This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Bayesian Nonparametric Classification for Incomplete Data
With a High Missing Rate: an Application to Semiconductor Manufacturing Data

Sewon Park Department of Statistics, Seoul National University Kyeongwon Lee Department of Statistics, Seoul National University Da-Eun Jeong Samsung Electronics Heung-Kook Ko Samsung Electronics Jaeyong Lee Department of Statistics, Seoul National University
Abstract

During the semiconductor manufacturing process, predicting the yield of the semiconductor is an important problem. Early detection of defective product production in the manufacturing process can save huge production cost. The data generated from the semiconductor manufacturing process have characteristics of highly non-normal distributions, complicated missing patterns and high missing rate, which complicate the prediction of the yield. We propose Dirichlet process - naive Bayes model (DPNB), a classification method based on the mixtures of Dirichlet process and naive Bayes model. Since the DPNB is based on the mixtures of Dirichlet process and learns the joint distribution of all variables involved, it can handle highly non-normal data and can make predictions for the test dataset with any missing patterns. The DPNB also performs well for high missing rates since it uses all information of observed components. Experiments on various real datasets including semiconductor manufacturing data show that the DPNB has better performance than MICE and MissForest in terms of predicting missing values as percentage of missing values increases.


Key words: Missing Data; Imputation; Dirichlet process Mixture Model; Naive Bayes Model; Classification.

1 Introduction

Missing data occur widely in engineering problems and scientific research. Especially, handling missing value is often the first problem to consider in analyzing the data from the fields of the manufacturing process, social sciences, and biology. Prediction improvement of the yield of semiconductor manufacturing are the major issues for having stable and consistent manufacturing processes. For the purpose, a huge amount of data consisting of hundreds of variables or factors and millions of observations in the semiconductor manufacturing process is collected and analyzed. However, about 90 percent of data generated from manufacturing process is not recorded due to data storage limit. Metabolomics data contain typically 10-20 percent missing values, which are caused by biological factors such as metabolites being absent, technical reasons such as the limit of detection (LOD), and measurement error (Lee and Styczynski, 2018; Playdon et al., 2019). Furthermore, social scientists conduct empirical research to test or verify theoretical concepts and hypotheses through survey experiments. Some respondents unfortunately provide no information for the survey item. If the missing data are not dealt with appropriately, researchers may draw a wrong conclusion and waste valuable time and resources.

Rubin (1976) and Little and Rubin (2019) established three types of missingness mechanisms: (a) missing completely at random (MCAR), (b) missing at random (MAR), and (c) missing not at random (MNAR). When missingness is unrelated to both observed and unobserved variables, the data are called MCAR. When missingness only depends on observed data, the data are MAR. MAR is a more plausible assumption than MCAR. When missingness depends on unobserved data or the missing value itself, the data are said to be MNAR. An example of the MNAR is censored missing values caused by the LOD (Wei et al., 2018). In this paper, we present empirical comparisons between the DPNB and other competitors under all missingness mechanisms.

The simplest approach to missing data is to remove incomplete observations or cases. But this method may lead to biased conclusions and lose some useful information. Another approach is the imputation which substitutes missing data for plausible values generated from statistical learning. Imputing missing data may be a more reasonable way than discarding missing values but not necessarily give better results.

We review several popular imputation techniques for estimating missing values. The basic method is the mean imputation, which replaces missing values with the mean of the observed values in a certain variable. The mean imputation is simple to apply but underestimates the variance and produces biased results under both MAR and MNAR (Little and Rubin, 2019). Another method is the multivariate imputation by chained equations (MICE) developed by Buuren and Groothuis-Oudshoorn (2010). The MICE as a type of multiple imputations (Rubin, 1976) constructs separate conditional models for each incomplete variable and iteratively imputes the missing values. Gaussian mixture model (GMM) is one of the most widely used model-based imputation methods (Ghahramani and Jordan, 1994; Lin et al., 2006; Williams et al., 2007). Model-based methods assume the joint distribution of all variables in the data and estimate parameters of the distribution (Das et al., 2018). GMM on incomplete data imputes missing values using conditional distribution properties of a multivariate normal distribution.

Contrary to the model-based methods, there are imputation strategies based on machine learning algorithms that do not rely upon distributional assumption on the data. Burgette and Reiter (2010) and Stekhoven and Bühlmann (2012) designed decision tree-based imputation methods using classification and regression tree and random forest, respectively. Another non-parametric approach is K-Nearest Neighbors (KNN) based imputation (Troyanskaya et al., 2001). Caruana (2001) and Brás and Menezes (2007) improved the accuracy of estimated missing values using an iterative process. The aforementioned imputation methods from machine learning can deal with both numerical and categorical data as well as mixed-type data. Thus, these methods have been commonly used in various fields related to missing data problems.

Deep learning-based imputation methods have been proposed in different neural network architectures. Sharpe and Solly (1995), Gupta and Lam (1996), and Silva-Ramírez et al. (2011) reconstructed missing values by using feed-forward neural networks (FNN). Bengio and Gingras (1996), Che et al. (2018), and Kim and Chi (2018) introduced recurrent neural networks to handle incomplete sequential data. Vincent et al. (2008) and Gondara and Wang (2018) designed deep generative models combining denoising autoencoders (DAE) to create the clean output from the noisy input considered as missing data. Yeh et al. (2017) and Yoon et al. (2018) provided modified generative adversarial nets (GAN) for filling missing values or regions.

In this paper, we propose a new combined method that performs both imputation and classification tasks. Our proposed method has several advantages over the existing methods. First, the class conditional distribution is an infinite Gaussian mixture model instead of a Gaussian used in a standard naive Bayes classifier. The proposed classifier can construct flexible decision boundaries and learn various types of non-linear decision boundaries. Second, the proposed imputer is more accurate than other imputation methods on incomplete data with high missing rates. In other words, as the percentage of missing values increases, the imputation technique based on class conditional distribution outperforms state-of-the-art methods under both MCAR and MAR assumptions. While other imputation methods utilize partial observed information by missing patterns and algorithms, the DPNB uses all information of observed components. Thus, this gap will be more apparent when the rate of missing data is high. Third, the DPNB is capable of predicting the labels of a set of new cases which have any missing patterns from a predictive model. Imputation techniques described above focus on estimating missing values and cannot conduct subsequent statistical analyses such as classification or regression. Most classification approaches on incomplete data designed combinations of a predictive model and imputation strategy. They used to transform missing data to complete cases and build a classifier in the training phase. However, missing data may exist not only in the training set but also in the test set. Our method can address the test set which has missing values with any missing patterns. Experiments also show that the proposed method gives better classification accuracy on multiple datasets with high missing rates from the UCI Machine Learning Repository.

The remainder of the paper is structured as follows. Section 2 provides preliminary notions of imputation through finite Gaussian mixture models and the Dirichlet process prior as a key factor in our proposed method. In section 3, we propose an extension of the well-known Mixture Discriminant Analysis (MDA), which uses an infinite Gaussian mixture model on incomplete data. Section 4 shows empirical results on the accuracy of imputation and classification in different settings. A real data example is presented in Section 5. We end this paper in section 6 with conclusions.

2 Preliminaries

In this section, we review the Gaussian mixture models on missing data and the Dirichlet process, which are core concepts of our proposed method.

2.1 Gaussian mixture models for missing Data

Algorithms for Gaussian mixture models on missing data have been studied in the last few decades. Ghahramani and Jordan (1994) used the Expectation-Maximization (EM) algorithm to find parameter values and missing components maximizing the likelihood of Gaussian mixture models. Zhang and Everson (2004) developed a Bayesian approach of mixture models using Gibbs sampler. This method utilizes full conditional distributions to obtain the joint posterior distribution of parameters and missing values. Williams et al. (2007) introduced variational inference based on the mean-field approximation for Bayesian mixture models. Both missing values and parameters are iteratively updated until the evidence lower bound (ELBO) converges.

We focus on the estimation of a Gaussian mixture model, permitting simultaneous inference of missing values through Gibbs sampling is a Markov Chain Monte Carlo (MCMC). Let xi,i=1,,n\textbf{x}_{i},i=1,\ldots,n be nn independent pp-dimensional observations from the mixture distribution consists of HH Gaussian components. We partition an observation into two components xi={xioi,ximi}\textbf{x}_{i}=\{\textbf{x}_{i}^{o_{i}},\textbf{x}_{i}^{m_{i}}\}, where oi{1,2,,p}o_{i}\subset\{1,2,\ldots,p\} is an index set of observed variables and mi{1,2,,p}m_{i}\subset\{1,2,\ldots,p\} is an index set of missing variables . That is, xioi\textbf{x}_{i}^{o_{i}} and ximi\textbf{x}_{i}^{m_{i}} indicate the observed components and missing components from the iith observation xi\textbf{x}_{i}, respectively. We can express the mixture distribution as follows:

f(xi)=h=1Hwh𝒩p(xi|𝝁h,𝚺h)=h=1Hwh𝒩p([xioiximi]|[𝝁hoi𝝁hmi],[Σhoi,oiΣhoi,miΣhmi,oiΣhmi,mi])f(\textbf{x}_{i})=\sum_{h=1}^{H}w_{h}\mathcal{N}_{p}(\textbf{x}_{i}|\boldsymbol{\mu}_{h},\boldsymbol{\Sigma}_{h})=\sum_{h=1}^{H}w_{h}\,\mathcal{N}_{p}\left(\begin{bmatrix}\textbf{x}_{i}^{o_{i}}\\ \textbf{x}_{i}^{m_{i}}\end{bmatrix}\Bigg{|}\begin{bmatrix}\boldsymbol{\mu}_{h}^{o_{i}}\\ \boldsymbol{\mu}_{h}^{m_{i}}\end{bmatrix},\begin{bmatrix}\Sigma_{h}^{o_{i},o_{i}}&\Sigma_{h}^{o_{i},m_{i}}\\ \Sigma_{h}^{m_{i},o_{i}}&\Sigma_{h}^{m_{i},m_{i}}\end{bmatrix}\right) (1)

Here, whw_{h} is the non-negative mixing proportion and sum up to one, i.e., h=1Hwh=1\sum_{h=1}^{H}w_{h}=1.

To implement the Gibbs sampling for mixture models with incomplete data, we need the full conditional posterior distributions of model parameters and missing values. In this paper we only cover the full conditional posterior for missing values and skip other parameters of mixture models; see the paper by Franzén (2006) for more details. If the data come from a multivariate normal distribution with mean 𝝁\boldsymbol{\mu} and covariance 𝚺\boldsymbol{\Sigma}, a full conditional density for missing values based on the observed data can derive easily the following equations (2) by using a special property of multivariate normal distribution.

f(ximi|xioi,others)𝒩|mi|(ximi;𝝁mi|oi,𝚺mi|oi),\displaystyle f(\textbf{x}_{i}^{m_{i}}|\textbf{x}_{i}^{o_{i}},\text{others})\sim\mathcal{N}_{|m_{i}|}(\textbf{x}_{i}^{m_{i}};\boldsymbol{\mu}^{m_{i}|o_{i}},\boldsymbol{\Sigma}^{m_{i}|o_{i}}), (2)
𝝁mi|oi=𝝁mi+𝚺mi,oi(𝚺oi,oi)1(xioi𝝁oi),\displaystyle\boldsymbol{\mu}^{m_{i}|o_{i}}=\boldsymbol{\mu}^{m_{i}}+\boldsymbol{\Sigma}^{m_{i},o_{i}}(\boldsymbol{\Sigma}^{o_{i},o_{i}})^{-1}(\textbf{x}_{i}^{o_{i}}-\boldsymbol{\mu}^{o_{i}}),
𝚺mi|oi=𝚺mi,mi𝚺mi,oi(𝚺oi,oi)1𝚺oi,mi.\displaystyle\boldsymbol{\Sigma}^{m_{i}|o_{i}}=\boldsymbol{\Sigma}^{m_{i},m_{i}}-\boldsymbol{\Sigma}^{m_{i},o_{i}}(\boldsymbol{\Sigma}^{o_{i},o_{i}})^{-1}\boldsymbol{\Sigma}^{o_{i},m_{i}}.

Missing values are filled with samples drawn from its full conditional distribution and combined with the fixed observed values to update subsequently other parameters from full conditionals. This updating process is called the multivariate normal imputation, first implemented by Schafer (1997), which is one of the multiple imputation methods. In the case of the mixture models, we assume all data points are generated from the mixture of multivariate normal distributions. A observation with the missing values belong to hhth mixure component is imputed by sampling full conditional distribution for ximi\textbf{x}_{i}^{m_{i}} given 𝝁h\boldsymbol{\mu}_{h} and 𝚺h\boldsymbol{\Sigma}_{h} associated with hhth mixture component. Mixture components that indicate clusters are determined by latent auxiliary variables. See Zhang and Everson (2004).

2.2 Dirichlet process

Dirichlet process (DP) introduced by Ferguson (1973) is the most popular Bayesian nonparametric model and used in many applications for clustering in the last two decades. Let 𝒳\mathcal{X} be a measurable space and \mathcal{B} the Borel σ\sigma-field of subsets of 𝒳\mathcal{X}. Then we say that the random probability measure GG on (𝒳,)(\mathcal{X},\mathcal{B}) follows a Dirichlet process with a concentration parameter α>0\alpha>0 and a baseline probability measure G0G_{0}, denoted by GDP(α,G0)G\sim DP(\alpha,G_{0}), if for every finite disjoint partition A1,AkA_{1},\ldots A_{k} of 𝒳\mathcal{X},

(G(A1),,G(Ak))Dir(αG0(A1),,αG0(Ak)),(G(A_{1}),\ldots,G(A_{k}))\sim\text{Dir}(\alpha G_{0}(A_{1}),\ldots,\alpha G_{0}(A_{k})),

where Dir denotes the Dirichlet distribution. Dirichlet process can be represented in three different ways: (1) pólya urn scheme (Blackwell et al., 1973) (2) Chinese restaurant process (Aldous, 1985) (3) stick-breaking process (Sethuraman, 1994).

Sethuraman (1994) also showed that its realizations are discrete almost surely, even if G0G_{0} is a continuous distribution. The discreteness of the DP implies that it is unsuitable prior for data generated from continuous distributions. To eliminate this drawback, Antoniak (1974) adopted Dirichlet process mixture models (DPMM) having the following hierarchical model formulations: for i=1,,n,i=1,\ldots,n,

xi|θi\displaystyle\textbf{x}_{i}|\theta_{i} indf(xi|θi),\displaystyle\stackrel{{\scriptstyle ind}}{{\sim}}f(\textbf{x}_{i}|\theta_{i}),
θi\displaystyle\theta_{i} iidG,\displaystyle\stackrel{{\scriptstyle iid}}{{\sim}}G,
G\displaystyle G DP(α,G0),\displaystyle\sim DP(\alpha,G_{0}),

where ff is a parametric density function. The DPMM can be expressed as a limit of finite mixture models, where the number of mixture components is taken to infinity (Teh et al., 2006). For example, if a base measure G0G_{0} is a multivariate normal Inverse-Wishart conjugate prior and f(xi|θi)f(\textbf{x}_{i}|\theta_{i}) is multivariate normal, the DPMM can be an infinite normal mixture model. The advantage of the DPMM is the number of mixture components is not fixed in advance of fitting the data and automatically infer the number of clusters from the data unlike finite normal mixture models. Various algorithms have been developed for posterior inference of the DPMM such as marginal Gibbs sampling (MacEachern, 1994; Escobar, 1994; Escobar and West, 1995; Bush and MacEachern, 1996; Neal, 2000), conditional Gibbs sampling (Ishwaran and James, 2001; Walker, 2007; Kalli et al., 2011; Ge et al., 2015), split-merge MCMC sampling (Jain and Neal, 2004), sequential updating and greedy search (SUGS) algorithms (Wang and Dunson, 2011) and variational approximation (Blei et al., 2006).

3 Proposed model

In this section we propose a new classification approach for handling incomplete data based on the Dirichlet process mixture model. We call the proposed method as Dirichlet Process-Naive Bayes model (DPNB).

3.1 Generative model with DPMM

Generative models employ the Bayes theorem to build a classifier. Let the input or feature vector be X and the class labels be YY. We need the density of X conditioned on the class kk, P(X=x|Y=k)P(\textbf{X}=\textbf{x}|Y=k) and the prior probability, P(Y=k)P(Y=k) to compute the posterior probability:

P(Y=k|X=x)=P(Y=k)P(X=x|Y=k)l=1KP(Y=l)P(X=x|Y=l).P(Y=k|\textbf{X}=\textbf{x})=\dfrac{P(Y=k)\cdot P(\textbf{X}=\textbf{x}|Y=k)}{\sum_{l=1}^{K}P(Y=l)\cdot P(\textbf{X}=\textbf{x}|Y=l)}. (3)

Then, input x is assigned to the class having the highest posterior probability. Typical examples of generative models include linear discriminant analysis (LDA), quadratic discriminant, analysis (QDA), and Gaussian naive Bayes (GNB). The three models assume multivariate normal densities for P(X=x|Y=k)P(\textbf{X}=\textbf{x}|Y=k). Another example is to assume that the class-conditional density of X is a finite mixture of normals. This method is called mixture discriminant analysis (MDA) (Hastie and Tibshirani, 1996; Fraley and Raftery, 2002). Instead of using a finite mixture of normals, we propose to use the Dirichlet process mixture model, as the number of clusters is automatically determined and not limited.

We consider a binary classification problem and assume class labels Y{0,1}Y\in\{0,1\} has a binomial distribution with parameters nn and pp. If we have KK (>2)(>2) classes, Y{1,,K}Y\in\{1,\ldots,K\} is assumed to have a multinomial distribution with nn and p=(p1,,pK)\textbf{p}=(p_{1},\ldots,p_{K}). The distribution of X conditioned on the class kk is modeled by the following hierarchical formulation for DPMM:

xi|yi=k\displaystyle\textbf{x}_{i}|y_{i}=k ind𝒩(xik;𝝁ik,𝚺ik),\displaystyle\stackrel{{\scriptstyle ind}}{{\sim}}\mathcal{N}(\textbf{x}_{i}^{k};\boldsymbol{\mu}_{i}^{k},\boldsymbol{\Sigma}_{i}^{k}), (4)
(𝝁ik,𝚺ik)\displaystyle(\boldsymbol{\mu}_{i}^{k},\boldsymbol{\Sigma}_{i}^{k}) iidGk,i=1,2,,nk,\displaystyle\stackrel{{\scriptstyle iid}}{{\sim}}G_{k},\quad i=1,2,\ldots,n_{k},
Gk\displaystyle G_{k} DP(α,Mk),k=0,1,\displaystyle\sim DP(\alpha,M_{k}),\quad k=0,1,

where xik\textbf{x}_{i}^{k} is the iith feature vector and 𝝁k\boldsymbol{\mu}^{k} and 𝚺k\boldsymbol{\Sigma}^{k} are parameters of a normal distribution which belongs to only class kk. Here, the base measure for class kk, MkM_{k} is the conjugate multivariate normal–inverse Wishart distribution, i.e. Mk:=𝒩(𝝁k;𝐦0k,𝚺k/τ0k)𝒲(𝚺k;𝐁0k,ν0k)M_{k}:=\mathcal{N}(\boldsymbol{\mu}^{k};\mathbf{m}_{0}^{k},\boldsymbol{\Sigma}^{k}/\tau_{0}^{k})\cdot\mathcal{I}\mathcal{W}(\boldsymbol{\Sigma}^{k};\mathbf{B}_{0}^{k},\nu_{0}^{k}). Then the class-conditional density of X is given by

P(X=x|Y=k)=h=1whk𝒩(x;𝝁hk,𝚺hk),k{0,1}P(\textbf{X}=\textbf{x}|Y=k)=\sum_{h=1}^{\infty}w_{h}^{k}\mathcal{N}(\textbf{x};\boldsymbol{\mu}_{h}^{k},\boldsymbol{\Sigma}_{h}^{k}),\quad k\in\{0,1\} (5)

using posterior samples for mixing proportion and cluster specific parameters. We use the improved slice sampler suggested by Ge et al. (2015) to generate samples from the posterior distribution. We describe the detailed MCMC algorithm for the DPMM on incomplete data in Appendix A. Since YY has a binomial distribution, marginal probabilities over classes are estimated by

P(Y=k)=nkn,k{0,1},P(Y=k)=\frac{n_{k}}{n},\quad k\in\{0,1\}, (6)

where nkn_{k} denotes the number of observations belonging to the class kk. Putting (5) and (6) together in the equation (3), we can produce the following posterior probabilities for classes:

P(Y=k|X=x)=nk×h=1wh𝒩(x;𝝁hk,𝚺hk)l{0,1}(nl×h=1wh𝒩(x;𝝁hl,𝚺hl)).P(Y=k|\textbf{X}=\textbf{x})=\frac{n_{k}\times\sum_{h=1}^{\infty}w_{h}\mathcal{N}(\textbf{x};\boldsymbol{\mu}_{h}^{k},\boldsymbol{\Sigma}_{h}^{k})}{\sum_{l\in\{0,1\}}\left(n_{l}\times\sum_{h=1}^{\infty}w_{h}\mathcal{N}(\textbf{x};\boldsymbol{\mu}_{h}^{l},\boldsymbol{\Sigma}_{h}^{l})\right)}.

We compute posterior probabilities for all kk and choose only one class associated with the highest probability.

Both LDA and QDA assume multivariate normal densities for P(X=x|Y=k)P(\textbf{X}=\textbf{x}|Y=k) and GNB assumes that the features, x1,x2,,xp\textbf{x}_{1},\textbf{x}_{2},\ldots,\textbf{x}_{p} are conditional independent on YY and P(xj|Y=k)P(\textbf{x}_{j}|Y=k) is univariate normal density:

P(X=x|Y=k)=j=1pP(xj|Y=k).P(\textbf{X}=\textbf{x}|Y=k)=\prod_{j=1}^{p}P(\textbf{x}_{j}|Y=k).

Since three generative models are designed to different covariance structures of multivariate normal distributions, they have their respective classifiers. The decision boundaries for LDA are linear, while those for QDA are quadratic. GNB has not necessarily linear classification according to the data. However, MDA allows close approximation of not only linear but also nonlinear decision boundaries since mixture models can approximate arbitrary continuous distributions (Fraley and Raftery, 2002; Wang et al., 2010). Figure 1 shows that DPNB more accurately approximates most decision boundaries than do LDA, QDA, and GNB. In other words, DPNB is a much more flexible classifier than them. Furthermore, the classifier of DPNB provides comparable results with that of Support Vector Machine (SVM) and Random Forest (RF).

Refer to caption
(a) Linearly separable dataset
Refer to caption
(b) Circle dataset
Refer to caption
(c) Two moons dataset
Refer to caption
(d) XOR dataset
Refer to caption
(e) Two spirals dataset
Figure 1: Comparison of decision boundaries of the DPNB model using the approximate prediction rule and five classifiers.

3.2 Imputation and prediction strategy

The DPNB can make inferences simultaneously on both unknown quantities and missing values. The same imputation approach described in the previous subsection 2.1 is capable of being applied in the mixture models with infinite components. For estimating (5), DPNB should divide the data into subsets where each subset belongs to only one class according to the inference process. Then it substitutes separately missing values with plausible values generated from samplers (2) based on respective subsets. We expect that imputation based on subsets is more accurate than that based on the full data. Because it is easy to find easily probable values estimated from homogeneous input data which belongs to only one class. If we spotlight imputation problems in the classification data, complete subsets are put together into a complete full dataset. In practice, the empirical results on 4 UCI datasets provide that this split and merge imputer has better performance than state-of-the-art imputation algorithms as the missing rate increases.

In reality, missing data may occur both training and test dataset. DPNB can make predictions for classes even if all elements of some features in the test set are absent. Let a new observed input vector denote xo\textbf{x}_{\star}^{o_{\star}} and a new predicted class label YY^{\star}. The prediction rule for DPNB is given by

P(Y=k|xo)=P(Y=k)P(xo|Y=k)l{0,1}P(Y=l)P(xo|Y=l),k{0,1}.P(Y^{\star}=k|\textbf{x}_{\star}^{o_{\star}})=\dfrac{P(Y^{\star}=k)\cdot P(\textbf{x}_{\star}^{o_{\star}}|Y^{\star}=k)}{\sum_{l\in\{0,1\}}P(Y^{\star}=l)\cdot P(\textbf{x}_{\star}^{o_{\star}}|Y^{\star}=l)},\quad k\in\{0,1\}. (7)

We need to compute the predictive density of xo\textbf{x}_{\star}^{o_{\star}} conditioned on the class kk to complete the equation (7), the posterior probability of the new class label. It can be approximated by using the Monte Carlo Integration from the posterior samples in the training process as shown in (8). Therefore, DPNB can build classifiers regardless of missingness and missing rate on both training set and test set. Experiments also support that it predicts classes more accurately than do competitive models.

P(xo|Y=k)\displaystyle P(\textbf{x}_{\star}^{o_{\star}}|Y^{\star}=k) =P(xo,xm|Y=k)𝑑xm\displaystyle=\int P(\textbf{x}_{\star}^{o_{\star}},\textbf{x}_{\star}^{m_{\star}}\,|Y^{\star}=k)\,\,d\textbf{x}_{\star}^{m_{\star}} (8)
=P(xo,xm|𝝁,𝚺,Y=k)π(𝝁,𝚺|Y=k)𝑑𝝁𝑑𝚺𝑑xm\displaystyle=\int P(\textbf{x}_{\star}^{o_{\star}},\textbf{x}_{\star}^{m_{\star}}\,|\,\boldsymbol{\mu},\boldsymbol{\Sigma},Y^{\star}=k)\pi(\boldsymbol{\mu},\boldsymbol{\Sigma}\,|Y^{\star}=k)\,d\boldsymbol{\mu}\,d\boldsymbol{\Sigma}d\textbf{x}_{\star}^{m_{\star}}
=hπhkP(xo,xm|𝝁h,𝚺h,Y=k)π(𝝁h,𝚺h|Y=k)d𝝁hd𝚺hdxm\displaystyle=\int\sum_{h}\pi_{h}^{k}\,P(\textbf{x}_{\star}^{o_{\star}},\textbf{x}_{\star}^{m_{\star}}\,|\boldsymbol{\mu}_{h},\boldsymbol{\Sigma}_{h},Y^{\star}=k)\pi(\boldsymbol{\mu}_{h},\boldsymbol{\Sigma}_{h}\,|Y^{\star}=k)\,d\boldsymbol{\mu}_{h}\,d\boldsymbol{\Sigma}_{h}\,d\textbf{x}_{\star}^{m_{\star}}
1Mj=1MhjπhjkP(xo,xm|𝝁hj,𝚺hj,Y=k)dxm(MC integration)\displaystyle\approx\frac{1}{M}\sum_{j=1}^{M}\sum_{h_{j}}\pi_{h_{j}}^{k}\int P(\textbf{x}_{\star}^{o_{\star}},\textbf{x}_{\star}^{m_{\star}}\,|\boldsymbol{\mu}_{h_{j}},\boldsymbol{\Sigma}_{h_{j}},Y^{\star}=k)\,d\textbf{x}_{\star}^{m_{\star}}\,\,(\because\text{MC integration})
=1Mj=1MhjπhjkP(xo|𝝁hjo,𝚺hjo,o,Y=k)\displaystyle=\frac{1}{M}\sum_{j=1}^{M}\sum_{h_{j}}\pi_{h_{j}}^{k}P(\textbf{x}_{\star}^{o_{\star}}\,|\boldsymbol{\mu}_{h_{j}}^{o_{\star}},\boldsymbol{\Sigma}_{h_{j}}^{o_{\star},o_{\star}},Y^{\star}=k)
=1Mj=1Mhjπhjk𝒩|o|(xo|[𝝁hjk]o,[𝚺hjk]o,o),\displaystyle=\frac{1}{M}\sum_{j=1}^{M}\sum_{h_{j}}\pi_{h_{j}}^{k}\cdot\mathcal{N}_{|o_{\star}|}\left(\textbf{x}_{\star}^{o_{\star}}\,|\,[\boldsymbol{\mu}_{h_{j}}^{k}]^{o_{\star}},[\boldsymbol{\Sigma}_{h_{j}}^{k}]^{o_{\star},o_{\star}}\right),

where MM is the number of posterior samples generated by MCMC.

4 Experiments

In this section, we evaluate both the imputation and prediction performance of the DPNB model and competitors using multiple datasets. First, we assess the imputation accuracy of our proposed methods with state-of-the-art imputation techniques in various settings. Second, we quantitatively compare the prediction accuracy of the DPNB model and benchmark classification algorithms based on multiple incomplete datasets in different conditions. In all experiments we apply various missing rates for the covariates (from 10% to 60%) and two missingness scenarios: missing completely at random (MCAR), missing at random (MAR).

We conduct experiments on four real-life datasets from UCI Machine Learning Repository Dua and Graff (2017): Ecoli, Wine, Breast Cancer Wisconsin (Diagnostic), and Wine Quality datasets. Specifically, the Ecoli dataset contains 336 observations with 8 features and multiple classes. We transform multi-class labels into binary labels which have positive (type im) and negative (the rest). Two discrete variables, Lip and Chg are also removed. The red wine quality dataset contains 1599 observations on 11 attributes of wine and 6 wine quality classes. We divide them into 3 groups for quality: Excellent (7)(\geq 7), Good (6)(6), and Poor (5)(\leq 5). The rest of the multiple datasets are originally used in this experiment. Those datasets have only continuous input variables. The summary of UCI datasets is given in Table 1. Here, the imbalanced ratio (IR) is defined by

IR=maxC𝒜|C|minC𝒜|C|,\text{IR}=\dfrac{\max_{C\in\mathcal{A}}|C|}{\min_{C\in\mathcal{A}}|C|},

where 𝒜\mathcal{A} is the set of all classes. The higher the imbalance ratio is, the bigger disproportion exists between majority class and minority class.

Dataset # Samples # Features # Classes Class distribution IR
Ecoli 336 5 2 (259/77) 3.4
Breast Cancer Wisconsin (Diagnostic) 569 30 2 (357/212) 1.7
Wine 178 13 3 (59/71/48) 1.5
Wine Quality 1599 11 3 (744/638/217) 3.4
Table 1: Detailed information of UCI datasets.

4.1 Imputation performance

For each UCI dataset, we generate 100 different incomplete datasets removing 10% to 60% of the complete values based on the MCAR or MAR missing assumptions. We make use of the normalized root mean squared errors (NRMSE) as the imputation accuracy measure along with its standard deviation across the 100 replicated datasets. The NRMSE is defined as

NRMSE=mean((XtrueXimp)2)Var(Xtrue),\text{NRMSE}=\sqrt{\dfrac{\text{mean}((X^{\text{true}}-X^{\text{imp}})^{2})}{\text{Var}(X^{\text{true}})}},

where XtrueX^{\text{true}} is the original data and XimpX^{\text{imp}} the imputed data.

We compare our proposed methods with several popular imputation methods such as multivariate imputation by chained equations using the predictive mean matching; Buuren and Groothuis-Oudshoorn (2010) (denoted by MICE), random forest based imputation; Stekhoven and Bühlmann (2012) (denoted by RF), KNN based imputation with the optimal number of neighbors which minimizes cross-validation errors; Troyanskaya et al. (2001) (denoted by KNN), imputer utilizing deep denoising autoencoders; Gondara and Wang (2018) (denoted by MIDA), and imputation technique by adapting generative adversarial nets; Yoon et al. (2018) (denoted by GAIN).

Figure 2 shows that the imputer of the DPNB model performs well in most datasets with either the lowest or the second lowest average NRMSE values across 100 replicates irrespective of the missingness mechanism (see Table 9 of Appendix B for details). In particular, we find out from both Table 2 and Table 3 that the proposed model is more accurate than other imputation algorithms as missing rates increase. The DPNB model can fill missing values via different covariance structures constructed by all available observations. However, methods to form conditional distributions using RF and MICE are less accurate than the DPNB because they utilize partial observations and variables instead of full information. Deep learning-based imputation methods such as GAIN and MIDA have poor performance due to model complexity relative to the number of observed data points. Among all models, the worst-performing method is the KNN based imputation.

Refer to caption
Figure 2: Imputation error for DPNB and competitive models on four UCI datasets with different missing assumptions and missing rates. Curves represent the average of NRMSEs for 100 randomly generated missingness datasets from the four UCI datasets.

Method Missing Rate 10% 20% 30% 40% 50% 60% RF 0.09 (0.038) 0.099 (0.0258) 0.11 (0.019) 0.121 (0.0168) 0.136 (0.0156) 0.161 (0.0167) MICE 0.066 (0.029) 0.08 (0.0217) 0.093 (0.0163) 0.115 (0.0181) 0.138 (0.0194) 0.183 (0.0173) KNN 0.219 (0.0356) 0.284 (0.027) 0.347 (0.0215) 0.392 (0.0166) 0.441 (0.0188) 0.479 (0.0175) MIDA 0.156 (0.0346) 0.184 (0.0304) 0.189 (0.0245) 0.215 (0.0223) 0.239 (0.0271) 0.265 (0.0221) GAIN 0.113 (0.0228) 0.125 (0.0183) 0.13 (0.0137) 0.142 (0.0175) 0.156 (0.0165) 0.184 (0.0204) DPNB 0.085 (0.0522) 0.098 (0.0431) 0.107 (0.0362) 0.114 (0.029) 0.134 (0.0315) 0.142 (0.0211)

Table 2: Average normalized mean squared errors with estimated standard errors in parentheses from 100 replications for Breast Cancer Wisconsin (Diagnostic) dataset with varying percentage of missing under MCAR. The best method for each data set is given in bold.

Method Missing Rate 10% 20% 30% 40% 50% 60% RF 0.343 (0.0272) 0.395 (0.0244) 0.453 (0.0215) 0.511 (0.0195) 0.563 (0.0216) 0.614 (0.0226) MICE 0.487 (0.0285) 0.506 (0.0192) 0.532 (0.0156) 0.556 (0.0144) 0.579 (0.0128) 0.601 (0.0123) KNN 0.461 (0.0413) 0.536 (0.0285) 0.563 (0.0192) 0.638 (0.0188) 0.629 (0.0148) 0.707 (0.0221) MIDA 0.577 (0.0266) 0.583 (0.0225) 0.587 (0.0179) 0.599 (0.013) 0.605 (0.011) 0.613 (0.0106) GAIN 0.569 (0.0537) 0.566 (0.0443) 0.578 (0.044) 0.586 (0.0573) 0.597 (0.0562) 0.628 (0.0462) DPNB 0.381 (0.0277) 0.428 (0.0271) 0.463 (0.0207) 0.5 (0.0186) 0.526 (0.0182) 0.551 (0.0146)

Table 3: Average normalized mean squared errors with estimated standard errors in parentheses from 100 replications for Wine Quality dataset with varying percentage of missing under MAR. The best method for each data set is given in bold.

4.2 Predictive performance

We perform 10-fold cross-validation to obtain an estimate of the classification accuracy using simulated missing datasets generated from the previous subsection 4.1. The cross-validation process is repeated 10 times for fair comparisons. Since four UCI datasets have various class distributions and binary or multi-class classification problems, we use three types of classification performance metrics to measure the performance of classifiers:(a) accuracy rate, (b) area under the ROC curve (AUC), and (c) F1 score. The classification metrics corresponding to datasets are also shown in Table 4.

Ecoli Breast Cancer Wine Wine Quality
F1 score AUC Accuracy rate F1 score
Table 4: Classification performance measures of four UCI datasets.

We compare the DPNB model with other prediction models on the incomplete dataset. Other approaches predict test cases after the imputation stage. Some details of the procedures regarding post-imputation prediction are as follows. First, we divide the dataset with missing values into training and test sets. Imputation algorithms utilized in subsection 4.1 fill missing values of training sets. Second, we build the support vector machines (SVM) with radial basis function (RBF) kernels as a benchmark classifier on the imputed training datasets. Third, missing values in test sets are replaced with the mean of the features of the imputed training set. Finally, we make predictions on the imputed test set using the trained model. We call all competing methods types of imputation algorithms used in both training and test phases.

Figure 3 provides that the DPNB model comes up with the best performance except for the Breast Cancer Wisconsin (Diagnostic) dataset, where it is competitive. Both Table 5 and Table 6 also show that the DPNB has a better classification performance than the others on a dataset for higher missing rates. As missing rates increase, the prediction accuracy of the DPNB is not greatly reduced unlike other competitive approaches. This supports that the DPNB is less affected by both missing data mechanisms and proportions missing proportions than other methods. See Table 10 of Appendix B for further details about these results.

Refer to caption
Figure 3: Classification performance for DPNB and competitive models on four UCI datasets with different missing assumptions and missing rates. Curves represent the average of 10 times repeated cross-validation scores such as accuracy, AUC, and F1-score obtained from each method.

Method Missing Rate 10% 20% 30% 40% 50% 60% RF 0.995 (0.0011) 0.992 (0.0016) 0.989 (0.0032) 0.981 (0.0057) 0.972 (0.007) 0.952 (0.0077) MICE 0.995 (0.0011) 0.993 (0.002) 0.99 (0.0024) 0.984 (0.0059) 0.979 (0.0079) 0.967 (0.0041) KNN 0.995 (8e-04) 0.993 (0.0017) 0.99 (0.0019) 0.987 (0.0044) 0.985 (0.0029) 0.979 (0.0038) MIDA 0.994 (9e-04) 0.992 (0.0025) 0.989 (0.0029) 0.981 (0.0056) 0.975 (0.0074) 0.96 (0.0057) GAIN 0.995 (0.0013) 0.992 (0.0023) 0.989 (0.002) 0.983 (0.0059) 0.978 (0.0052) 0.965 (0.0059) DPNB 0.993 (0.0025) 0.992 (0.0029) 0.991 (0.002) 0.989 (0.0027) 0.987 (0.0037) 0.984 (0.004)

Table 5: Mean and standard deviation (in parentheses) for the average AUCs of 10-fold cross-validations on Breast Cancer Wisconsin (Diagnostic) dataset with varying missing rate under MCAR. The best method for each experiment is given in bold.

Method Missing Rate 10% 20% 30% 40% 50% 60% RF 0.994 (0.0017) 0.992 (0.0015) 0.989 (0.0028) 0.979 (0.005) 0.971 (0.0069) 0.953 (0.0083) MICE 0.994 (0.0012) 0.993 (0.001) 0.99 (0.0028) 0.984 (0.0042) 0.977 (0.007) 0.97 (0.0055) KNN 0.995 (0.0015) 0.992 (0.0013) 0.991 (0.0022) 0.986 (0.0047) 0.98 (0.0052) 0.981 (0.0057) MIDA 0.995 (0.0015) 0.991 (0.0019) 0.989 (0.0025) 0.982 (0.0037) 0.972 (0.0061) 0.96 (0.0127) GAIN 0.994 (0.002) 0.991 (0.0014) 0.99 (0.0031) 0.984 (0.0039) 0.978 (0.0062) 0.97 (0.0068) DPNB 0.991 (0.0024) 0.991 (0.0019) 0.99 (0.0029) 0.989 (0.0031) 0.988 (0.0032) 0.988 (0.003)

Table 6: Mean and standard deviation (in parentheses) for the average AUCs of 10-fold cross-validations on Breast Cancer Wisconsin (Diagnostic) dataset with varying missing rate under MAR. The best method for each experiment is given in bold.

5 Applications

The main goal of our study is to impute and predict the class of given data with high missing rate. We now apply the DP-Naive Bayes model to a semiconductor manufacturing dataset provided by Samsung Electronics’ DS division, which consists of manufacturing operation variables and the semiconductor quality variable related to defect rates. This dataset also includes many missing values and aims to reduce defective wafers made up of semiconductor materials. These challenges are expected to further illustrate the features of the DP-Naive Bayes model.

For our analysis, we used the data of 2218 wafer records and 60 manufacturing process variables, discarding features with missing rates larger than 97.5% and categorical values. 4(a) depicts the observed and missing values of the pre-processing data in colored tiles and white tiles, respectively. It is made up of mostly white tiles. Missing values in the dataset account for about 95% of all values and most manufacturing operation variables have over 90% missing rate, as can be seen in the 4(b). We can say that the data is MCAR since missing values seem to be pretty randomly scattered and occur at random in the manufacturing processes. Furthermore, the target metrology variable related to defect rates is highly imbalanced, with most records falling in the “No defect” class. The imbalance ratio of this dataset is roughly 9.4.

Refer to caption
(a)
Refer to caption
(b)
Figure 4: (a) Missing values (white) of a semiconductor manufacturing dataset, (b) percentage of missing values in each variable.

5.1 Imputation performance

To assess the imputation accuracy based on the real world dataset, we consider three scenarios by deleting artificially 50, 100, and 500 complete values in the semiconductor dataset under MCAR assumption. In every cases, we made 100 replicates, respectively. For performance comparisons, we then compute the average of normalized root mean squared errors obtained by each method in all cases.

The results of the DPNB and competing models are provided in Figure 5 and Table 7. They demonstrate that the DPNB model provides more accurate values and consistent performance through both minimum mean and the lowest standard deviation values over 100 replications of NRMSEs. As shown in the Breast Cancer example, both RF and MICE yield similar imputation performance with the DPNB model. Imputation methods based on deep architectures including MIDA and GAIN have difficulty selecting appropriate parameters to prevent the problem of overfitting. These methods have then lower imputation accuracy than the DPNB, RF and MICE.

Refer to caption
Figure 5: Boxplots of NRMSEs over 100 replications for semiconductor dataset with varying number of artificial missing values.
#50 #100 #500
RF 0.0226 (0.00452) 0.0226 (0.00296) 0.0233 (0.00163)
MICE 0.023 (0.00457) 0.0233 (0.00287) 0.0234 (0.0014)
KNN 0.0284 (0.00562) 0.0309 (0.00566) 0.0325 (0.00442)
MIDA 0.031 (0.00513) 0.0312 (0.00394) 0.0312 (0.00394)
GAIN 0.029 (0.00617) 0.0281 (0.00418) 0.0281 (0.00418)
DPNB 0.0219 (0.00423) 0.0222 (0.00282) 0.0224 (0.0013)
Table 7: Average of NRMSEs over 100 replications for semiconductor dataset with different numbers of artificial missing values under MCAR. Estimated standard errors of NRMSEs are shown in parentheses. The best method for each experiment is given in bold.

5.2 Predictive performance

In this experiment, we randomly partitioned the real dataset into training and test sets with different ratios. The training set was set from 50% to 90% of the overall dataset and the test set was accordingly set from 50% to 10% of it. For example, if 90% of the data is used as the training set, then the remaining 10% is used as the test set. Likewise, we created 100 replicate datasets in all cases. Furthermore, the bootstrapping-based oversampling technique that replicates observations from minority class was applied into training sets in order to balance the two classes. Note that the same indices of observations selected from the minority class were used in all methods given percentages of the training sets.

We add the SVM with a linear kernel as another base classifier. We will name competitive methods by combining imputation techniques and kernels of support vector machine, e.g., “RF+L” means that the RF imputes missing values of training sets and the SVM with a linear kernel is fitted on the imputed training sets and predicts the test set values. The results of this experiment are given in Table 8. The columns of competing methods represent their average F1-score over 100 replicate datasets, normalized by the average F1-score of the DPNB model. The values higher than one indicate that the methods provide better performance than the DPNB and values lower than one indicate worse performance in imbalanced settings.

RF+L RF+R MICE+L MICE+R KNN+L KNN+R MIDA+L MIDA+R GAIN+L GAIN+R DPNB 50% / 50% 0.9516 (3) 0.6638 (9) 0.8618 (6) 0.0349 (11) 0.8359 (7) 0.7016 (8) 0.6508 (10) 0.9813 (2) 0.8891 (5) 0.9461 (4) 1 (1) 60% / 40% 0.9829 (3) 0.7192 (8) 0.8833 (6) 0.0282 (11) 0.8559 (7) 0.7067 (9) 0.6434 (10) 0.9786 (4) 0.9262 (5) 1.0096 (1) 1 (2) 70% / 30% 0.9845 (5) 0.7248 (9) 0.9186 (6) 0.0427 (11) 0.8752 (7) 0.7423 (8) 0.7174 (10) 1.0419 (1) 1.0051 (2) 0.9851 (4) 1 (3) 80% / 20% 1.0184 (1) 0.6591 (10) 0.9391 (6) 0.0453 (11) 0.9019 (7) 0.7314 (9) 0.7814 (8) 0.9965 (4) 1.0124 (2) 0.971 (5) 1 (3) 90% / 10% 0.9982 (2) 0.6281 (10) 0.9621 (6) 0.0298 (11) 0.8186 (7) 0.7394 (8) 0.7329 (9) 0.9841 (3) 0.9662 (5) 0.9825 (4) 1 (1) Average nomalized F1-score 0.987 0.679 0.913 0.036 0.857 0.724 0.705 0.996 0.960 0.979 1.000 Average rank 2.8 9.2 6.0 11.0 7.0 8.4 9.4 2.8 3.8 3.6 2.0

Table 8: Normalized average F1-scores over 100 replications for the semiconductor dataset with different ratios between training and test sets. The rank of the methods judged by normalized F1-scores is shown in parentheses. The best method for each experiment is given in bold.

As shown in Table 8, the DPNB method performs well in practice since it ranks high in almost all cases. In particular, it yields the best prediction performance on two out of the five ratios. We can say that the DPNB model provides stable performance by obtaining the highest average normalized F1-score and average rank and stable inference in the incomplete data with high missing rates. However, since the semiconductor dataset has very high missing rates, the DPNB does not provide remarkable performance against other methods as shown in the previous section. The methods using the KNN imputer have poor performance irrespective of the type of classifiers. We also present the actual average F1-score in Table 11 of Appendix B for more details.

6 Conclusions

We propose a new method for classification problems with incomplete data based on the Dirichlet process and naive Bayes model, DPNB. The DPNB method is free from the distribution assumption and constructs flexible imputer and classifier. The flexibility and effect of the DPNB model are verified by various experiments. Moreover, the DPNB model suffers less from the overfitting problem, which frequently occurs using a flexible model, by considering proper prior. The DPNB model shows stable and better performance than other methods on experiments even missing rate is high. As an improvement to the DPNB model, we would like to address the problem with computational time. Due to the limit of the MCMC algorithm, the DPNB model takes a longer time compared to other models. In future studies, we would like to propose ways to reduce computational time for the DPNB model such as variational methods.

Acknowledgement

This work was supported by Samsung Electronics Co., Ltd. (IO210216-08417-01).

References

  • (1)
  • Aldous (1985) Aldous, D. J. (1985). Exchangeability and related topics, École d’Été de Probabilités de Saint-Flour XIII—1983, Springer, pp. 1–198.
  • Antoniak (1974) Antoniak, C. E. (1974). Mixtures of dirichlet processes with applications to bayesian nonparametric problems, The annals of statistics pp. 1152–1174.
  • Bengio and Gingras (1996) Bengio, Y. and Gingras, F. (1996). Recurrent neural networks for missing or asynchronous data, Advances in neural information processing systems, pp. 395–401.
  • Blackwell et al. (1973) Blackwell, D., MacQueen, J. B. et al. (1973). Ferguson distributions via pólya urn schemes, The annals of statistics 1(2): 353–355.
  • Blei et al. (2006) Blei, D. M., Jordan, M. I. et al. (2006). Variational inference for dirichlet process mixtures, Bayesian analysis 1(1): 121–143.
  • Brás and Menezes (2007) Brás, L. P. and Menezes, J. C. (2007). Improving cluster-based missing value estimation of dna microarray data, Biomolecular engineering 24(2): 273–282.
  • Burgette and Reiter (2010) Burgette, L. F. and Reiter, J. P. (2010). Multiple imputation for missing data via sequential regression trees, American journal of epidemiology 172(9): 1070–1076.
  • Bush and MacEachern (1996) Bush, C. A. and MacEachern, S. N. (1996). A semiparametric bayesian model for randomised block designs, Biometrika 83(2): 275–285.
  • Buuren and Groothuis-Oudshoorn (2010) Buuren, S. v. and Groothuis-Oudshoorn, K. (2010). mice: Multivariate imputation by chained equations in r, Journal of statistical software pp. 1–68.
  • Caruana (2001) Caruana, R. (2001). A non-parametric em-style algorithm for imputing missing values., AISTATS.
  • Che et al. (2018) Che, Z., Purushotham, S., Cho, K., Sontag, D. and Liu, Y. (2018). Recurrent neural networks for multivariate time series with missing values, Scientific reports 8(1): 1–12.
  • Das et al. (2018) Das, S., Datta, S. and Chaudhuri, B. B. (2018). Handling data irregularities in classification: Foundations, trends, and future challenges, Pattern Recognition 81: 674–693.
  • Dua and Graff (2017) Dua, D. and Graff, C. (2017). UCI machine learning repository.
    http://archive.ics.uci.edu/ml
  • Escobar (1994) Escobar, M. D. (1994). Estimating normal means with a dirichlet process prior, Journal of the American Statistical Association 89(425): 268–277.
  • Escobar and West (1995) Escobar, M. D. and West, M. (1995). Bayesian density estimation and inference using mixtures, Journal of the american statistical association 90(430): 577–588.
  • Ferguson (1973) Ferguson, T. S. (1973). A bayesian analysis of some nonparametric problems, The annals of statistics pp. 209–230.
  • Fraley and Raftery (2002) Fraley, C. and Raftery, A. E. (2002). Model-based clustering, discriminant analysis, and density estimation, Journal of the American statistical Association 97(458): 611–631.
  • Franzén (2006) Franzén, J. (2006). Bayesian inference for a mixture model using the gibbs sampler, MResearch Report 1.
  • Ge et al. (2015) Ge, H., Chen, Y., Wan, M. and Ghahramani, Z. (2015). Distributed inference for dirichlet process mixture models, International Conference on Machine Learning, pp. 2276–2284.
  • Ghahramani and Jordan (1994) Ghahramani, Z. and Jordan, M. I. (1994). Supervised learning from incomplete data via an em approach, Advances in neural information processing systems, pp. 120–127.
  • Gondara and Wang (2018) Gondara, L. and Wang, K. (2018). Mida: Multiple imputation using denoising autoencoders, Pacific-Asia Conference on Knowledge Discovery and Data Mining, Springer, pp. 260–272.
  • Gupta and Lam (1996) Gupta, A. and Lam, M. S. (1996). Estimating missing values using neural networks, Journal of the Operational Research Society 47(2): 229–238.
  • Hastie and Tibshirani (1996) Hastie, T. and Tibshirani, R. (1996). Discriminant analysis by gaussian mixtures, Journal of the Royal Statistical Society: Series B (Methodological) 58(1): 155–176.
  • Ishwaran and James (2001) Ishwaran, H. and James, L. F. (2001). Gibbs sampling methods for stick-breaking priors, Journal of the American Statistical Association 96(453): 161–173.
  • Jain and Neal (2004) Jain, S. and Neal, R. M. (2004). A split-merge markov chain monte carlo procedure for the dirichlet process mixture model, Journal of computational and Graphical Statistics 13(1): 158–182.
  • Kalli et al. (2011) Kalli, M., Griffin, J. E. and Walker, S. G. (2011). Slice sampling mixture models, Statistics and computing 21(1): 93–105.
  • Kim and Chi (2018) Kim, Y.-J. and Chi, M. (2018). Temporal belief memory: Imputing missing data during rnn training., In Proceedings of the 27th International Joint Conference on Artificial Intelligence (IJCAI-2018).
  • Lee and Styczynski (2018) Lee, J. Y. and Styczynski, M. P. (2018). Ns-knn: A modified k-nearest neighbors approach for imputing metabolomics data, Metabolomics 14(12): 153.
  • Lin et al. (2006) Lin, T. I., Lee, J. C. and Ho, H. J. (2006). On fast supervised learning for normal mixture models with missing information, Pattern Recognition 39(6): 1177–1187.
  • Little and Rubin (2019) Little, R. J. and Rubin, D. B. (2019). Statistical analysis with missing data, Vol. 793, John Wiley & Sons.
  • MacEachern (1994) MacEachern, S. N. (1994). Estimating normal means with a conjugate style dirichlet process prior, Communications in Statistics-Simulation and Computation 23(3): 727–741.
  • Neal (2000) Neal, R. M. (2000). Markov chain sampling methods for dirichlet process mixture models, Journal of computational and graphical statistics 9(2): 249–265.
  • Playdon et al. (2019) Playdon, M. C., Joshi, A. D., Tabung, F. K., Cheng, S., Henglin, M., Kim, A., Lin, T., van Roekel, E. H., Huang, J., Krumsiek, J. et al. (2019). Metabolomics analytics workflow for epidemiological research: Perspectives from the consortium of metabolomics studies (comets), Metabolites 9(7): 145.
  • Rubin (1976) Rubin, D. B. (1976). Inference and missing data, Biometrika 63(3): 581–592.
  • Schafer (1997) Schafer, J. L. (1997). Analysis of incomplete multivariate data, CRC press.
  • Sethuraman (1994) Sethuraman, J. (1994). A constructive definition of dirichlet priors, Statistica sinica pp. 639–650.
  • Sharpe and Solly (1995) Sharpe, P. K. and Solly, R. (1995). Dealing with missing values in neural network-based diagnostic systems, Neural Computing & Applications 3(2): 73–77.
  • Silva-Ramírez et al. (2011) Silva-Ramírez, E.-L., Pino-Mejías, R., López-Coello, M. and Cubiles-de-la Vega, M.-D. (2011). Missing value imputation on missing completely at random data using multilayer perceptrons, Neural Networks 24(1): 121–129.
  • Stekhoven and Bühlmann (2012) Stekhoven, D. J. and Bühlmann, P. (2012). Missforest—non-parametric missing value imputation for mixed-type data, Bioinformatics 28(1): 112–118.
  • Teh et al. (2006) Teh, Y. W., Jordan, M. I., Beal, M. J. and Blei, D. M. (2006). Hierarchical dirichlet processes, Journal of the American Statistical Association 101(476): 1566–1581.
  • Troyanskaya et al. (2001) Troyanskaya, O., Cantor, M., Sherlock, G., Brown, P., Hastie, T., Tibshirani, R., Botstein, D. and Altman, R. B. (2001). Missing value estimation methods for dna microarrays, Bioinformatics 17(6): 520–525.
  • Vincent et al. (2008) Vincent, P., Larochelle, H., Bengio, Y. and Manzagol, P.-A. (2008). Extracting and composing robust features with denoising autoencoders, Proceedings of the 25th international conference on Machine learning, pp. 1096–1103.
  • Walker (2007) Walker, S. G. (2007). Sampling the dirichlet mixture model with slices, Communications in Statistics—Simulation and Computation® 36(1): 45–54.
  • Wang et al. (2010) Wang, C., Liao, X., Carin, L., Dunson, D. B. and Blei, D. (2010). Classification with incomplete data using dirichlet process priors., Journal of Machine Learning Research 11(12).
  • Wang and Dunson (2011) Wang, L. and Dunson, D. B. (2011). Fast bayesian inference in dirichlet process mixture models, Journal of Computational and Graphical Statistics 20(1): 196–216.
  • Wei et al. (2018) Wei, R., Wang, J., Su, M., Jia, E., Chen, S., Chen, T. and Ni, Y. (2018). Missing value imputation approach for mass spectrometry-based metabolomics data, Scientific reports 8(1): 1–10.
  • Williams et al. (2007) Williams, D., Liao, X., Xue, Y., Carin, L. and Krishnapuram, B. (2007). On classification with incomplete data, IEEE transactions on pattern analysis and machine intelligence 29(3): 427–436.
  • Yeh et al. (2017) Yeh, R. A., Chen, C., Yian Lim, T., Schwing, A. G., Hasegawa-Johnson, M. and Do, M. N. (2017). Semantic image inpainting with deep generative models, Proceedings of the IEEE conference on computer vision and pattern recognition, pp. 5485–5493.
  • Yoon et al. (2018) Yoon, J., Jordon, J. and Van Der Schaar, M. (2018). Gain: Missing data imputation using generative adversarial nets, arXiv preprint arXiv:1806.02920 .
  • Zhang and Everson (2004) Zhang, J. and Everson, R. (2004). Bayesian estimation and classification with incomplete data using mixture models, 2004 International Conference on Machine Learning and Applications, 2004. Proceedings., IEEE, pp. 296–303.

Appendix A Slice sampling for the DPMM on incomplete data

Based on the model (4), approximate posterior of X conditioned on the class kk is computed using an improved slice sampler for the DPMM presented by Ge et al. (2015). However, since observations xi:={xioi,ximi},i=1,,n\textbf{x}_{i}:=\{\textbf{x}_{i}^{o_{i}},\textbf{x}_{i}^{m_{i}}\},\,i=1,\ldots,n contain missing values, we additionally embed updating steps (2) for estimating the missing values in the slice sampling algorithm for the DPMM. We can summarize the MCMC algorithm for X conditioned on the class kk as follows:


[Step 1] For each cluster hh, sample mixing proportions wk:=(w1k,w2k,,wHk,wk)\textbf{w}^{k}:=(w_{1}^{k},w^{k}_{2},\ldots,w^{k}_{H},w_{*}^{k})

[wk|others]Dir(π1k,,πHk,α),[\,\textbf{w}^{k}\,|\,\text{others}\,]\sim\text{Dir}(\pi_{1}^{k},\ldots,\pi_{H}^{k},\alpha),

where πhk:=|{i:zik=h}|\pi_{h}^{k}:=|\{i:z_{i}^{k}=h\}| and HH is the number of (current) clusters.


[Step 2] Sample auxilrary variables and set the minimum

[uik|others]𝒰(0,wzikk),i=1,,nk,\displaystyle[\,u_{i}^{k}\,|\,\text{others}\,]\sim\mathcal{U}(0,w^{k}_{z^{k}_{i}}),\quad\forall i=1,\ldots,n_{k},
uk=miniuik.\displaystyle u_{*}^{k}=\min_{i}u_{i}^{k}.

[Step 3] Generate new clusters through stick-breaking processes until wk<ukw_{*}^{k}<u_{*}^{k}

HkHk+1,νHkkBeta(1,α),\displaystyle H^{k}\leftarrow H^{k}+1,\quad\nu^{k}_{H^{k}}\sim\text{Beta}(1,\alpha),
wHkk=wk×νHkk,wkwk×(1νHkk),\displaystyle w^{k}_{H^{k}}=w^{k}_{*}\times\nu^{k}_{H^{k}},\quad w_{*}^{k}\leftarrow w_{*}^{k}\times(1-\nu^{k}_{H^{k}}),
(𝝁Hkk,𝚺Hkk)𝒩(𝝁k;𝐦0k,𝚺k/τ0k)𝒲(𝚺k;𝐁0k,ν0k),\displaystyle(\boldsymbol{\mu}^{k}_{H^{k}},\boldsymbol{\Sigma}^{k}_{H^{k}})\sim\mathcal{N}(\boldsymbol{\mu}^{k};\mathbf{m}_{0}^{k},\boldsymbol{\Sigma}^{k}/\tau_{0}^{k})\cdot\mathcal{I}\mathcal{W}(\boldsymbol{\Sigma}^{k};\mathbf{B}_{0}^{k},\nu_{0}^{k}),

where wkw_{*}^{k} is the remaining stick length.


[Step 4] For each observation 𝐱ik\mathbf{x}^{k}_{i}, sample the assignment variable zikz^{k}_{i}

p(zi=h|others)I(whkuik)𝒩(xik;𝝁ik,𝚺ik).\displaystyle p(z_{i}=h\,|\,\text{others})\propto I(w_{h}^{k}\geq u^{k}_{i})\cdot\mathcal{N}(\textbf{x}_{i}^{k};\boldsymbol{\mu}_{i}^{k},\boldsymbol{\Sigma}_{i}^{k}).

for h=1,,Hkh=1,\ldots,H^{k}.


[Step 5] For each cluster hh, sample cluster parameters (𝝁hk,𝚺hk)(\boldsymbol{\mu}^{k}_{h},\boldsymbol{\Sigma}^{k}_{h})

p(𝝁hk,𝚺hk|others)𝒩(𝝁hk;𝐦0k,𝚺k/τ0k)𝒲(𝚺hk;𝐁0k,ν0k){i:zik=h}𝒩(xik;𝝁ik,𝚺ik).\displaystyle p(\boldsymbol{\mu}^{k}_{h},\boldsymbol{\Sigma}^{k}_{h}\,|\,\text{others})\propto\mathcal{N}(\boldsymbol{\mu}^{k}_{h};\mathbf{m}_{0}^{k},\boldsymbol{\Sigma}^{k}/\tau_{0}^{k})\cdot\mathcal{I}\mathcal{W}(\boldsymbol{\Sigma}^{k}_{h};\mathbf{B}_{0}^{k},\nu_{0}^{k})\prod_{\{i:\,z_{i}^{k}=h\}}\mathcal{N}(\textbf{x}_{i}^{k};\boldsymbol{\mu}_{i}^{k},\boldsymbol{\Sigma}_{i}^{k}).

[Step 6] For each observation 𝐱ik\mathbf{x}_{i}^{k} that belongs to specific hhth mixure component, sample the missing values 𝐱imi\mathbf{x}_{i}^{m_{i}}

[ximi|others]𝒩|mi|((𝝁hk)mi|oi,(𝚺hk)mi|oi),\displaystyle[\,\textbf{x}_{i}^{m_{i}}\,|\,\text{others}\,]\sim\mathcal{N}_{|m_{i}|}((\boldsymbol{\mu}_{h}^{k})^{m_{i}|o_{i}},(\boldsymbol{\Sigma}_{h}^{k})^{m_{i}|o_{i}}),
(𝝁hk)mi|oi=(𝝁hk)mi+(𝚺hk)mi,oi((𝚺hk)oi,oi)1(xioi(𝝁hk)oi),\displaystyle(\boldsymbol{\mu}_{h}^{k})^{m_{i}|o_{i}}=(\boldsymbol{\mu}_{h}^{k})^{m_{i}}+(\boldsymbol{\Sigma}_{h}^{k})^{m_{i},o_{i}}((\boldsymbol{\Sigma}_{h}^{k})^{o_{i},o_{i}})^{-1}(\textbf{x}_{i}^{o_{i}}-(\boldsymbol{\mu}_{h}^{k})^{o_{i}}),
(𝚺hk)mi|oi=(𝚺hk)mi,mi(𝚺hk)mi,oi((𝚺hk)oi,oi)1(𝚺hk)oi,mi.\displaystyle(\boldsymbol{\Sigma}_{h}^{k})^{m_{i}|o_{i}}=(\boldsymbol{\Sigma}_{h}^{k})^{m_{i},m_{i}}-(\boldsymbol{\Sigma}_{h}^{k})^{m_{i},o_{i}}((\boldsymbol{\Sigma}_{h}^{k})^{o_{i},o_{i}})^{-1}(\boldsymbol{\Sigma}_{h}^{k})^{o_{i},m_{i}}.

[Step 7] Repeat steps 1-6 until convergence.

Appendix B Full simulation results

In this section we present additional simulation results for imputation and predictive performance reported in section 4 and section 5.

Dataset Model MCAR MAR 10% 20% 30% 40% 50% 60% 10% 20% 30% 40% 50% 60% Ecoli RF 0.705 (0.0564) 0.746 (0.0398) 0.804 (0.0341) 0.864 (0.0391) 0.889 (0.0323) 0.984 (0.0367) 0.709 (0.055) 0.761 (0.0467) 0.818 (0.0385) 0.868 (0.0345) 0.924 (0.0463) 1.02 (0.0857) MICE 0.786 (0.0589) 0.814 (0.0346) 0.847 (0.0279) 0.885 (0.0229) 0.888 (0.022) 0.941 (0.0225) 0.791 (0.055) 0.822 (0.0371) 0.857 (0.0297) 0.887 (0.0244) 0.909 (0.0385) 0.962 (0.0598) KNN 0.737 (0.0619) 0.781 (0.0427) 0.841 (0.0386) 0.897 (0.0369) 0.9 (0.0339) 0.971 (0.0393) 0.744 (0.0553) 0.794 (0.0383) 0.849 (0.0343) 0.912 (0.0319) 0.924 (0.0486) 0.978 (0.0851) MIDA 0.898 (0.0416) 0.908 (0.0415) 0.927 (0.0295) 0.936 (0.0229) 0.949 (0.0243) 0.987 (0.0247) 0.891 (0.0429) 0.916 (0.0431) 0.926 (0.0287) 0.936 (0.0233) 0.946 (0.0269) 0.966 (0.0259) GAIN 0.824 (0.0675) 0.839 (0.052) 0.865 (0.0477) 0.898 (0.0489) 0.95 (0.0801) 1.058 (0.1221) 0.833 (0.068) 0.85 (0.0498) 0.866 (0.0402) 0.906 (0.0565) 0.945 (0.0784) 1.008 (0.0888) DPNB 0.616 (0.0521) 0.643 (0.0422) 0.682 (0.0383) 0.717 (0.0265) 0.743 (0.0287) 0.793 (0.0256) 0.629 (0.0481) 0.641 (0.0425) 0.682 (0.0327) 0.719 (0.0268) 0.75 (0.0278) 0.793 (0.0253) Breast Cancer RF 0.09 (0.038) 0.099 (0.0258) 0.11 (0.019) 0.121 (0.0168) 0.136 (0.0156) 0.161 (0.0167) 0.091 (0.03) 0.099 (0.024) 0.11 (0.0183) 0.117 (0.0174) 0.131 (0.016) 0.15 (0.0143) MICE 0.066 (0.029) 0.08 (0.0217) 0.093 (0.0163) 0.115 (0.0181) 0.138 (0.0194) 0.183 (0.0173) 0.065 (0.0236) 0.077 (0.0206) 0.091 (0.017) 0.106 (0.0162) 0.127 (0.0176) 0.173 (0.0158) KNN 0.219 (0.0356) 0.284 (0.027) 0.347 (0.0215) 0.392 (0.0166) 0.441 (0.0188) 0.479 (0.0175) 0.233 (0.0357) 0.279 (0.0261) 0.355 (0.0199) 0.389 (0.0206) 0.457 (0.0178) 0.479 (0.0169) MIDA 0.156 (0.0346) 0.184 (0.0304) 0.189 (0.0245) 0.215 (0.0223) 0.239 (0.0271) 0.265 (0.0221) 0.165 (0.0351) 0.18 (0.0305) 0.191 (0.0245) 0.213 (0.0243) 0.225 (0.0223) 0.259 (0.0225) GAIN 0.113 (0.0228) 0.125 (0.0183) 0.13 (0.0137) 0.142 (0.0175) 0.156 (0.0165) 0.184 (0.0204) 0.115 (0.0235) 0.122 (0.0164) 0.127 (0.0141) 0.134 (0.013) 0.149 (0.0159) 0.172 (0.0195) DPNB 0.085 (0.0522) 0.098 (0.0431) 0.107 (0.0362) 0.114 (0.029) 0.134 (0.0315) 0.142 (0.0211) 0.087 (0.0585) 0.101 (0.0464) 0.109 (0.0384) 0.115 (0.0298) 0.124 (0.0291) 0.144 (0.0208) Wine RF 0.232 (0.0361) 0.236 (0.0326) 0.25 (0.0293) 0.275 (0.027) 0.293 (0.0236) 0.315 (0.025) 0.225 (0.036) 0.241 (0.0315) 0.253 (0.0288) 0.271 (0.0259) 0.289 (0.0209) 0.317 (0.0229) MICE 0.273 (0.0521) 0.281 (0.0352) 0.297 (0.031) 0.311 (0.0275) 0.327 (0.023) 0.352 (0.018) 0.272 (0.0456) 0.283 (0.0348) 0.294 (0.0304) 0.31 (0.0291) 0.326 (0.0243) 0.353 (0.0234) KNN 0.249 (0.052) 0.259 (0.0486) 0.287 (0.04) 0.315 (0.0358) 0.338 (0.038) 0.361 (0.0409) 0.261 (0.0576) 0.268 (0.0446) 0.31 (0.0416) 0.305 (0.0338) 0.373 (0.0363) 0.372 (0.0368) MIDA 0.319 (0.0604) 0.324 (0.039) 0.336 (0.0321) 0.349 (0.0257) 0.359 (0.0228) 0.37 (0.0208) 0.309 (0.0511) 0.325 (0.0345) 0.334 (0.0287) 0.345 (0.0256) 0.357 (0.0253) 0.372 (0.024) GAIN 0.279 (0.0382) 0.288 (0.0317) 0.312 (0.0293) 0.333 (0.0354) 0.352 (0.032) 0.375 (0.0312) 0.274 (0.0429) 0.295 (0.0332) 0.312 (0.0293) 0.326 (0.0315) 0.349 (0.0286) 0.376 (0.0298) DPNB 0.222 (0.0391) 0.227 (0.0305) 0.229 (0.0237) 0.234 (0.0166) 0.241 (0.0185) 0.242 (0.0169) 0.222 (0.0413) 0.228 (0.0284) 0.23 (0.0217) 0.232 (0.0216) 0.24 (0.0178) 0.242 (0.0187) Wine Quality RF 0.346 (0.0263) 0.404 (0.0258) 0.463 (0.0225) 0.52 (0.0243) 0.566 (0.0216) 0.62 (0.0224) 0.343 (0.0272) 0.395 (0.0244) 0.453 (0.0215) 0.511 (0.0195) 0.563 (0.0216) 0.614 (0.0226) MICE 0.486 (0.031) 0.511 (0.0193) 0.535 (0.018) 0.56 (0.0136) 0.584 (0.0148) 0.607 (0.0097) 0.487 (0.0285) 0.506 (0.0192) 0.532 (0.0156) 0.556 (0.0144) 0.579 (0.0128) 0.601 (0.0123) KNN 0.473 (0.0377) 0.526 (0.0291) 0.598 (0.0207) 0.592 (0.0167) 0.67 (0.0195) 0.642 (0.0134) 0.461 (0.0413) 0.536 (0.0285) 0.563 (0.0192) 0.638 (0.0188) 0.629 (0.0148) 0.707 (0.0221) MIDA 0.577 (0.027) 0.587 (0.0209) 0.594 (0.0166) 0.6 (0.0143) 0.611 (0.012) 0.617 (0.0102) 0.577 (0.0266) 0.583 (0.0225) 0.587 (0.0179) 0.599 (0.013) 0.605 (0.011) 0.613 (0.0106) GAIN 0.564 (0.0562) 0.567 (0.0457) 0.576 (0.0438) 0.587 (0.0423) 0.6 (0.0443) 0.624 (0.0477) 0.569 (0.0537) 0.566 (0.0443) 0.578 (0.044) 0.586 (0.0573) 0.597 (0.0562) 0.628 (0.0462) DPNB 0.383 (0.029) 0.429 (0.0311) 0.471 (0.0267) 0.507 (0.022) 0.532 (0.0189) 0.553 (0.0149) 0.381 (0.0277) 0.428 (0.0271) 0.463 (0.0207) 0.5 (0.0186) 0.526 (0.0182) 0.551 (0.0146)

Table 9: Average of NRMSEs over 100 replications for four UCI datasets with different missingness patterns and missing rates. Estimated standard errors of NRMSEs are shown in parentheses.

Dataset Metric Model MCAR MAR 10% 20% 30% 40% 50% 60% 10% 20% 30% 40% 50% 60% Ecoli F1 score RF 0.737 (0.027) 0.695 (0.0304) 0.63 (0.0378) 0.575 (0.0777) 0.537 (0.0516) 0.391 (0.065) 0.743 (0.0286) 0.666 (0.0371) 0.642 (0.0418) 0.529 (0.0747) 0.483 (0.0656) 0.29 (0.0859) MICE 0.728 (0.0268) 0.688 (0.0308) 0.619 (0.0337) 0.556 (0.0577) 0.523 (0.0565) 0.348 (0.0627) 0.737 (0.029) 0.669 (0.0296) 0.63 (0.0548) 0.543 (0.0605) 0.446 (0.0758) 0.294 (0.0826) KNN 0.732 (0.0334) 0.682 (0.0288) 0.621 (0.0262) 0.547 (0.0587) 0.51 (0.0399) 0.356 (0.1007) 0.732 (0.0284) 0.666 (0.0434) 0.626 (0.0322) 0.523 (0.0628) 0.445 (0.095) 0.31 (0.0815) MIDA 0.733 (0.0309) 0.719 (0.0498) 0.672 (0.0291) 0.63 (0.0475) 0.606 (0.0408) 0.499 (0.0483) 0.754 (0.0237) 0.696 (0.0301) 0.665 (0.0387) 0.626 (0.0377) 0.542 (0.0463) 0.429 (0.067) GAIN 0.747 (0.0248) 0.702 (0.0365) 0.664 (0.016) 0.611 (0.0571) 0.581 (0.0361) 0.459 (0.072) 0.745 (0.0237) 0.68 (0.0376) 0.653 (0.0379) 0.609 (0.0609) 0.533 (0.0725) 0.381 (0.0545) DPNB 0.758 (0.0182) 0.744 (0.0302) 0.699 (0.0225) 0.678 (0.034) 0.643 (0.0429) 0.589 (0.0302) 0.767 (0.0254) 0.745 (0.0202) 0.707 (0.0242) 0.681 (0.0457) 0.63 (0.0384) 0.551 (0.0715) Breast Cancer AUC RF 0.995 (0.0011) 0.992 (0.0016) 0.989 (0.0032) 0.981 (0.0057) 0.972 (0.007) 0.952 (0.0077) 0.994 (0.0017) 0.992 (0.0015) 0.989 (0.0028) 0.979 (0.005) 0.971 (0.0069) 0.953 (0.0083) MICE 0.995 (0.0011) 0.993 (0.002) 0.99 (0.0024) 0.984 (0.0059) 0.979 (0.0079) 0.967 (0.0041) 0.994 (0.0012) 0.993 (0.001) 0.99 (0.0028) 0.984 (0.0042) 0.977 (0.007) 0.97 (0.0055) KNN 0.995 (8e-04) 0.993 (0.0017) 0.99 (0.0019) 0.987 (0.0044) 0.985 (0.0029) 0.979 (0.0038) 0.995 (0.0015) 0.992 (0.0013) 0.991 (0.0022) 0.986 (0.0047) 0.98 (0.0052) 0.981 (0.0057) MIDA 0.994 (9e-04) 0.992 (0.0025) 0.989 (0.0029) 0.981 (0.0056) 0.975 (0.0074) 0.96 (0.0057) 0.995 (0.0015) 0.991 (0.0019) 0.989 (0.0025) 0.982 (0.0037) 0.972 (0.0061) 0.96 (0.0127) GAIN 0.995 (0.0013) 0.992 (0.0023) 0.989 (0.002) 0.983 (0.0059) 0.978 (0.0052) 0.965 (0.0059) 0.994 (0.002) 0.991 (0.0014) 0.99 (0.0031) 0.984 (0.0039) 0.978 (0.0062) 0.97 (0.0068) DPNB 0.993 (0.0025) 0.992 (0.0029) 0.991 (0.002) 0.989 (0.0027) 0.987 (0.0037) 0.984 (0.004) 0.991 (0.0024) 0.991 (0.0019) 0.99 (0.0029) 0.989 (0.0031) 0.988 (0.0032) 0.988 (0.003) Wine Accuracy rate RF 0.96 (0.0143) 0.936 (0.0205) 0.917 (0.0302) 0.86 (0.0174) 0.806 (0.0301) 0.743 (0.019) 0.966 (0.01) 0.95 (0.0155) 0.919 (0.0168) 0.867 (0.0148) 0.788 (0.0316) 0.687 (0.0451) MICE 0.963 (0.0115) 0.935 (0.0179) 0.921 (0.0169) 0.878 (0.0219) 0.827 (0.0217) 0.784 (0.0243) 0.969 (0.0095) 0.943 (0.0183) 0.913 (0.0167) 0.867 (0.0238) 0.805 (0.0376) 0.754 (0.02) KNN 0.961 (0.0133) 0.948 (0.0137) 0.925 (0.0187) 0.888 (0.0223) 0.853 (0.0323) 0.823 (0.0228) 0.96 (0.0079) 0.944 (0.0093) 0.922 (0.019) 0.875 (0.02) 0.832 (0.0294) 0.782 (0.0247) MIDA 0.959 (0.0128) 0.941 (0.0148) 0.925 (0.0193) 0.883 (0.0158) 0.827 (0.0123) 0.801 (0.0257) 0.964 (0.0124) 0.948 (0.0163) 0.925 (0.0179) 0.873 (0.0225) 0.831 (0.0353) 0.761 (0.0277) GAIN 0.961 (0.0113) 0.947 (0.0162) 0.927 (0.0221) 0.889 (0.0234) 0.851 (0.0196) 0.807 (0.0245) 0.969 (0.0134) 0.946 (0.0137) 0.924 (0.0198) 0.882 (0.016) 0.817 (0.0249) 0.767 (0.0185) DPNB 0.968 (0.0117) 0.961 (0.0157) 0.952 (0.0111) 0.932 (0.0186) 0.91 (0.0219) 0.875 (0.0202) 0.973 (0.0107) 0.964 (0.0125) 0.95 (0.0181) 0.939 (0.016) 0.897 (0.0202) 0.862 (0.0298) Wine Quality F1 Score RF 0.587 (0.0087) 0.544 (0.0133) 0.525 (0.0149) 0.498 (0.018) 0.448 (0.0167) 0.401 (0.0192) 0.594 (0.0118) 0.544 (0.0122) 0.528 (0.0114) 0.496 (0.014) 0.445 (0.0133) 0.401 (0.0148) MICE 0.587 (0.0145) 0.548 (0.0099) 0.515 (0.0163) 0.481 (0.0263) 0.438 (0.0271) 0.403 (0.0137) 0.592 (0.0108) 0.548 (0.0091) 0.529 (0.0108) 0.486 (0.02) 0.437 (0.0219) 0.404 (0.0284) KNN 0.592 (0.0115) 0.55 (0.0097) 0.52 (0.0114) 0.493 (0.0117) 0.466 (0.0238) 0.419 (0.0146) 0.594 (0.0105) 0.555 (0.0106) 0.539 (0.017) 0.499 (0.0139) 0.457 (0.0245) 0.427 (0.0176) MIDA 0.589 (0.011) 0.548 (0.0159) 0.526 (0.0147) 0.494 (0.0259) 0.459 (0.031) 0.423 (0.0186) 0.592 (0.0102) 0.552 (0.0088) 0.525 (0.0143) 0.5 (0.0164) 0.456 (0.0221) 0.421 (0.0261) GAIN 0.589 (0.0106) 0.551 (0.0124) 0.515 (0.0183) 0.489 (0.0198) 0.448 (0.0273) 0.407 (0.0161) 0.586 (0.012) 0.55 (0.0097) 0.526 (0.0166) 0.493 (0.025) 0.453 (0.0243) 0.409 (0.0152) DPNB 0.597 (0.0092) 0.566 (0.0077) 0.547 (0.014) 0.528 (0.0149) 0.51 (0.0082) 0.488 (0.0078) 0.596 (0.0108) 0.567 (0.0111) 0.544 (0.0104) 0.527 (0.0135) 0.511 (0.013) 0.487 (0.0108)

Table 10: Average of 10 times repeated 10-folds cross-validated metrics for four UCI datasets with different missingness patterns and missing rates. Estimated standard errors of the average cross-validated scores are shown in parentheses.
50% 60% 70% 80% 90%
RF+L 0.161 (0.032) 0.168 (0.0306) 0.168 (0.0322) 0.182 (0.0344) 0.177 (0.0467)
RF+R 0.112 (0.0632) 0.123 (0.0585) 0.124 (0.0647) 0.118 (0.0655) 0.111 (0.0789)
MICE+L 0.146 (0.0346) 0.151 (0.0322) 0.157 (0.0365) 0.168 (0.0409) 0.171 (0.0595)
MICE+R 0.006 (0.0118) 0.005 (0.0117) 0.007 (0.0163) 0.008 (0.0186) 0.005 (0.0194)
KNN+L 0.141 (0.0234) 0.146 (0.0324) 0.149 (0.0362) 0.161 (0.0446) 0.145 (0.0587)
KNN+R 0.118 (0.0492) 0.121 (0.0528) 0.127 (0.0527) 0.131 (0.0531) 0.131 (0.0618)
MIDA+L 0.11 (0.0306) 0.11 (0.0338) 0.122 (0.0403) 0.14 (0.0489) 0.13 (0.0641)
MIDA+R 0.166 (0.0256) 0.167 (0.0324) 0.178 (0.0215) 0.178 (0.0232) 0.174 (0.0334)
GAIN+L 0.15 (0.0316) 0.158 (0.0316) 0.171 (0.0289) 0.181 (0.0343) 0.171 (0.0602)
GAIN+R 0.16 (0.0387) 0.172 (0.0216) 0.168 (0.0347) 0.174 (0.0291) 0.174 (0.0353)
DPNB 0.169 (0.0223) 0.171 (0.0182) 0.171 (0.0248) 0.179 (0.0309) 0.177 (0.042)
Table 11: Average of F1-scores over 100 replications for the semiconductor dataset with different ratios between training and test sets. Estimated standard errors of F1-scores are shown in parentheses.