Skip to main content
Log in

A semiparametric multiply robust multiple imputation method for causal inference

  • Published:
Metrika Aims and scope Submit manuscript

Abstract

Evaluating the impact of non-randomized treatment on various health outcomes is difficult in observational studies because of the presence of covariates that may affect both the treatment or exposure received and the outcome of interest. In the present study, we develop a semiparametric multiply robust multiple imputation method for estimating average treatment effects in such studies. Our method combines information from multiple propensity score models and outcome regression models, and is multiply robust in that it produces consistent estimators for the average causal effects if at least one of the models is correctly specified. Our proposed estimators show promising performances even with incorrect models. Compared with existing fully parametric approaches, our proposed method is more robust against model misspecifications. Compared with fully non-parametric approaches, our proposed method does not have the problem of curse of dimensionality and achieves dimension reduction by combining information from multiple models. In addition, it is less sensitive to the extreme propensity score estimates compared with inverse propensity score weighted estimators and augmented estimators. The asymptotic properties of our method are developed and the simulation study shows the advantages of our proposed method compared with some existing methods in terms of balancing efficiency, bias, and coverage probability. Rubin’s variance estimation formula can be used for estimating the variance of our proposed estimators. Finally, we apply our method to 2009–2010 National Health Nutrition and Examination Survey to examine the effect of exposure to perfluoroalkyl acids on kidney function.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  • Byers T, Nestle M, McTiernan A, Doyle C, Currie-Williams A, Gansler T, Thun M (2002) American cancer society guidelines on nutrition and physical activity for cancer prevention: reducing the risk of cancer with healthy food choices and physical activity. CA A Cancer J Clin 52(2):92–119

    Article  Google Scholar 

  • Calle EE, Kaaks R (2004) Overweight, obesity and cancer: epidemiological evidence and proposed mechanisms. Nat Rev Cancer 4(8):579–591

    Article  Google Scholar 

  • Cao W, Tsiatis AA, Davidian M (2009) Improving efficiency and robustness of the doubly robust estimator for a population mean with incomplete data. Biometrika 96(3):723–734

    Article  MathSciNet  MATH  Google Scholar 

  • Chen S, Haziza D (2017) Multiply robust imputation procedures for the treatment of item nonresponse in surveys. Biometrika 104(2):439–453

    MathSciNet  MATH  Google Scholar 

  • Chen S, Haziza D (2019) On the nonparametric multiple imputation with multiply robustness. Stat Sin 29:2035–2053

    MATH  Google Scholar 

  • Cohen HW, Hailpern SM, Fang J, Alderman MH (2006) Sodium intake and mortality in the NHANES II follow-up study. Am J Med 119(3):275-e7

    Article  Google Scholar 

  • De Luna X, Waernbaum I, Richardson TS (2011) Covariate selection for the nonparametric estimation of an average treatment effect. Biometrika 98(4):861–875

    Article  MathSciNet  MATH  Google Scholar 

  • Devroye LP, Wagner TJ (1977) The strong uniform consistency of nearest neighbor density estimates. Ann Stat 5(3):536–540

  • Duan X, Yin G (2017) Ensemble approaches to estimating the population mean with missing response. Scand J Stat 44(4):899–917

    Article  MathSciNet  MATH  Google Scholar 

  • French SA, Hennrikus DJ, Jeffery RW (1996) Smoking status, dietary intake, and physical activity in a sample of working adults. Health Psychol 15(6):448–454

    Article  Google Scholar 

  • Graham JW, Olchowski AE, Gilreath TD (2007) How many imputations are really needed? some practical clarifications of multiple imputation theory. Prev Sci 8(3):206–213

    Article  Google Scholar 

  • Han P (2018) Calibration and multiple robustness when data are missing not at random. Stat Sin 28(4):1725–1740

    MathSciNet  MATH  Google Scholar 

  • Han P, Wang L (2013) Estimation with missing data: beyond double robustness. Biometrika 100(2):417–430

    Article  MathSciNet  MATH  Google Scholar 

  • Healy GN, Matthews CE, Dunstan DW, Winkler EAH, Owen N (2011) Sedentary time and cardio-metabolic biomarkers in us adults: Nhanes 2003–06. Eur Heart J 32(5):590–597

    Article  Google Scholar 

  • Hebert JR, Kabat GC (1990) Differences in dietary intake associated with smoking status. Eur J Clin Nutr 44(3):185–193

    Google Scholar 

  • Heitjan DF, Little RJ (1991) Multiple imputation for the fatal accident reporting system. J Roy Stat Soc: Ser C (Appl Stat) 40(1):13–29

    MATH  Google Scholar 

  • Hernan M, Robins J (2020) Causal inference: What if. boca raton: Chapman & hill/crc

  • Holland PW (1986) Statistics and causal inference. J Am Stat Assoc 81(396):945–960

    Article  MathSciNet  MATH  Google Scholar 

  • Hsu CH, Long Q, Li Y, Jacobs E (2014) A nonparametric multiple imputation approach for data with missing covariate values with application to colorectal adenoma data. J Biopharm Stat 24(3):634–648

    Article  MathSciNet  Google Scholar 

  • Kang JD, Schafer JL et al (2007) Demystifying double robustness: a comparison of alternative strategies for estimating a population mean from incomplete data. Stat Sci 22(4):523–539

    MathSciNet  MATH  Google Scholar 

  • Kataria A, Trasande L, Trachtman H (2015) The effects of environmental chemicals on renal function. Nat Rev Nephrol 11(10):610

    Article  Google Scholar 

  • Kim J, Haziza D (2014) Doubly robust inference with missing data in survey sampling. Stat Sin 24(1):375–394

    MathSciNet  MATH  Google Scholar 

  • Levey AS, Stevens LA, Schmid CH, Zhang Y, Castro AF III, Feldman HI, Kusek JW, Eggers P, Van Lente F, Greene T (2009) A new equation to estimate glomerular filtration rate. Ann Intern Med 150(9):604–612

    Article  Google Scholar 

  • Little RJ, Rubin DB (2019) Statistical analysis with missing data, vol 793. John Wiley & Sons, New York

    MATH  Google Scholar 

  • Long Q, Hsu C, Li Y (2012) Doubly robust nonparametric multiple imputation for ignorable missing data. Stat Sin 22:149–172

    Article  MathSciNet  MATH  Google Scholar 

  • Lu CY (2009) Observational studies: a review of study designs, challenges and strategies to reduce confounding. Int J Clin Pract 63(5):691–697

    Article  Google Scholar 

  • Maura M, Boyle P, La Vecchia C, Decarli A, Talamini R, Franceschi S (1998) Population attributable risk for breast cancer: diet, nutrition, and physical exercise. JNCI J Natl Cancer Inst 90(5):389–394

    Article  Google Scholar 

  • Nielsen SF (2003) Proper and improper multiple imputation. Int Stat Rev 71(3):593–607

    Article  MATH  Google Scholar 

  • Pearl J (2009) Causal inference in statistics: an overview. Stat Surv 3:96–146

    Article  MathSciNet  MATH  Google Scholar 

  • Rubin DB (1974) Estimating causal effects of treatments in randomized and nonrandomized studies. J Educ Psychol 66(5):688–701

    Article  Google Scholar 

  • Rubin DB (1987) Multiple imputation for nonresponse in surveys. John Wiley & Sons, New York

    Book  MATH  Google Scholar 

  • Rubin DB (1990) Formal mode of statistical inference for causal effects. J Stat Plan Inference 25(3):279–292

    Article  Google Scholar 

  • Rubin DB (2005) Causal inference using potential outcomes: design, modeling, decisions. J Am Stat Assoc 100(469):322–331

    Article  MathSciNet  MATH  Google Scholar 

  • Rubin DB, Schenker N (1991) Multiple imputation in health-are databases: an overview and some applications. Stat Med 10(4):585–598

    Article  Google Scholar 

  • Schafer J (1999) Multiple imputation: a primer. Stat Methods Med Res 8(1):3–15

    Article  Google Scholar 

  • Schafer JL (1999) Multiple imputation: a primer. Stat Methods Med Res 8(1):3–15

    Article  Google Scholar 

  • Shankar A, Xiao J, Ducatman A (2011) Perfluoroalkyl chemicals and chronic kidney disease in us adults. Am J Epidemiol 174(8):893–900

    Article  Google Scholar 

  • Silverman BW (1978) Weak and strong uniform consistency of the kernel estimate of a density and its derivatives. Ann Stat 6(1):177–184

  • Stone CJ (1977) Consistent nonparametric regression. Ann Stat 5(4):595–620

  • Van der Vaart AW (2000) Asymptotic statistics, vol 3. Cambridge University Press, Cambridge

    Google Scholar 

  • Watkins DJ, Josson J, Elston B, Bartell SM, Shin H-M, Vieira VM, Savitz DA, Fletcher T, Wellenius GA (2013) Exposure to perfluoroalkyl acids and markers of kidney function among children and adolescents living near a chemical plant. Environ Health Perspect 121(5):625–630

    Article  Google Scholar 

  • Zhang S (2019) Multiply robust empirical likelihood inference for missing data and causal inference problems. University of Waterloo, Waterloo

    Google Scholar 

  • Zhao J, Hinton P, Chen J, Jiang J (2020) Causal inference for the effect of environmental chemicals on chronic kidney disease. Comput Struct Biotechnol J 18:93–99

    Article  Google Scholar 

Download references

Acknowledgements

S. Chen was supported by the National Institute on Minority Health and Health Disparities (NIMHD) at National Institutes of Health (NIH) (1R21MD014658-01A1) and the Oklahoma Shared Clinical and Translational Resources (U54GM104938) with an Institutional Development Award (IDeA) from National Institute of General Medical Sciences. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health. The work of D. Haziza was supported by grants from the Natural Sciences and Engineering Research Council of Canada.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sixia Chen.

Ethics declarations

Conflict of interest statement

On behalf of all authors, the corresponding author states that there is no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

1.1 A: Regularity conditions

Before providing sketched proofs of Theorems 1 and 2, we will first provide some necessary regularity conditions.

Let \(\hat{Z}_{i}^{(g)}=\left( \hat{Z}_{1i}^{(g)},\hat{Z}_{2i}^{(g)}\right) \) where \(\hat{Z}_{1i}^{(g)}=\hat{m}_{g,i}\) and \(\hat{Z}_{2i}^{(g)}=\hat{p}_{g,i}\). Next, let \(\beta _{0g}^{*(k)}\),\(\alpha ^{*(j)}\), and \(\eta _{m_g}^{*}\) be the probability limits for the estimators \(\hat{\beta }_{0g}^{(k)}\), \(\;\hat{\alpha }^{(j)}\), and \(\;\hat{\eta }_{m_g}\), respectively.

Denote

$$\begin{aligned} Z_i^{*(g)}=(Z_{1i}^{*(g)},Z_{2i}^{*(g)}) \end{aligned}$$

with

$$\begin{aligned} Z_{1i}^{*(g)}=\hat{U}_{m_{g},i}^{*\top }\; \frac{\eta _{m_g}^{*2}}{\eta ^{*T}_{p}\eta ^{*}_{m_1}} ,\;\; Z_{2i}^{*(1)}=\hat{U}_{p,i}^{*\top }\; \frac{\eta _{p}^{*2}}{\eta ^{*T}_{p}\eta ^{*}_{p}},\;\text {and}\;\; Z_{2i}^{*(0)}=1-\left( \hat{U}_{p,i}^{*\top }\; \frac{\eta _{p}^{*2}}{\eta ^{*T}_{p}\eta ^{*}_{p}}\right) \end{aligned}$$

where

$$\begin{aligned} \hat{U}_{m_{g},i}^{*}=\{m_g^{(1)}(X_i;\beta _{0g}^{*(1)}),...,m_g^{(K)}(X_i;\beta _{0g}^{*(K)})\}^\top \end{aligned}$$

and

$$\begin{aligned} \hat{U}_{p,i}^*=\{p^{(1)}(X_i;\alpha ^{*(1)}),...,p^{(J)}(X_i;\alpha ^{*(J)})\}^\top . \end{aligned}$$

Finally, let \(f_g(Z^{*(g)})\) be the density function of \(Z^{*(g)}\). We assume the following regularity conditions necessary for proving Theorem 1 and Theorem 2. Conditions (C1) and (C2) apply to each propensity score model, \(j=1,...,J\), and each pair of outcome regression models, \(k=1,...,K\).

  1. (C1)

    \(\hat{\alpha }^{(j)}\) is the unique solution of \(S_p^{(j)}(\alpha ^{(j)})=0\) and \(\;\hat{\beta }_{0g}^{(k)}\) is the unique solution of \(S_{m_g}^{(k)}(\beta _{0g}^{(k)})=0\) where \(S_p^{(j)}(\alpha ^{(j)})\) and \(S_{m_g}^{(k)}\) are as defined in Sect. 3.

  2. (C2)

    \(S_p^{(j)}(\alpha ^{(j)})\) converges almost surely to \(S_p^{*(j)}(\alpha ^{(j)})=E\{S_p^{(j)}(\alpha ^{(j)})\}\), uniformly in \(\alpha ^{(j)}\), and \(S_p^{*(j)}(\alpha ^{(j)})=0\) has a unique solution \(\alpha ^{*(j)}\). Also, \(S_{m_g}^{(k)}(\beta _{0g}^{(k)})\) converges almost surely to \(S_{m_g}^{*(k)}(\beta _{0g}^{(k)})=E\{S_{m_g}^{(k)}(\beta _{0g}^{(k)})\}\), uniformly in \(\beta _{0g}^{(k)}\), and \(S_{m_g}^{*(k)}(\beta _{0g}^{(k)})=0\) has a unique solution in \(\beta _{0g}^{*(k)}\).

  3. (C3)

    \(E(Y_g^2)<\infty \) and \(E\{\mu _g^{2}(Z^{*(g)})\}<\infty \), where \(\mu _g^{2}(Z^{*(g)})=E(Y_g\vert Z^{*(g)})\).

  4. (C4)

    \(H/n=o(1)\) and \(\text {log}(n)/H=o(1)\).

  5. (C5)

    \(f_g(Z^{*(g)})\) and \(\pi _g(Z^{*(g)})\) are continuous and bounded away from 0 in the compact support of \(Z^{*(g)}\).

The consistency of \(\hat{\alpha }^{(j)}\) and \(\hat{\beta }_{0g}^{(k)}\) is ensured by Conditions (C1) and (C2). These conditions are satisfied for most linear (and generalized linear) models. Condition (C3) is useful for deriving the asymptotic expansion and normality of \(\hat{\tau }_{MRMI}\). Condition (C4) is used to control the asymptotic order of H. Condition (C5), a common condition in nonparametric statistics, helps avoid extreme values of the propensity and density scores.

1.2 B: Sketched proof of Theorem 1

Let \(\hat{\alpha }=(\hat{\alpha }^{(1)},...,\hat{\alpha }^{(J)})\), \(\hat{\beta }_{0g}=(\hat{\beta }_{0g}^{(1)},...,\hat{\beta }_{0g}^{(K)})\), \(\hat{\alpha }^*=(\hat{\alpha }^{*(1)},...,\hat{\alpha }^{*(J)})\), and \(\hat{\beta }^*_{0g}=(\hat{\beta }_{0g}^{*(1)},...,\hat{\beta }_{0g}^{*(K)})\). According to (C1), (C2), and Van der Vaart (2000), it can be shown that \(\hat{\alpha }\rightarrow ^{p}\alpha ^*\) and \(\hat{\beta }_{0g}\rightarrow ^{p}\beta _{0g}^{*}\).

Assume that one of the pairs of outcome regression models is correct, say \(m_1^{(1)}(X_i;\beta _{01}^{(1)})\) and \(m_0^{(1)}(X_i;\beta _{00}^{(0)})\). Then, we have \(\beta _{0g}^{*(1)}=\beta _{0g}\) and \(\;\eta _{m_g}^{*}=(1,0,...,0)^{T}\), which implies that \(Z_{1i}^{*(g)}=m_g(X_i;\beta _{0g})\).

It follows that

$$\begin{aligned} \begin{aligned} E\left( Y_1\vert T,Z^{*(1)}\right)&= E\left\{ E\left( Y_1\vert T,Z^{*(1)},X\right) \vert T,Z^{*(1)}\right\} \\&=E\left\{ E\left( Y_1\vert Z^{*(1)},X\right) \vert T,Z^{*(1)}\right\} \\&=E\left( Z_1^{*(1)}\vert T,Z^{*(1)}\right) =Z_1^{*(1)}=E\left( Y_1\vert Z^{*(1)}\right) \end{aligned} \end{aligned}$$

If one of the propensity score models is correctly specified, say \(p_1^{(1)}(X_i;\alpha ^{(1)})\), then \(\alpha ^{*(1)}=\alpha \),\(\;\eta _p^*=(1,0,...,0)^\top \) and \(\;Z_{2i}^{*(1)}=p_1(X_i;\alpha )\). Therefore,

$$\begin{aligned} E\left( Y_1 \vert T,Z^{*(1)}\right) =E\left( Y_1\vert Z^{*(1)}\right) \end{aligned}$$

due to the fact that \(Y_1\) is independent of T given \(Z_2^{*(1)}\).

According to Devroye and Wagner (1977) and Silverman (1978), it can be shown that

$$\begin{aligned} \sum _{j \in R_H^{(1)}(i)} \frac{1}{H}Y_{1j}\rightarrow ^{p}E\left( Y_{1i} \vert Z_i^{*(1)},T_i=1\right) \; \; \; \; \; \end{aligned}$$
(B.1)

and

$$\begin{aligned} \frac{1}{H}\sum _{j \in R_H^{(1)}(i)}(Y_{1j}-\bar{Y}_{1R_H^{(1)}(i)})^2\rightarrow ^{p}V(Y_{1i} \vert Z_i^{*(1)},T_i=1) \; \; \; \; \; \end{aligned}$$
(B.2)

uniformly for \(i\in s\) as \(n \rightarrow \infty \), \(L \rightarrow \infty \), \(H \rightarrow \infty \), and conditions (C4) and (C5). In addition, according to Chebyshev’s inequality, (B.1), and (B.2) we have

$$\begin{aligned}&Pr\left( \left| \frac{1}{n} \sum _{i=1}^n(1-T_i)\frac{1}{L}\sum _{l=1}^LY_{1i}^{*(l)}-E\left\{ (1-T_i)E(Y_{1i}^{*(l)}\vert Y_1,X,T)\right\} \right| > \epsilon \right) \nonumber \\&\quad \le \epsilon ^{-2}V\left\{ \frac{1}{n}\sum _{i=1}^n(1-T_i)\frac{1}{L}\sum _{l=1}^L Y_{1i}^{*(l)}\right\} \nonumber \\&\quad =\epsilon ^{-2}\left[ V \left\{ \frac{1}{n}\sum _{i=1}^n(1-T_i)E\left( Y_{1i}^{*(l)}\vert Y_1,X,T\right) \right\} \right. \nonumber \\&\qquad \left. +E \left\{ \frac{1}{n^2}\sum _{i=1}^n\left( 1-T_i\right) \frac{1}{L}V\left( Y_{1i}^{*(l)}\vert Y_1,X,T\right) \right\} \right] \nonumber \\&\quad =\epsilon ^{-2} \left[ V\left\{ \frac{1}{n}\sum _{i=1}^n \left( 1-T_i\right) \sum _{j \in R_H^{(1)}(i)}\frac{1}{H}Y_{1j}\right\} \right. \nonumber \\&\quad \left. + E\left\{ \frac{1}{n^2}\sum _{i=1}^n(1-T_i) \frac{1}{L} \frac{1}{H}\sum _{j \in R_H^{(1)}(i)}\left( Y_{1j}-\bar{Y}_{1R_H^{(1)}(i)}\right) ^2\right\} \right] \nonumber \\&=\epsilon ^{-2}\left[ V\left\{ \frac{1}{n}\sum _{i=1}^n \left( 1-T_i\right) E(Y_{1i}\vert Z_i^{*(1)},T_i=1)\right\} \right. \nonumber \\&\quad \left. + E\left\{ \frac{1}{n^2}\sum _{i=1}^n(1-T_i)\frac{1}{L}V(Y_{1i} \vert Z_i^{*(1)},T_i=1)\right\} \right] \nonumber \\&\quad +o(n^{-1})+o(n^{-1}L^{-1})\nonumber \\&=o(n^{-1})+o(n^{-1}L^{-1}) \; \; \; \; \; \end{aligned}$$
(B.3)

According to (B.3), we have

$$\begin{aligned} \frac{1}{n}\sum _{i=1}^n \left( 1-T_i\right) \frac{1}{L}\sum _{l=1}^LY_{1i}^{*(l)} \rightarrow ^p E\{(1-T_i)E(Y_{1i}^{*(l)}\vert Y_1,X,T)\}\; \; \; \; \end{aligned}$$
(B.4)

as \(n \rightarrow \infty \), \(L \rightarrow \infty \), and \(H \rightarrow \infty \). Therefore, if at least one of the propensity score models or one of the pairs of regression models is correctly specified, according to (B.1) and (B.4), we have

$$\begin{aligned} \begin{aligned} \hat{\mu }_{1MRMI}&=\frac{1}{n}\sum _{i=1}^n \left\{ T_iY_{1i}+(1-T_i)\frac{1}{L}\sum _{l=1}^LY_{1i}^{*(l)}\right\} \\&\rightarrow ^p E(T_iY_{1i})+E \left\{ (1-T_i)E(Y_{1i}^{*(l)}\vert Y_1,X,T)\right\} \\&=E(T_iY_{1i})+E\left\{ (1-T_i)E \left( \sum _{j \in R_H^{(l)(1)}(i)}\frac{1}{H}Y_{1i}^{*(l)}\vert X,Y_1,T \right) \right\} \\&=E(T_iY_{1i})+E\left\{ (1-T_i)\sum _{j \in R_H^{(1)}(i)}\frac{1}{H}Y_{1j}\right\} \rightarrow E(T_iY_{1i})\\&+E\left\{ (1-T_i)E\left( Y_{1i}\vert Z_i^{*(1)},T_i=1\right) \right\} \\&=E(T_iY_{1i})+E\left\{ (1-T_i)Y_{1i}\right\} \\&=E(Y_1) \end{aligned} \end{aligned}$$

as \(n \rightarrow \infty \), \(L \rightarrow \infty \), and \(H \rightarrow \infty \), where we used the consistent probability weights argument from Stone (1977) in the derivation.

It follows by similar argument that \(E(\hat{\mu }_{0MRMI}) \rightarrow ^p E(Y_0)\) as \(n \rightarrow \infty \), \(L \rightarrow \infty \), and \(H \rightarrow \infty \).

Then

$$\begin{aligned} E(\hat{\tau })=E(\hat{\mu }_{1MRMI})-E(\hat{\mu }_{0MRMI})\rightarrow ^p E(Y_1)-E(Y_0) \end{aligned}$$

as \(n \rightarrow \infty \), \(L \rightarrow \infty \), and \(H \rightarrow \infty \). Therefore, Theorem 1 is proven.

1.3 C: Sketched proof of Theorem 2

We can write \(\hat{\mu }_{1MRMI}\) as

$$\begin{aligned} \hat{\mu }_{1MRMI}=\hat{\mu }_1+\hat{\mu }_{1MRMI}-\hat{\mu }_1 \end{aligned}$$

where

$$\begin{aligned} \hat{\mu }_1=\frac{1}{n}\sum _{i=1}^n\left\{ T_iY_{1i}+(1-T_i)\sum _{j\in R_H^{(1)}(i)}\frac{1}{H}Y_{1j}\right\} \; \; \; \; \; \end{aligned}$$
(C.1)

and

$$\begin{aligned} \hat{\mu }_{1MRMI}-\hat{\mu }_1=\frac{1}{n}\sum _{i=1}^n\left\{ (1-T_i)\left( \frac{1}{L}\sum _{l=1}^L Y_{1i}^{*(1)(l)}-\frac{1}{H}\sum _{j \in R_H^{(1)}(i)}Y_{1j}\right) \right\} \; \; \; \; \; \end{aligned}$$
(C.2)

Then, using a first-order Taylor expansion of \(\hat{\mu }_1\) about \(\theta _1=\theta ^*_1\), we obtain

$$\begin{aligned} \hat{\mu }_1=\hat{\mu }_1^*+E\left( \frac{\partial \hat{\mu }_1^*}{\partial \theta _1}\right) (\hat{\theta _1}-\theta ^*_1)+o_p(n^{-\frac{1}{2}}) \; \; \; \; \; \end{aligned}$$
(C.3)

where \(\partial \hat{\mu }_1^*/\partial \theta _1\) is \(\partial \hat{\mu }_1/\partial \theta _1\) evaluated at \(\theta ^*_1\). Because \(\theta _1\) is the solution of the estimating equation \(S_{m_1p}(\theta _1)=0\), one can show that

$$\begin{aligned} \hat{\theta _1}-\theta ^*_1=-\left[ E\left\{ \frac{\partial S_{m_1p}(\theta ^*_1)}{\partial \theta _1}\right\} \right] ^{-1}S_{m_1p}(\theta ^*_1)+o_p(n^{-\frac{1}{2}}) \; \; \; \; \; \end{aligned}$$
(C.4)

Plugging (C.4) into (C.3) yields

$$\begin{aligned} \hat{\mu }_1=\hat{\mu }_1^*+A_1S_{m_1p}(\theta ^*_1) \; \; \; \; \; \end{aligned}$$
(C.5)

where \(A_1\) is defined in Theorem 2 of Sect. 4. Following a similar argument to that of Long et al. (2012), given regularity conditions (C1)–(C5), one can show that

$$\begin{aligned} \hat{\mu }_1^*-\mu _1=Q_1+Q_2+Q_3+o_p(n^{-\frac{1}{2}}) \; \; \; \; \; \end{aligned}$$
(C.6)

where

$$\begin{aligned}&Q_1=\frac{1}{n}\sum _{i=1}^n\left\{ \psi _1\left( Z_i^{*(1)}\right) -\mu _1\right\} \;,\; Q_2=\frac{1}{n}\sum _{i=1}^n T_i\left\{ Y_{1i}-\psi _1\left( Z_i^{*(1)}\right) \right\} ,\; \text {and} \\&Q_3=\frac{1}{n}\sum _{i=1}^n \frac{1-\pi _1\left( Z_i^{*(1)}\right) }{\pi _1\left( Z_i^{*(1)}\right) }T_i\left\{ Y_{1i}-\psi _1\left( Z_i^{*(1)}\right) \right\} . \end{aligned}$$

Because \(\hat{\mu }_{1MRMI}-\hat{\mu }_1\) and \(\hat{\mu }_1\) are asymptotically independent, we have

$$\begin{aligned} \begin{aligned} V(\hat{\mu }_{1MRMI}-\hat{\mu }_1)&=V\left\{ E(\hat{\mu }_{1MRMI}-\hat{\mu }_1\vert X, Y_1,T)\right\} \\&=E\left\{ V(\hat{\mu }_{1MRMI}-\hat{\mu }_1\vert X, Y_1,T)\right\} \\&=\frac{1}{nL}E\left\{ \frac{1}{n}\sum _{i=1}^n(1-T_i)V(Y_{1i}^{*(l)} \vert Y_1,X,T)\right\} \\&=\frac{1}{nL}E\left\{ \frac{1}{n}\sum _{i=1}^n(1-T_i)\frac{1}{H}\sum _{j \in R_H^{(1)}(i)}(Y_{1j}-\bar{Y}_{1R_{H^{(1)}(i)}})^2 \right\} \\&=O\left( \frac{1}{nL}\right) \; \; \; \; \; \end{aligned} \end{aligned}$$
(C.7)

where \(\bar{Y}_{1R_{H^{(1)}(i)}}=H^{-1}\sum _{j \in R_H^{(1)}(i)}Y_{1j}\). According to the asymptotic independence between \(\hat{\mu }_{1MRMI}-\hat{\mu }_1\) and \(\hat{\mu }_1\), \(Q_1\) and \(Q_2\), \(Q_1\) and \(Q_3\), Eqs. (C.1), (C.2), (C.5)–(C.7), and regularity condition (C3), we have

$$\begin{aligned} \begin{aligned} \hat{\mu }_{1MRMI}&=\frac{1}{n}\sum _{i=1}^n \left[ \frac{T_i}{\pi _1\left( Z_i^{*(1)}\right) }Y_{1i}+\left\{ 1-\frac{T_i}{\pi _1\left( Z_i^{*(1)}\right) }\right\} \right. \\&\left. \psi _1\left( Z_i^{*(1)}\right) +A_1s_1(X_i;T_i;\theta ^*_1)\right] \\&\quad +o_p(n^{-\frac{1}{2}}) \end{aligned} \end{aligned}$$

as \(n \rightarrow \infty \) and \(L \rightarrow \infty \).

We may apply a similar argument for \(\hat{\mu }_{0MRMI}\) to obtain

$$\begin{aligned} \begin{aligned} \hat{\mu }_{0MRMI}&=\frac{1}{n}\sum _{i=1}^n \left[ \frac{1-T_i}{\pi _0\left( Z_i^{*(0)}\right) }Y_{0i}+\left\{ 1-\frac{1-T_i}{\pi _0\left( Z_i^{*(0)}\right) }\right\} \right. \\&\quad \left. \psi _0\left( Z_i^{*(0)}\right) +A_0s_0(X_i;T_i;\theta ^*_0)\right] \\&\quad +o_p(n^{-\frac{1}{2}}) \end{aligned} \end{aligned}$$

as \(n \rightarrow \infty \) and \(L \rightarrow \infty \).

Then, because \(\hat{\tau }_{MRMI}=\hat{\mu }_{1MRMI}-\hat{\mu }_{0MRMI}\), we have

$$\begin{aligned} \begin{aligned} \hat{\tau }_{MRMI}&=\frac{1}{n}\sum _{i=1}^n \Bigg [\left\{ \frac{T_i}{\pi _1\left( Z_i^{*(1)}\right) }Y_{1i}+\left\{ 1-\frac{T_i}{\pi _1\left( Z_i^{*(1)}\right) }\right\} \psi _1\left( Z_i^{*(1)}\right) +A_1s_1(X_i;T_i;\theta ^*_1)\right\} \\&-\left\{ \frac{1-T_i}{\pi _0\left( Z_i^{*(0)}\right) }Y_{0i}+\left\{ 1-\frac{1-T_i}{\pi _0\left( Z_i^{*(0)}\right) }\right\} \psi _0\left( Z_i^{*(0)}\right) +A_0s_0(X_i;T_i;\theta ^*_0)\right\} \Bigg ]+o_p(n^{-\frac{1}{2}}) \end{aligned} \end{aligned}$$

Then, by the Central Limit Theorem, we obtain

$$\begin{aligned} n^{\frac{1}{2}}(\hat{\tau }_{MRMI}-\tau )\rightarrow ^dN(0,\sigma ^2_{MRMI}) \end{aligned}$$

with \(\sigma ^2_{MRMI}\) defined in Theorem 2 of Sect. 4.

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gochanour, B., Chen, S., Beebe, L. et al. A semiparametric multiply robust multiple imputation method for causal inference. Metrika 86, 517–542 (2023). https://doi.org/10.1007/s00184-022-00883-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00184-022-00883-0

Keywords

Navigation