.e., v = hV (paV ; GV (zV )). As a have on V is to switch the response function among all pos- result, by updating the generator, since parameter space ΘV sible response functions from PAV to V in the domain of approximately represents the domain of UV , we can explic- RV . It is known that we can use P (u) to express the joint itly traverse possible distributions over UV no matter what distribution P (v) as well as all ACEs (Tian and Pearl 2000). dimensions and domains it has. The relationship between UV and RV that we can Based on above discussions, to parameterize a causal similarly use P (r) to do the same, where R denotes the set model, we define an implicit generative model for each of all response-function variables. As a result, by expressing V ∈ V as follows. P (v) and the ACE using P (r), we can compute the lower or Definition 3. For a causal model ∀V ∈ V, v = upper bound of the ACE by searching for the P (r) that min- fV (pa , uV ), its parameterized version is given by imizes or maximizes the ACE, subject to that P (v) agrees V with the observational data. This would give us a linear pro- ∀V ∈ V, v = hV (paV ; GV (zV )) gramming problem with P (r) as variables. where generators G (z ) contain parameters that are to be As can be seen, if PA and/or V are continuous variables, V V V learned from data. there will be an infinite number of response functions from PAV to V , resulting a linear programming problem with in- For simplicity, we denote the parameterized causal model finite variables P (r). Thus, the above method is limited to as v = G(z) where z denotes the set of noise terms for all categorical endogenous variables and cannot directly extend endogenous variables. to the continuous domain. We tackle this challenge by adopt- Encoding independence assumptions. It is worth not- ing the generative adversarial learning framework and rep- ing that, since ΘV is a representation of UV , it should resenting candidate causal models in a parameter space such inherit the independence relationship between UV and that it can be searched using state-of-the-art optimization al- other exogenous variables. That is to say, ΘV1 and ΘV2 gorithms. Consider to estimate response functions from PAV should be (in)dependent if UV1 and UV2 are known to be to V by neural networks with a certain network structure. (in)dependent. We address this issue by using the same ran-

This means that we can define the network parameters as dom noise for generators GV1 and GV2 if UV1 and UV2 the response-function variables for summarizing all possi- are dependent. More formally, as shown in (Tian and Pearl ble mappings from PAV ,UV to V that could be estimated 2002), any causal graph can be decomposed into a num- by a fixed-architecture neural network. We then use the im- ber of disjoint components, called c-components, such that plicit generative model to generate the distribution for the any pair of exogenous variables are correlated if they belong response-function variable such that the domain of the dis- to the same component and independent if they belong to tribution is represented as the parameter space of the gener- different components. C-component factorization (Tian and ative model. We parameterize the causal model by express- Pearl 2002) provides a way to group all correlated variables ing it with response-function variables. Finally, the param- together. For example, in Figure1, A, B belong to one c- eterzied causal model is used to formulate an adversarial component and C belongs to another c-component. In our learning problem for computing the bounds of the ACE. method, we adopt the c-component factorization to decom- In the following, we explain our method in details. pose the causal graph into c-components. Then, we share the same noise term among generators corresponding to vari- Parameterizing Causal Models ables within each c-component. Specifically, for each endogenous variable V , a neural net- Optimizing ACEs work v = hV (paV ; θV ) with input paV and parameters θV ∈ ΘV is used as a universal estimator of response func- Our objective is to find causal models that bound the ACE tions from PAV to V , i.e., we treat ΘV as the response vari- of interest. Given a parameterized causal model defined in able. Thus, the domain of UV is estimated by parameter Definition3, computing the ACE from it is straightforward. 0 space ΘV of the neural network, and distribution P (uV ) is For any intervention do(a ), we directly perform it to modify correspondingly represented by the distribution over all pa- the parameterized causal model as: rameters values, i.e., P (θV ). Then, to traverse distributions 0 a = a ; ∀V 6= A, v = hV (paV ; GV (zV )). over UV is simulated by traversing distributions over ΘV . As a special case, if PAV = ∅, then we directly let v = θV Then, we estimate the value of an ACE of A on B by sam- to represent a trivial mapping. pling B from the intervened parameterized causal model. To generate different distributions for θV , we adopt We denote it as ACE(G;a1,a0) since the estimated value the implicit generative model (Mohamed and Lakshmi- depends on all generators. As a result, we can learn the narayanan 2016). An implicit generative model defines a lower bound by minimizing ACE(G;a1,a0), and learn the stochastic procedure to generate data by transforming some upper bound by maximizing ACE(G;a1,a0) or minimizing random noise to the data via some deterministic function. In −ACE(G;a1,a0). Meanwhile, we want the causal models searched in above learning process to be confined to those agree with a given observational distribution P (v). The parameterized causal model allows us to compare the generated distribution with the observational distribution, and use the discrepancy to up- date generators. Similar to generative adversarial networks (GAN) (Goodfellow et al. 2014), we make use of a discrim- inator D to distinguish observational data from that gener- ated by the parameterized causal model. We train the dis- criminator to maximize its probability of assigning correct labels, which is then used as a measure of the discrepancy between the generated distribution and the observational dis- tribution. Symbolically, it is given by max V (G, D) where D Figure 2: The network architecture with three generators V (G, D)=Ev∼P (v)[log D(v)]+Ez∼P (z)[log(1−D(G(z)))]. GA,GB,GC for the causal graph shown in Figure1. A dis- criminator is used to measure the difference between the When the generated distribution is equivalent to the obser- generated data and the real data. vational distribution, we reach the theoretical minimal value of maxD V (G, D). We denote this value by m, which varies for different versions of the GAN model (e.g., it is equal to U, but assume that we know A and B are confounded. Thus, − log 4 in the vanilla GAN (Goodfellow et al. 2014)). we define three neural networks (generators): GC , GA, and Combining above two partial objectives, to obtain the G . By conducting the c-composition factorization, we ob- lower bound (similarly for the upper bound), we would like B tain two c-components: {C} and {A, B}. So, generator GC to learn generators G that minimize ACE(G;a1,a0) subject uses one noise term z1 and generators GA and GB share an- to that maxD V (G, D) ≤ m+η. Here η is a threshold which other noise term z . To construct the parameterized causal specifies how close we want the generated distribution is to 2 model, since C has no parent, GC (z1) directly generates c to the observational distribution. Ideally η should be set to 0 but represent a trivial mapping. On the other hand, a is generated in practice we can set it to a small value to allow some room by the response function, a neural network h (c; G (z )) for imperfect fitting. By introducing the Lagrange multiplier V A 2 with c as the input and GA(z2) as the parameters. Similarly, λ, this constrained optimization problem can be converted b is generated by the neural network h (a; G (z )) with a to an unconstrained optimization problem, given by V B 2 as the input and GB(z2) as the parameters. As a result, the n  o parameterized causal model is given by min max ACE(G;a1,a0) + λ max V (G, D) − m − η . G λ≥0 D c = fC (uC ) c = GC (z1) parameterized Since the first term is irrelevant to D, we can pull maxD out a = fA(c, uA) ======⇒ a = hV (c; GA(z2)) of the sum. Finally, the optimization problem is defined as: b = fB(a, uB) b = hV (a; GB(z2)) Problem 1. Given a causal graph and the data, the lower The model structure is shown in Figure2, and the ACE of A bound (similarly for the upper bound) of the ACE of A on B on B is given by is computed by solving the optimization ACE(G;a1,a0) =Ez2∼P (z2) [hV (a1; GB(z2))] min max max {ACE(G;a1,a0) + λ (V (G, D) − m − η)} . G λ≥0 D − Ez2∼P (z2) [hV (a0; GB(z2))] . The pseudocode of above procedure is given in Algorithm 1 in the supplementary file. Implementation Considerations The training procedure is as follows. We continually sam- In practice, the generative adversarial learning does not ple mini-batches of noise samples z. For each noise sample, guarantee to converge to the global optimum (Goodfellow we compute the expressions of B prior and post to the inter- et al. 2014), which means that we may not be able to find vention do(a) based on the parameterized causal model. The the optimal solution to Problem1. However, the proposed average of the post-intervention expressions of B is used bounding method is still meaningful even if the optimal so- to compute ACE(G;a1,a0), and the average of the prior- lution cannot be obtained. Recall that the fundamental of intervention expressions of B is used to compute V (G, D). the bounding method is to search for all the causal models Then, we recurrently update the discriminator, λ, and gener- that agree with the observational distribution and find the ators based on the gradients of the objective function. maximal/minimal ACE among them. Thus, we can treat the optimizing process as a constructive way of estimating the Example ACE. That is to say, after the discriminator loss V (G, D) We use the toy example shown in Figure1 to demonstrate becomes stable, each intermediate solution provides a feasi- how to formulate the optimization problem given a causal ble estimation of the ACE, which can be used to construct graph. In general, the causal model of this example consists the bounds. Formally, we examine the intermediate solu- of three structural equations: c = fC (uC ), a = fA(c, uA), tions where constraint maxD V (G, D) ≤ m + η is satisfied. b = fB(a, uB). We cannot observe the hidden confounder Then, we take multiple intermediate solutions to compute the and , and use the one sided confidence in- then compute the ACE based on the coefficient of the terval to estimate the bound. This estimation is meaningful treatment variable. as the causal models that agree with the observational distri- • Instrumental variable estimation: We implement this bution rarely fall outside the interval. Experiments show that method following the classic instrumental variable for- this approach can provide good estimations to the ACE. mula (Bowden and Turkington 1984). Our method relies on the input of the causal graph. Many algorithms have been proposed to discovery the causal • Propensity score adjusted regression: We adopt the structure with possible hidden confounding, such as the propensity score adjusted regression explained in (Abdia well-known Fast Causal Inference (FCI) algorithm (Spirtes, et al. 2017) and follow the method in (Hirano and Imbens Meek, and Richardson 1995) and its extension tsFCI (Entner 2004) to handle continuous variables. and Hoyer 2010). We can adopt these algorithms for learn- Experimental Settings In the implementation of our ing the causal graph from data in practice. method, we use one hidden layer with 16 nodes for all gener- ators GV (·) and neural networks hV (·; ·), and use ReLU as Linear Causal Models: A Special Case the activation function. For the discriminator, we adopt the In some situations we may assume that the mapping from framework of Wasserstein GAN (Arjovsky, Chintala, and PAV to V for each variable V is a certain type of equation. Bottou 2017), which leverages the Wasserstein distance be- For example, linear causal models assume that all structural tween observational and generated distributions. The bene- equations in the model are linear. In this case, we can adopt fits of using the Wasserstein GAN are to stabilize the training a simplified generative model to encode this assumption. and provide a meaningful loss metric to indicate properties Specifically, for each variable V , we define the response of convergence. In addition, the adaptive gradient clipping function as the inner product between a parameter vector (Belghazi et al. 2018) is also used to stabilize the training. T The threshold η in the constraint is set to 0.001, and we take and the input, i.e., v = GV (zV ) · [paV , 1] . The parame- (1) (2) (|paV |) 50 solutions satisfying the constraint to compute the mean ter vector GV (zV ) = [θV , θV , ··· , θV , gV (zV )] con- (1) (2) and variance. The upper bound is computed as mean + std, sists of two part: one is a set of parameters θV , θV , ··· that and the lower bound is computed as mean − std. are irrelevant to the noise, and the other is a neural network gV (zV ) that maps the noise to a single variable. Upon the Synthetic Data definition of the response function, the joint distribution and We manually build a causal graph (shown in Figure3) with the ACE could be expressed, and the optimization problem 5 continuous endogenous variables, Z, X, W, V, Y . We as- could be formulated and solved similarly to Problem1. sume that the exogenous variables associated with X and Encoding the equation assumption can reduce the search V are confounded by a hidden variable U1; the exogenous space for finding causal models and shrink the bound- variables associated W and V are confounded by another ing range. The following proposition shows that, for lin- hidden variable U2; and other pairs of exogenous variables ear causal models, if there exists an instrumental variable are independent. To completely specify the causal model, all for ACE of A on B, i.e., a variable that affects A and af- exogenous variables are Gaussian or mixed-Gaussian noises fects B only by influencing A, then the ACE estimated from with zero mean and variance= 0.1. Problem1 would converge to a fixed value as the generated To illustrate the capability of coping with nonlinear causal distribution converges to the observational distribution. This model, we design the causal model as follows. result is consistent to the well-known instrumental variable z z formula (Bowden and Turkington 1984). Please refer to the u1 = 1, u2 = 2, z = Uniform(θ1, θ2) + z supplementary file for the proof. x x x w w 2 w x = θ0 + θ1 z + θ2 u1 + x, w = θ0 + θ1 x + θ2 u2 + w Proposition 1. Let C be an instrumental variable for ACE v = θv + θvu + θvu +  , y = θy + θyw + θyv +  of A on B, then both bounds computed from Problem1 will 0 1 1 2 2 v 0 1 2 y cov(B,C) where 1, 2, z, x, w, v, y are independent Gaussian converge to cov(A,C) (a1 − a0) if the generated distribution converges to the observational distribution. noises. After specifying the causal model, we generate 10,000 examples without hidden variables U1,U2 which are then used by all methods for estimating the ACE. The Experiments ground truth of the ACE is directly obtained by performing We implement and evaluate the proposed bounding method. the intervention on the causal model. For instrumental vari- We use both synthetic data and a real-world dataset, Adult able estimation, Z is treated as the instrumental variable. (Dheeru and Karra Taniskidou 2017). In addition to our method, we also evaluate following baseline methods for in- Results The experimental results are shown in Figure4. ferring the ACE. Previous bounding and estimation meth- ACEs are obtained by performing different interventions, ods (Wu et al. 2019; Louizos et al. 2017) are not included as i.e., E[Y |do(x1)] − E[Y |do(x0)] with different x1 and x0. they cannot handle continuous or high dimension treatment For demonstration, we select five interventions and report attributes. the corresponding ACEs. We see that our upper bound and lower bound cover the ground truth in all interventions. • Linear/: We build a linear/logistic re- However, other baseline methods cannot produce accurate gression on the outcome using all observed variables, and estimations and fall outside the bounds in most cases. This U1 V race age sex native_country U2

marital_status Z X W Y

edu_level Figure 3: The causal graph of synthetic data: Z, X, W, V, Y U ,U are observed variables and 1 2 are hidden variables. occupation hours_per_week workclass relationship

4 Ground truth income Upper bound

3 Lower bound

2 Propensity score Instrumental variable

1 Figure 5: The causal graph of the Adult dataset. 0

−1 Ground truth Lower bound Propensity score

0.7 Upper bound Logistic regression Average Causal Effect Average −2 0.6 −3 0.5 −4 0.4

x1=−3 x1=−2 x1=−1 x1=0 x1=1

Intervention 0.3 0.2

Figure 4: Average causal effects with different interventions 0.1 Average Causal Effect Average

(x0 = 2) on the nonlinear synthetic dataset. 0.0 −0.1 −0.2

may be due to the fact that their assumptions (e.g., strong −0.3 x0=1 x0=2 x0=3 x0=4 x0=5 x0=6 x0=7 x0=8 x0=9 x0=10 ignorability for propensity score, and linear assumption for Intervention linear regression and instrumental variable formula) do not hold in our general setting. Especially, note that when x1 = Figure 6: Average causal effects with different interventions 1, our bounds can help estimate the correct sign of the ACE, (x1 = 16) on the Adult dataset. while other methods produce opposite estimations. We also evaluate the special linear case as described in the previous section and obtain similar results. Please refer variable for the ACE of edu level on income, the in- to the supplementary file for details. strumental variable estimation is not performed. Results The experimental results are shown in Figure6. Adult Dataset ACEs are evaluated by performing different interventions We further conduct evaluations on the real-world Adult E[Y |do(X = x1)] − E[Y |do(X = x0)], and we report 10 dataset. The Adult dataset consists of 48,842 tuples with 11 interventions where x1 = 16 and x0 ranges from 1 to 10. attributes. We make use of the causal graph in (Zhang, Wu, In general, the ground truth decreases from 0.11 to 0.04 as and Wu 2018) shown in Figure5. With all the attributes ob- x0 increases. It falls in the range of the upper bound and served, we evaluate the ACE of edu level (which ranges lower bound in all interventions, which validate the efficacy from 0 to 16 with 0 the lowest level) on income (1 if of our method. However, other baseline methods including > 50K and 0 if ≤ 50K) as the ground truth by using the the logistic regression and propensity score cannot produce code from (Xu et al. 2019). accurate estimations and fall outside the bounds in all cases. In order to simulate hidden confounders, we deliberately hide 5 attributes (denoted by dotted nodes in Figure5), and Conclusions treat the remaining attributes (denoted by solid nodes) as ob- We proposed a bounding method for estimating average served to compute the ACE. In this setting, we can easily causal effects (ACEs) from observational data with hidden see that native country and race are two hidden con- confounding. The method parameterizes the causal model founders. Meanwhile, marital status is also a hidden using implicit generative models, and builds an adversarial confounder since it is a confounder between edu level network to formulate a constrained optimization problem for and occupation while the latter is an intermediate on a computing the bounds. We showed that encoding the linear causal path from edu level to income, resulting con- assumption can make the bounds converge to a fixed value. founding effects of edu level on income. As a result, Experiments using both synthetic and real-world datasets the ACE of edu level on income is unidentifiable. Our showed that our method provides more accurate estimations method is implemented using the sub-graph induced by the than several widely used causal inference methods. observed attributes. All baseline methods are computed us- Reproducibility. Our software together with the datasets ing observed attributes only. Since there is no instrumental used in this paper are available at this link. Ethics Statement Hirano, K.; and Imbens, G. W. 2004. The propensity score Causal inference is a fundamental technique for inferring with continuous treatments. Applied Bayesian modeling and causal effects among variables. In past years, it has been causal inference from incomplete-data perspectives 226164: widely adopted by computer scientists for addressing ethical 73–84. issues in artificial intelligence such as fair machine learning. Hu, H.; and Kerschberg, L. 2018. Evolving Medical On- These works usually rely on analyzing causal effects of sen- tologies Based on Causal Inference. In 2018 IEEE/ACM sitive attributes such as gender or race on decisions such as ASONAM’18, 954–957. IEEE. hiring or admission. Such analysis can be performed either on the training data so that the data could modified before Kallus, N.; Mao, X.; and Zhou, A. 2018. Interval estima- it is used for training, or on model outputs so that it could tion of individual-level causal effects under unobserved con- arXiv preprint arXiv:1810.02894 be used to regulate the learning procedure. However, when founding. . hidden confounders exist, how to estimating causal effects Kocaoglu, M.; Snyder, C.; Dimakis, A. G.; and Vishwanath, is a big challenge. Failing to correctly estimate causal ef- S. 2018. CausalGAN: Learning Causal Implicit Generative fects may undermine these causal-based fair machine learn- Models with Adversarial Training. In International Confer- ing techniques. The contribution of our paper is that it pro- ence on Learning Representations. vides a reliable approach for bounding the causal effects un- Kusner, M. J.; Loftus, J.; Russell, C.; and Silva, R. 2017. der the existence of hidden confounders. It can be potentially Counterfactual fairness. In Advances in NeurIPS’17, 4066– used to assess fairness of data or machine learning models 4076. and facilitate the research in fair machine learning. Li, Y.; Kuang, K.; Li, B.; Cui, P.; Tao, J.; Yang, H.; and Wu, Acknowledgments F. 2020. Continuous Treatment Effect Estimation via Gen- erative Adversarial De-confounding. In Proceedings of the This work was supported in part by NSF 1646654, 1920920, 2020 KDD Workshop on Causal Discovery, 4–22. PMLR. and 1946391. Louizos, C.; Shalit, U.; Mooij, J. M.; Sontag, D.; Zemel, R.; References and Welling, M. 2017. Causal effect inference with deep latent-variable models. In Advances in NeurIPS’17, 6446– Abdia, Y.; Kulasekera, K.; Datta, S.; Boakye, M.; and Kong, 6456. M. 2017. Propensity scores based methods for estimating average treatment effect and average treatment effect among Madras, D.; Creager, E.; Pitassi, T.; and Zemel, R. 2019. treated: A comparative study. Biometrical Journal 59(5): Fairness through causal awareness: Learning causal latent- 967–985. variable models for biased data. In Proceedings of FAT’19, Arjovsky, M.; Chintala, S.; and Bottou, L. 2017. Wasserstein 349–358. ACM. gan. arXiv preprint arXiv:1701.07875 . Miles, C. H.; Kanki, P.; Meloni, S.; and Tchetgen, E. J. T. Avin, C.; Shpitser, I.; and Pearl, J. 2005. Identifiability of 2015. On partial identification of the pure direct effect. arXiv path-specific effects. In Proceedings of IJCAI’05, 357–363. preprint arXiv:1509.01652 . Balke, A.; and Pearl, J. 1997. Bounds on treatment ef- Mohamed, S.; and Lakshminarayanan, B. 2016. Learn- fects from studies with imperfect compliance. Journal of ing in implicit generative models. arXiv preprint the American Statistical Association 92(439): 1171–1176. arXiv:1610.03483 . Belghazi, M. I.; Baratin, A.; Rajeswar, S.; Ozair, S.; Ben- Pearl, J. 2009. Causality. Cambridge university press. gio, Y.; Courville, A.; and Hjelm, R. D. 2018. Mine: Pearl, J. 2010. An introduction to causal inference. The mutual information neural estimation. arXiv preprint international journal of 6(2). arXiv:1801.04062 . Rubin, D. B. 2005. Causal inference using potential out- Bowden, R.; and Turkington, D. 1984. Instrumental Vari- comes: Design, modeling, decisions. Journal of the Ameri- ables. Cambridge, UK: Cambridge University Press. can Statistical Association 100(469): 322–331. Chiappa, S. 2019. Path-specific counterfactual fairness. In Shalit, U.; Johansson, F. D.; and Sontag, D. 2017. Estimat- Proceedings of AAAI’19, volume 33, 7801–7808. ing individual treatment effect: generalization bounds and Dheeru, D.; and Karra Taniskidou, E. 2017. UCI Ma- algorithms. In Proceedings of the 34th International Confer- chine Learning Repository. University of California, Irvine, ence on Machine Learning-Volume 70, 3076–3085. JMLR. School of Information and Computer Sciences. org. Entner, D.; and Hoyer, P. O. 2010. On causal discovery from Shpitser, I.; and Pearl, J. 2008. Complete identification data using FCI. Probabilistic graphical models methods for the causal hierarchy. Journal of Machine Learn- 121–128. ing Research 9(Sep): 1941–1979. Goodfellow, I.; Pouget-Abadie, J.; Mirza, M.; Xu, B.; Spirtes, P.; Meek, C.; and Richardson, T. 1995. Causal in- Warde-Farley, D.; Ozair, S.; Courville, A.; and Bengio, ference in the presence of latent variables and selection bias. Y. 2014. Generative adversarial nets. In Advances in In Proceedings of the Eleventh conference on Uncertainty in NeurIPS’14, 2672–2680. artificial intelligence, 499–506. Tian, J.; and Pearl, J. 2000. Probabilities of causation: Bounds and identification. In Proceedings of UAI’00, 589– 598. Morgan Kaufmann Publishers Inc. Tian, J.; and Pearl, J. 2002. A general identification condi- tion for causal effects. In AAAI/IAAI, 567–573. Wang, M. D.; and Wu, H. 2019. Tutorial: Causal Inference in Biomedical Data Analytics–Basics and Recent Advances. In Proceedings of ACM-BCB’19, 558–558. Wu, Y.; Zhang, L.; and Wu, X. 2019. Counterfactual fair- ness: Unidentification, bound and algorithm. In Proceedings of IJCAI’19, 10–16. Wu, Y.; Zhang, L.; Wu, X.; and Tong, H. 2019. PC-Fairness: A Unified Framework for Measuring Causality-based Fair- ness. In Advances in NeurIPS’19, 3399–3409. Xu, D.; Wu, Y.; Yuan, S.; Zhang, L.; and Wu, X. 2019. Achieving causal fairness through generative adversarial networks. In Proceedings of IJCAI’19. Yoon, J.; Jordon, J.; and van der Schaar, M. 2018. GANITE: Estimation of individualized treatment effects using genera- tive adversarial nets . Zhang, J.; and Bareinboim, E. 2018. Fairness in decision- making–the causal explanation formula. In 32nd AAAI Con- ference on Artificial Intelligence. Zhang, L.; Wu, Y.; and Wu, X. 2018. Causal modeling-based discrimination discovery and removal: Criteria, bounds, and algorithms. IEEE Transactions on Knowledge and Data En- gineering 31(11): 2035–2050. Zhao, J.; Runfola, D. M.; and Kemper, P. 2017. Simulation study in quantifying heterogeneous causal effects. In Winter Simulation Conference, 2650–2661.