Skip to content
BY 4.0 license Open Access Published by De Gruyter November 22, 2022

Causal effect on a target population: A sensitivity analysis to handle missing covariates

  • Bénédicte Colnet EMAIL logo , Julie Josse , Gaël Varoquaux and Erwan Scornet EMAIL logo

Abstract

Randomized controlled trials (RCTs) are often considered the gold standard for estimating causal effect, but they may lack external validity when the population eligible to the RCT is substantially different from the target population. Having at hand a sample of the target population of interest allows us to generalize the causal effect. Identifying the treatment effect in the target population requires covariates to capture all treatment effect modifiers that are shifted between the two sets. Standard estimators then use either weighting (IPSW), outcome modeling (G-formula), or combine the two in doubly robust approaches (AIPSW). However, such covariates are often not available in both sets. In this article, after proving L 1 -consistency of these three estimators, we compute the expected bias induced by a missing covariate, assuming a Gaussian distribution, a continuous outcome, and a semi-parametric model. Under this setting, we perform a sensitivity analysis for each missing covariate pattern and compute the sign of the expected bias. We also show that there is no gain in linearly imputing a partially unobserved covariate. Finally, we study the substitution of a missing covariate by a proxy. We illustrate all these results on simulations, as well as semi-synthetic benchmarks using data from the Tennessee student/teacher achievement ratio (STAR), and a real-world example from critical care medicine.

1 Introduction

1.1 Context

Randomized controlled trials (RCTs) are often considered the gold standard for estimating causal effects [1]. Yet, they may lack external validity, when the population eligible to the RCT is substantially different from the target population of the intervention policy [2]. Indeed, if there are treatment effect modifiers with a different distribution in the target population than that in the trial, some form of adjustment of the causal effects measured on the RCT is necessary to estimate the causal effect in the target population. Using covariates present in both RCT and an observational sample of the target population, this target population average treatment effect (ATE) can be identified and estimated with a variety of methods [3,4, 5,6,7, 8,9,10, 11,12,13, 14,15], reviewed in refs [16,17].

In this context, two main approaches exist to estimate the target population ATE from an RCT. The inverse probability of sampling weighting (IPSW) reweights the RCT sample so that it resembles the target population with respect to the necessary covariates for generalization, while the G-formula models the outcome, using the RCT sample, with and without treatment conditionally on the same covariates, and then marginalizes the model to the target population of interest. These two methods can be combined in a doubly robust approach – augmented inverse probability of sampling weighting (AIPSW) – which enjoys better statistical properties. These methods rely on covariates to capture the heterogeneity of the treatment and the population distributional shift. But the datasets describing the RCT and the target population are seldom acquired as part of a homogeneous effort, and as a result, they come with different covariates [6,18,19, 20,21,22]. Restricting the analysis to the covariates in common raises the risk of omitting an important one leading to identifiability issues. Controlling biases due to unobserved covariates are of crucial importance for causal inference, where it is known as sensitivity analysis [23,24,25].

1.2 Prior work

The problem of missing covariates is central in causal inference as, in an observational study, one can never prove that there is no hidden confounding. In that setting, sensitivity analysis strives to assess how far confounding would affect the conclusion of a study (e.g., would the ATE be of a different sign with such a hidden confounder). Such approaches date back to a study on the effect of smoking on lung cancer [23], and have been further developed for both parametric [24,25,26, 27,28] and semi-parametric situations [29,30]. Typically, the analysis translates expert judgment into mathematical expression of how much the confounding affects the treatment assignment and the outcome, and finally how much the estimated treatment effect is biased. In practice, the expert must usually provide sensitivity parameters that reflect plausible properties of the missing confounder. Classic sensitivity analysis, dedicated to ATE estimation from observational data, use as sensitivity parameters the impact of the missing covariate on treatment assignment probability along with the strength on the outcome of the missing confounder. However, given that these quantities are hardly directly transposable when it comes to generalization, these approaches cannot be directly applied to estimate the population treatment effect. These parameters have to be respectively replaced by the covariate shift and the strength of a treatment effect modifier. Existing sensitivity analysis methods for generalization usually consider a completely unobserved covariate. Ref. [31] rely on a logistic model for sampling probability and a linear generative model of the outcome. Ref. [32] propose a sensitivity analysis assuming a model on the identification bias of the conditional average treatment effect. Very recent works propose two other approaches: (i) Ref. [33] rely on the IPSW estimator and bound the error on the density ratio and then derive the bias on the ATE following the spirit of ref. [25]; (ii) Ref. [34] present a method with very few assumptions on the data generative process leading to three sensitivity parameters, including the variance of the treatment effect. As the analysis starts from two data sets, the missing covariate can also be partially observed in one of the two data sets, which opens the door to new dedicated methods, in addition to sensitivity methods for totally missing covariates. Following this observation, refs [35,36] handle the case where a covariate is present in the RCT but not in the observational data set, and establish a sensitivity analysis under the hypothesis of a linear generative model for the outcome. When the missing covariate is partially observed, practitioners sometimes impute missing values based on other observed covariates, though this approach is poorly documented. For example, [19] impute a partially observed covariate in a clinical study using a range of plausible distributions. Imputation has also been used in the context of individual participant data in meta-analysis [37,38].

1.3 Contributions

In this work, we investigate the problem of a missing covariate that affects the identifiability of the target population average treatment effect (ATE), a common situation when combining different data sources. This work comes after the identifiability assessment, that is, we consider that the necessary set of covariates to generalize is known, but a necessary covariate is totally or partially missing. Section 2 recalls the context along with the generic notations and assumptions used when coming to generalization. In Section 3, we quantify the bias due to unobserved covariates under the assumption of a semi-parametric generative process, considering a linear conditional average treatment effect (CATE) and under a transportability assumption of links between covariates in both populations. This bias is not estimator specific and remains valid for the IPSW, G-formula, and AIPSW estimators. We also prove that a linear imputation of a partially missing covariate cannot replace a sensitivity analysis. As mentioned in Section 1, and unlike classic sensitivity analysis, several missing data patterns can be observed: either totally missing or missing in one of the two sets. Therefore, Section 3 provides sensitivity analysis frameworks for all the possible missing data patterns, including the case of a proxy variable that would replace the missing one. These results can be useful for users as they may be tempted to consider the intersection of common covariates between the RCT and the observational data. We detail how the different patterns involve either one or two sensitivity parameters. To give users an interpretable analysis, and due to the specificity of the sensitivity parameters at hands, we propose an adaptation of sensitivity maps [24] that are commonly used to communicate sensitivity analysis results. Section 4 presents an extensive synthetic simulation analysis that illustrates theoretical results along with a semi-synthetic data simulation using the Tennessee student/teacher achievement ratio (STAR) experiment evaluating the effect of class size on children performance in elementary schools [39]. Finally, Section 5 provides a real-world analysis to assess the effect of acid tranexomic on the disability rating score (DRS) for trauma patients when a covariate is totally missing.

2 Problem setting: generalizing a causal effect

This section recalls the complete case context and identification assumptions. Any reader familiar with the notations and willing to jump to the sensitivity analysis can directly go to Section 3.

2.1 Notations

Notations are grounded on the potential outcome framework [1]. We model each observation in the RCT or observational population as described by a random tuple ( X i , Y i ( 0 ) , Y i ( 1 ) , A i , S i ) for i { 1 , , n } drawn from a distribution ( X , Y ( 0 ) , Y ( 1 ) , A , S ) R p × R 2 × { 0 , 1 } 2 , such that the observations are iid. For each observation, X i is a p -dimensional vector of covariates, A i denotes the binary treatment assignment (with A i = 1 if treated and A i = 0 otherwise), Y i ( a ) is the continuous outcome had the subject been given treatment a (for a { 0 , 1 } ), and S i is a binary indicator for RCT eligibility (i.e., meet the RCT inclusion and exclusion criteria) and willingness to participate if being invited to the trial ( S i = 1 if eligible and S i = 0 if not). Assuming consistency of potential outcomes, and no interference between treated and nontreated subject (SUTVA assumption), we denote by Y i = A i Y i ( 1 ) + ( 1 A i ) Y i ( 0 ) the observed outcome under treatment assignment A i .

Assuming the potential outcomes are integrable, we define the conditional average treatment effect (CATE):

x X , τ ( x ) = E [ Y ( 1 ) Y ( 0 ) X = x ] ,

and the population average treatment effect (ATE):

τ = E [ Y ( 1 ) Y ( 0 ) ] = E [ τ ( X ) ] .

Unless explicitly stated, all expectations are taken with respect to all variables involved in the expression. We model the patients belonging to an RCT sample of size n and in an observational data sample of size m by n + m independent random tuples: { X i , Y i ( 0 ) , Y i ( 1 ) , A i , S i } i = 1 n + m , where the RCT samples i = 1 , , n are identically distributed according to P ( X , Y ( 0 ) , Y ( 1 ) , A , S S = 1 ) , and the observational data samples i = n + 1 , , n + m are identically distributed according to P ( X , Y ( 0 ) , Y ( 1 ) , A , S ) . We also denote = { 1 , , n } the index set of units observed in the RCT study, and O = { n + 1 , , n + m } the index set of units observed in the observational study.

For each RCT sample i , we observe ( X i , A i , Y i , S i = 1 ) , while for observational data i O , we consider the setting where we only observe the covariates X i , which is a common case in practice. A typical data set is presented in Figure 1.

Figure 1 
                  Typical data structure, where a covariate would be available in the RCT, but not in the observational data set (left) or the reverse situation (right). In this specific example, 
                        
                           
                           
                              obs
                              =
                              
                                 {
                                 
                                    1
                                    ,
                                    2
                                 
                                 }
                              
                           
                           {\rm{obs}}=\left\{1,2\right\}
                        
                      (
                        
                           
                           
                              mis
                              =
                              
                                 {
                                 
                                    3
                                 
                                 }
                              
                           
                           {\rm{mis}}=\left\{3\right\}
                        
                     ), corresponds to common (resp. different) covariates in the two datasets.
Figure 1

Typical data structure, where a covariate would be available in the RCT, but not in the observational data set (left) or the reverse situation (right). In this specific example, obs = { 1 , 2 } ( mis = { 3 } ), corresponds to common (resp. different) covariates in the two datasets.

Because the RCT sample and observational data do not follow the same covariate distribution, the ATE τ is different from the RCT’s (or sample[1]) average treatment effect τ 1 , which can be expressed as follows:

τ τ 1 , where τ 1 E [ Y ( 1 ) Y ( 0 ) S = 1 ] .

This difference is the core of the lack of external validity introduced in the beginning of the work, but formalized with a mathematical expressions[2]. Throughout the article, we denote μ a ( x ) E [ Y ( a ) X = x ] the conditional mean outcome under treatment a { 0 , 1 } (also called responses surfaces). and e 1 ( x ) P ( A = 1 X = x , S = 1 ) the propensity score in the RCT population. This function is imposed by the trial characteristics and is usually a constant denoted by e 1 (other cases include stratified RCT trials).

For notational clarity, estimators are indexed by the number of observations used for their computation. For instance, response surfaces can be estimated using controls and treated individuals in the RCT to obtain, respectively, μ ˆ 0 , n and μ ˆ 1 , n . Similarly, we denote by τ ˆ n an estimator of τ depending only on the RCT samples (e.g., the difference-in-means estimator), and by τ ˆ n , m an estimator computed using both datasets.

2.2 Identifiability (or causal) assumptions

The consistency of treatment assignment assumption ( Y = A Y ( 1 ) + ( 1 A ) Y ( 0 ) ) has already been introduced in Section 2. To ensure the internal validity of the RCT, we need to assume randomization of treatment assignment and positivity of trial treatment assignment.

Assumption 1

(Treatment randomization within the RCT) a { 0 , 1 } , Y ( a ) A S = 1 , X .

In some cases, the trial is said to be completely randomized, that is, a { 0 , 1 } , Y ( a ) A S = 1 , thus removing any potential stratification of the treatment assignment.

Assumption 2

(Positivity of trial treatment assignment) η 1 > 0 , x X , η 1 e 1 ( x ) 1 η 1 .

Under these two assumptions, along with the SUTVA assumption (see, e.g., [1]), the most classical difference-in-means estimator is consistent for τ 1 . To generalize the RCT estimate to the target population, three additional assumptions are required for the identification of the target population ATE τ .

Assumption 3

(Representativity of observational data) For all i O , X i P ( X ) , where P is the target population distribution.

Then, a key assumption concerns the set of covariates that allows the identification of the target population treatment effect. This implies a conditional independence relation being called the ignorability assumption on trial participation or S-ignorability [3,5,8,11,20,21,36,41,42].

Assumption 4

(Ignorability assumption on trial participation – [5]) Y ( 1 ) Y ( 0 ) S X .

Assumption 4 indicates that covariates X needed to generalize correspond to covariates being both treatment effect modifiers and subject to a distributional shift between the RCT sample and the target population. Different strategies have been proposed to assess whether a treatment effect is constant and usually relies on marginal variance, CDFs, or quantiles comparison [43]. Other techniques are possible such as comparing Var [ Y X obs , A = 1 , S = 1 ] to Var [ Y X obs , A = 0 , S = 1 ] , to assess whether an important treatment effect modifier is missing. In our work, we assume that the user is aware of which variables are treatment effect modifiers and subject to a distributional shift. We call these covariates as key covariates.

Assumption 5

(Positivity of trial participation – [5]) There exists a constant c such that for all x with probability 1, P ( S = 1 X = x ) c > 0 .

2.3 Estimation strategies

To transport the ATE, several methods exist: the G-formula [6,44,45], inverse propensity weighting score (IPSW) [4,13,44], and the augmented IPSW (AIPSW) estimators. Note that other methods exist, such as calibration [15,46]. For example, the G-formula estimator consists in modeling the expected values for each potential outcome, conditional on the covariates.

Definition 1

(G-formula – [47]) The G-formula is denoted τ ˆ G , n , m , and defined as follows:

(1) τ ˆ G , n , m = 1 m i = n + 1 n + m ( μ ˆ 1 , n ( X i ) μ ˆ 0 , n ( X i ) ) ,

where μ ˆ a , n ( X i ) is an estimator of μ a ( X i ) obtained on the RCT sample. These intermediary estimates are called nuisance components.

Beyond causal assumptions stated earlier, the behavior of the G-formula estimator strongly depends on that of the surface response estimators μ ˆ a , n for a { 0 , 1 } . To analyze the G-formula, we introduce following assumptions on the consistency of the nuisance parameters μ ˆ 0 , n and μ ˆ 1 , n .

Assumption 6

(Consistency of surface response estimators) Denote μ ˆ 0 , n (respectively μ ˆ 1 , n ) an estimator of μ 0 (respectively μ 1 ). Let D n the RCT sample, so that

(H1-G) For a { 0 , 1 } , E [ μ ˆ a , n ( X ) μ a ( X ) D n ] p 0 when n ,

(H2-G) For a { 0 , 1 } , there exist C 1 , N 1 so that for all n N 1 , almost surely, E [ μ ˆ a , n 2 ( X ) D n ] C 1 .

Proposition 1

(Informal – L 1 -consistency of G-formula, IPSW, and AIPSW) Under causal assumptions (Assumptions 1, 2, 3, 4, and 5) and Assumption 6, the G-formula is L 1 -consistent (asymptotically unbiased). In appendix, we recall definitions of IPSW and AIPSW estimators and give the precise conditions under which L 1 -consistency of those estimators is achieved.

Proofs and a more formal statement are presented in Sections A and B. The sensitivity analysis presented later holds for any L 1 -consistent estimator.

3 Impact of a missing key covariate for a linear CATE

3.1 Situation of interest: a missing covariate in one dataset

We study the common situation where both data sets (RCT and observational) contain a different subset of the total covariates X . X can be decomposed as X = X mis X obs , where X obs denotes the covariates that are present in both data sets, the RCT and the observational study. X mis denotes the covariates that are either partially observed in one of the two data sets or totally unobserved in both data sets. We do not consider (sporadic) missing data problems as in [48], but only cases where the covariate is totally observed or not per data sources. We denote by obs (resp. mis) the index set of observed (resp. missing) covariates. An illustration of a typical data set is presented in Figure 1, with an example of two missing data patterns.

In our context, due to (partially) unobserved covariates, estimators of the target population ATE may be implemented on X obs only. To make the notations clear, we add a subscript obs on any estimator applied on the set X obs rather than X . Such estimators may suffer from bias due to Assumption 4 violation, that is:

Y ( 1 ) Y ( 0 ) S X but Y ( 1 ) Y ( 0 ) ⊥̸ S X obs .

We denote τ ˆ n , m , obs any generalization estimator (G-formula, IPSW, AIPSW) applied on the covariate set X obs rather than X .

3.2 Expression of the missing covariate bias

3.2.1 Model and hypothesis

To analyze the effect of a missing covariate, we introduce a nonparametric generative model. In particular, we focus on zero-mean additive-error representation, where the CATE depends linearly on X . We admit that there exist δ R p , σ R + , and a function g : X R , such that:

(2) Y = g ( X ) + A X , δ + ε , where ε N ( 0 , σ 2 ) ,

assuming τ ( X ) X , δ . In appendix (see Section D), we prove why this assumption on the generative model for Y does not come with a loss of generality.

Under this model, the average treatment effect (ATE) takes the following form:

τ = E [ Y ( 1 ) Y ( 0 ) X = x ] f ( x ) d x = δ , x f ( x ) d x = δ T E [ X ] .

Only variables that are both treatment effect modifier ( δ j 0 ) and subject to a distributional change between the RCT and the target population are necessary to generalize the ATE. If some of these key covariates are missing, the estimation of the target population ATE will be biased. Our goal here is to express the bias of a missing variable on the transported ATE. But first, we have to specify a context in which a certain permanence of the relationship between X obs and X mis in the two data sets holds. Therefore, we introduce the transportability of covariate relationship assumption.

Assumption 7

(Transportability of covariate relationship) The distribution of X is Gaussian, that is, X N ( μ , Σ ) , and transportability of Σ is true, that is, X S = 1 N ( μ R C T , Σ ) .

This assumption, and in particular, the transportability of Σ , is of major importance for the sensitivity analysis we develop below. Indeed, as soon as the correlation pattern changes in amplitude and sign between the two populations, the sensitivity analysis can be invalidated. The plausibility of Assumption 7 can be partially assessed through a statistical test on Σ obs , obs , for example, a Box’s M test [49], supported with vizualizations [50]. A discussion can be found in the experimental study (Section 4) and in appendix (Section G), showing that this assumption is plausible in many situations.

3.2.2 Main result

Theorem 1

Assume that Assumptions 1, 2, 3, 4, and 5 (identifiability) hold, along with Model (2) and (7) (sensitivity model). Let B be the following quantity:

(3) B = j mis δ j ( E [ X j ] E [ X j S = 1 ] Σ j , obs Σ obs , obs 1 ( E [ X obs ] E [ X obs S = 1 ] ) ) ,

where Σ obs , obs is the submatrix of Σ composed of rows and columns corresponding to variables present in both data sets. Similarly, Σ j , obs is composed of the jth row of Σ and has columns corresponding to variables present in both data sets. Consider a procedure τ ˆ n , m that estimates τ with no asymptotic bias (e.g., the G-formula introduced in Definition 1under Assumption 6). Let τ ˆ n , m , obs be the same procedure but trained on observed data only. Then

(4) lim n , m E [ τ ˆ n , m , obs ] τ = B .

Proof is given in appendix (see Section C).

3.2.2.1. Comment on L 1 -consistency

Theorem 1 is valid for any L 1 -consistent generalization estimator. In particular, we provide in appendix the detailed assumptions (similar as Assumption 6) under which two other popular estimators, IPSW and AIPSW, are asymptotically unbiased (see Section A). Note that most of the existing works on estimating the target population causal effect focus on identification or establish consistency for parametric models or oracle estimators which are not bona fide estimation procedures as they require knowledge of some population data-generation mechanisms [4,5,13,21,45,51,52]. To our knowledge, no general L 1 -consistency results for the G-formula, IPSW, and AIPSW procedures are available in a nonparametric case, when either the CATE or the weights are estimated from the data without prior knowledge.

3.2.2.2. What if outcomes are also available in the observational sample?

Who can do more can do less; therefore, this outcome covariate could be dropped and the analysis conducted without it. But alternative strategies exist. First, the outcome in the observational data – even if present in only one of the treatment group – would allow to test for the presence or absence of a missing treatment effect modifier [17] (see their Section 4.2), and therefore its strength. Moreover this would allow to rely on strategies to diminish the variance of the estimates [34]. Finally, the assumption of a linear CATE could be reconsidered and softened, but we let this question to future work.

3.3 Sensitivity analysis

The above theoretical bias B (see equation (3)) can be used to translate expert judgments about the strength of missing covariates, which corresponds to sensitivity analysis. In the rest of our work, we exemplify Theorem 1 in scenarios for which there is a totally unobserved covariate (Section 3.3.1), a missing covariate in RCT (Section 3.3.2.1), or a missing covariate in the observational sample (Section 3.3.2.1). Section 3.3.3 completes the previous sections presenting an adaptation to sensitivity maps. Finally, Section 3.3.4 details the imputation case, and Section 3.3.5 presents the case of a proxy variable. All these methods rely on different assumptions recalled in Table 1.

Table 1

Summary of the assumptions and results pointer for all the sensitivity methods according to the missing covariate pattern when the generative outcome is semi-parametric with a linear CATE (2)

Missing covariate pattern Assumption(s) required Procedure’s label
Totally unobserved covariate X mis X obs 1
Partially observed in observational study Assumption 7 2
Partially observed in RCT No assumption 3
Proxy variable Assumptions 7 and 8 5

3.3.1 Sensitivity analysis when a key covariate is totally unobserved

When a covariate is totally unobserved, a common and natural assumption is to assume independence between this covariate and the observed ones [24]. Although strong, this assumption allows us to estimate the identification bias.

Corollary 1

(Sensitivity model) Assume that Model (2) holds, along with Assumptions 1, 2, 3, 4, 5, and 7. Assume also that X mis X obs , X mis X obs S = 1 . Consider a procedure τ ˆ n , m that estimates τ with no asymptotic bias. Let τ ˆ n , m , obs be the same procedure but trained on observed data only. Then

lim n , m E [ τ ˆ n , m , obs ] τ = δ mis Δ m ,

where Δ m = E [ X mis ] E [ X mis S = 1 ] .

Corollary 1 is a direct consequence of Theorem 1, particularized for the case where X obs X mis and X obs X mis S = 1 . In this expression, Δ m and δ mis are called the sensitivity parameters. To estimate the bias implied by an unobserved covariate, we have to determine how strongly X mis is a treatment effect modifier (through δ mis ), and how strongly it is linked to the trial inclusion (through the shift between the trial sample and the target population Δ m = E [ X mis ] E [ X mis S = 1 ] ). Table 2 summarizes the similarities and differences with approaches of [24,31] approaches, and our approach.

Table 2

Summary of the differences among [24] method, being a prototypical method for sensitivity analysis for observational data and hidden counfounding, [31] method, and our method

[24] [31] Sensitivity model
Assumption on covariates X mis X obs X mis X obs X mis X obs
Model on Y Linear model Linear model Linear CATE (2)
Other assumption Model on A (logit) Model on S (logit)
First sensitivity parameter Strength on Y , using δ mis Strength on Y , using δ mis Strength on Y , using δ mis
Second sensitivity parameter Strength on A (logit’s coefficient) Strength on S (logit’s coefficient) Δ m : shift of X mis

In the setting of Corollary 1, sensitivity analysis can be carried out using Procedure 1 described below . To represent the bias magnitude as a function of the sensitivity parameters, we develop a graphical aid adapted from sensitivity maps [24,30], see Section 3.3.3.

Procedure 1: A totally unobserved covariate
init : δ mis [ ] ; // Define a range for plausible δ mis values
init : Δ m [ ] ; // Define a range for plausible Δ m values
Compute all possible bias δ mis Δ m (see Lemma 1)
return Sensitivity map

A partially observed covariate could always be removed so that this sensitivity analysis could be conducted for every missing data patterns (the variable being missing in the RCT or in the observational data). However, dropping a partially observed covariate (i) is inefficient as it discards available information and, (ii) amounts to considering the variable as totally unobserved, which, in turn, leads us to assume independence between observed and unobserved covariates, a very strong hypothesis. Therefore, in the following subsections, we propose methods that use the partially observed covariate – when available – to improve the bias estimation.

3.3.2 Sensitivity analysis when a key covariate is partially observed

When partially available, we propose to use X mis to have a better estimate of the bias. Unlike mentioned earlier, this approach does not need the partially observed covariate to be independent of all other covariates, but rather captures the dependencies from the data.

3.3.2.1. Observed in observational study

Suppose one key covariate X mis is observed in the observational study, but not in the RCT. Under Assumption 7, the asymptotic bias of any L 1 -consistent estimator τ ˆ n , m , obs is derived in Theorem 1. The quantitative bias is informative as it depends only on the regression coefficients δ and on the shift in expectation between covariates. Indeed, the bias term can be decomposed as follows:

B = δ mis X mis ’s strength ( E [ X mis ] E [ X mis S = 1 ] Shift of  X mis : Δ m Σ mis , obs Σ obs , obs 1 ( E [ X obs ] E [ X obs S = 1 ] ) Can be estimated from the data ) .

By using the observational study where the necessary covariates are all observed, one can estimate the covariance term Σ mis , obs Σ obs , obs 1 together with the shift for the observed set of covariates. Unfortunately, the remaining parameters δ mis , corresponding to the coefficient of the missing covariates in the complete linear model, and Δ m = E [ X mis ] E [ X mis S = 1 ] are not identifiable from the observed data. These two parameters correspond respectively to the strength of the treatment effect modifier and the distributional shift of the missing covariate. These two quantities are used as sensitivity parameters to estimate a plausible range of the bias (see Procedure 2). Simulations illustrate how these sensitivity parameters can be used, along with graphical visualization derived from sensitivity maps (see Section 4).

Procedure 2: Observed in observational
init : δ mis [ ] ; // Define a range for plausible δ mis values
init : Δ mis [ ] ; // Define a range for plausible Δ mis values
Estimate Σ obs , obs , Σ mis , obs , and E [ X obs ] on the observational dataset;
Estimate E [ X obs S = 1 ] on the RCT dataset;
Compute all possible biases for the predefined ranges of δ mis and Δ mis , according to Theorem 1.
return Sentivity map

3.3.2.1.1 Data-driven approach to determine sensitivity parameter. Note that guessing a good range for the shift Δ mis is probably easier than giving a range for the coefficients δ mis . We propose a data-driven method to estimate δ mis . First, learn a linear model of X mis from observed covariates X obs on the observational data, then impute the missing covariate in the trial, and finally obtain δ ˆ mis with a Robinson procedure on the imputed trial data [53,54,55]. The Robinson procedure is recalled in Appendix (see Section E). This method is used in the semi-synthetic simulation (see Section 4.2).

3.3.2.1.2 Observed in the RCT. The method we propose here was already developed in refs [35,36], and we briefly recall its principle in this section. Note that we extend this method by considering a semi-parametric model (2), while they considered a completely linear model. For this missing covariate pattern, only one sensitivity parameter is necessary. As the RCT is the complete data set, the regression coefficients δ of (2) can be estimated for all the key covariates, leading to an estimate δ ˆ mis for the partially unobserved covariate. Refs [35,36] showed that:

(5) τ = δ obs , E [ X obs ] + δ mis , E [ X mis ] Unknown .

In this case, and as the influence of X mis as a treatment effect modifier can be estimated from the data through δ ˆ mis , only one sensitivity parameter is needed, namely, E [ X mis ] . Therefore, we assume to be given a range of plausible values for E [ X mis ] , for example, according to a domain expert prior.

Note that δ mis can be estimated following a Robinson procedure. This allows extending [36]’s work to the semi-parametric case. Softening even more the parametric assumption where only X mis is additive in the CATE is a natural extension, but out of the scope of the present work.

Procedure 3: Observed in RCT
init : E [ X mis ] [ ] ; // Define a range for plausible values of E [ X mis ]
Estimate δ with the Robinson procedure, that is:
Run a nonparametric regression Y X on the RCT, and denote
m ˆ n ( x ) = E [ Y X = x , S = 1 ] the obtained estimator;
Define the transformed features Y ˜ = Y m ˆ n ( X ) and Z ˜ = ( A e 1 ( X ) ) X .
Estimate δ ˆ running the OLS regression on Y ˜ Z ˜ ;
Estimate E [ X obs ] on the observational dataset;
Compute all possible biases for the range of E [ X mis ] according to (5).
return Sensitivity map

3.3.3 Vizualization: sensitivity maps

From now on, each of the sensitivity method suppose to translate sensitivity parameter(s) and to compute the range of bias associated. A last step is to communicate or visualize the range of biases, which is slightly more complicated when there are two sensitivity parameters. Sensitivity map is a way to aid such judgement [24,30]. It consists in having a two-dimensional plot, each of the axis representing the sensitivity parameter, and the solid curve is the set of sensitivity parameters that leads to an estimate that induces a certain bias’ threshold. Here, we adapt this method to our settings with several changes. Because coefficients interpretation is hard, a typical practice is to translate a regression coefficient into a partial R 2 . For example, ref. [24], a prototypical example, proposes to interpret the two parameters with partial R 2 . In our case, a close quantity can be used:

(6) R 2 V [ δ mis X mis ] V j obs δ ˆ j X j ,

where the denominator term is obtained when regressing Y on X obs . If this R 2 coefficient is close to 1, then the missing covariate has a similar influence on Y compared to other covariates. On the contrary, if R 2 is close to 0, then the impact of X mis on Y as a treatment effect modifier is small compared to other covariates. But in our case one of the sensitivity parameters is really palpable as it is the covariate shift Δ m . We advocate keeping the regression coefficient and shift as sensitivity parameter rather than a R 2 to help practitioners as it allows to keep the sign of the bias, which can be in favor of the treatment and help interpreting the sensitivity analysis. Furthermore, even if postulating an hypothetical value of a coefficient is tricky, when the covariate is partially observed, an imputation procedure can be proposed to have a grasp of the coefficient’s true value.

In Figure 2, we present a glimpse of the simulation result, to introduce the principle of the sensitivity map, with on the left the representation using R 2 and on the right a representation keeping the raw sensitivity parameters. In this plot, we consider the covariate X 3 to be missing, so that we represent what would be the bias if we missed X 3 . The associated sensitivity parameters are represented on each axis. In other words, the sensitivity map shows how strong an unobserved key covariate would need to be to induce a bias that would force to reconsider the conclusion of the study because the bias is above a certain threshold, which is represented by the blue line. For example, in our simulation set-up, X 3 is below the threshold as illustrated in Figure 2. The threshold can be proposed by expert, and here, we proposed the absolute difference between τ ˆ n , m , obs and the RCT estimate τ ˆ 1 as a natural quantity. In particular, we observe that keeping the sign of the sensitivity parameter allows to be even more confident on the direction of the bias.

Figure 2 
                     Sensitivity maps: 
                           
                              
                              
                                 
                                    
                                       X
                                    
                                    
                                       3
                                    
                                 
                              
                              {X}_{3}
                           
                         is supposed to be a missing covariate. (Left) Regular sensitivity map showing how strong an key covariate would need to be to induce a bias of 
                           
                              
                              
                                 ∼
                                 6
                              
                               \sim 6
                           
                         in function of the two sensitivity parameters 
                           
                              
                              
                                 
                                    
                                       Δ
                                    
                                    
                                       m
                                    
                                 
                              
                              {\Delta }_{m}
                           
                         and partial 
                           
                              
                              
                                 
                                    
                                       R
                                    
                                    
                                       2
                                    
                                 
                              
                              {R}^{2}
                           
                         when a covariate is totally unobserved. (Right) The exact same simulation data are represented, while using rather 
                           
                              
                              
                                 
                                    
                                       δ
                                    
                                    
                                       mis
                                    
                                 
                              
                              {\delta }_{{\rm{mis}}}
                           
                         than the partial 
                           
                              
                              
                                 
                                    
                                       R
                                    
                                    
                                       2
                                    
                                 
                              
                              {R}^{2}
                           
                        , and superimposing the heatmap of the bias which allows to reveal the general landscape along with the sign of the bias.
Figure 2

Sensitivity maps: X 3 is supposed to be a missing covariate. (Left) Regular sensitivity map showing how strong an key covariate would need to be to induce a bias of 6 in function of the two sensitivity parameters Δ m and partial R 2 when a covariate is totally unobserved. (Right) The exact same simulation data are represented, while using rather δ mis than the partial R 2 , and superimposing the heatmap of the bias which allows to reveal the general landscape along with the sign of the bias.

3.3.4 Partially observed covariates: imputation

Another practically appealing solution is to impute the partially observed covariate, based on the complete data set (whether it is the RCT or the observational one) following Procedure 4. We analyze theoretically in this section the bias of such procedure in Corollary 2 and show that there is no gain in linearly imputing the partially observed covariate.

To ease the mathematical analysis, we focus on a G-formula estimator based on oracles quantities: the best imputation function and the surface responses are assumed to be known. While these are not available in practice, they can be approached with consistent estimates of the imputation functions and the surface responses. The precise formulations of our oracle estimates are given in Definitions 2 and 3.

Definition 2

(Oracle estimator when covariate is missing in the observational data set) Assume that the RCT is complete and that the observational sample contains one missing covariate X mis . We assume that we know

  1. the true response surfaces μ 1 and μ 0 .

  2. the true linear relation between X mis as a function of X obs .

Our oracle estimate τ ˆ G , , m , imp consists in applying the G-formula with the true response surfaces μ 1 and μ 0 (I) on the observational sample, in which the missing covariate has been imputed by the best (linear) function (II).

Definition 3

(Oracle estimator when covariate is missing in the RCT data set) Assume that the observational sample is complete and that the RCT contains one missing covariate X mis . We assume that we know

  1. the true linear relation between X mis as a function of X obs , which leads to the optimal imputation X ˆ mis ,

  2. the conditional expectations, E [ Y ( a ) X obs , X ˆ mis , S = 1 ] , for a { 0 , 1 } .

Our oracle estimate τ ˆ G , , , imp consists in optimally imputing the missing variable X mis in the RCT (I). Then, the G-formula is applied to the observational sample, with the surface responses that have been perfectly fitted on the completed RCT sample.

Corollary 2

(Oracle bias of imputation in a Gaussian setting) Assume that the CATE is linear (2) and that Assumption 7holds. Let B be the following quantity:

B = δ mis ( E [ X mis ] E [ X mis S = 1 ] Σ j , obs Σ obs , obs 1 ( E [ X obs ] E [ X obs S = 1 ] ) ) .

  1. Complete RCT. Assume that the RCT is complete and that the observational data set contains a missing covariate X mis . Consider the oracle estimator τ ˆ G , , m , imp in Definition 2. Then,

    τ lim m E [ τ ˆ G , , m , imp ] = B

  2. Complete Observational. Assume that the observational data set is complete and that the RCT contains a missing covariate X mis . Consider the oracle estimator τ ˆ G , , , imp in Definition 3. Then,

    τ E [ τ ˆ G , , , imp ] = B

Derivations are detailed in appendix (see Subsection C.2). Corollary 2 highlights that there is no gain in linearly imputing the missing covariate compared to dropping it. Simulations (Section F) show that the average bias of a finite-sample imputation procedure is similar to the bias of τ ˆ G , , , obs .

Procedure 4: Linear imputation
Model X mis a linear combination of X obs on the complete data set;
Impute the missing covariate with X ˆ mis with the previous fitted model;
Compute τ ˆ with the G-formula using the imputed data set X obs X ˆ mis ;
return τ ˆ

3.3.5 Using a proxy variable in place of the missing covariate

Another solution is to use a so-called proxy variable. The impact of a proxy in the case of a linear model is documented in econometrics [56,57, 58,59]. An example of a proxy variable is the height of children as a proxy for their age. Note that in this case, even if the age is present in one of the two datasets, only the children’s height is kept in for this method.

Here, we propose a framework to handle a missing key covariate with a proxy variable and estimate the bias reduction accounting for the additional noise brought by the proxy.

Assumption 8

(Proxy framework) Assume that X mis X obs , and that there exists a proxy variable, X prox such that

X prox = X mis + η

where E [ η ] = 0 , Var [ η ] = σ prox 2 , and Cov ( η , X mis ) = 0 . In addition, we suppose that Var [ X mis ] = Var [ X mis S = 1 ] = σ mis 2 .

Definition 4

Let τ ˆ G , n , m , prox be the G-formula estimator, where X mis is substituted by X prox as detailed in assumption 8.

Lemma 1

Assume that the generative linear model (2) holds, along with Assumption 7and the proxy framework (8). Then the asymptotic bias of τ ˆ G , n , m , prox is:

lim n , m E [ τ ˆ G , n , m , prox ] τ = δ mis Δ m 1 σ mis 2 σ mis 2 + σ prox 2 .

We denote δ ˆ prox the estimated coefficient for X prox . Such an estimation can be obtained using a Robinson procedure when regressing Y on the set X obs X prox .

Corollary 3

The asymptotic bias in Lemma 1can be estimated using the following expression:

lim n , m E [ τ ˆ G , n , m , prox ] τ = δ ˆ prox ( E [ X prox ] E [ X prox S = 1 ] ) σ prox 2 σ mis 2 .

Proofs of Lemma 1 and Corollary 3 are detailed in Appendix (Proof C.3). Note that, as expected, the average bias reduction strongly depends on the quality of the proxy. In the limit case, if σ prox 0 so that the correlation between the proxy and the missing variable is one, then the bias is null. In general, if σ prox σ mis , then the proxy variable does not diminish the bias.

Finally, we propose a practical approach in Procedure 5. Note that it requires to have a range of possible σ prox values. We recommend to use the data set on which the proxy along with the partially unobserved covariate are present, and to obtain an estimation of this quantity on this subset.

Procedure 5: Proxy variable
init : σ prox [ ] ; // Define a range for plausible σ prox values
if X mis is in RCT then
init : Δ mis [ ] ; / / Define a range for plausible Δ mis values Estimate δ mis with the Robinson procedure (see Procedure 3 for details) ; Compute all possible biases for the range of σ prox according to Lemma C.3 .
else
Estimate δ prox with the Robinson procedure (see Procedure 3 for details) ; Estimate E [ X prox ] and E [ X prox S = 1 ] ; Compute all possible bias for range of σ prox according to Corollary 3 .
return Biases’s range

4 Synthetic and semi-synthetic simulations

More information on simulation settings can be found in Appendix, see Section F

4.1 Synthetic simulations

While results presented in Section 3 apply to any function g (see (2)), we choose g as a linear function to illustrate our findings. All simulations are available on github[3], and include nonlinear forms for g .

4.1.1 Simulations parameters

We use a similar simulation framework as in refs [15,16], where five covariates are generated independently, except for X 1 and X 5 whose correlation is set at 0.8, except when explicitly mentioned. We simulate marginals as X j N ( 1 , 1 ) for all j = 1 , , 5 . The trial selection process is defined using a logistic regression model, such that:

(7) logit { P ( S = 1 X ) } = β s , 0 + β s , 1 X 1 + + β s , 5 X 5 .

This selection process implies that the variance–covariance matrix in the RCT sample and in the target population may be different depending on the (absolute) value of the coefficients β s . In our simulation set-up, the overall variance–covariance structure is kept identical as visualized on Figure 3. The outcome is generated according to a linear model, following Model 2, that is,

(8) Y ( a ) = β 0 + β 1 X 1 + + β 5 X 5 + a ( δ 1 X 1 + + δ 5 X 5 ) + ε with ε N ( 0 , 1 ) .

In this simulation, we set β = ( 5 , 5 , 5 , 5 , 5 ) , and other parameters as described in Table 3.

Figure 3 
                     
                        Variance–covariance preservation in the simulation set-up highlighted with pairwise covariance ellipses for one realization of the simulation (package heplots).
Figure 3

Variance–covariance preservation in the simulation set-up highlighted with pairwise covariance ellipses for one realization of the simulation (package heplots).

Table 3

Simulations parameters

Covariates X 1 X 2 X 3 X 4 X 5
Treatment effect modifier Yes Yes Yes No No
Linked to trial inclusion Yes No Yes Yes No
δ δ 1 = 30 δ 2 = 30 δ 3 = 10 δ 4 = 0 δ 5 = 0
β s β s , 1 = 0.4 β s , 2 = 0 β s , 3 = 0.3 β s , 4 = 0.3 β s , 5 = 0
. X 1 X 2 X 1 X 3 X 1 X 4 X 1 X 5 ⊥̸ X 1

First, a sample of size 10,000 is drawn from the covariate distribution. From this sample, the selection model (7) is applied, which leads to an RCT sample of size n 2,800 . Then, the treatment is generated according to a Bernoulli distribution with probability equal to e 1 = 0.5 . Finally, the outcome is generated according to (8). The observational sample is obtained by drawing a new sample of size m = 10,000 from the covariate distribution. In this setting, the ATE equals τ = j = 1 5 δ j E [ X j ] = j = 1 5 δ j = 50 . Besides, the sample selection ( S = 1 ) in (7) is biased toward lower values of X 1 (and indirectly X 5 ) and higher values of X 3 . This situation illustrates a case where τ 1 τ . Empirically, we obtain τ 1 44 .

4.1.2 Illustration of Theorem 1

Figure 4 presents results of a simulation with 100 repetitions with no missing covariates (see none in the figure), and the impact of missing covariate(s) when using the G-formula or the IPSW to generalize. The theoretical bias from Theorem 1 is also represented.

Figure 4 
                     
                        Illustration of Theorem 
                        1: Simulation results for the linear model with missing covariate(s) when generalizing the treatment effect using the G-formula (Definition 1) or IPSW (see Definition A1 in appendix) estimators on the set of observed covariates. Missing covariate are indicated on the x-axis. The theoretical bias (orange dot) is obtained from Theorem 1. Simulations are repeated 100 times.
Figure 4

Illustration of Theorem 1: Simulation results for the linear model with missing covariate(s) when generalizing the treatment effect using the G-formula (Definition 1) or IPSW (see Definition A1 in appendix) estimators on the set of observed covariates. Missing covariate are indicated on the x-axis. The theoretical bias (orange dot) is obtained from Theorem 1. Simulations are repeated 100 times.

The absence of covariates X 2 , X 4 , and/or X 5 does not affect ATE generalization because these covariates are not simultaneously treatment effect modifiers and shifted (between the RCT sample and the target population). In addition, the signs of the biases depend on the signs of the coefficients associated with the missing variables, as highlighted by settings for which X 1 and X 3 are missing. As shown in Theorem 1, variables acting on Y without being treatment effect modifiers and linked to trial inclusion can help to reduce the bias, if correlated to a (partially-) unobserved key covariate. This is stressed out in our experiment by comparing the settings for which X 1 , X 5 are missing and the one where only X 1 is missing.

4.1.3 A totally unobserved covariate (from Section 3.3.1)

To illustrate this case, the missing covariate has to be supposed independent of all the others. Here, we consider X 3 . Then, according to Lemma 1, the two sensitivity parameters δ mis and the shift Δ m can be used to produce a sensitivity map for the bias on the transported ATE. Procedure 1 summarizes the different steps, and the sensitivity map’s output result is presented in Figure 2.

4.1.4 A missing covariate in the RCT (from Section 3.3.2.1)

In this case, we need to specify ranges of values for the two sensitivity parameters δ mis and Δ m . The experimental protocol is designed such that all covariates are successively partially missing in the RCT. Because each missing variable implies a different landscape due to the dependence relation to other covariates (as stated in Theorem 1), each variable requires a different heatmap (except if covariates are all independent). Results are depicted in Figure 5. Figure 5 illustrates the benefit of Protocol 2 accounting for other correlated covariates, and compared to a protocol assuming independent covariates. Indeed, X 1 and X 2 are strong treatment effect modifiers (see Table 3, where δ 1 = δ 2 ), but X 1 is correlated with other completely observed covariates, which allows to “lower” the bias if X 1 is completely removed from the analysis compared to a similar covariate that would be independent of all other covariates. This is highlighted with a nonsymmetric bias landscape for X 1 in Figure 5. As a consequence, for a same value of δ mis value, a guessed shift of Δ mis = 0.25 allows to conclude on a lower bias on the map for X 1 , while it would not be the case for covariate X 2 (which is completely independent).

Figure 5 
                     
                        Simulations results when applying procedure 2: Heatmaps with a blue curve showing how strong an unobserved key covariate would need to be to induce a bias of 
                           
                              
                              
                                 
                                    
                                       τ
                                    
                                    
                                       1
                                    
                                 
                                 −
                                 τ
                                 
                                 ∼
                                 
                                 −
                                 6
                              
                              {\tau }_{1}-\tau \hspace{0.33em} \sim \hspace{0.33em}-6
                           
                         in function of the two sensitivity parameters 
                           
                              
                              
                                 
                                    
                                       Δ
                                    
                                    
                                       m
                                    
                                 
                              
                              {\Delta }_{m}
                           
                         and 
                           
                              
                              
                                 
                                    
                                       δ
                                    
                                    
                                       mis
                                    
                                 
                              
                              {\delta }_{{\rm{mis}}}
                           
                         when a covariate is totally unobserved. Each heatmap illustrates a case where the covariate would be missing (indicated on the top of the map), given all other covariates. The cross indicate the coordinate of true sensitivity parameters, in adequation with the bias empirically observed in Figure 4. The bias landscape depends on the dependence of the covariate with other observed covariates, as illustrated with an asymmetric heatmap when 
                           
                              
                              
                                 
                                    
                                       X
                                    
                                    
                                       1
                                    
                                 
                              
                              {X}_{1}
                           
                         is partially observed, due to the presence of 
                           
                              
                              
                                 
                                    
                                       X
                                    
                                    
                                       5
                                    
                                 
                              
                              {X}_{5}
                           
                        .
Figure 5

Simulations results when applying procedure 2: Heatmaps with a blue curve showing how strong an unobserved key covariate would need to be to induce a bias of τ 1 τ 6 in function of the two sensitivity parameters Δ m and δ mis when a covariate is totally unobserved. Each heatmap illustrates a case where the covariate would be missing (indicated on the top of the map), given all other covariates. The cross indicate the coordinate of true sensitivity parameters, in adequation with the bias empirically observed in Figure 4. The bias landscape depends on the dependence of the covariate with other observed covariates, as illustrated with an asymmetric heatmap when X 1 is partially observed, due to the presence of X 5 .

4.1.5 A missing covariate in the observational data (from Section 3.3.2.1)

In this case, we need to specify a range for the values of only one sensitivity parameter, namely, E [ X mis ] (see (5)). In our experimental protocol, we assume that X 1 is missing and apply Procedure 3 . Results are presented in Table 4.

Table 4

Simulations results when applying procedure 3: Results of the simulation considering X 1 being partially observed in the RCT, and using the sensitivity method of [35], but with a Robinson procedure to handle semi-parametric generative functions. When varying the sensitivity parameters, the estimated ATE is close to the true ATE ( τ = 50 ) when the sensitivity parameter is closer to the true one ( E [ X mis ] = 1 ). The results are presented for 100 repetitions

Sensitivity parameter E [ X mis ] 0.8 0.9 1.0 1.1 1.2
Empirical average τ ˆ G , n , m , obs 44 47 50 53 56
Empirical standard deviation τ ˆ G , n , m , obs 0.4 0.4 0.3 0.3 0.4

Simulations illustrating imputation (Corollary 2) and usage of a proxy (Lemma 1) are available in appendix, in Section F.

4.1.6 Violation of Assumption 7

To assess the impact of a lack of transportability of the variance–covariance matrix (Assumption 7), we propose to observe the effect of an increasing (in absolute value) coefficient involved in the sampling process (equation (7)). We observe that the bigger the coefficient, the bigger the deviations from the theory, as expected. To illustrate this phenomenon, we associate the logistic regression coefficient (the further away from the zero, the more Assumption 7 is unvalidated) with the p -value of a Box-M test assessing if the variance covariance matrix from the two sources are different. Empirically, the bias is still well estimated by procedures described in Section 3 even if the p -value is lower than 0.05. Results are available in Figures 6 and 7.

Figure 6 
                     Empirical link between the logistic regression coefficient for sampling bias 
                           
                              
                              
                                 
                                    
                                       β
                                    
                                    
                                       s
                                       ,
                                       1
                                    
                                 
                              
                              {\beta }_{s,1}
                           
                         and the p-value of a Box-M test. The average p-value is computed by repeating 50 times the simulation. We recall that in Figure 4, 
                           
                              
                              
                                 
                                    
                                       β
                                    
                                    
                                       s
                                       ,
                                       1
                                    
                                 
                                 ≔
                                 −
                                 0.4
                              
                              {\beta }_{s,1}:= -0.4
                           
                        .
Figure 6

Empirical link between the logistic regression coefficient for sampling bias β s , 1 and the p-value of a Box-M test. The average p-value is computed by repeating 50 times the simulation. We recall that in Figure 4, β s , 1 0.4 .

Figure 7 
                     
                        Impact of poor transportability of the variance–covariance matrix which is simulated with a decreasing coefficient 
                           
                              
                              
                                 
                                    
                                       β
                                    
                                    
                                       s
                                       ,
                                       1
                                    
                                 
                              
                              {\beta }_{s,1}
                           
                        , responsible of the covariate shift between the RCT sample and the observational sample. The lower 
                           
                              
                              
                                 
                                    
                                       β
                                    
                                    
                                       s
                                       ,
                                       1
                                    
                                 
                              
                              {\beta }_{s,1}
                           
                        , the higher the absolute empirical bias (boxplots), and the higher the difference between the predictions given by Theorem 1 (orange dots) compared to the effective empirical biases (boxplots).
Figure 7

Impact of poor transportability of the variance–covariance matrix which is simulated with a decreasing coefficient β s , 1 , responsible of the covariate shift between the RCT sample and the observational sample. The lower β s , 1 , the higher the absolute empirical bias (boxplots), and the higher the difference between the predictions given by Theorem 1 (orange dots) compared to the effective empirical biases (boxplots).

4.2 A semi-synthetic simulation: the STAR experiment

The semi-synthetic experiment is a mean to evaluate the methods on (semi) real data where neither the data generation process nor the distribution of the covariates is under control.

4.2.1 Simulation details

We use the data from a randomized controlled trial, the STAR study. This RCT is a pioneering randomized study from the domain of education [58], started in 1985, and designed to estimate the effects of smaller classes in primary school, on the children’s grades. This experiment showed a strong payoff to smaller classes [60]. In addition, the effect has been shown to be heterogeneous [39], where class sizes have a larger effect for minority students and those on subsidized lunch. For our purposes, we focus on the same subgroup of children, same treatment (small versus regular classes), and same outcome (average of all grades at the end) as in ref. [61].

A total of 4,218 children are concerned by the treatment randomization, with treatment assignment at first grade only. On the whole data, we estimated an average treatment effect of 12.80 additional points on the grades (95% CI [10.41–15.2]) with the difference-in-means estimator. We consider this estimate as the ground truth τ as it is the global RCT. Then, we generate a random sample of 500 children to serve as the observational study. From the rest of the data, we sample a biased RCT according to a logistic regression that defines probability for each class to be selected in the RCT, and using only the variable g1surban informing on the neighborhood of the school, which can be considered as a proxy for the socioeconomic status. The final selection is performed using a Bernoulli procedure, which leads to 563 children in the RCT. The resulting RCT is such that τ ˆ 1 is 4.85 (95% CI [ 2.07 –11.78]), which is underestimated. This is due to the fact that that the selection is performed toward children that benefit less from the class size reduction according to previous studies [39,60,61]. When generalizing the ATE with the G-formula on the full set of covariates, estimating the nuisance components with a linear model, and estimating the confidence intervals with a stratified bootstrap (1,000 repetitions), the target population ATE is recovered with an estimate of 13.05 (95% CI [5.07–22.11]) Not including the covariate on which the selection is performed (g1surban) leads to a biased generalized ATE of 5.87 (95% CI [ 1.47 –12.82]). These results are presented in Figure 8, along with AIPSW estimates. The IPSW is not represented due to a too large variance.

Figure 8 
                     
                        Simulated STAR data: True target population ATE estimation using all the STAR’s RCT data is represented (difference-in-means). This is highlighted with a red dashed line to represent the ground truth. The ATE estimate of a biased RCT (difference-in-means) is also represented showing a lower treatment effect due to a covariate shift along the covariate g1surban. Two estimators are used for the generalization, the G-formula (Definition 1) and the AIPSW (Definition A2); both relying on linear or logistic models for the nuisance components. The generalized ATE is either estimated with all covariates (blue) or with all covariates except
                        g1surban (orange). The confidence intervals are estimated with a stratified bootstrap (1,000 repetitions). Similar results are obtained when nuisance components are estimated with random forest.
Figure 8

Simulated STAR data: True target population ATE estimation using all the STAR’s RCT data is represented (difference-in-means). This is highlighted with a red dashed line to represent the ground truth. The ATE estimate of a biased RCT (difference-in-means) is also represented showing a lower treatment effect due to a covariate shift along the covariate g1surban. Two estimators are used for the generalization, the G-formula (Definition 1) and the AIPSW (Definition A2); both relying on linear or logistic models for the nuisance components. The generalized ATE is either estimated with all covariates (blue) or with all covariates except g1surban (orange). The confidence intervals are estimated with a stratified bootstrap (1,000 repetitions). Similar results are obtained when nuisance components are estimated with random forest.

4.2.2 Application of the sensitivity methods

We now successively consider two different missing covariate patterns to apply the methods presented in Section 3.3.2.

4.2.2.1 Considering g1surban is missing in the observational study

[35]’s method (recalled in Section 3.3.2.1) can be applied, if we are given a set of plausible values for E [ g1surban ] . Specifying the following range ] 2.1 , 2.7 [ (containing the true value for E [ g1surban ] ) leads to a range for the generalized ATE of ] 9.5 , 16.7 [ . Recalling that the ground truth is 12.80 (95% CI[10.41–15.2]), the estimated range has a good overlap with the ground truth. In other words, with this specification of the range, a user would correctly conclude that without this key variable, the generalized ATE is probably underestimated.

4.2.2.2 Considering g1surban is missing in the RCT

Figure 9 illustrates the method when the missing covariate is in the RCT data set (see Procedure 2). This method relies on Assumption 7, which we test with a Box M-test on Σ (though in practice such a test could only be performed on Σ obs , obs ). Including only numerical covariates would reject the null hypothesis ( p -value = 0.034). Note that beyond violating Assumption 7, some variables are categorical (eg race and gender). Further discussions about violation of this assumption are available in Appendix (Section G).

Figure 9 
                     
                        Sensitivity analysis of STAR data: considering the covariate g1surban is missing in the RCT. The black cross indicates that the point estimate value for the bias would be an expert that have the true sensitivity values (
                           
                              
                              
                                 −
                                 6.4
                              
                              -6.4
                           
                        ), and the true bias value is represented by the red line (
                           
                              
                              
                                 −
                                 7.08
                              
                              -7.08
                           
                        ). Dashed lines correspond to the 95% confidence intervals around the estimated sensitivity parameters.
Figure 9

Sensitivity analysis of STAR data: considering the covariate g1surban is missing in the RCT. The black cross indicates that the point estimate value for the bias would be an expert that have the true sensitivity values ( 6.4 ), and the true bias value is represented by the red line ( 7.08 ). Dashed lines correspond to the 95% confidence intervals around the estimated sensitivity parameters.

In this application, applying recommendations from Section 3.3.2.1 (see paragraph entitled Data-driven approach to determine sensitivity parameter) allows us to obtain δ g1surban 11 . We consider that the shift is correctly given by domain expert, and so the true shift is taken with uncertainty corresponding to the 95% confidence interval of a difference in mean. Finally, Figure 9 allows to conclude on a negative bias, that is, E [ τ ˆ n , m , obs ] τ . Note that our method underestimates a bit the true bias, with an estimated bias of 6.4 when the true bias is 7.08 , delimited with the continue red curve on the top right.

5 Application on critical care data

A motivating application of our work is the generalization to a French target population – represented by the Traumabase registry – of the CRASH-3 trial [62], evaluating tranexamic acide (TXA) to prevent death from traumatic brain injury (TBI).

5.1 CRASH-3

A total of 175 hospitals in 29 different countries participated to the randomized and placebo-controlled trial, called CRASH-3 [63], where adults with TBI suffering from intracranial bleeding were randomly administrated TXA [62]. The inclusion criteria of the trial are patients with a Glasgow Coma scale (GCS)[4] score of 12 or lower or any intracranial bleeding on CT scan, and no major extracranial bleeding. The outcome we consider in this application is the disability rating scale (DRS) after 28 days of injury in patients treated within 3 hours of injury. Such an index is a composite ordinal indicator ranging from 0 to 29, the higher the value, the stronger the disability. This outcome can be considered as a secondary outcome. This outcome has some drawbacks in the sense that TXA diminishes the probability to die from TBI and therefore may increase the number of high DRS values [64]. Therefore, to avoid a censoring or truncation due to death, we keep all individuals and set the DRS score of deceased ones to 30. The difference-in-means estimator gives an ATE of 0.29 with [95% CI 0.80 −0.21]), therefore not giving a significant evidence of an effect of TXA on DRS.

5.2 Traumabase

To improve decisions and patient care in emergency departments, the Traumabase group, comprising 23 French Trauma centers, collects detailed clinical data from the scene of the accident to the release from the hospital. The resulting database, called the Traumabase, comprises 23,000 trauma admissions to date and is continually updated. In this application, we consider only the patients suffering from TBI, along with considering an imputed database. The Traumabase comprises a large number of missing values, and this is why we used a multiple imputation via chained equations (MICE) [65] prior to applying our methodology.

5.3 Predicting the treatment effect on the Traumabase data

We want to generalize the treatment effect to the French patients – represented by the Traumabase data base. Six covariates are present at baseline, with age, sex, time since injury, systolic blood pressure, Glasgow Coma scale score (GCS), and pupil reaction. Sex is not considered in the final sensitivity analysis as a noncontinuous covariate, and pupil reaction is considered as continuous ranging from 0 to 2. However, an important treatment effect modifier is missing, that is, the time between treatment and the trauma. For example, ref. [66] reveals a 10% reduction in treatment effectiveness for every 20 min increase in time to treatment (TTT). In addition, TTT is probably shifted between the two populations. Therefore, this covariate breaks Assumption 4 (ignorability on trial participation), and we propose to apply the methods developed in Section 3.

5.4 Sensitivity analysis

The concatenated data set with the RCT and observational data contains 12,496 observations (with n = 8,977 and m = 7,743 ). Considering a totally missing covariate, we apply Procedure 1. We assume that time to treatment (TTT) is independent of all other variables, for example, the ones related to the patient baseline characteristics (e.g., age) or to the severity of the trauma (e.g., the Glasgow score). Clinicians support this assumption as the time to receive the treatment depends on the time for the rescuers to come to the accident area, and not on the other patient characteristics. We first estimate the target population treatment effect with the set of observed variables and the G-formula estimator, leading to an estimated ATE τ ˆ n , m , obs of 0.08 (95% CI [ 0.50 −0.44]). The nuisance parameters are estimated using random forests, and the confidence interval with nonparametric stratified bootstrap. As the omission of the TTT variable could affect this conclusion, the sensitivity analysis gives insights on the potential bias. We apply the method relative to a completely missing covariate (Section 3.3.1).

A common practice in sensitivity analysis is to use observed covariates as benchmark to guess the impact of an unobserved covariates. For example, the Glasgow score is also suspected to be a treatment effect modifier and is shifted between the two populations. We place it on a sensitivity map (Figure 10) along with the true corresponding values for δ glasgow and Δ glasgow . As the Traumabase contain more individuals with a higher Glasgow score, a positive shift is reported. In addition, the higher the Glasgow score, the higher the effect (low DRS), so that δ glasgow < 0 . Finally, removing the Glasgow score from the analysis would lead to τ ˆ obs , n , m > τ . The sensitivity map does not allow to conclude that this bias is big enough compared to the confidence intervals previously mentioned for τ ˆ obs , n , m . Is the TTT a stronger or more shifted covariate than the Glasgow score? Previous publications have suggested a huge impact of TTT, and therefore, one could expect a bigger impact on the bias. In Figure 11, we represent a sensitivity map for TTT that could be drawn by domain experts. Here, sensitivity parameters are guessed. For example, one can suspect that treatment is given on average 20 minutes earlier in the Traumabase (e.g., interviewing nurses and doctors in Trauma centers), and the coefficient δ TTT is inferred from a previous work on TXA. In Figure 11, one can see that not observing TTT has a bigger impact on the bias than not observing the Glasgow score (almost 10 times bigger), suggesting another conclusion: a positive and significant effect of TXA on the Traumabase population, if the sensitivity parameters are correctly guessed. Also, as soon as there is a risk of a treatment given later than in the CRASH3 trial, this sensitivity map would help raising an alarm on a negative effect on the Traumabase population.

Figure 10 
                  
                     Sensitivity map if the Glasgow score covariate was missing: the true corresponding values for 
                        
                           
                           
                              
                                 
                                    δ
                                 
                                 
                                    
                                       
                                       glasgow
                                       
                                    
                                 
                              
                           
                           {\delta }_{\text{glasgow}}
                        
                      and 
                        
                           
                           
                              
                                 
                                    Δ
                                 
                                 
                                    
                                       
                                       glasgow
                                       
                                    
                                 
                              
                           
                           {\Delta }_{\text{glasgow}}
                        
                      are computed with respectively a Robinson procedure and a mean difference. Intervals correspond to 95% confidence intervals.
Figure 10

Sensitivity map if the Glasgow score covariate was missing: the true corresponding values for δ glasgow and Δ glasgow are computed with respectively a Robinson procedure and a mean difference. Intervals correspond to 95% confidence intervals.

Figure 11 
                  
                     Sensitivity map for TTT: Intervals represent plausible parameters range, with a treatment given on average 10–30 min earlier in the Traumabase, and an heterogeneous coefficient inspired from [66].
Figure 11

Sensitivity map for TTT: Intervals represent plausible parameters range, with a treatment given on average 10–30 min earlier in the Traumabase, and an heterogeneous coefficient inspired from [66].

6 Conclusion

In this work, we have studied sensitivity analyses for causal-effect generalization to assess the impact of a partially unobserved confounder (either in the RCT or in the observational data set) on the ATE estimation. In particular:

  1. To go beyond the common requirement that the unobserved confounder is independent from the observed covariates, we instead assume that their covariance is transported (Assumption 7). Our simulation study (4) shows that even with a slightly deformed covariance, the proposed sensitivity analysis procedure gives useful estimates of the bias.

  2. Leveraging the high interpretability of our sensitivity parameter, our framework concludes on the sign of the estimated bias. This sign is important as accepting a treatment effect highly depends on the direction of the generalization shift. We integrate the aforementioned methods into the existing sensitivity map visualization, using a heatmap to represent the sign of the estimated bias.

  3. Our procedures use a sensitivity parameter with a direct interpretation: the shift in expectation Δ m of the missing covariate between the RCT and the observational data. We hope that this will ease practical applications of sensitivity analyses by domain experts.

Our proposal inherits limitations from the more standard sensitivity analysis methods with observational data, namely, the semi-parametric assumption of the outcome model along with an hypothesis on covariate structures (Gaussian inputs). Therefore, future extensions of this work could explore ways to relax either the parametric assumption or the distributional assumption to support more robust sensitivity analyses. Another possible extension to a missing binary covariate could be deduced from this work, in the case where this covariate is independent of the others in both populations.

Acknowledgments

We would like to acknowledge helpful discussions with Drs Marine Le Morvan, Daniel Malinsky, and Shu Yang. We also would like to acknowledge the insights, discussions, and medical expertise from the Traumabase group and physicians, in particular, Drs François-Xavier Ageron and Tobias Gauss. In addition, none of the data analysis part could have been done without the help of Dr. Ian Roberts and the CRASH-3 group, who shared with us the clinical trial data. Finally, we thank the reviewers for their careful reading allowing to deeply improve this research work.

  1. Funding information: Authors are all funded by their respective employer (Inria or École polytechnique).

  2. Conflict of interest: Authors state no conflict of interest.

  3. Data availability statement: The datasets generated and/or analysed are available in the [unobserved-covariate] repository, https://github.com/BenedicteColnet/unobserved-covariate. The CRASH-3 and Traumabase data sets are not available due to privacy reasons.

Appendix A Estimators of the target population ATE

In this section, we grant assumptions presented in Section 2.1 and study the asymptotic behavior – and in particular, the L 1 -consistency – of three estimators: the G-formula, the IPSW, and the AIPSW.

A.1 G-formula

The G-formula procedure and its consistency assumption are detailed in the core text, see Section 3, and in particular, Definition 1 and Assumption 6. Here, we present the theorem for consistency.

Theorem A1

(G-formula consistency) Consider the G-formula estimator in Definition 1along with Assumptions 1, 2, 3, 4, 5 (identifiability), and Assumption 6 (consistency), then the G-formula estimator converges toward τ in L 1 norm,

τ ˆ G , n , m L 1 n , m τ .

A.2 IPSW

Another approach, called inverse propensity weighting score (IPSW), consists in weighting the RCT sample so that is ressembles the target population distribution.

Definition A1

(Inverse propensity weighting score – IPSW – [5,13]) The IPSW estimator is denoted τ ˆ IPSW , n , m , and defined as follows:

(A1) τ ˆ IPSW , n , m = 1 n i = 1 n n m Y i α ˆ n , m ( X i ) A i e 1 ( X i ) 1 A i 1 e 1 ( X i ) ,

where α ˆ n , m is an estimate of the odd ratio of the indicatrix of being in the RCT:

α ( x ) = P ( i i O , X i = x ) P ( i O i O , X i = x ) .

This intermediary quantity to estimate, α ( . ) , is called a nuisance component.

Similar to the G-formula, we introduce here an assumption on the behavior of the nuisance component α to carry out the mathematical analysis of the IPSW.

Assumption A1

(Consistency assumptions – IPSW) Denoting by n m α ˆ n , m ( x ) the estimated weights on the set of observed covariates X , the following conditions hold,

  1. (H1-IPSW) sup x X n m α ˆ n , m ( x ) f X ( x ) f X S = 1 ( x ) = ε n , m a . s . 0 , when n , m ,

  2. (H2-IPSW) for all n , m large enough E [ ε n , m 2 ] exists and E [ ε n , m 2 ] a . s . 0 , when n , m ,

  3. (H3-IPSW) Y is square integrable.

Theorem A2

(IPSW consistency) Consider the IPSW estimator in Definition A1 along with Assumptions 1, 2, 3, 4, 5 (identifiability), and A1 (consistency). Then, τ ˆ IPSW , n , m converges toward τ in L 1 norm,

τ ˆ IPSW , n , m L 1 n , m τ .

Theorem A2 establishes the consistency of IPSW in a more general framework than that of [4,5,7, 13,21], assuming neither oracle estimator nor parametric assumptions on α ( . ) .

A.3 AIPSW

The model for the expectation of the outcomes among randomized individuals (used in the G-estimator in Definition 1) and the model for the probability of trial participation (used in IPSW estimator in Definition A1) can be combined to form an augmented IPSW estimator (AIPSW) that has a doubly robust statistical property.

Definition A2

(Augmented IPSW - AIPSW – [45]) The AIPSW estimator is denoted τ ˆ AIPSW , n , m , and defined as

τ ˆ AIPSW , n , m = 1 n i = 1 n n m α ˆ n , m ( X i ) A i ( Y i μ ˆ 1 , n ( X i ) ) e 1 ( X i ) ( 1 A i ) ( Y i μ ˆ 0 , n ( X i ) ) 1 e 1 ( X i ) + 1 m i = n + 1 m + n ( μ ˆ 1 , n ( X i ) μ ˆ 0 , n ( X i ) ) .

Recently, it has been shown that the AIPSW estimator can be derived from the influence function of the parameter τ [45]. Under additional conditions on the rate of convergence of the nuisance parameters, it is possible to obtain asymptotic normality results[5]. As in this work we only require L 1 -consistency for the sensitivity analysis to hold, we therefore do not detail asymptotic normality conditions.

To prove AIPSW consistency, we make the following assumptions on the nuisance parameters.

Assumption A2

(Consistency assumptions - AIPSW) The nuisance parameters are bounded, and more particularly:

  1. (H1-AIPSW) There exists a function α 0 bounded from above and below (from zero), satisfying

    lim m , n sup x X n m α ˆ n , m ( x ) 1 α 0 ( x ) = 0 ,

  2. (H2-AIPSW) There exist two bounded functions ξ 1 , ξ 0 : X R , such that a { 0 , 1 } ,

    lim n + sup x X ξ a ( x ) μ ˆ a , n ( x ) = 0 .

Theorem A3

(AIPSW consistency) Consider the AIPSW estimator in Definition A2, along with Assumptions 1, 2, 3, 4, 5hold (identifiability), and Assumption A2 (consistency). Considering that estimated surface responses μ ˆ a , n ( . ) where a { 0 , 1 } are obtained following a cross-fitting estimation, then if Assumption 6 or Assumption A1 also holds, then τ ˆ AIPSW , n , m converges toward τ in L 1 norm,

τ ˆ AIPSW , n , m L 1 n , m τ .

B L 1 -convergence of G-formula, IPSW, and AIPSW

This appendix contains the proofs of theorems given in Section A. We recall that this work completes and details existing theoretical work performed by ref. [13] on IPSW (but focused on a so-called nested-trial design and assuming parametric model for the weights) and from [68] developing results within the semi-parametric theory.

B.1 L 1 -convergence of G-formula

This section contains the proof of Theorem A1, which assumes Assumption 6. For the state of clarity, we recall here Assumption 6. Denoting μ ˆ 0 , n ( . ) and μ ˆ 1 , n ( . ) estimators of μ 0 ( . ) and μ 1 ( . ) respectively, and D n the RCT sample, so that

  1. (H1-G) For a { 0 , 1 } , E [ μ ˆ a , n ( X ) μ a ( X ) D n ] p 0 when n ,

  2. (H2-G) For a { 0 , 1 } , there exist C 1 , N 1 so that for all n N 1 , almost surely, E [ μ ˆ a , n 2 ( X ) D n ] C 1 .

Proof of Theorem A1

In this proof, we largely rely on a oracle estimator τ ˆ G , , m (built with the true response surfaces), defined as follows:

τ ˆ G , , m = 1 m i = n + 1 n + m μ 1 ( X i ) μ 0 ( X i ) .

The central idea of this proof is to compare the actual G-formula τ ˆ G , n , m – on which the nuisance parameters are estimated on the RCT data – with the oracle.

B.1.1 L 1 -convergence of the surface responses

For the proof, we will require that the estimated surface responses μ ˆ 1 , n ( . ) and μ ˆ 0 , n ( . ) converge toward the true ones in L 1 . This is implied by assumptions (H1-G) and (H2-G). Indeed, for all n > 0 and all a { 0 , 1 } , thanks to the triangle inequality and linearity of expectation, we have

E [ μ ˆ a , n ( X ) μ a ( X ) D n ] E [ μ ˆ a , n ( X ) D n ] + E [ μ a ( X ) D n ] = E [ μ ˆ a , n ( X ) D n ] ( ) + E [ μ a ( X ) ] ( ) .

First, note that the quantity ( ) is upper bounded thanks to assumption (H2-G), using Jensen’s inequality. Also note that the quantity ( ) is upper bounded because the potential outcomes are integrables, that is, E [ Y ( 1 ) ] and E [ Y ( 0 ) ] exist (see Section 2.1).

Therefore, E [ μ ˆ a , n ( X ) μ a ( X ) D n ] is upper bounded. Consequently, using (H2-G) and a generalization of the dominated convergence theorem, one has

E [ μ ˆ a , n ( X ) μ a ( X ) ] = E [ E [ μ ˆ a , n ( X ) μ a ( X ) D n ] ] n 0 ,

which implies

a { 0 , 1 } , μ ˆ a , n ( X ) L 1 n μ a ( X ) .

L 1 -convergence of τ ˆ G , n , m toward τ

For all m , n > 0 ,

τ ˆ G , n , m τ ˆ G , , m = 1 m i = n + 1 n + m ( μ ˆ 1 , n ( X i ) μ ˆ 0 , n ( X i ) ) ( μ 1 ( X i ) μ 0 ( X i ) ) = 1 m i = n + 1 n + m ( μ ˆ 1 , n ( X i ) μ 1 ( X i ) ) ( μ ˆ 0 , n ( X i ) μ 0 ( X i ) ) .

Therefore, taking the expectation of the absolute value on both sides, and using the triangle inequality and the fact that observations are iid,

E [ τ ˆ G , n , m τ ˆ G , , m ] = E 1 m i = n + 1 n + m ( μ ˆ 1 , n ( X i ) μ 1 ( X i ) ) ( μ ˆ 0 , n ( X i ) μ 0 ( X i ) ) E [ μ ˆ 1 , n ( X ) μ 1 ( X ) ] + E [ μ ˆ 0 , n ( X ) μ 0 ( X ) ] .

Note that this last inequality can be obtained because different observations are used to ( i ) build the estimated surface responses μ ˆ a , n (for a { 0 , 1 } ) and ( i i ) to evaluate these estimators. Indeed, the proof would be much more complex if the sum was taken over the n observations used to fit the models. Due to the L 1 -convergence of each of the surface response when n (see the first part of the proof), we have

lim n E [ τ ˆ G , n , m τ ˆ G , , m ] = 0 .

In other words,

(A2) m , τ ˆ G , n , m L 1 n τ ˆ G , , m .

This equality is true for any m and intuitively can be understood as the fitted response surfaces μ ˆ a , n ( . ) can be very close to the true ones as soon as n is large enough. Then, the G-formula estimator, no matter the size of the observational data set, is close to the oracle one in L 1 . Hence, one can deduce a result on the difference between τ and the G-formula,

E [ τ ˆ G , n , m τ ] E [ τ ˆ G , n , m τ ˆ G , , m ] + E [ τ ˆ G , , m τ ] .

According to the weak law of large number, we have

τ ˆ G , , m L 1 m τ .

Combining this result with equation (A2), we have

τ ˆ G , n , m L 1 n , m τ ,

which concludes the proof.□

B.2 L 1 -convergence of IPSW

This section provides the proof of Theorem A2, and for the sake of clarity, we recall Assumption A1. Denoting n m α ˆ n , m ( x ) , the estimated weights on the set of covariates X , the following conditions hold,

  1. (H1-IPSW) sup x X n m α ˆ n , m ( x ) f X ( x ) f X S = 1 ( x ) = ε n , m a . s . 0 , when n , m ,

  2. (H2-IPSW) we have for all n , m large enough E [ ε n , m 2 ] exists and E [ ε n , m 2 ] a . s . 0 , when n , m ,

  3. (H3-IPSW) Y is square integrable.

Proof of Theorem A2

First, we consider an oracle estimator τ ˆ IPSW , n that is based on the true ratio f X ( x ) f X S = 1 ( x ) , that is,

τ ˆ IPSW , n = 1 n i = 1 n Y i f X ( X i ) f X S = 1 ( X i ) A i e 1 ( X i ) 1 A i 1 e 1 ( X i ) .

Note that [21] also considers such an estimator and document its consistency (see their appendix). Indeed, assuming the finite variance of Y , the strong law of large numbers (also called Kolmogorov’s law) allows us to state that:

(A3) τ ˆ IPSW , n a . s . E Y f X ( X ) f X S = 1 ( X ) A e 1 ( X ) 1 A 1 e 1 ( X ) S = 1 = τ , as n .

Now, we need to prove that this result also holds for the estimate τ ˆ IPSW , n , m where the weights are estimated from the data. To this aim, we first use the triangle inequality comparing τ ˆ IPSW , n , m with the oracle IPSW:

τ ˆ IPSW , n , m τ ˆ IPSW , n = 1 n i = 1 n A i Y i e 1 ( X i ) ( 1 A i ) Y i 1 e 1 ( X i ) n α ˆ n , m ( X i ) m f X ( X i ) f X S = 1 ( X i ) 1 n i = 1 n A i Y i e 1 ( X i ) ( 1 A i ) Y i 1 e 1 ( X i ) n α ˆ n , m ( X i ) m f X ( X i ) f X S = 1 ( X i ) Triangular inequality = 1 n i = 1 n A i Y i e 1 ( X i ) ( 1 A i ) Y i 1 e 1 ( X i ) n α ˆ n , m ( X i ) m f X ( X i ) f X S = 1 ( X i ) ε n , m n i = 1 n A i Y i e 1 ( X i ) ( 1 A i ) Y i 1 e 1 ( X i ) Assumption 9 (H1-IPSW)

Taking the expectation on the previous inequality gives,

E [ τ ˆ IPSW , n , m τ ˆ IPSW , n ] E ε n , m 1 n i = 1 n A i Y i e 1 ( X i ) ( 1 A i ) Y i 1 e 1 ( X i ) E [ ε n , m 2 ] E 1 n i = 1 n A i Y i e 1 ( X i ) ( 1 A i ) Y i 1 e 1 ( X i ) 2 C.S., square integ, and H3-IPSW E [ ε n , m 2 ] E 1 n i = 1 n 2 Y min ( η 1 , 1 η 1 ) 2 Assumption 2 and triangular inequality E [ ε n , m 2 ] E 1 n i = 1 n 4 Y 2 min ( η 1 2 , ( 1 η 1 ) 2 ) Jensen = E [ ε n , m 2 ] 2 E [ Y 2 ] min ( η 1 , ( 1 η 1 ) ) iid

Therefore, using (H2-IPSW),

(A4) E [ τ ˆ IPSW , n , m τ ˆ IPSW , n ] 0 , as n , m .

Finally, note that

E [ τ ˆ IPSW , n , m τ ] E [ τ ˆ IPSW , n , m τ ˆ IPSW , n ] + E [ τ ˆ IPSW , n τ ] .

The second right-hand side term tends to be zero by the weak law of large numbers (same reasoning as for the G-formula) and the first term tends to zero using (A4), which leads to

τ ˆ IPSW , n , m L 1 n , m τ .

B.3 L 1 convergence of AIPSW

The proof of Theorem A3 is based on Assumption A2 and either Assumption 6 or Assumption A1. Therefore the proof contains two parts. For clarity, we recall here Assumption A2:

  1. (H1-AIPSW) There exists a function α 0 bounded from above and below (from zero), satisfying

    lim m , n sup x X n m α ˆ n , m ( x ) 1 α 0 ( x ) = 0 ,

  2. (H2-AIPSW) There exist two bounded functions ξ 1 , ξ 0 : X R , such that a { 0 , 1 } ,

    lim n + sup x X ξ a ( x ) μ ˆ a , n ( x ) = 0 ,

    and, for all i { 1 , , n } ,

    lim n + sup x X ξ a ( x ) μ ˆ a , n k ( i ) ( x ) = 0 .

Proof of Theorem A3

Note that the cross-fitting procedure supposes to divide the data into K evenly sized folds, where K is typically set to 5 or 10 (e.g., see [69]). Let k ( . ) be a mapping from the sample indices i = 1 , , n to the K evenly sized data folds, and fit μ ˆ 0 , n ( . ) and μ ˆ 1 , n ( . ) with cross-fitting over the K folds using methods tuned for optimal predictive accuracy. For i { 1 , , n } , μ ˆ 0 , n k ( i ) ( . ) and μ ˆ 1 , n k ( i ) ( . ) denote response surfaces fitted on all folds except the k ( i ) th. Let us also denote by μ ˆ 0 , n ( . ) and μ ˆ 1 , n ( . ) , the surface responses estimated using the whole data set.

First case – Assumption 6

Grant Assumption 6. Here, we show that, due to this assumption, surface responses are consistently estimated. Recall that the AIPSW estimator τ ˆ AIPSW , n , m is defined as follows:

τ ˆ AIPSW , n , m = 1 n i = 1 n n m α ˆ n , m ( X i ) A i ( Y i μ ˆ 1 , n k ( i ) ( X i ) ) e 1 ( X i ) A n , m 1 n i = 1 n n m α ˆ n , m ( X i ) ( 1 A i ) ( Y i μ ˆ 0 , n k ( i ) ( X i ) ) 1 e 1 ( X i ) B n , m + 1 m i = n + 1 m + n ( μ ˆ 1 , n ( X i ) μ ˆ 0 , n ( X i ) ) C n , m

Note that τ ˆ AIPSW , n , m is composed of three terms, where the last C m , n corresponds to the G-formula τ ˆ G , n , m . Now, considering E [ τ ˆ AIPSW , n , m τ ] , and using the triangle inequality and linearity of the expectation,

(A5) E [ τ ˆ AIPSW , n , m τ ] E [ A n , m ] + E [ B n , m ] + E [ τ ˆ G , n , m τ ] .

Because Assumption 6 holds and according to Theorem A1, we have

(A6) E [ τ ˆ G , n , m τ ] 0 , when n , m .

Now, consider the term A n , m , so that,

A n , m = 1 n i = 1 n n m α ˆ n , m ( X i ) 1 α 0 ( X i ) A i ( Y i μ ˆ 1 , n k ( i ) ( X i ) ) e 1 ( X i ) A n , m , 1 + 1 n i = 1 n 1 α 0 ( X i ) A i ( Y i μ ˆ 1 , n k ( i ) ( X i ) ) e 1 ( X i ) A n , m , 2 .

Regarding A n , m , 1 , we have

E [ A n , m , 1 ] 1 η 1 sup x X n m α ˆ n , m ( x ) 1 α 0 ( x ) ( E [ A i Y i μ ˆ 1 , n k ( i ) ( X i ) ] ) 1 η 1 sup x X n m α ˆ n , m ( x ) 1 α 0 ( x ) ( E [ Y ( 1 ) ] + E [ ξ 1 ( X ) ] + ε ) ,

which tends to zero according to (H1-AIPSW). Regarding A n , m , 2 , by the weak law of large numbers,

1 n i = 1 n 1 α 0 ( X i ) A i ( Y i μ ˆ 1 , n k ( i ) ( X i ) ) e 1 ( X i ) L 1 n E 1 α 0 ( X i ) A i ( Y i μ ˆ 1 , n k ( i ) ( X i ) ) e 1 ( X i ) = E 1 α 0 ( X i ) E [ ( Y i μ ˆ 1 , n k ( i ) ( X i ) ) X i , D n k ( i ) ] = E 1 α 0 ( X i ) E [ μ 1 ( X ) μ ˆ 1 , n k ( i ) ( X ) D n k ( i ) ] ,

where

E 1 α 0 ( X i ) E [ μ 1 ( X ) μ ˆ 1 , n k ( i ) ( X ) D n k ( i ) ] sup x X 1 α 0 ( x ) E [ E [ μ 1 ( X ) μ ˆ 1 , n k ( i ) ( X ) D n k ( i ) ] ] ,

which tends to zero according to Assumption 6. Therefore

(A7) A n , m L 1 n 0 .

Using equations (A6) and (A7) in (A5) along with the L 1 -convergence of the G-formula toward τ allows us to conclude that

τ ˆ AIPSW , n , m L 1 n , m τ .

Second case - Assumption A1

Grant Assumption A1. Here, we show that, due to this assumption, weights are consistently estimated. Note that the AIPSW estimate can be rewritten as follows:

τ ˆ AIPSW , n , m = 1 n i = 1 n n m α ˆ n , m ( X i ) A i Y i e 1 ( X i ) ( 1 A i ) Y i 1 e 1 ( X i ) D n , m 1 n i = 1 n n m α ˆ n , m ( X i ) f X ( X i ) f X S = 1 ( X i ) A i μ ˆ 1 , n k ( i ) ( X i ) e 1 ( X i ) E n , m + 1 n i = 1 n n m α ˆ n , m ( X i ) f X ( X i ) f X S = 1 ( X i ) ( 1 A i ) μ ˆ 0 , n k ( i ) ( X i ) 1 e 1 ( X i ) F n , m 1 n i = 1 n f X ( X i ) f X S = 1 ( X i ) A i μ ˆ 1 , n k ( i ) ( X i ) e 1 ( X i ) ( 1 A i ) μ ˆ 0 , n k ( i ) ( X i ) 1 e 1 ( X i ) G n + 1 m i = n + 1 m + n ( μ ˆ 1 , n ( X i ) μ ˆ 0 , n ( X i ) ) . C n , m .

Again, using the expectation and the triangle inequality, one has,

(A8) E [ τ ˆ AIPSW , n , m τ ] E [ D n , m τ ] + E [ E n , m ] + E [ F n , m ] + E [ G n + C n , m ] .

Note that the term D n , m corresponds to the IPSW estimator (Definition A1). According to Assumption A1 and Theorem A2, E [ D n , m τ ] converges to 0 as n , m . Now, we study the convergence of each of the remaining terms in equation (A8).

B.3.1 Considering E n , m and F n , m

Let us now consider the term E n , m . First, note that, according to Assumption A2 (H2-AIPSW), the estimated surface responses are uniformly bounded for n large enough, that is, there exists μ M > 0 such that, for all a { 0 , 1 } , for all n large enough,

sup x X μ ˆ a , n ( x ) μ M .

It follows that, for all n large enough,

E n , m 1 n i = 1 n n m α ˆ n , m ( X i ) f X ( X i ) f X S = 1 ( X i ) 2 i = 1 n A i μ ˆ 1 , n k ( i ) ( X i ) e 1 ( X i ) 2 Cauchy-Schwarz 1 n i = 1 n n m α ˆ n , m ( X i ) f X ( X i ) f X S = 1 ( X i ) 2 1 η 1 i = 1 n ( μ ˆ 1 , n k ( i ) ( X i ) ) 2 Assumption 2 1 n i = 1 n n m α ˆ n , m ( X i ) f X ( X i ) f X S = 1 ( X i ) 2 μ M η 1 Assumption 10 (H2-AIPSW) 0 , when n , m . Assumption 9

The reasoning is the same for the term F n , m , which also converges uniformly toward 0 when n , m .

B.3.2 Considering G n and C n , m

By Assumption (H2-AIPSW), for all ε > 0 , for all n large enough, for all x X ,

μ ˆ 1 , n ( x ) [ ξ 1 ( x ) ε , ξ 1 ( x ) + ε ] .

Therefore, for all n large enough, and for all m ,

1 m i = n + 1 m + n μ ˆ 1 , n ( X i ) 1 m i = n + 1 m + n ξ 1 ( X i ) 1 m i = n + 1 m + n μ ˆ 1 , n ( X i ) ξ 1 ( X i ) ε .

Consequently,

C n , m 1 m i = n + 1 m + n ξ 1 ( X i ) + 1 m i = n + 1 m + n ξ 0 ( X i ) 2 ε .

Therefore,

C n , m E [ ξ 1 ( X ) ] + E [ ξ 0 ( X ) ] 2 ε + 1 m i = n + 1 m + n ξ 1 ( X i ) E [ ξ 1 ( X ) ] + 1 m i = n + 1 m + n ξ 0 ( X i ) E [ ξ 0 ( X ) ] .

Hence, by the law of large numbers,

C n , m L 1 n , m E [ ξ 1 ( X ) ] E [ ξ 0 ( X ) ] .

We can apply the same reasoning for the term G n , by taking into account the fact that it uses a cross-fitting strategy. By Assumption A2 (H2-AIPSW), for all ε > 0 , for all n large enough, for all x X , for all i { 1 , , n } ,

μ ˆ 1 , n k ( i ) ( x ) [ ξ 1 ( x ) ε , ξ 1 ( x ) + ε ] .

By using this inequality, we obtain

1 n i = 1 n f X ( X i ) f X S = 1 ( X i ) A i e 1 ( X i ) μ ˆ 1 , n k ( i ) ( X i ) 1 n i = 1 n f X ( X i ) f X S = 1 ( X i ) A i e 1 ( X i ) ξ ( X i ) ε η 1 sup x X 1 α 0 ( x ) .

Besides, by the law of large numbers,

lim n 1 n i = 1 n f X ( X i ) f X S = 1 ( X i ) A i e 1 ( X i ) ξ 1 ( X i ) = E f X ( X i ) f X S = 1 ( X i ) A i e 1 ( X i ) ξ a ( X ) = E [ ξ a ( X ) ] .

Consequently, as mentioned earlier,

G n , m L 1 n , m E [ ξ 0 ( X ) ] E [ ξ 1 ( X ) ] .

Finally,

C n , m + G n , m L 1 n , m 0 ,

which concludes the proof.□

C Proofs for the missing covariate setting

This section gathers proofs related to the case where key covariates (treatment effect modifiers with distributional shift) are missing. In particular, this appendix contains the proofs of results presented in Section 3.

C.1 Proof of Theorem 1

Proof

Theorem 1 is essentially a statement about the observed distribution. One can first derived what is the partial identification of τ under the observed distribution τ obs , that is,

τ obs = E [ E [ Y ( 1 ) Y ( 0 ) X obs = x obs , S = 1 ] ] = E [ E [ δ , X X obs = x obs , S = 1 ] ] Linear CATE = E [ E [ δ , X obs + δ , X mis X obs = x obs , S = 1 ] ] X = ( X mis , X obs ) = E [ δ , X obs ] + E [ E [ δ , X mis X obs = x obs , S = 1 ] ] . Ignorability

As the covariates X are assumed to be a Gaussian vector distributed as N ( μ , Σ ) , and considering the assumption on the variance–covariance matrix (Assumption 7), one can have an explicit expression of the conditional expectation [70].

E [ X mis X obs = x obs ] = E [ X mis ] + Σ mis , obs ( Σ obs , obs ) 1 ( x obs E [ X obs ] ) .

Therefore, plugging this expression into τ obs and comparing it to τ ,

τ τ obs = δ , E [ X mis ] E [ X mis S = 1 ] Σ mis , obs Σ obs , obs 1 ( E [ X obs ] E [ X obs S = 1 ] = j mis δ j ( E [ X j ] E [ X j S = 1 ] Σ j , obs Σ obs , obs 1 ( E [ X obs ] E [ X obs S = 1 ] ) .

Note that the last row is only a different way to write the scalar product into a sum.

Then, any L 1 -consistent estimator h a t τ n , m , obs of τ on the observed set of covariates will follow

lim n , m E [ τ ˆ n , m , obs ] τ = j mis δ j ( E [ X j ] E [ X j S = 1 ] Σ j , obs Σ obs , obs 1 ( E [ X obs ] E [ X obs S = 1 ] ) .

C.2 Imputation

This part contains the proof of Corollary 2.

Proof

This proof is divided into two parts, depending on the missing covariate pattern.□

C.2.1 Consider the RCT as the complete dataset

We assume that the linear link between the missing covariate X mis and the observed one X obs in the trial population is known, so is the true response surfaces μ 1 ( . ) and μ 0 ( . ) . We consider the estimator τ ˆ G , , m , imp based on the two previous oracles quantities. We denote by c 0 , , c # o b s the coefficients linking X obs and X mis in the trial, so that, on the event S = 1 ,

(A9) X mis = c 0 + j obs c j X j + ε ,

where ε is a Gaussian noise satisfying E [ ε X obs ] = 0 almost surely. Since we assume that the true link between X mis and X obs is known (i.e., we know the coefficients c 0 , , c d ), the imputation of the missing covariate on the observational sample writes

(A10) X ˆ mis c 0 + j obs c j X j .

We denote X ˜ the imputed data set composed of the observed covariates and the imputed one in the observational sample. The expectation of the oracle estimator τ ˆ G , , m , imp is defined as follows:

E [ τ ˆ G , , m , i m p ] = E 1 m i = n + 1 n + m μ 1 ( X ˜ i ) μ 0 ( X ˜ i ) By definition of  τ ˆ G , , m , imp = E 1 m i = n + 1 n + m δ , X ˜ i Linear CATE (2) = E 1 m i = n + 1 n + m j obs δ j X j , i + δ mis X ˆ m i s , i .

Because of the finite variance of X obs and X ˆ mis , the law of large numbers allows to state that:

lim m E [ τ ˆ G , , m , imp ] = j obs δ j E [ X j ] + δ mis E [ X ˆ mis ] .

Due to Assumption 7, the distribution of the vector X is Gaussian in both populations, and one can use the conditional expectation for a multivariate gaussian law to write the conditional expectation in the trial population, that is,

(A11) E [ X mis X obs , S = 1 ] = E [ X mis S = 1 ] + Σ mis , obs Σ obs , obs 1 ( X obs E [ X obs S = 1 ] ) .

Combining (A9) and (A11), one can obtain:

(A12) c 0 + j obs c j X j = E [ X mis S = 1 ] + Σ mis , obs Σ obs , obs 1 ( X obs E [ X obs S = 1 ] ) .

Now, we can compute,

E [ X ˆ mis ] = E [ c 0 + j obs c j X j ] = E [ E [ X mis S = 1 ] + Σ mis , obs Σ obs , obs 1 ( X obs E [ X obs S = 1 ] ) ] ( 20 ) = E [ X mis S = 1 ] + Σ mis , obs Σ obs , obs 1 ( E [ X obs ] E [ X obs S = 1 ] ) .

This last result allows to conclude that,

lim m E [ τ ˆ G , , m , i m p ] = j obs δ j E [ X j ] + δ mis ( E [ X mis S = 1 ] + Σ mis , obs Σ obs , obs 1 ( E [ X obs ] E [ X obs S = 1 ] ) ) .

Finally, as τ = j = 1 p δ j E [ X j ] ,

τ lim m E [ τ ˆ G , , m , i m p ] = δ mis ( E [ X mis ] E [ X mis S = 1 ] Σ mis , obs Σ obs , obs 1 ( E [ X obs ] E [ X obs S = 1 ] ) ) ,

which concludes this part of the proof.

C.2.2 Consider the observational data as the complete data set

We assume here that the true relations between X mis and X obs is known, and the true response model is also known. We denote by τ G , , , imp the estimator based on these two quantities.

More precisely, we denote by c 0 , , c # o b s the coefficients linking X obs and X mis in the observational population, so that

(A13) X mis = c 0 + j obs c j X j + ε ,

where ε is a Gaussian noise satisfying E [ ε X obs ] = 0 almost surely.

As the estimator is an oracle, the relation in (A13) is used to impute the missing covariate in the observational sample, so that

(A14) X ˆ mis c 0 + j obs c j X j .

We denote X ˜ the imputed data set composed of the observed covariates and the imputed one in the trial population. Note that the X ˆ mis is a linear combination of X obs in the trial population and thus a measurable function of X obs . This property is used below and labelled as (A14). As τ G , , , imp is an oracle, one have:

E [ τ G , , , imp ] = E [ E [ Y ( 1 ) Y ( 0 ) X ˜ , S = 1 ] ] = E [ E [ Y ( 1 ) Y ( 0 ) X ˆ mis , X obs , S = 1 ] ] = E [ E [ Y ( 1 ) Y ( 0 ) X obs , S = 1 ] ] ( 22 ) = E j obs δ j X j + δ mis E [ X mis X obs , S = 1 ] . ( 2 ) = j obs δ j E [ X j ] + δ mis E [ E [ X mis X obs , S = 1 ] ] = j obs δ j E [ X j ] + δ mis ( E [ X mis S = 1 ] + Σ mis , obs Σ obs , obs 1 ( E [ X obs ] E [ X obs S = 1 ] ) ) . ( 19 )

Finally, as τ = j = 1 p δ j E [ X j ] ,

τ E [ τ G , , , imp ] = δ mis ( E [ X mis ] E [ X mis S = 1 ] Σ mis , obs Σ obs , obs 1 ( E [ X obs ] E [ X obs S = 1 ] ) ) ,

which concludes this part of the proof.

C.3 Proxy variable

Proof of Lemma 1

Recall that we denote τ ˆ G , n , m , prox the G-formula estimator using X prox instead of X mis in the G-formula. The derivations of τ ˆ G , n , m , prox give:

E [ τ ˆ G , n , m , prox ] = E [ E [ Y X obs , X prox , S = 1 , A = 1 ] E [ Y X obs , X prox , S = 1 , A = 0 ] ] Definition of τ ˆ G , n , m , prox = E [ E [ g ( X ) + δ , X X obs , X prox , S = 1 ] E [ g ( X ) X obs , X prox , S = 1 ] ] = E [ E [ δ , X X obs , X prox , S = 1 ] ] = j obs δ j E [ X j ] + δ mis E [ E [ X mis X obs , X prox , S = 1 ] ] Linearity of  Y  (2) = j obs δ j E [ X j ] + δ mis E [ E [ X mis X prox , S = 1 ] ] X mis X obs  (8) and Assumption 1 .

The framework of the proxy variable (8) allows to have an expression of the conditional expectation of X mis [70]:

E [ E [ X mis X prox , S = 1 ] ] = E [ X mis S = 1 ] + Cov ( X mis , X prox ) V [ X prox ] ( X prox E [ X prox S = 1 ] ) ,

where

V [ X prox ] = V [ X mis + η ] = V [ X mis ] + V [ η ] + 2 Cov ( η , X mis ) = 0 ( 8 ) = σ mis 2 + σ prox 2

and

Cov ( X mis , X prox ) = E [ X mis X prox ] E [ X prox ] 2 E [ X mis ] = E [ X prox ] = E [ X prox 2 η X prox ] E [ X prox ] 2 = E [ X prox 2 ] E [ X prox ] 2 E [ η X prox ] = V [ X prox ] E [ η X mis ] E [ η 2 ] = σ mis 2 + σ prox 2 0 σ prox 2 = σ mis 2 .

Therefore, we have

E [ E [ X mis X prox , S = 1 ] ] = E [ X mis S = 1 ] + σ mis 2 σ mis 2 + σ prox 2 ( X prox E [ X prox S = 1 ] ) ,

which allows us to complete the first derivation:

E [ τ ˆ G , n , m , prox ] = j obs δ j E [ X j ] + δ mis E E [ X mis S = 1 ] + σ mis 2 σ mis 2 + σ prox 2 ( X prox E [ X prox S = 1 ] ) = j obs δ j E [ X j ] + δ mis E [ X mis S = 1 ] + σ mis 2 σ mis 2 + σ prox 2 ( E [ X prox ] E [ X prox S = 1 ] ) = j obs δ j E [ X j ] + δ mis E [ X mis S = 1 ] + σ mis 2 σ mis 2 + σ prox 2 ( E [ X mis ] E [ X mis S = 1 ] ) ,

since E [ X prox S = 1 ] = E [ X mis S = 1 ] and E [ X prox ] = E [ X mis ] . Recalling that τ = δ j E [ X j ] , the final form of the bias of τ ˆ G , n , m , prox can be obtained as follows:

τ E [ τ ˆ G , n , m , prox ] = δ mis ( E [ X mis ] E [ X mis S = 1 ] ) 1 σ mis 2 σ mis 2 + σ prox 2 .

Proof of Corollary 3

Note that the final expression of the bias obtained in the previous proof cannot be estimated in all missing covariate patterns. For example, if X mis is partially observed in the RCT, then an estimate of δ mis can be computed, and therefore, the bias can be estimated. But in all other missing covariate pattern, a temptation is to estimate δ prox from the regression of Y against X = ( X obs , X prox ) with an OLS procedure. Ref. [59] details the infinite sample estimate of such a coefficient:

lim n , m E [ δ ˆ prox ] = δ mis σ mis 2 σ mis 2 + σ prox 2 .

Note that the quantity σ mis 2 σ mis 2 + σ prox 2 is always lower than 1; therefore, if δ mis 1 , then δ ˆ prox underestimates δ mis . This phenomenon is called the attenuation bias. This point is documented by ref. [59], and is due to heteroscedasticity in the plug-in regression:

Cov [ X prox , ε ] = Cov [ X mis + η , ε δ mis η ] = δ mis σ η 2 0 .

This asymptotic estimate can be plugged-in into the previous bias estimation:

τ E [ τ ˆ G , n , m , prox ] = δ ˆ prox ( E [ X prox ] E [ X prox S = 1 ] ) σ prox 2 σ mis 2 .

D Toward a semi-parametric model

This section completes Model 2 and justifies why this the assumption of a linear CATE is somewhat natural when considering a continuous outcome Y .

For a continuous outcome Y , the outcome model can be written with two terms, a baseline and the CATE. Indeed, when focusing on zero-mean additive-error representations leads to assume that the potential outcomes are generated according to:

(A15) Y ( A ) = μ ( A , X ) + ε A ,

for some function μ L 2 ( { 0 , 1 } × X R ) and a noise ε A satisfying E [ ε A X ] = 0 almost surely.

Lemma A1

Assume that the nonparametric generative model of equation (A15) holds, then there exists a function g : X R such that

(A16) Y ( A ) = g ( X ) + A τ ( X ) + ε A , where τ ( X ) E [ Y ( 1 ) Y ( 0 ) X ] .

Lemma A1 follows from rewriting equation (A15), accounting for the fact that A is binary and Y R . Such a decomposition is often used in the literature [55]. This model allows to have a simpler expression of the treatment effect without any additional assumptions due to the discrete nature of A . In other words, this model enables placing independent functional form on the CATE τ ( X ) , sometimes relying on the idea that the CATE is smoother, while the baseline response can be more complex [71]. In the context of the sensitivity analysis, this model has the interest of highlighting treatment effect modifier variables, such as variables that intervene in the CATE τ ( X ) .

E Robinson procedure

This appendix recall the so-called Robinson procedure that aims at estimating the CATE coefficients δ in a semi-parametric equation such as (2). This method was developed by [53] and has been further extended [69,54,55]. Such a procedure is called a R-learner, where the R denotes Robinson or Residuals. We recall the procedure,

  1. Run a nonparametric regressions Y X using a parametric or non parametric method. The best method can be chosen with a cross-validation procedure. We denote m ˆ n ( x ) = E [ Y X = x ] the estimator obtained.

  2. Define the transformed features Y ˜ = Y m ˆ n ( X ) and Z ˜ = ( A e 1 ( X ) ) X , using the previous procedure m ˆ n .

  3. Estimate δ ˆ n running the OLS regression on the transformed features Y ˜ Z ˜ .

If the nonparametric regressions of m ( x ) satisfies E [ ( m ˆ ( X ) m ( X ) ) 2 ] 1 2 = o P 1 n 1 4 , then the procedure to estimate δ is n -consistent and asymptotically normal,

n ( δ ˆ δ ) N ( 0 , V R ) , V R = Var [ Z ˜ ] 1 Var [ Z ˜ Y ˜ ] Var [ Z ˜ ] 1 .

See refs [54,69] for details.

F Synthetic simulation – extension

This section completes the synthetic simulation presented in Section 4.

F.1 Simulation parameters

Parameters chosen highlight different covariate roles and strength importance. In this setting, covariates X 1 , X 2 , and X 3 are the so-called treatment effect modifiers due to a nonzero δ coefficients, and X 1 , X 3 , and X 4 are shifted from the RCT sample and the target population distribution due to a nonzero β s coefficient. Therefore, covariates X 1 and X 3 are necessary to generalize the treatment effect in both groups. Because in the simulation X 2 and X 4 are independent, the set X 1 and X 3 is also sufficient to generalize. Only X 2 has the same marginal distribution in the RCT sample and in the observational study. Note the amplitude and sign of different coefficients used, along with dependence between variables allows to illustrate several phenomena. For example X 3 is less shifted in between the two samples compared to X 1 because β s , 1 β s , 3 .

F.2 Additional comments on Figure 4

Note that depending on the correlation strength between X 1 and X 5 , the missing of X 1 can lead to different coefficients estimations when using the G-formula estimation, and different bias on the ATE. Table A1 illustrates this situation, where the higher the correlation, the higher the error on the coefficients estimations, but the lower the bias on the ATE when only X 1 is missing.

Table A1

Coefficients estimated in the simulation: Simulation with X 1 as the missing covariate repeated 100 times, means of estimated coefficients for X 5 and bias on ATE using the Robinson procedure

ρ X 1 , X 5 δ 5 δ ˆ 5 τ τ ˆ obs g
0.05 6.34 8.24
0.5 16.78 6.35
0.95 28.56 0.93

F.3 Imputation

When a covariate is partially observed, at temptation is to impute the missing part with a model learned on the complete part as detailed in procedure 4. Section 3 illustrates Corollary 2, as it shows that linear imputation does not diminish the bias compared to a case where the generalization is performed using only the restricted set of observed covariates. In Figure A1, we simulated all the missing covariate patterns (in RCT or in observational) considering X 1 is partially missing, with varying correlation strength between X 5 and X 1 , and fitting a linear imputation model. Imputation does not lead to a lower bias than totally removing the partially observed covariate. Therefore, in case of a partially missing covariate, we advocate running a sensitivity analysis rather than a linear imputation.

Figure A1 
                     
                        Simulations results when imputing (procedure 4): Results when imputing 
                           
                              
                              
                                 
                                    
                                       X
                                    
                                    
                                       1
                                    
                                 
                              
                              {X}_{1}
                           
                         with a linear model fitted on the complete data set (either the RCT or the observational). All the missing covariate patterns are simulated using either the G-formula or the IPSW estimators. The impact of the correlation between 
                           
                              
                              
                                 
                                    
                                       X
                                    
                                    
                                       1
                                    
                                 
                              
                              {X}_{1}
                           
                         and 
                           
                              
                              
                                 
                                    
                                       X
                                    
                                    
                                       5
                                    
                                 
                              
                              {X}_{5}
                           
                         is investigated. Each simulation is repeated 100 times. All procedures have a similar bias as the procedure ignoring the partially missing covariate (totally.missing), so that a linear imputation (procedure 4) improves neither the bias nor the variance.
Figure A1

Simulations results when imputing (procedure 4): Results when imputing X 1 with a linear model fitted on the complete data set (either the RCT or the observational). All the missing covariate patterns are simulated using either the G-formula or the IPSW estimators. The impact of the correlation between X 1 and X 5 is investigated. Each simulation is repeated 100 times. All procedures have a similar bias as the procedure ignoring the partially missing covariate (totally.missing), so that a linear imputation (procedure 4) improves neither the bias nor the variance.

F.4 Proxy variable

Finally and to illustrate Lemma 1, the simulation is extended to replace X 1 by a proxy variable, generated following (8) with a varying σ prox . The generalized ATE is estimated with the G-formula. The experiments are repeated 20 times per σ prox values. Results are presented in Figure A2. Whenever σ prox is small compared to σ mis (which is equal to one in this simulation), therefore the bias is small.

Figure A2 
                     
                        Simulation results for proxy variable (procedure 5) Simulation when a key covariate is replaced by a proxy following the proxy-framework (see Assumption 8). The theoretical bias (1) is represented along with the empirical values obtained when generalizing the ATE with the plugged-in G-formula estimator.
Figure A2

Simulation results for proxy variable (procedure 5) Simulation when a key covariate is replaced by a proxy following the proxy-framework (see Assumption 8). The theoretical bias (1) is represented along with the empirical values obtained when generalizing the ATE with the plugged-in G-formula estimator.

G Homogeneity of the variance–covariance matrix

Recall that Assumption 7 states that the covariance matrices in both data sets are identical. This assumption, which may appear to be very restrictive, can be partially tested on the set of observed covariates. In this section, we present such a test [Box’s M-test 49], which illustrates the validity of Assumption 7 on some particular data set. Taking one step further, we study the impact of Assumption 7 violation on the resulting estimate.

G.1 Statistical test and visualizations

Ref. [50] presents available tests to assess if covariance matrices from two data sample are equal. Despite its sensitivity to violation, Box’s M-test [49] can be used test the equality. In particular, the package heplots contains the tests and visualizations in R. The command line to perform the test is detailed below.

library(heplots)
boxM(data[, c("X1", "X2", "X3", "X4")], group = data$S)

Even if we cannot bring a general rule to know if the covariance matrices are equal, we can display some examples in which Assumption 7 holds. For instance, [50] report that the skull data is an example of a real data set with multiple sources where there are substantial differences among the means of groups, but little evidence for heterogeneity of their covariance matrices.

G.1.1 Semi-synthetic experiment: STAR

While doing the semi-synthetic experiment on the STAR data set, the Box M-test rejects the null hypothesis when considering only numerical covariates (age, g1freelunch, gkfreelunch, and g1surban) with a p-value of 0.022. This indicates that the preservation of the variance–covariance structure between the two simulated sources does not hold. To help support conclusions, one can visualize how the variance covariance matrix vary in between the two sources, as presented in Figure A3, supporting that the changes in the variance–covariance are not very strong.

Figure A3 
                        
                           Pairwise data ellipses for the STAR data, centered at the origin. This view allows to compare the variances and covariances for all pairs of variables.
Figure A3

Pairwise data ellipses for the STAR data, centered at the origin. This view allows to compare the variances and covariances for all pairs of variables.

G.1.2 Traumabase and CRASH-3

Note that this part’s purpose is only to illustrate the principle as the application performed in Section 5 relies on the independence between the time to treatment and all other covariates, and not Assumption 7.

One can inspect how far the variance and covariance change in between the two sources. Pairwise data ellipses are presented in Figure A4 for CRASH-3 and Traumabase patients, suggesting rather strong difference in the variance–covariance matrix. As expected Box M-test largely rejects the null hypothesis.

Figure A4 
                        
                           Pairwise data ellipses for the CRASH-3 and Traumabase data, centered at the origin. CRASH-3 data are in blue and Traumabase data in red. This view allows to compare the variances and covariances for all pairs of variables. While the mean are really different in the two sources, the variances and covariances are not so different.
Figure A4

Pairwise data ellipses for the CRASH-3 and Traumabase data, centered at the origin. CRASH-3 data are in blue and Traumabase data in red. This view allows to compare the variances and covariances for all pairs of variables. While the mean are really different in the two sources, the variances and covariances are not so different.

It is interesting to note that in some cases the variance covariance matrix is identical in between two populations. For example, we tested whether the two major trauma centers in France present heterogeneity in the variance–covariance matrix, and the Box M test does not reject the null hypothesis.

G.2 Extension of the simulations

Simulations presented in Section 4 can be extended to illustrate empirically the consequences of a poorly specified Assumption 7. Suppose X 1 is the unobserved covariate and that the variance–covariance matrix is not the same in the randomized population ( S = 1 ) as in the target population. But the heterogeneities in between the two sources can be different in their nature, affecting covariates depending or not from X 1 . We can imagine two situations, a situation (A) where the link in between X 1 and X 5 is different in the two sources, and another situation (B) where the link in between X 2 and X 3 is not the same. The situation is illustrated in Figure A5(a) and (b) with pairwise data ellipses. Note that with n = 1,000 and m = 10,000 , a Box-M test largely rejects the null hypothesis with a similar statistic value for both situations. When computing the bias according to Theorem 1 and repeating the experiment 50 times, empirical evidence is made that the localization of the heterogeneity impacts or not the bias computation. As presented in Figure A5(c), situation A affects the bias computation, when situation B keeps the bias estimation valid.

Figure A5 
                     
                        Effect of a different variance–covariance matrix on the ATE estimation, where heterogeneity between the two variance–covariance matrix is introduced as presented in (a) and (b). (c) The impact on the estimated average treatment effect (ATE). Situations A and B result in a similar statistics when using a Box-M test, but leads to very different impact on the bias estimation as visible on (c). The simulations are repeated 50 times, with a similar outcome generative model as in (8), and 
                           
                              
                              
                                 n
                                 =
                                 
                                    
                                    1,000
                                    
                                 
                              
                              n=\hspace{0.1em}\text{1,000}\hspace{0.1em}
                           
                         and 
                           
                              
                              
                                 m
                                 =
                                 
                                    
                                    10,000
                                    
                                 
                              
                              m=\hspace{0.1em}\text{10,000}\hspace{0.1em}
                           
                        . (a) Situation A – Centered pairwise data ellipses. (b) Situation B – Centered pairwise data ellipses. (c) ATE estimation in the two situations where 
                           
                              
                              
                                 
                                    
                                       
                                          
                                             τ
                                          
                                          
                                             ˆ
                                          
                                       
                                    
                                    
                                       G
                                       ,
                                       obs
                                    
                                 
                              
                              {\hat{\tau }}_{G,{\rm{obs}}}
                           
                         is estimated considering 
                           
                              
                              
                                 
                                    
                                       X
                                    
                                    
                                       1
                                    
                                 
                              
                              {X}_{1}
                           
                         is missing and denoted ATE.uncomplete, while the bias 
                           
                              
                              
                                 B
                              
                              B
                           
                         is estimated following Theorem 1 giving ATE.corrected (
                           
                              
                              
                                 
                                    
                                       
                                          
                                             τ
                                          
                                          
                                             ˆ
                                          
                                       
                                    
                                    
                                       G
                                       ,
                                       obs
                                    
                                 
                                 +
                                 
                                    
                                       B
                                    
                                    
                                       ˆ
                                    
                                 
                              
                              {\hat{\tau }}_{G,{\rm{obs}}}+\hat{B}
                           
                        ).
Figure A5

Effect of a different variance–covariance matrix on the ATE estimation, where heterogeneity between the two variance–covariance matrix is introduced as presented in (a) and (b). (c) The impact on the estimated average treatment effect (ATE). Situations A and B result in a similar statistics when using a Box-M test, but leads to very different impact on the bias estimation as visible on (c). The simulations are repeated 50 times, with a similar outcome generative model as in (8), and n = 1,000 and m = 10,000 . (a) Situation A – Centered pairwise data ellipses. (b) Situation B – Centered pairwise data ellipses. (c) ATE estimation in the two situations where τ ˆ G , obs is estimated considering X 1 is missing and denoted ATE.uncomplete, while the bias B is estimated following Theorem 1 giving ATE.corrected ( τ ˆ G , obs + B ˆ ).

G.3 Recommendations

Our current recommendations when considering the Assumption 7 is, first, to visualize the heterogeneity of variance–covariance matrix with pairwise data ellipses on Σ obs , obs . A statistical test such as a Box-M test can be applied on Σ obs , obs . We also want to emphasize that a statistical test depends on the size of the data sample, when what really matters in this assumption for the sensitivity analysis to be valid is the permanence of covariance structure of the missing covariates with the strongly correlated observed covariates. Simulations presented in Figure A5 is somehow an empirical pathological case where the variance–covariance matrix are equivalently different when considering a statistical test, but leads to different consequences on the validity of Theorem 1, and therefore the sensitivity analysis.

G.4 Comment about the notations

The notations used in this work is inherited from the generalization literature and reflects the idea of a plausibility to be sampled from a target superpopulation. The point of view of two population with support inclusion is equivalent for our purpose. Still, thinking to the problem of a sampling bias, then Assumption 7 imposes unusual restrictions for P ( X S = 0 ) , that is a subpopulation of the target population. As we do not do any inference on that population and as it has no practical interpretation, we do not discuss this in this work.

References

[1] Imbens GW, Rubin DB. Causal inference in statistics, social, and biomedical sciences. Cambridge UK: Cambridge University Press; 2015. 10.1017/CBO9781139025751Search in Google Scholar

[2] Rothwell PM. External validity of randomised controlled trials: “to whom do the results of this trial apply?”. The Lancet. 2005;365:82–93. 10.1016/S0140-6736(04)17670-8Search in Google Scholar PubMed

[3] Imbens G, Hotz J, Mortimer J. Predicting the efficacy of future training programs using past. J Econometrics. 2005;125(1–2):241–70. 10.1016/j.jeconom.2004.04.009Search in Google Scholar

[4] Cole SR, Stuart EA. Generalizing evidence from randomized clinical trials to target populations: The ACTG 320 trial. Am J Epidemiol. 2010;172:107–15. 10.1093/aje/kwq084Search in Google Scholar PubMed PubMed Central

[5] Stuart EA, Cole SR, Bradshaw CP, Leaf PJ. The use of propensity scores to assess the generalizability of results from randomized trials. J R Stat Soc A (Stat Soc). 2011;174:369–86. 10.1111/j.1467-985X.2010.00673.xSearch in Google Scholar PubMed PubMed Central

[6] Pearl J, Bareinboim E. Transportability of causal and statistical relations: A formal approach. Proc AAAI Confer Artif Intelligence. 2011 Aug;25(1). Available from: https://www.semanticscholar.org/paper/Transportability-of-Causal-and-Statistical-A-Formal-Pearl-Bareinboim/09bc36898974d5d41936d698426880d0f9ed29f5. 10.1109/ICDMW.2011.169Search in Google Scholar

[7] Bareinboim E, Pearl J. A general algorithm for deciding transportability of experimental results. J Causal Inference. 2013;1(1):107–34. 10.1515/jci-2012-0004Search in Google Scholar

[8] Tipton E. Improving generalizations from experiments using propensity score subclassification: assumptions, properties, and contexts. J Educ Behav Stat. 2013;38:239–66. 10.3102/1076998612441947Search in Google Scholar

[9] Bareinboim E, Tian J, Pearl J. Recovering from selection bias in causal and statistical inference. Proceedings of the AAAI Conference on Artificial Intelligence; 2014. Vol. 28(1). https://doi.org/10.1609/aaai.v28i1.9074.10.1145/3501714.3501740Search in Google Scholar

[10] Pearl J, Bareinboim E. External validity: From Do-Calculus to transportability across populations. Stat Sci. 2014;29(4):579–95. 10.1214/14-STS486. Search in Google Scholar

[11] Kern H, Stuart E, Hill J, Green D. Assessing methods for generalizing experimental impact estimates to target populations. J Res Educ Effectiveness. 2016 01;9:1–25. 10.1080/19345747.2015.1060282Search in Google Scholar PubMed PubMed Central

[12] Bareinboim E, Pearl J. Causal inference and the data-fusion problem. Proce National Academy Sci. 2016;113(27):7345–52. Available from: https://www.pnas.org/content/113/27/7345. 10.1073/pnas.1510507113Search in Google Scholar PubMed PubMed Central

[13] Buchanan AL, Hudgens MG, Cole SR, Mollan KR, Sax PE, Daar ES, et al. Generalizing evidence from randomized trials using inverse probability of sampling weights. J R Stat Soc A (Stat Soc). 2018;181:1193–209. 10.1111/rssa.12357Search in Google Scholar PubMed PubMed Central

[14] Stuart EA, Ackerman B, Westreich D. Generalizability of randomized trial results to target populations: design and analysis possibilities. Res Social Work Practice. 2018;28(5):532–7. 10.1177/1049731517720730Search in Google Scholar PubMed PubMed Central

[15] Dong L, Yang S, Wang X, Zeng D, Cai J. Integrative analysis of randomized clinical trials with real world evidence studies. 2020. arXiv: http://arXiv.org/abs/arXiv:200301242. Search in Google Scholar

[16] Colnet B, Mayer I, Chen G, Dieng A, Li R, Varoquaux G, et al. Causal inference methods for combining randomized trials and observational studies: a review; 2020. Search in Google Scholar

[17] Degtiar I, Rose S. A review of generalizability and transportability. Annual Review of Statistics and Its Application. 2021. 10.1146/annurev-statistics-042522-103837Search in Google Scholar

[18] Susukida R, Crum R, Stuart E, Ebnesajjad C, Mojtabai R. Assessing sample representativeness in randomized control trials: application to the national institute of drug abuse clinical trials network. Addiction. 2016 01;111:1226–34. 10.1111/add.13327Search in Google Scholar PubMed PubMed Central

[19] Lesko CR, Cole SR, Hall HI, Westreich D, Miller WC, Eron JJ, et al. The effect of antiretroviral therapy on all-cause mortality, generalized to persons diagnosed with HIV in the USA, 2009-1. Int J Epidemiol. 2016 01;45(1):140–50. 10.1093/ije/dyv352. Search in Google Scholar PubMed PubMed Central

[20] Stuart EA, Rhodes A. Generalizing treatment effect estimates from sample to population: a case study in the difficulties of finding sufficient data. Eval Rev. 2017;41(4):357–88. 10.1177/0193841X16660663Search in Google Scholar PubMed PubMed Central

[21] Egami N, Hartman E. Covariate selection for generalizing experimental results: application to a large-scale development program in Uganda. J R Stat Soc A (Stat Soc). 2021;184(4):1524–48.10.1111/rssa.12734Search in Google Scholar

[22] Li F, Buchanan AL, Cole SR. Generalizing trial evidence to target populations in non-nested designs: applications to AIDS clinical trials. J R Stat Soc Ser C Appl Stat. 2022;71:669–97.10.1111/rssc.12550Search in Google Scholar PubMed PubMed Central

[23] Cornfield J, Haenszel W, Hammond EC, Lilienfeld AM, Shimkin MB, Wynder EL. Smoking and lung cancer: recent evidence and a discussion of some questions. J Natl Cancer Inst. 1959 01;22(1):173–203. 10.1093/jnci/22.1.173. Search in Google Scholar

[24] Imbens G. Sensitivity to exogeneity assumptions in program evaluation. Am Econ Rev. 2003;93:126–32. 10.1257/000282803321946921Search in Google Scholar

[25] Rosenbaum P. Sensitivity analysis in observational studies. Wiley StatsRef: Statistics Reference Online, vol. 4; 2005. Search in Google Scholar

[26] Dorie V, Harada M, Carnegie N, Hill J. A flexible, interpretable framework for assessing sensitivity to unmeasured confounding. Stat Medicine. 2016 Sep;35(20):3453–70. 10.1002/sim.6973Search in Google Scholar PubMed PubMed Central

[27] Ichino A, Nannicini T, Mealli F. From temporary help jobs to permanent employment: what can we learn from matching estimators and their sensitivity? J Appl Econom. 2008 04;23:305–27. 10.1002/jae.998Search in Google Scholar

[28] Cinelli C, Hazlett C. Making sense of sensitivity: extending omitted variable bias. J R Stat Soc B. 2020 February;82(1):39–67. Available from: https://ideas.repec.org/a/bla/jorssb/v82y2020i1p39-67.html. 10.1111/rssb.12348Search in Google Scholar

[29] Franks A, D’Amour A, Feller A. Flexible sensitivity analysis for observational studies without observable implications. J Am Stat Assoc. 2019;115(532):1–38.10.1080/01621459.2019.1604369Search in Google Scholar

[30] Veitch V, Zaveri A. Sense and sensitivity analysis: simple post-hoc analysis of bias due to unobserved confounding. Part of Advances in Neural Information Processing Systems 33 (NeurIPS 2020). 2020. Search in Google Scholar

[31] Andrews I, Oster E. A simple approximation for evaluating external validity bias. Econ Lett. 2019;178:58–62. Available from: https://www.sciencedirect.com/science/article/pii/S0165176519300655. 10.3386/w23826Search in Google Scholar

[32] Dahabreh IJ, Robins JM, Haneuse SJPA, Saeed I, Robertson SE, Stuart EA, et al. Sensitivity analysis using bias functions for studies extending inferences from a randomized trial to a target population. In Sarah R, Jon S, Elizabeth S, Miguel H (Eds.). Extending inferences from a randomized trial to a new target population: Extending inferences from a trial to a new target population. Statistics in Medicine. 39. 2019. 10.1002/sim.8426.2019. Search in Google Scholar PubMed

[33] Nie X, Imbens G, Wager S. Covariate balancing sensitivity analysis for extrapolating randomized trials across locations; 2021. Search in Google Scholar

[34] Huang M, Egami N, Hartman E, Miratrix L. Leveraging population outcomes to improve the generalization of experimental results; 2021. Search in Google Scholar

[35] Nguyen TQ, Ebnesajjad C, Cole SR, Stuart EA. Sensitivity analysis for an unobserved moderator in RCT-to-target-population generalization of treatment effects. Ann Appl Stat. 2017;11(1):225–47. 10.1214/16-AOAS1001Search in Google Scholar

[36] Nguyen T, Ackerman B, Schmid I, Cole S, Stuart E. Sensitivity analyses for effect modifiers not observed in the target population when generalizing treatment effects from a randomized controlled trial: assumptions, models, effect scales, data scenarios, and implementation details. Plos One. 2018 12;13:e0208795. 10.1371/journal.pone.0208795Search in Google Scholar PubMed PubMed Central

[37] Resche-Rigon M, White I, Bartlett J, Peters SAE, Thompson S. Multiple imputation for handling systematically missing confounders in meta-analysis of individual participant data. Stat Med. 2013 07;32:4890–905. 10.1002/sim.5894Search in Google Scholar PubMed

[38] Jolani S, Debray T, Koffijberg H, van Buuren S, Moons K. Imputation of systematically missing predictors in an individual participant data meta-analysis: a generalized approach using MICE. Stat Med. 2015;34(11):1841–63. 10.1002/sim.6451Search in Google Scholar PubMed

[39] Krueger AB. Experimental estimates of education production functions. Quarterly J Econ. 1999;114(2):497–532. Available from: https://ideas.repec.org/a/oup/qjecon/v114y1999i2p497-532.html. 10.3386/w6051Search in Google Scholar

[40] Miratrix LW, Sekhon JS, Theodoridis AG, Campos LF. Worth weighting? How to think about and use weights in survey experiments. Political Analysis. 2018;26(3):275–91. 10.1017/pan.2018.1.Search in Google Scholar

[41] Hartman E, Grieve R, Ramsahai R, Sekhon JS. From sample average treatment effect to population average treatment effect on the treated: combining experimental with observational studies to estimate population treatment effects. J R Stat Soc A (Stat Soc). 2015;178(3):757–78. Available from: https://rss.onlinelibrary.wiley.com/doi/abs/10.1111/rssa.12094. 10.1111/rssa.12094Search in Google Scholar

[42] Pearl J. Generalizing experimental findings. J Causal Infer. 2015;3(2):259–66. 10.1515/jci-2015-0025. Search in Google Scholar

[43] Ding P, Feller A, Miratrix L. Randomization inference for treatment effect variation. J R Stat Soc B. 2016 June;78(3):655–71. https://ideas.repec.org/a/bla/jorssb/v78y2016i3p655-671.html. 10.1111/rssb.12124Search in Google Scholar

[44] Lesko CR, Buchanan AL, Westreich D, Edwards JK, Hudgens MG, Cole SR. Generalizing study results: a potential outcomes perspective. Epidemiology. 2017;28:553–61. 10.1097/EDE.0000000000000664Search in Google Scholar PubMed PubMed Central

[45] Dahabreh IJ, Robins JM, Haneuse SJ, Hernán MA. Generalizing causal inferences from randomized trials: counterfactual and graphical identification; 2019. arXiv: http://arXiv.org/abs/arXiv:190610792. Search in Google Scholar

[46] Chattopadhyay A, Cohn ER, Zubizarreta JR. One-step weighting to generalize and transport treatment effect estimates to a target population; 2022. https://arxiv.org/abs/2203.08701. Search in Google Scholar

[47] Dahabreh IJ, Robertson SE, Tchetgen EJT, Stuart EA, Hernán MA. Generalizing causal inferences from individuals in randomized trials to all trial-eligible individuals. Biometrics. 2019;75:685–94. https://onlinelibrary.wiley.com/doi/abs/10.1111/biom.13009. 10.1111/biom.13009Search in Google Scholar PubMed

[48] Mayer I, Josse J, Group T. Generalizing treatment effects with incomplete covariates; 2021. Available from: https://arxiv.org/abs/2104.12639. Search in Google Scholar

[49] Box GEP. A general distribution theory for a class of likelihood criteria. Biometrika. 1949 12;36(3–4):317–46. 10.1093/biomet/36.3-4.317. Search in Google Scholar

[50] Friendly M, Sigal M. Visualizing tests for equality of covariance matrices. Am Statist. 2020;74(2):144–55. 10.1080/00031305.2018.1497537. Search in Google Scholar

[51] Lunceford JK, Davidian M. Stratification and weighting via the propensity score in estimation of causal treatment effects: a comparative study. Statistics in medicine. 2004;23:2937–60. 10.1002/sim.7231Search in Google Scholar PubMed

[52] Correa J, Tian J, Bareinboim E. Generalized adjustment under confounding and selection biases. Proceedings of the AAAI Conference on Artificial Intelligence; 2018. Vol. 32(1). https://ojs.aaai.org/index.php/AAAI/article/view/12125. 10.1609/aaai.v32i1.12125Search in Google Scholar

[53] Robinson P. Root-N-consistent semiparametric regression. Econometrica. 1988;56(4):931–54. https://EconPapers.repec.org/RePEc:ecm:emetrp:v:56:y:1988:i:4:p:931-54. 10.2307/1912705Search in Google Scholar

[54] Wager S. STATS 361: Causal inference. 2020. https://web.stanford.edu/∽swager/teaching.html.Search in Google Scholar

[55] Nie X, Wager S. Quasi-Oracle estimation of heterogeneous treatment effects. Biometrika. 2020 09;108:299 319. 10.1093/biomet/asaa076Search in Google Scholar

[56] Chen X, Hong H, Tamer E. Measurement error models with auxiliary data. Rev Econ Studies. 2005 02;72:343–66. 10.1111/j.1467-937X.2005.00335.xSearch in Google Scholar

[57] Chen X, Hong H, Nekipelov D. Measurement error models; 2007. https://www.semanticscholar.org/paper/MEASUREMENT-ERROR-MODELS-Chen-Hong/543cc793a1d900e138fa9b132fae7dd8b65dad3d.Search in Google Scholar

[58] Angrist JD, Pischke JS. Mostly harmless econometrics: an empiricistas companion. Economics Books, Princeton University Press; 2009. 10.1515/9781400829828Search in Google Scholar

[59] Wooldridge JM. Introductory econometrics: a modern approach (4th ed., international student ed.). Nelson Education; 2009. Search in Google Scholar

[60] Finn JD, Achilles CM. Answers and questions about class size: a statewide experiment. Am Educ Res J. 1990;27(3):557–77. 10.3102/00028312027003557. Search in Google Scholar

[61] Kallus N, Puli AM, Shalit U. Removing hidden confounding by experimental grounding. In Advances in Neural Information Processing Systems; 2018. p. 10888–97. Search in Google Scholar

[62] CRASH-3. Effects of tranexamic acid on death, disability, vascular occlusive events and other morbidities in patients with acute traumatic brain injury (CRASH-3): a randomised, placebo-controlled trial. The Lancet. 2019;394(10210):1713–23. 10.1016/S0140-6736(19)32233-0. Search in Google Scholar PubMed PubMed Central

[63] Dewan Y, Komolafe E, Mejìa-Mantilla J, Perel P, Roberts I, Shakur-Still H. CRASH-3: Tranexamic acid for the treatment of significant traumatic brain injury: study protocol for an international randomized, double-blind, placebo-controlled trial. Trials. 2012 06;13:87. 10.1186/1745-6215-13-87Search in Google Scholar PubMed PubMed Central

[64] Brenner A, Arribas M, Cuzick J, Jairath V, Stanworth S, Ker K, et al. Outcome measures in clinical trials of treatments for acute severe haemorrhage. Trials. 2018;19:533. 10.1186/s13063-018-2900-4Search in Google Scholar PubMed PubMed Central

[65] van Buuren S. Flexible imputation of missing data. Second Edition. Boca Raton, FL: Chapman and Hall/CRC; 2018. https://stefvanbuuren.name/fimd/. 10.1201/9780429492259Search in Google Scholar

[66] Mansukhani R, Frimley L, Shakur-Still H, Sharples L, Roberts I. Accuracy of time to treatment estimates in the CRASH-3 clinical trial: impact on the trial results. Trials. 2020 07;21:1–8. 10.1186/s13063-020-04623-5Search in Google Scholar PubMed PubMed Central

[67] Kennedy EH. Semiparametric theory and empirical processes in causal inference. In He H, Wu P, Chen D (Eds.), Statistical causal inferences and their applications in public health research. New York: Springer. 2016:141–67. 10.1007/978-3-319-41259-7 8 (arxiv:1510.04740).Search in Google Scholar

[68] Dahabreh IJ, Robertson SE, Steingrimsson JA, Stuart EA, Hernán MA. Extending inferences from a randomized trial to a new target population. Stat Med. 2020;39(14):1999–2014. 10.1002/sim.8426Search in Google Scholar PubMed

[69] Chernozhukov V, Chetverikov D, Demirer M, Duflo E, Hansen C, Newey W, et al. Double/debiased machine learning for treatment and structural parameters. Econom J. 2018;21(1):C1–C68. https://doi.org/10.1111/ectj.12097.10.3386/w23564Search in Google Scholar

[70] Ross SM. A first course in probability. 5th ed. Upper Saddle River, N.J.: Prentice Hall; 1998. Search in Google Scholar

[71] Gao Z, Hastie T. Estimating heterogeneous treatment effects for general responses; 2021. https://arxiv.org/abs/2103.04277. Search in Google Scholar

Received: 2021-11-02
Revised: 2022-09-15
Accepted: 2022-09-15
Published Online: 2022-11-22

© 2022 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 28.4.2024 from https://www.degruyter.com/document/doi/10.1515/jci-2021-0059/html
Scroll to top button