Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

The Discovery of Novel Biomarkers Improves Breast Cancer Intrinsic Subtype Prediction and Reconciles the Labels in the METABRIC Data Set

  • Heloisa Helena Milioli,

    Affiliations Priority Research Centre for Bioinformatics, Biomarker Discovery and Information-Based Medicine, Hunter Medical Research Institute, New Lambton Heights, NSW, Australia, School of Environmental and Life Science, The University of Newcastle, Callaghan, NSW, Australia

  • Renato Vimieiro,

    Affiliations Priority Research Centre for Bioinformatics, Biomarker Discovery and Information-Based Medicine, Hunter Medical Research Institute, New Lambton Heights, NSW, Australia, Centro de Informática, Universidade Federal de Pernambuco, Recife, PE, Brazil

  • Carlos Riveros,

    Affiliations Priority Research Centre for Bioinformatics, Biomarker Discovery and Information-Based Medicine, Hunter Medical Research Institute, New Lambton Heights, NSW, Australia, School of Electrical Engineering and Computer Science, The University of Newcastle, Callaghan, NSW, Australia

  • Inna Tishchenko,

    Affiliations Priority Research Centre for Bioinformatics, Biomarker Discovery and Information-Based Medicine, Hunter Medical Research Institute, New Lambton Heights, NSW, Australia, School of Electrical Engineering and Computer Science, The University of Newcastle, Callaghan, NSW, Australia

  • Regina Berretta,

    Affiliations Priority Research Centre for Bioinformatics, Biomarker Discovery and Information-Based Medicine, Hunter Medical Research Institute, New Lambton Heights, NSW, Australia, School of Electrical Engineering and Computer Science, The University of Newcastle, Callaghan, NSW, Australia

  • Pablo Moscato

    Pablo.Moscato@newcastle.edu.au

    Affiliations Priority Research Centre for Bioinformatics, Biomarker Discovery and Information-Based Medicine, Hunter Medical Research Institute, New Lambton Heights, NSW, Australia, School of Electrical Engineering and Computer Science, The University of Newcastle, Callaghan, NSW, Australia

Abstract

Background

The prediction of breast cancer intrinsic subtypes has been introduced as a valuable strategy to determine patient diagnosis and prognosis, and therapy response. The PAM50 method, based on the expression levels of 50 genes, uses a single sample predictor model to assign subtype labels to samples. Intrinsic errors reported within this assay demonstrate the challenge of identifying and understanding the breast cancer groups. In this study, we aim to: a) identify novel biomarkers for subtype individuation by exploring the competence of a newly proposed method named CM1 score, and b) apply an ensemble learning, as opposed to the use of a single classifier, for sample subtype assignment. The overarching objective is to improve class prediction.

Methods and Findings

The microarray transcriptome data sets used in this study are: the METABRIC breast cancer data recorded for over 2000 patients, and the public integrated source from ROCK database with 1570 samples. We first computed the CM1 score to identify the probes with highly discriminative patterns of expression across samples of each intrinsic subtype. We further assessed the ability of 42 selected probes on assigning correct subtype labels using 24 different classifiers from the Weka software suite. For comparison, the same method was applied on the list of 50 genes from the PAM50 method.

Conclusions

The CM1 score portrayed 30 novel biomarkers for predicting breast cancer subtypes, with the confirmation of the role of 12 well-established genes. Intrinsic subtypes assigned using the CM1 list and the ensemble of classifiers are more consistent and homogeneous than the original PAM50 labels. The new subtypes show accurate distributions of current clinical markers ER, PR and HER2, and survival curves in the METABRIC and ROCK data sets. Remarkably, the paradoxical attribution of the original labels reinforces the limitations of employing a single sample classifiers to predict breast cancer intrinsic subtypes.

Introduction

Breast cancer has been perceived as several distinct diseases characterised by intrinsic aberrations, heterogeneous behaviour and divergent clinical outcome [1]. The classification of breast cancer in discernible molecular subtypes has motivated translational researchers in the past decades towards the design of patient prognosis and the development of tailored treatments [2]. In this scenario, the analysis of breast tumours using microarray data has significantly improved the disease taxonomy and the discovery of new biomarkers for implementation in clinical practice [36]. In the early 2000s, five intrinsic subtypes were proposed: luminal A, luminal B, HER2-enriched, normal-like and basal-like breast tumours [7, 8]. Following this initial molecular taxonomy, further sub-classifications of breast cancer in distinct entities have been suggested [911].

The transcriptomic patterns observed across subtypes has given us insight into the molecular complexity and inherent alterations in tumour cells modelling the breast cancer heterogeneity and unpredicted outcome [12, 13]. Strikingly, intrinsic gene lists have been explored to reliably assign breast tumour samples into formal molecular subtypes, survival rate and treatment outline [3, 7, 8, 1418]. Recently, Parker and colleagues [16] proposed a list of 50 genes that together with the Prediction Analysis for Microarrays (PAM) classification algorithm [19] aimed at identifying subtypes and enlarging the prognostic information with high potential for validation in clinical settings [16, 20, 21]. The resulting technique, called the PAM50 method, has been widely applied to categorize tumours into one of the five classical intrinsic subtypes.

Although independent cohorts attempted to identify molecular subtypes, the chosen microarray-based Single Sample Predictor (SSP) model revealed unreliable assignments and modest agreement between studies [21, 22]. In fact, the perceived inability of some analytical methods to deal with the challenges of processing high-dimensional data, in addition to the difficulties on validating independent/unpaired technologies may limit the precise characterisation of the subtypes [21, 23, 24]. Therefore, novel methods are urgently needed in order to provide better tumour stratification and accurate biomarkers identification [25, 26]. In this scenario, the high quality of the microarray gene expression data set processed by the Molecular Taxonomy of Breast Cancer International Consortium (METABRIC) [27], with over 2000 samples, offers a unique opportunity to refine and expand the list of transcripts that best discriminate intrinsic subtypes. A precise classification of breast tumours, consequently, would lead to improvements in the valuation of the disease, currently guided by oestrogen and progesterone receptor (ER and PR) status, and HER2 amplification [24, 28].

In this report, we focus on the use of a ranking feature method based on the newly proposed CM1 score [29] to identify probe sets that appear naturally from the METABRIC breast cancer data set. For doing so, we use the entire set of 48803 probes as an alternative to the selection from pre-existing literature as performed by other authors [15, 16]. Moreover, the quality of the probes for predicting subtypes is carefully appraised in the METABRIC data set (Illumina BeadArray) and further validated in different studies (Affymetrix GeneChip) accessed through the Research Online Cancer Knowledgebase (ROCK) interface [30]. However, instead of relying on a single method to assign sample subtype, as suggested by Parker et al. (2009) [16] with the PAM50 method, we explore an ensemble learning. Our analysis is based on the performance of a large set of classification models from the Weka software suite [31]; a technique previously recommended by Ravetti and Moscato [32]. The classifiers are used in combination with the list of probes selected using CM1 score and, alternatively, with the 50 genes from the PAM50 commercial assay [16]. We also compute several statistical measures to determine the power of both lists on predicting breast cancer subtypes. Ultimately, we correlate the study outcomes within current clinical information and survival analysis.

Materials and Methods

Data sets description

The METABRIC microarray data set used in this study is hosted by the European Bioinformatics Institute (EBI) and deposited in the European Genome-Phenome Archive (EGA) at http://www.ebi.ac.uk/ega/, under accession number EGAS00000000083. It consists of transcriptomic information (cDNA microarrays profiling) processed on the Illumina HT-12 v3 platform (Illumina_Human_WG-v3), as described in [27]. The log2-normalised gene expression values of primary tumours were divided into two subsets by METABRIC: discovery (997 samples) and validation (989 samples), which were respectively used as training and test sets in our experiments. The original study collected and analysed data under the approval of the ethics Institutional Review Board (details in [27]). The use of this data for research was also approved by the Human Ethics Research Committee (HREC) of The University of Newcastle, Australia, (approval number: H-2013–0277).

The second data set is publicly available in ROCK online portal [30] at http://rock.icr.ac.uk/, under data source access GSE47561. This source integrates ten data studies (GSE2034, GSE11121, GSE20194, GSE1456, GSE2603, GSE6532, GSE20437, E-TABM-185, GSE7390, GSE5847) performed on the Affymetrix Human Genome U133A Array (HG-U133A) platform. The matrix contains log2 RMA re-normalised gene expression data in a unique comprehensive report of 1570 samples. Thus, the GSE47561 data set was used as a second validation set to test our method.

In brief, both METABRIC and ROCK data sets have information on patients’ long-term clinical and pathological outcomes, including the sample assignment into intrinsic subtypes (luminal A, luminal B, HER2-enriched, normal-like, and basal-like) according to the PAM50 method [16]. The METABRIC data set has a more comprehensive description of patient clinical features, whereas the ROCK data set presents no standardized information across the ten different studies.

Study Design and Computing Resources

In this study, we propose a systematic approach that aims at improving breast cancer subtype prediction. The systematic approach is built based on feature selection and data mining concepts. We first compute the CM1 score—using the microarray mRNA expression values—to rank the whole set of probes based on their discriminative power across breast cancer subtypes. We then select the top 10 probes that best represent each intrinsic subtype. The quality of this selection is assessed using a set of classifiers from the Weka software suite with the METABRIC and ROCK data sets, followed by the statistical analysis. The process flow is depicted in Fig 1, and further explained in the remainder of this section.

thumbnail
Fig 1. The step-by-step process.

The image shows the method steps based on CM1 score and ensemble learning. The METABRIC discovery set is used to compute the CM1 score, based on the original labels previously assigned with the PAM50 method. This step has an output of 42 discriminative probes selected, the CM1 list. The following step involve the sample subtype classification based on a 10-fold cross-validation. Samples in the METABRIC discovery set are considered to train 24 classifiers using the CM1 list and, alternatively, the PAM50 list. The samples are partitioned into ten folds; then a model is built using 90% of samples, which is used to predict the labels of the remaining 10%. After the ten turns are finished, the level of association between the predicted and original METABRIC labels is computed using several statistics. In the training-test setting, labels of samples in the METABRIC validation set and ROCK set are predicted with the models built in the discovery. Statistics measurements are again computed to assess the model performance on predicting breast cancer subtypes. In both classification steps, the new labels are attributed based on the consensus of the majority of the classifiers. Finally, the results or new labels are compared against the clinical data, the current markers ER, PR and HER2, and survival curves.

https://doi.org/10.1371/journal.pone.0129711.g001

Selection of biomarkers using the CM1 score

The CM1 score is a supervised univariate method used to measure the difference in expression levels of samples in two different classes [29]. In this study, it is used as a ranking feature to select a subset of highly discriminative probes for each breast cancer intrinsic subtype. Let X and Y be a partition of a set of samples into two classes, with X the class of interest and Y the remaining classes. A sample either belongs to X or to Y. For each probe i we compute the CM1 score as: (1) where is the average expression value of the probe i for samples in class X, is the average expression value of the probe i for samples in class Y; max{yi} and min{yi} are the minimum and maximum expression values of the probe i for samples in the class Y. Eq 1 can be interpreted as the normalised difference between the averages of expression values in the class X and Y. The normalisation is proportional to the range of values in Y.

To define the most discriminative probes for each breast cancer subtype (luminal A, luminal B, HER2-enriched, normal-like and basal-like), we computed the CM1 score for each of 48803 probes taking the subtype of interest and the remaining ones. This results in 5 lists of 48803 CM1 scores.

Considering the fact that Parker et al. (2009) [16] were able to define the five breast cancer classes based on 50 genes, for each subtype we chose the 10 most important probes (5 with the greatest positive CM1 score values—indicating up-regulated probes relative to the other subtypes –, and 5 with the smallest negative values – representing down-regulation). This set is referred to as the balanced top ten in this paper. Collecting the balanced top ten lists of all subtypes leads to a new set of 42 unique Illumina probes, meaning that 8 probes appear in multiple subtypes. This list is hereafter called the CM1 list.

Assessment of the quality of the CM1 list based on ensemble learning

The quality of the CM1 list for distinguishing subtypes was assessed using a list of well-known classifiers available in the Weka data mining software suite [31]. It uses different types of classifiers such as bayesian, functions, lazy, meta, rule-based and decision trees. Each classifier was trained with a subset of the data comprising all samples in the METABRIC discovery set and the 42 probes in the CM1 list using both 10-fold cross-validation and training-test setting. In the 10-fold cross-validation, the samples are first partitioned into ten folds; then a model is built using 90% of samples, which is thereafter used to predict the labels of the remaining 10%. After the ten turns are finished, the level of association between the predicted and original METABRIC labels is computed using Cramer’s V [33]. In the training-test setting, labels of samples in the METABRIC validation set and ROCK data are predicted using models built with the samples in the discovery set. The new labels were attributed based on the consensus of the majority of the classifiers (i.e. more than 50% percent), and whenever such condition was not achieved samples were marked as inconsistent (INC).

A similar approach was performed with the PAM50 list to serve as baseline for comparing the results obtained with the 42 probes from the CM1 list. The 50 genes identified by Parker et al. (2009) [16] were mapped to Illumina probes by Curtis et al. (2012) [27], following strict criteria. Only genes and corresponding probe with perfect annotation [34] on the Illumina HT-12 v3 BeadChip were considered. Probes containing SNPs, multiple targets or mismatches, or lying in repeat-masked regions were discarded. Finally, a total of 48 probes corresponding to genes in the PAM50 list were selected to conduct the classification experiments as described for the CM1 list. For Affymetrics HG-U133A, the CM1 and PAM50 lists were mapped according to ‘genefu’ R package, using Entrez Gene ID as reference. For instance, the 42 probes from the CM1 list matched 33 probes, whereas the 48 from PAM50 list paired 43 probes in the Affymetrix platform. In case of multiple mappings the probe with the most variation was selected according to the ‘genefu’ instructions. Before testing the classifiers in ROCK data set, the Affymetrix and Illumina expression levels were min-max normalised.

The statistical analysis

Cramer’s V.

Given a r × c contingency table describing the association between the original labels and those predicted by the majority of classifiers, Cramer’s V measures the level of association between those two nominal variables. The statistic ranges from 0, representing no association between the two variables, to 1, representing complete association. Cramer’s V is computed using Eq 2. (2) where N is the number of samples in the data set, and χ2 is Pearson’s chi-squared value.

Average sensitivity (AS).

The average sensitivity (AS) [31] was also computed to assess the performance of classifiers with both lists. The AS is the average proportion of accurately classified samples of each subtype. Considering a r × c contingency table associating initial and predicted labels, the average sensitivity of a classifier is given by Eq 3. (3) where r is the number of classes (subtypes), nii is the number of samples of class i correctly predicted as i, and ni is the number of samples of class i (row marginals).

Fleiss’ kappa.

The consensus of the different classification methods concerning the samples’ labels was measured by the popular interrater reliability metric Fleiss’ kappa [35, 36]. The statistic was used to gauge not only the agreement among classifiers trained with the different probe sets, but also between the labels assigned by the majority of classifiers and the original METABRIC labels. It also quantifies the agreement between predicted labels using the CM1 and PAM50 lists.

Assuming a s × c contingency table informing how many times each of the c classes were assigned to each of the s samples in the k different sample labellings, the Fleiss’ kappa statistic is computed as defined by Eq 4. (4) where nij contains the number of times sample i was assigned label j, ∑j nij = k, and pj = (∑i nij)/sk is the probability with which the label j is assigned to a sample.

Kappa values range from to +1, which, according to Landis and Koch’s division [37], can be interpreted in the following manner: (1) values below zero are considered poor agreement; (2) values between zero and 0.2 are considered slight agreement; (3) 0.21 ≤ κ ≤ 0.40 is fair agreement; (4) 0.41 ≤ κ ≤ 0.60 moderate agreement; (5) 0.61 ≤ κ ≤ 0.80 substantial agreement; and (6) 0.81 ≤ κ ≤ 1 is regarded as an almost perfect agreement.

Adjusted Rand Index.

The agreement between pairs of sample labellings was also quantified using this metric. It ranges between 0 to 1, where 1 indicates an almost perfect concordance between the two compared bipartitions, and 0 a complete discordance between them. The Adjusted Rand Index is a version of Rand index corrected for chance when the partitions are picked at random [38, 39]. Given a r × c contingency table between two labelling R and C, it can be measured by: (5) where 1 ≤ ir, 1 ≤ jc, and nij is an entry of the contingency table representing the number of samples that are in class Ri in the partition R and Cj in the partition C, ni and nj are the table’s marginals.

Survival analysis

The survival analysis for each breast cancer subtype is performed using Cox proportional hazards model from the package survival in the R software [40, 41]. Only patients who either died due to the disease or are still alive are considered for model estimation. The clinical parameters relevant for the survival study are chosen in correspondence with Curtis et al. (2012) [27]: age at the time of diagnosis, tumor size, tumor grade, the number of positive lymph nodes and ER status according to immunohistochemistry. Since the probability model based on the observations available at certain time points becomes less and less reliable with the increasing time, the median survival lines based on the last 10 observations are plotted in dash. Due to the compilation of ten different studies and the existence of significant gaps in patients’ clinical information, the survival curves in the ROCK data set are not representative across subtypes. In particular, the number of patients with information about overall survival and disease free survival is limited to only 405, with no specification on the cause of death (i.e. if due to disease or not).

Results

Section description and resources

To understand the results described in this section, we introduce the sequence of our approach which combines the CM1 score and ensemble learning. First, we detail the selection of discriminative probes ranked according to the CM1 score; calculated for each of the five breast cancer subtypes. Second, we show the quality of our probes by using 24 classification models based on a 10-fold cross-validation and training-test setting in the METABRIC and ROCK data sets. The same approach is also performed with the list of 50 genes used in the PAM50 method. In addition, statistical analysis are reported to determine the power of both lists on predicting breast cancer subtypes. Finally, we demonstrate the consistency between the new labels assigned with current clinical markers ER, PR and HER2, and survival curves. The step-by-step approach is detailed in the Materials and Methods section.

Using the CM1 list to differentiate the five intrinsic breast cancer subtypes

The CM1 score was applied to rank the set of 48803 probes for each of the five subtypes in the METABRIC discovery data set (Supporting Information S1 Table). It is important to remark that this method used the original PAM50 subtypes attributed to samples in the METABRIC discovery set. The purpose of doing so is to provide a better molecular characterisation of each class using the wealth of the METABRIC transcriptomic data, besides improving the breast cancer subtype prediction. The probes with the top five negative and top five positive CM1 scores were selected for each subtype. Here, we aimed at obtaining 50 probes that appear naturally from a rich and unique data set. We would then be able to compare such a list with the list of 50 genes embedded in the PAM50 method [16]—the PAM50 list. The final list comprising the union of the top ranked probes is displayed in Table 1, and their CM1 scores and ranks in each subtype in Table 2. Some of the 50 probes selected, however, discriminate more than one subtype and resulted in a list of 42 unique elements, the CM1 list. Our selection includes 30 novel biomarkers, while the remaining 12 genes are common with the PAM50 list.

The effectiveness of the CM1 list for segregating the five subtypes is depicted in Fig 2. The figure shows the expression values of the top five negative and top five positive ranked probes for each subtype across 997 samples in the METABRIC discovery set. For instance, the ten probes selected for the basal-like subtype—the most representative class—expose a consistent separation between samples from this class and the remaining ones. The second heat map in Fig 3 illustrates the expression levels of unique probes from the CM1 list in the Illumina platform, in which rows represent probes and columns represent samples. Rows and columns were ordered according to gene expression similarity using a memetic algorithm [27]. This image also exposes the overall discriminative power of our list for distinguishing samples of the five subtypes.

thumbnail
Fig 2. The gene expression profile of the balanced top ten probes selected for each of the five breast cancer intrinsic subtypes across 997 samples from the discovery set.

The annotated genes are defined for each subtype as an intrinsic, highly discriminative, signature. Samples were ordered according to the gene expression similarities in each breast cancer subtype. Colours represent the selected genes and sample subtypes: luminal A (yellow), luminal B (green), HER2-enriched (purple), normal-like (blue), and basal-like (red).

https://doi.org/10.1371/journal.pone.0129711.g002

thumbnail
Fig 3. Gene expression patterns of the 42 probes selected using the CM1 score.

The heat map diagram exhibit 42 probes (rows) and 997 samples (columns) from the discovery set ordered according to gene expression similarity, based on a memetic algorithm [27]. The labels highlighted on top show the sample distribution according to the ER positive and negative status. It also illustrates the original PAM50 subtypes luminal A (yellow), luminal B (green), HER2-enriched (purple), normal-like (blue), and basal-like (red) in the METABRIC discovery set. Two probes in the CM1 list refer to the same gene, HPN, which was then appended with the corresponding Illumina probe ID.

https://doi.org/10.1371/journal.pone.0129711.g003

A detailed description of our 42 probes in the context of the literature can be found in Supporting Information S1 Text. Among them we highlight seven, targeting the following transcripts: AURKB, CCL15, C6orf211, GABRP, IGF2BP3, PSAT1, and TFF3. Fig 4 shows the box plot of their expression levels across intrinsic subtypes in the METABRIC discovery and validation sets, and the ROCK set. We emphasized these transcripts due to the remarkable differential expression behaviour across the five classes. Besides, they are novel potential markers for breast cancer subtyping, not considered by Parker et al [16]. Box plots of expression levels for all transcripts in the CM1 list in the METABRIC discovery and validation and ROCK data sets are provided in Supporting Information S1 Fig. Even though those probes were selected from the METABRIC discovery set only, their variation across subtypes in the validation set and ROCK test set are also impressive.

thumbnail
Fig 4. The mRNA log2 normalised expression values of 7 novel highly discriminative biomarkers across the five intrinsic subtypes in the METABRIC discovery and validation sets, and ROCK set.

The box plot uncover the values of 997 samples in the METABRIC discovery set, 989 in the validation set, and 1570 in the ROCK test set.

https://doi.org/10.1371/journal.pone.0129711.g004

The ensemble of classifiers reveal high levels of agreement between CM1 and PAM50 lists

After applying the ensemble learning, several statistical measures were computed as referred in Materials and Methods. The main purpose of the statistics is to determine the performance of the 24 classification methods from the Weka software suite. In other words, we investigate the consistency of intrinsic subtype labels attributed by the majority of classifiers having as input either the CM1 or PAM50 lists. The quality of both lists was estimated according to the Cramer’s V statistic and the Average Sensitivity. Additionally, we computed the popular interrater reliability metric Fleiss’ kappa to establish the consensus of sample labelling across different classifiers. This metric was used to gauge the agreement among classifiers trained with CM1 and PAM50 lists against the original labels in the data sets, and between the labels assigned by the majority of classifiers using both lists. Ultimately, we applied the Adjusted Rand Index to quantify the agreement between pairs of samples that are either in the same class or in different classes according to both lists.

Average Cramer’s V statistic and Average Sensitivity to measure the performance of individual classifiers.

We determined the performance of the ensemble learning (Supporting Information S2 Table, and S3 Table) with two measures: Cramer’s V statistic and Average Sensitivity (Table 3). Cramer’s V is used to measure the strength of association among variables in the row and column, given a contingency table (Tables 4, 5 and 6). The rows represent the original PAM50 labels and the columns the subtypes assigned by the majority of the classifiers in the ensemble. For instance, Cramer’s V statistic showed an average association between original and predicted subtypes of 0.73±0.06 and 0.63±0.04 in the METABRIC discovery and validation sets respectively with the CM1 list; and 0.75±0.06 and 0.64±0.04 with PAM50 list. Expanding the validation process using the ROCK test set, Cramer’s V ranged from 0.57±0.06 with the CM1, and 0.58±0.05 using PAM50 list.

thumbnail
Table 3. The ensemble learning overall performance on assigning labels to samples in the METABRIC discovery and validation sets, and ROCK test set.

https://doi.org/10.1371/journal.pone.0129711.t003

thumbnail
Table 4. Contingency tables for predicted labels using the 24 classifiers trained with the CM1 list.

https://doi.org/10.1371/journal.pone.0129711.t004

thumbnail
Table 5. Contingency tables for predicted labels using the 24 classifiers trained with the PAM50 list.

https://doi.org/10.1371/journal.pone.0129711.t005

thumbnail
Table 6. Contingency tables for predicted labels using the 24 classifiers trained with CM1 and PAM50 lists.

https://doi.org/10.1371/journal.pone.0129711.t006

The Average Sensitivity statistic was used to characterize the average proportion of accurately labelled samples in each subtype. Considering the analysis with CM1 list, the measure was 0.76±0.06 in the METABRIC discovery set and 0.64±0.04 in the validation set; and with PAM50 list was 0.78±0.07 and 0.65±0.05, respectively. Likewise, the average sensitivity calculated for the ROCK test set was 0.67±0.07 using the CM1 and 0.69±0.08 with PAM50 list. A complete table containing the performance of all individual classification methods is available in the (Supporting Information S2 Table and S3 Table).

The levels of agreement explained by interrater reliability metric Fleiss’ kappa.

Fleiss’ kappa was computed to assess the reliability of agreement between two raters, as displayed in Table 7. We initially compared the agreement Among classifiers which indicates the overall performance of classifiers alone. We then compared Predicted vs Original, that is, the agreement between subtypes assigned by the majority of classifiers using CM1 and PAM50 lists compared to the original PAM50 labels in the METABRIC discovery and validation sets, and ROCK test set. We also calculated the kappa between labels attributed by the majority of classifiers using both lists, CM1 vs PAM50. We refer to the Materials and Methods section for an interpretation of κ values. For instance, the high levels of agreement between two raters reflect more than what would be expected by chance.

thumbnail
Table 7. Agreement of the 24 classifiers on assigning labels to samples in the data sets measured by Fleiss’ kappa statistic.

https://doi.org/10.1371/journal.pone.0129711.t007

Considering the agreement of the ensemble of classifiers, there was a substantial agreement in both METABRIC discovery and validation sets, and ROCK test set (Table 7). Fleiss’ kappa was 0.73, 0.75 and 0.63 with the CM1 list for METABRIC discovery, validation and ROCK data sets, respectively. Values obtained with the PAM50 list were 0.72, 0.73 and 0.6, respectively. By comparing the subtypes predicted by the majority of classifiers and original PAM50 labels, there was an almost perfect agreement with CM1 (κ = 0.81) and PAM50 (κ = 0.84) lists in the discovery set. In the validation and ROCK sets, on the other hand, labels showed only a moderate agreement for both lists (κ ≃ 0.6). Strikingly, the Fleiss’ kappa between subtypes predicted using the CM1 and PAM50 lists (κ = 0.86, 0.83, and 0.8 in the METABRIC discovery, validation, and ROCK sets, respectively) revealed an almost perfect agreement. This statistical measure confirm our visual analysis of the contingency tables as they find strong relationship across the subtype labels in each data set. A detail of the agreement among classifiers by intrinsic subtype is shown in (Supporting Information S4 Table).

The agreement according to the Adjusted Rand Index.

The agreement between the different sample labellings was also scrutinized using the Adjusted Rand Index measure (Table 8). The values obtained with the CM1 list were 0.757 in the METABRIC discovery and 0.426 in the validation sets, and 0.453 in the ROCK test set. For PAM50 list, the values were 0.792, 0.457 and 0.507, respectively. Similar to Fleiss’ kappa, the agreement between labels predicted with CM1 and PAM50 lists is higher than the agreement with the original labels. The Adjusted Rand Index values were 0.822, 0.788 and 0.642 for the three data sets, respectively. The numbers obtained with this measure also revealed remarkable concordance of CM1 and PAM50 lists assigned labels.

thumbnail
Table 8. Agreement measured by the Adjusted Rand Index between different samples’ labellings.

https://doi.org/10.1371/journal.pone.0129711.t008

The use of an ensemble learning with the CM1 list improves the subtype distribution in the METABRIC and ROCK data sets

The number of samples in each original PAM50 subtype is markedly different across the METABRIC sets (Fig 5). In the discovery set, there is a clear abundance of luminal A and B subtypes, precisely 73.62% of all samples. In contrast, the proportion of luminals in the validation set is only 48.14%. The ratio of luminal A to luminal B samples changed from 1.74 in the discovery to 1.14 in the validation set. However, when the CM1 or PAM50 lists are used in conjunction with the ensemble of classifiers, samples in the discovery and validation sets are more homogeneously distributed. The percentage of samples in the discovery set labelled as luminal A and B using CM1 and PAM50 lists are 73.53% and 73.72%, respectively. These proportions match the original number (73.62%). On the other hand, in the validation set the CM1 and PAM50 lists assigned a total of 64% and 63.19% luminal samples, against the 48.14% previously mentioned. The distribution of subtypes also become more similar to the discovery set. Likewise, ROCK test set also changed the pattern of class distribution after the performance of the ensemble of classifiers. The differences in class distributions might not be attributed to the randomisation procedure used by the studies as the performance of the ensemble of classifiers with both lists reconcile the distribution of subtypes.

thumbnail
Fig 5. Class distribution in the METABRIC discovery and validation sets, and in the ROCK set.

The bars represent the number of samples in each breast cancer subtype. In the first row, the labels refer to the original assignment using the PAM50 method. The following rows show the new labels attributed using an ensemble of 24 classifiers with PAM50 and CM1 lists, respectively. Samples were classified as inconsistent if there was no consensus between the majority of classifiers as to what should be the correct subtype.

https://doi.org/10.1371/journal.pone.0129711.g005

We summarize the similarities and differences in subtypes distribution (graphically displayed in Fig 5) by computing the square root of the Jensen-Shannon divergence [225]. This is a true metric of distance between probability distributions. Its plot in Fig 6 shows the similarity between all possible pairs of data sets based on their distribution of subtype labels (Supporting Information S4 Table). It can be observed that the original labels are the most divergent ones, especially in the METABRIC validation and ROCK test sets. The high similarity of samples distribution among subtypes based on the assignments with CM1 or PAM50 lists is evident. Such similarity was not expected for the ROCK set as the ensemble of classifiers was trained with METABRIC discovery (Illumina platform data) and tested in the ROCK set (Affymetrix platform data). The limited number of probes matching Illumina and Affymetrix in both lists (as described in Materials and Methods) seems not to affect the performance of the ensemble learning. Yet the divergences in the original class distributions might not be attributed to the randomisation procedure used by the consortium. These results point out to the relative strength and robustness of a set of classifiers compared to single methods to predict breast cancer subtype labels. They also indicate that there is an issue to be considered by researchers when using the original PAM50 labels from the METABRIC study for analysing data and building predictive models.

thumbnail
Fig 6. Similarity between subtypes distribution in the METABRIC discovery and validation sets, and in the ROCK set.

The image shows the similarity between the subtypes distribution for METABRIC discovery (MD) and validation (MD) sets, and ROCK test set (RS). The labels were assigned in the original data sets using the PAM50 method, and relabelled in this study with an ensemble learning using PAM50 and CM1 lists. The similarity is measured using the square root of the Jensen-Shannon divergence. Darker shades represent more similar distributions, while lighter shades refer to divergent patterns. The diagonal shows the darkest color as each data set is the closest to itself. According to this image, labels assigned using an ensemble learning with CM1 and PAM50 lists are highly similar, and both exhibit lower levels of agreement with the original labels assigned using a single classifier (PAM), or PAM50 method.

https://doi.org/10.1371/journal.pone.0129711.g006

Breast cancer intrinsic subtypes show different clinical markers distribution and survival curves

Given the heterogeneity among breast cancer patients and the intricate assignment of PAM50 labels in the original METABRIC data set, we further investigated whether significant differences exist in the analysis of current clinical markers (ER, PR and HER2). Figs 7, 8 and 9 show, respectively, the distribution of the ER, PR and HER2 across intrinsic subtypes in the METABRIC discovery and validation sets, considering the original PAM50 labels and the labels assigned by ensemble of classifiers using CM1 and PAM50 lists. The new subtype labelling markedly improves the status of the clinical markers in the METABRIC data set. For instance, the ER marker distribution across subtypes shows an important decrease in the number of HER2-enriched and basal-like samples that are ER-positive according to the original PAM50 labels. The PR marker, likewise, varies the distribution when predicted labels based on the ensemble of classifiers using either CM1 and PAM50 list are compared with the original labels. HER2 amplification has a particular behaviour across all subtypes. Under the new subtype labels, the distribution of the three clinical markers becomes more consistent with what is expected according to the literature for each class: luminal A (ER+ and/or PR+, HER2−); luminal B (ER+ and/or PR+, HER2±); HER2-enriched (ER−, PR− and HER2+); and basal-like (ER−, PR−, HER2−) [226].

thumbnail
Fig 7. ER marker distribution across subtypes in the METABRIC data sets.

(A) Discovery and (B) Validation. The bars represent the number of samples with ER positive and negative in the five intrinsic subtypes, based on the patients’ clinical information. The top row is based on the original subtype labels obtained with the PAM50 list and a single classifier (PAM). Middle and bottom rows are based on the labels obtained by Ensemble Learning using the PAM50 and CM1 lists, respectively.

https://doi.org/10.1371/journal.pone.0129711.g007

thumbnail
Fig 8. PR marker distribution across subtypes in the METABRIC data set.

(A) Discovery and (B) Validation. The bars represent the number of samples with PR positive and negative distributed in the five intrinsic subtypes, based on the patients’ clinical information. The top row is based on the original subtype labels obtained with the PAM50 list and a single classifier (PAM). Middle and bottom rows are based on the labels obtained by Ensemble Learning using the PAM50 and CM1 lists, respectively.

https://doi.org/10.1371/journal.pone.0129711.g008

thumbnail
Fig 9. HER2 distribution across subtypes in the METABRIC data sets.

(A) Discovery and (B) Validation. The bars represent the number of samples with HER2 amplification (positive or negative) for each intrinsic subtype based on the patients’ clinical information. The top row is based on the original subtype labels obtained with the PAM50 list and a single classifier (PAM). Middle and bottom rows are based on the labels obtained by Ensemble Learning using the PAM50 and CM1 lists, respectively.

https://doi.org/10.1371/journal.pone.0129711.g009

Subsequently, we illustrate the survival curves for all breast cancer subtypes using Cox proportional hazards model, as described in Materials and Methods. The curves were plotted based on the original PAM50 labels and those assigned by the majority of classifiers. For generating the survival curves, we included the most relevant clinical variables as covariates: grade, size, age at diagnosis, number of lymph nodes positive, and ER status (immunohistochemistry) [27]. This analysis revealed different curves in the METABRIC discovery and validation sets (Fig 10). For instance, luminal B and basal-like subtypes show a similar pattern across data sets. Luminal A, HER2-enriched and normal-like, on the other hand, have a more consistent survival pattern when the CM1 and PAM50 lists are used in conjunction with the ensemble learning. Taken as a whole, the results of this section support the increased robustness of labels assigned by the ensemble of classifiers with the CM1 or PAM50 lists, and point out to inconsistencies in the original subtype assignment in the METABRIC study.

thumbnail
Fig 10. The survival curves for METABRIC discovery and validation sets.

The survival curves for each breast cancer subtype are generated using Cox proportional hazards model based on the grade and size of the tumour, patient’s age, number of lymph nodes positive and ER status. Each curve represents the survival probability at a certain time after the diagnosis. Ticks on the curve correspond to the observations of patients who are still alive, while drops indicate the death. The probability curves based on the last 10 observations are plotted in dash. The top row is based on the original subtype labels obtained with the PAM50 list and a single classifier (PAM). Middle and bottom rows are based on the labels obtained by Ensemble Learning using the PAM50 and CM1 lists, respectively.

https://doi.org/10.1371/journal.pone.0129711.g010

Discussion

In this study, we exposed the power of the CM1 list for improving the breast cancer subtype prediction in the METABRIC and ROCK data sets. The CM1 score portrayed 30 novel genes as potential biomarkers, along with 12 well-established markers shared between CM1 and PAM50 lists. The 42 biomarkers have a great potential to differentiate breast cancer intrinsic subtypes. Among them, AGR3, HPN, ANKRD30A, AURKB, PROM1, VTCN1, CRYAB, CDK1, CDKN3, SERPINA3, SOX11, TRPV6, CLCA2, MUCL1, COL11A1, DARC, TFF3, IGF2BP3, IL33, SUSD3, PSAT1, and GABRP are reported in different studies associated with breast cancer; however not in the context of subtype differentiation. Noteworthy, the CM1 list revealed a set of probes for which little literature exists in relation to breast cancer subtypes: CDCA5, CCL15, COL17A1, GLYATL2, ROPN1, LINC00993 and C6orf211. Their expression levels vary across different subtypes, and are yet a new avenue to be explored. We also emphasize the 12 common genes (CEP55, ESR1, FOXA1, FOXC1, KRT17, MAPT, MELK, MMP11, NAT1, SFRP1, UBE2C, and UBE2T) due to their important role for breast cancer disease and intrinsic subtyping.

Within the application of an ensemble of classifiers, CM1 and PAM50 lists showed concordant predictive power for disease subtyping. In fact, there was an almost perfect agreement between the labellings obtained with the majority of classifiers using both lists; however different from the original labels. In this study, we want to highlight the weakness of relying in a single method to assign subtypes labels, as opposed to the power and robustness of an ensemble learning. We therefore discourage label assignments based on a single classifier and also suggests a thorough review of those intrinsic subtypes given the importance of such data sets to breast cancer research. The results indicate that there is an issue to be considered by researchers when using the original PAM50 labels for analysing data. The use of incorrect labels would lead to a plethora of misguided and misleading results by other investigators that use METABRIC or ROCK data sets.

In spite of luminals sharing the same origin and large molecular commonalities [227, 228], the ensemble of classifiers accurately predicted luminal samples in the METABRIC data set, and showed some ambiguity on assigning the subtype A or B for a small number of samples, specially in the ROCK data set. This may be a consequence of the reduced number of probes matching across Illumina and Affymetrix platforms. HER2-enriched notably improved label consistency in the ROCK data. Furthermore, the normal-like tumours received more often contradictory and inaccurate subtype labelling among both data sets. The poor overall outcome for this subtype is supported by the discussion that normal-like is an artefact of sample processing with high contamination of normal breast tissue [13, 16, 229]; however, still crucial to be elucidated. Ultimately, the basal-like subtype maintained the classification with a unique profile, markedly divergent from other subtypes [21, 22, 230]; even though this group has recently been partitioned into new fundamental classes [9, 10].

Overall, the new intrinsic subtype labels based on the CM1 list and ensemble learning revealed more accurate distributions of clinical markers (ER, PR and HER2) and survival curves, when compared to the original PAM50 labels in the METABRIC cohort and ROCK test set. Interestingly, the CM1 list shows ESR1 (ER) among the 42 probes, but brings other independent genes that are also relevant for overall predictions. Robust data sets like METABRIC have contributed to the understanding of breast cancer disease in terms of its molecular complexity and intrinsic alterations. The main limitations of the research in the field, nevertheless, is the uncertainty in the exact classification of intrinsic subtypes; over and above the discovery of molecular signatures and standard clinical biomarkers. Under consideration, a consistent taxonomy needs yet to be established prior to implementation in clinical practice. Additional research involving the genome, transcriptome, proteome, and epigenome, will lastly portray a true landscape of subtypes and contribute to breast cancer management.

Supporting Information

S1 Text. CM1 list and literature review.

The document shows the CM1 probe list along with an extensive literature review. The 42 CM1 biomarkers revealed a great potential to differentiate breast cancer intrinsic subtypes in the METABRIC and ROCK data sets. The 30 novel markers and 12 well-established genes vary the expression levels across different subtypes. The vast majority has been associated with breast cancer disease, either included or not in the subtyping context.

https://doi.org/10.1371/journal.pone.0129711.s001

(PDF)

S1 Fig. The mRNA log2 normalised expression values of 42 probes in the CM1 list across the five intrinsic subtypes in the METABRIC discovery and validation sets, and ROCK set.

Box plots illustrating the expression levels for all selected transcripts in the CM1 list in the METABRIC discovery and validation sets, and ROCK test set. The figure shows the probes differential behaviour across breast cancer intrinsic subtypes.

https://doi.org/10.1371/journal.pone.0129711.s002

(TIFF)

S1 Table. The CM1 score calculated for each breast cancer subtype.

Table listing the CM1 score used to rank the set of 48803 probes for each of the five breast cancer subtypes in the METABRIC discovery data set. In each case, we selected the top 10 highly discriminative probes (5 with the greatest positive CM1 score values—indicating up-regulated probes relative to the other subtypes, and 5 with the smallest negative values—representing down-regulation).

https://doi.org/10.1371/journal.pone.0129711.s003

(XLSX)

S2 Table. The performance of the classifiers using the CM1 list.

Table describing the performance of each classifier on the METABRIC discovery and validation sets, and ROCK test set using the CM1 list. It shows the percentage of correctly, incorrectly and not classified samples, Fleiss Kappa index, Cramer’s V, Average Sensitivity, and other values for classification. The 24 classifiers from the Weka software suite are also listed. The labels predicted by each classifier for all samples using CM1 list are defined as: 1—luminal A, 2—luminal B, 3—HER2-enriched, 4—normal-like, 5—basal-like. Count of predicted labels was obtained with the consensus of the majority of classifiers.

https://doi.org/10.1371/journal.pone.0129711.s004

(XLSX)

S3 Table. The performance of the classifiers using the PAM50 list.

Table describing the performance of each classifier on the METABRIC discovery and validation sets, and ROCK test set using the PAM50 list. It shows the percentage of correctly, incorrectly and not classified samples, Fleiss Kappa index, Cramer’s V, Average Sensitivity, and other values for classification. The 24 classifiers from the Weka software suite are also listed. The labels predicted by each classifier for all samples using CM1 list are defined as: 1—luminal A, 2—luminal B, 3—HER2-enriched, 4—normal-like, 5—basal-like. Count of predicted labels was obtained with the consensus of the majority of classifiers.

https://doi.org/10.1371/journal.pone.0129711.s005

(XLSX)

S4 Table. The agreement between sample labelling with Fleiss’ Kappa measure and the Jensen-Shannon divergence of two probability distributions.

Table containing the Fleiss’ Kappa agreement of labels for the METABRIC discovery and validation sets, and ROCK test set. It shows the overall agreement Among classifiers using CM1 and PAM50 lists, as well as the agreement for each subtype. The predicted—original are described in the table and contain the agreement between the mostly predicted and initial labels of samples; whereas the CM1—PAM50 show agreement between labels assigned by the majority of classifiers using CM1 and PAM50 lists. The file also has the Jensen-Shannon divergence between two probability distributions. Numbers represent the similarity between subtypes distribution for METABRIC discovery and validation sets, and ROCK test set. The similarity is measured using the square root of the Jensen-Shannon divergence.

https://doi.org/10.1371/journal.pone.0129711.s006

(XLSX)

Acknowledgments

This study makes use of data generated by the Molecular Taxonomy of Breast Cancer International Consortium. Funding for the project was provided by Cancer Research UK and the British Columbia Cancer Agency Branch.

PM is supported by Australian Research Council (ARC, http://www.arc.gov.au/) Future Fellowship FT120100060. This project is partially funded by ARC Discovery Project DP120102576, Australia.

PM and RB also acknowledge the support of Cancer Institute of New South Wales (http://www.cancerinstitute.org.au/), Big Data Big Impact Grant 13/DATA/1–03 “The integration of bioinformatics, chemoinformatics, and toxicogenomics methods: a new approach for the identification of combination tailored therapies and novel drug targets in breast cancer.”

Author Contributions

Conceived and designed the experiments: HHM RV CR PM. Performed the experiments: HHM RV CR IT. Analyzed the data: HHM RV CR IT. Contributed reagents/materials/analysis tools: HHM RV CR RB IT. Wrote the paper: HHM RV CR IT PM. Critical review of content: RB. First draft of manuscript: HHM.

References

  1. 1. Reis-Filho JS, Pusztai L. Gene expression profiling in breast cancer: classification, prognostication, and prediction. The Lancet. 2011;378(9805):1812–1823.
  2. 2. Portier BP, Gruver AM, Huba MA, Minca EC, Cheah AL, Wang Z, et al. From morphologic to molecular: established and emerging molecular diagnostics for breast carcinoma. N Biotechnol. 2012;29(6):665–81. pmid:22504737
  3. 3. van’t Veer LJ, Dai H, van de Vijver MJ, He YD, Hart AAM, Mao M, et al. Gene expression profiling predicts clinical outcome of breast cancer. Nature. 2002;415:530–536.
  4. 4. Prat A, Ellis MJ, Perou CM. Practical implications of gene-expression-based assays for breast oncologists. Nat Rev Clin Oncol. 2012;9(1):48–57.
  5. 5. Kelly CM, Bernard PS, Krishnamurthy S, Wang B, Ebbert MT, Bastien RR, et al. Agreement in risk prediction between the 21-gene recurrence score assay (Oncotype DX(R)) and the PAM50 breast cancer intrinsic Classifier in early-stage estrogen receptor-positive breast cancer. Oncologist. 2012;17(4):492–498. pmid:22418568
  6. 6. Dowsett M, Sestak I, Lopez-Knowles E, Sidhu K, Dunbier AK, Cowens JW, et al. Comparison of PAM50 risk of recurrence score with oncotype DX and IHC4 for predicting risk of distant recurrence after endocrine therapy. J Clin Oncol. 2013;31(22):2783–2790. pmid:23816962
  7. 7. Perou CM, Sørlie T, Eisen MB, van de Rijn M, Jeffrey SS, Rees CA, et al. Molecular portraits of human breast tumours. Nature. 2000;406(6797):747–752. pmid:10963602
  8. 8. Sørlie T, Perou CM, Tibshirani R, Aas T, Geisler S, Johnsen H, et al. Gene expression patterns of breast carcinomas distinguish tumor subclasses with clinical implications. Proc Natl Acad Sci U S A. 2001;98(19):10869–10874. pmid:11553815
  9. 9. Herschkowitz JI, Simin K, Weigman VJ, Mikaelian I, Usary J, Hu Z, et al. Identification of conserved gene expression features between murine mammary carcinoma models and human breast tumors. Genome Biol. 2007;8(5):R76. pmid:17493263
  10. 10. Prat A, Parker JS, Karginova O, Fan C, Livasy C, Herschkowitz JI, et al. Phenotypic and molecular characterization of the claudin-low intrinsic subtype of breast cancer. Breast Cancer Res. 2010;12(5):R68. pmid:20813035
  11. 11. Lehmann BD, Bauer JA, Chen X, Sanders ME, Chakravarthy AB, Shyr Y, et al. Identification of human triple-negative breast cancer subtypes and preclinical models for selection of targeted therapies. J Clin Invest. 2011;121(7):2750–2767. pmid:21633166
  12. 12. Nielsen TO, Parker JS, Leung S, Voduc D, Ebbert M, Vickery T, et al. A comparison of PAM50 intrinsic subtyping with immunohistochemistry and clinical prognostic factors in tamoxifen-treated estrogen receptor-positive breast cancer. Clin Cancer Res. 2010;16(21):5222–5232. pmid:20837693
  13. 13. Weigelt B, Baehner FL, Reis-Filho JS. The contribution of gene expression profiling to breast cancer classification, prognostication and prediction: a retrospective of the last decade. J Pathol. 2010;220(2):263–280. pmid:19927298
  14. 14. van de Vijver MJ, He YD, van’t Veer LJ, Dai H, Hart AAM, Voskuil DW, et al. A Gene-expression Signature as a Predictor of Survival in Breast Cancer. N Engl J Med. 2002;347(25):1999–2009. pmid:12490681
  15. 15. Hu Z, Fan C, Oh DS, Marron JS, He X, Qaqish BF, et al. The molecular portraits of breast tumors are conserved across microarray platforms. BMC Genomics. 2006;7(96).
  16. 16. Parker JS, Mullins M, Cheang MC, Leung S, Voduc D, Vickery T, et al. Supervised risk predictor of breast cancer based on intrinsic subtypes. J Clin Oncol. 2009;27(8):1160–1167. pmid:19204204
  17. 17. Bastien RR, Rodríguez-Lescure A, Ebbert MT, Prat A, Munárriz B, Rowe L, et al. PAM50 Breast Cancer Subtyping by RT-qPCR and Concordance with Standard Clinical Molecular Markers. BMC Med Genomics. 2012;5(44):1–12.
  18. 18. Sørlie T, Tibshirani R, Parker J, Hastie T, Marron JS, Nobel A, et al. Repeated observation of breast tumor subtypes in independent gene expression data sets. Proc Natl Acad Sci U S A. 2003;100(14):8418–8423. Available from: http://www.pnas.org/content/100/14/8418.abstract. pmid:12829800
  19. 19. Tibshirani R, Hastie T, Narasimhan B, Chu G. Diagnosis of multiple cancer types by shrunken centroids of gene expression. Proc Natl Acad Sci U S A. 2002;99(10):6567–6572.
  20. 20. Perou CM, Parker JS, Prat A, Ellis MJ, Bernard PS. Clinical implementation of the intrinsic subtypes of breast cancer. Lancet Oncol. 2010;11(8):718–719. pmid:20688274
  21. 21. Weigelt B, Mackay A, A’hern R, Natrajan R, Tan DS, Dowsett M, et al. Breast cancer molecular profiling with single sample predictors: a retrospective analysis. Lancet Oncol. 2010;11(4):339–349. pmid:20181526
  22. 22. Haibe-Kains B, Desmedt C, Loi S, Culhane AC, Bontempi G, Quackenbush J, et al. A three-gene model to robustly identify breast cancer molecular subtypes. J Natl Cancer Inst. 2012;104(4):311–325. pmid:22262870
  23. 23. Sotiriou C, Pusztai L. Gene-Expression Signatures in Breast Cancer. N Engl J Med. 2009;360:790–800. pmid:19228622
  24. 24. Weigelt B, Reis-Filho JS. Histological and molecular types of breast cancer: is there a unifying taxonomy? Nat Rev Clin Oncol. 2009;6(12):718–730. pmid:19942925
  25. 25. Colombo PE, Milanezi F, Weigelt B, Reis-Filho J. Microarrays in the 2010s: the contribution of microarray-based gene expression profiling to breast cancer classification, prognostication and prediction. Breast Cancer Res. 2010;13(212):1–15. Available from: http://dx.doi.org/10.1186/bcr2890.
  26. 26. Weigelt B, Pusztai L, Ashworth A, Reis-Filho JS. Challenges translating breast cancer gene signatures into the clinic. Nat Rev Clin Oncol. 2012;9(1):58–64.
  27. 27. Curtis C, Shah SP, Chin SF, Turashvili G, Rueda OM, Dunning MJ, et al. The genomic and transcriptomic architecture of 2,000 breast tumours reveals novel subgroups. Nature. 2012;486(7403):346–352. pmid:22522925
  28. 28. Ambs S. Prognostic Significance of Subtype Classification for Short- and Long-Term Survival in Breast Cancer: Survival Time Holds the Key. Plo Med. 2010;7:e1000281.
  29. 29. Marsden J, Budden D, Craig H, Moscato P. Language Individuation and Marker Words: Shakespeare and His Maxwell’s Demon. PLoS ONE. 2013;8(6):e66813. pmid:23826143
  30. 30. Ur-Rehman S, Gao Q, Mitsopoulos C, Zvelebil M. ROCK: a resource for integrative breast cancer data analysis. Breast Cancer Res Treat. 2013;139(3):907–921. pmid:23756628
  31. 31. Witten IH, Frank E. Data Mining: Practical Machine Learning Tools and Techniques. 2nd ed. San Francisco: Morgan Kaufmann; 2005.
  32. 32. Gómez Ravetti M, Moscato P. Identification of a 5-Protein Biomarker Molecular Signature for Predicting Alzheimer’s Disease. PLoS ONE. 2008 09;3(9):e3111. pmid:18769539
  33. 33. Liebetrau AM. Measures of association. 32. Sage; 1983.
  34. 34. Dunning MJ, Curtis C, Barbosa-Morais NL, Caldas C, Tavare S, Lynch AG. The importance of platform annotation in interpreting microarray data. Lancet Oncol. 2010;11(8):717. pmid:20688273
  35. 35. Fleiss JL. Measuring nominal scale agreement among many raters. Psychol Bull. 1971;76(5):378–382.
  36. 36. Fleiss JL, Levin B, Paik MC. 18. In: The Measurement of Interrater Agreement. John Wiley & Sons, Inc.; 2004. p. 598–626.
  37. 37. Landis JR, Koch GG. The measurement of observer agreement for categorical data. Biometrics. 1977;33(1):159–174. pmid:843571
  38. 38. Hubert L, Arabie P. Comparing partitions. Journal of Classification. 1985;2(1):193–218.
  39. 39. Vinh NX, Epps J, Bailey J. Information Theoretic Measures for Clusterings Comparison: Is a Correction for Chance Necessary? In: Proceedings of the 26th Annual International Conference on Machine Learning; 2009. p. 1073–1080.
  40. 40. Therneau TM. Modeling Survival Data: Extending the Cox Model. Springer Science & Business Media; 2000.
  41. 41. Kalbfleisch JD, Prentice RL. The Statistical Analysis of Failure Time Data. John Wiley & Sons; 2011.
  42. 42. Ciriello G, Sinha R, Hoadley K, Jacobsen A, Reva B, Perou C, et al. The molecular diversity of Luminal A breast tumors. Breast Cancer Res Treat. 2013;141(3):409–420. pmid:24096568
  43. 43. Hegyi K, Egervari K, Sandor Z, Mehes G. Aurora kinase B expression in breast carcinoma: cell kinetic and genetic aspects. Pathobiology. 2012;79(6):314–322. pmid:22688343
  44. 44. Ahn SG, Lee HM, Lee HW, Lee SA, Lee SR, Leem SH, et al. Prognostic Discrimination Using a 70-Gene Signature among Patients with Estrogen Receptor-Positive Breast Cancer and an Intermediate 21-Gene Recurrence Score. Int J Mol Sci. 2013;14(12):23685–23699. pmid:24304542
  45. 45. Fiskus W, Hembruff S, Rao R, Sharma P, Balusu R, Venkannagari S, et al. Co-treatment with vorinostat synergistically enhances activity of Aurora kinase inhibitor against human breast cancer cells. Breast Cancer Res Treat. 2012;135(2):433–444. pmid:22825030
  46. 46. Gully C, Zhang F, Chen J, Yeung J, Velazquez-Torres G, Wang E, et al. Antineoplastic effects of an Aurora B kinase inhibitor in breast cancer. Mol Cancer. 2010;9(1):42. pmid:20175926
  47. 47. Arbitrario J, Belmont B, Evanchik M, Flanagan WM, Fucini R, Hansen S, et al. SNS-314, a pan-Aurora kinase inhibitor, shows potent anti-tumor activity and dosing flexibility in vivo. Cancer Chem Pharmacol. 2010;65(4):707–717.
  48. 48. Hardwicke MA, Oleykowski CA, Plant R, Wang J, Liao Q, Moss K, et al. GSK1070916, a potent Aurora B/C kinase inhibitor with broad antitumor activity in tissue culture cells and human tumor xenograft models. Mol Cancer Ther. 2009;8(7):1808–1817. pmid:19567821
  49. 49. Ueki T, Nishidate T, Park JH, Lin ML, Shimo A, Hirata K, et al. Involvement of elevated expression of multiple cell-cycle regulator, DTL/RAMP (denticleless/RA-regulated nuclear matrix associated protein), in the growth of breast cancer cells. Oncogene. 2008;27(43):5672–5683. pmid:18542055
  50. 50. Kalous O, Conklin D, Desai A, Dering J, Goldstein J, Ginther C, et al. AMG 900, pan-Aurora kinase inhibitor, preferentially inhibits the proliferation of breast cancer cell lines with dysfunctional p53. Breast Cancer Res Treat. 2013;141(3):397–408. pmid:24091768
  51. 51. Bush TL, Payton M, Heller S, Chung G, Hanestad K, Rottman JB, et al. AMG 900, a Small-Molecule Inhibitor of Aurora Kinases, Potentiates the Activity of Microtubule-Targeting Agents in Human Metastatic Breast Cancer Models. Mol Cancer Ther. 2013;12(11):2356–2366. pmid:23990115
  52. 52. Romanelli A, Clark A, Assayag F, Chateau-Joubert S, Poupon MF, Servely JL, et al. Inhibiting Aurora Kinases Reduces Tumor Growth and Suppresses Tumor Recurrence after Chemotherapy in Patient-Derived Triple-Negative Breast Cancer Xenografts. Mol Cancer Ther. 2012;11(12):2693–2703. pmid:23012245
  53. 53. Sanchez-Bailon MP, Calcabrini A, Gomez-Dominguez D, Morte B, Martin-Forero E, Gomez-Lopez G, et al. Src kinases catalytic activity regulates proliferation, migration and invasiveness of MDA-MB-231 breast cancer cells. Cell Signal. 2012;24(6):1276–1286. Available from: http://www.ncbi.nlm.nih.gov/pubmed/22570868. pmid:22570868
  54. 54. Soncini C, Carpinelli P, Gianellini L, Fancelli D, Vianello P, Rusconi L, et al. PHA-680632, a novel Aurora kinase inhibitor with potent antitumoral activity. Clin Cancer Res. 2006;12(13):4080–4089. Available from: http://www.ncbi.nlm.nih.gov/pubmed/16818708. pmid:16818708
  55. 55. Mertsch S, Schmitz N, Jeibmann A, Geng JG, Paulus W, Senner V. Slit2 involvement in glioma cell migration is mediated by Robo1 receptor. J Neurooncol. 2008;87(1):1–7. pmid:17968499
  56. 56. Zhang W, Kawanishi M, Miyake K, Kagawa M, Kawai N, Murao K, et al. Association between YKL-40 and adult primary astrocytoma. Cancer. 2010;116(11):2688–2697. pmid:20499402
  57. 57. Carretero M, Ruiz-Torres M, Rodriguez-Corsino M, Barthelemy I, Losada A. Pds5B is required for cohesion establishment and Aurora B accumulation at centromeres. EMBO J. 2013;32(22):2938–2949. pmid:24141881
  58. 58. Nguyen MH, Koinuma J, Ueda K, Ito T, Tsuchiya E, Nakamura Y, et al. Phosphorylation and Activation of Cell Division Cycle Associated 5 by Mitogen-Activated Protein Kinase Play a Crucial Role in Human Lung Carcinogenesis. Cancer Res. 2010;70(13):5337–5347. pmid:20551060
  59. 59. Murphy KJ, Ter Horst JP, Cassidy AW, DeSouza IE, Morgunova M, Li C, et al. Temporal dysregulation of cortical gene expression in the isolation reared Wistar rat. J Neurochem. 2010;113(3):601–614. pmid:20096092
  60. 60. Inoda S, Hirohashi Y, Torigoe T, Nakatsugawa M, Kiriyama K, Nakazawa E, et al. Cep55/c10orf3, a tumor antigen derived from a centrosome residing protein in breast carcinoma. J Immunother. 2009;32(5):474–485. pmid:19609239
  61. 61. Colak D, Nofal A, Albakheet A, Nirmal M, Jeprel H, Eldali A, et al. Age-specific gene expression signatures for breast tumors and cross-species conserved potential cancer progression markers in young women. PLoS ONE. 2013;8(5):e63204. pmid:23704896
  62. 62. Martin KJ, Patrick DR, Bissell MJ, Fournier MV. Prognostic breast cancer signature identified from 3D culture model accurately predicts clinical outcome across independent datasets. PLoS ONE. 2008;3(8):e2994. pmid:18714348
  63. 63. Lin ML, Park JH, Nishidate T, Nakamura Y, Katagiri T. Involvement of maternal embryonic leucine zipper kinase (MELK) in mammary carcinogenesis through interaction with Bcl-G, a pro-apoptotic member of the Bcl-2 family. Breast Cancer Res. 2007;9(1):R17. pmid:17280616
  64. 64. Agnati LF, Genedani S, Leo G, Forni A, Woods AS, Filaferro M, et al. Abeta peptides as one of the crucial volume transmission signals in the trophic units and their interactions with homocysteine. Physiological implications and relevance for Alzheimer’s disease. J Neural Transm. 2007;114(1):21–31.
  65. 65. Canevari G, Re Depaolini S, Cucchi U, Bertrand JA, Casale E, Perrera C, et al. Structural insight into maternal embryonic leucine zipper kinase (MELK) conformation and inhibition toward structure-based drug design. Biochem. 2013;52(37):6380–6387.
  66. 66. Mahasenan KV, Li C. Novel inhibitor discovery through virtual screening against multiple protein conformations generated via ligand-directed modeling: a maternal embryonic leucine zipper kinase example. J Chem Inf Model. 2012;52(5):1345–1355. pmid:22540736
  67. 67. Hebbard LW, Maurer J, Miller A, Lesperance J, Hassell J, Oshima RG, et al. Maternal embryonic leucine zipper kinase is upregulated and required in mammary tumor-initiating cells in vivo. Cancer Res. 2010;70(21):8863–8873. pmid:20861186
  68. 68. Pickard MR, Green AR, Ellis IO, Caldas C, Hedge VL, Mourtada-Maarabouni M, et al. Dysregulated expression of Fau and MELK is associated with poor prognosis in breast cancer. Breast Cancer Res. 2009;11(4):R60. pmid:19671159
  69. 69. Warsow G, Struckmann S, Kerkhoff C, Reimer T, Engel N, Fuellen G. Differential network analysis applied to preoperative breast cancer chemotherapy response. PLoS ONE. 2013;8(12):e81784. pmid:24349128
  70. 70. Loussouarn D, Campion L, Leclair F, Campone M, Charbonnel C, Ricolleau G, et al. Validation of UBE2C protein as a prognostic marker in node-positive breast cancer. Br J Cancer. 2009;101(1):166–173. pmid:19513072
  71. 71. Rawat A, Gopal G, Selvaluxmy G, Rajkumar T. Inhibition of ubiquitin conjugating enzyme UBE2C reduces proliferation and sensitizes breast cancer cells to radiation, doxorubicin, tamoxifen and letrozole. Cell Oncol. 2013;36(6):459–467.
  72. 72. Psyrri A, Kalogeras KT, Kronenwett R, Wirtz RM, Batistatou A, Bournakis E, et al. Prognostic significance of UBE2C mRNA expression in high-risk early breast cancer. A Hellenic Cooperative Oncology Group (HeCOG) Study. Ann Oncol. 2012;23(6):1422–1427.
  73. 73. Taylor KJ, Sims AH, Liang L, Faratian D, Muir M, Walker G, et al. Dynamic changes in gene expression in vivo predict prognosis of tamoxifen-treated patients with breast cancer. Breast Cancer Res. 2010;12(3):R39. pmid:20569502
  74. 74. Parris TZ, Kovacs A, Aziz L, Hajizadeh S, Nemes S, Semaan M, et al. Additive effect of the AZGP1, PIP, S100A8 and UBE2C molecular biomarkers improves outcome prediction in breast carcinoma. Int J Cancer. 2014;134(7):1617–1629. pmid:24114735
  75. 75. Jäger D, Filonenko V, Gout I, Frosina D, Eastlake-Wade S, Castelli S, et al. NY-BR-1 is a Differentiation Antigen of the Mammary Gland. Appl Immunohistochem Mol Morphol. 2007;15(1):77–83. pmid:17536312
  76. 76. Woodard AH, Yu J, Dabbs DJ, Beriwal S, Florea AV, Elishaev E, et al. NY-BR-1 and PAX8 immunoreactivity in breast, gynecologic tract, and other CK7+ carcinomas: potential use for determining site of origin. Am J Clin Pathol. 2011;136(3):428–435. pmid:21846919
  77. 77. Varga Z, Theurillat JP, Filonenko V, Sasse B, Odermatt B, Jungbluth AA, et al. Preferential nuclear and cytoplasmic NY-BR-1 protein expression in primary breast cancer and lymph node metastases. Clin Cancer Res. 2006;12(9):2745–2751. pmid:16675566
  78. 78. Seil I, Frei C, Sültmann H, Knauer SK, Engels K, Jäger E, et al. The differentiation antigen NY-BR-1 is a potential target for antibody-based therapies in breast cancer. Int J Cancer. 2007;120(12):2635–2642. pmid:17330232
  79. 79. Balafoutas D, zur Hausen A, Mayer S, Hirschfeld M, Jaeger M, Denschlag D, et al. Cancer testis antigens and NY-BR-1 expression in primary breast cancer: prognostic and therapeutic implications. BMC Cancer. 2013;13:271. pmid:23731661
  80. 80. Giger O, Caduff R, O’Meara A, Diener PA, Knuth A, Jager D, et al. Frequent expression of the breast differentiation antigen NY-BR-1 in mammary and extramammary Paget’s disease. Pathol Int. 2010;60(11):726–734. pmid:20946522
  81. 81. Theurillat JP, Zurrer-Hardi U, Varga Z, Storz M, Probst-Hensch NM, Seifert B, et al. NY-BR-1 protein expression in breast carcinoma: a mammary gland differentiation antigen as target for cancer immunotherapy. Cancer Immunol Immunother. 2007;56(11):1723–1731. pmid:17410359
  82. 82. Theurillat JP, Zurrer-Hardi U, Varga Z, Barghorn A, Saller E, Frei C, et al. Distinct expression patterns of the immunogenic differentiation antigen NY-BR-1 in normal breast, testis and their malignant counterparts. Int J Cancer. 2008;122(7):1585–1591. pmid:18041742
  83. 83. Ikeda H, Taira N, Hara F, Fujita T, Yamamoto H, Soh J, et al. The estrogen receptor influences microtubule-associated protein tau (MAPT) expression and the selective estrogen receptor inhibitor fulvestrant downregulates MAPT and increases the sensitivity to taxane in breast cancer cells. Breast Cancer Res. 2010;12(3):R43. pmid:20579400
  84. 84. Valet F, de Cremoux P, Spyratos F, Servant N, Dujaric ME, Gentien D, et al. Challenging single- and multi-probesets gene expression signatures of pathological complete response to neoadjuvant chemotherapy in breast cancer: experience of the REMAGUS 02 phase II trial. Breast. 2013;22(6):1052–1059. pmid:24095610
  85. 85. Kotoula V, Kalogeras KT, Kouvatseas G, Televantou D, Kronenwett R, Wirtz RM, et al. Sample parameters affecting the clinical relevance of RNA biomarkers in translational breast cancer research. Virchows Arch. 2013;462(2):141–154. pmid:23262785
  86. 86. Spicakova T, O’Brien MM, Duran GE, Sweet-Cordero A, Sikic BI. Expression and silencing of the microtubule-associated protein Tau in breast cancer cells. Mol Cancer Ther. 2010;9(11):2970–2981. pmid:21062914
  87. 87. Tanaka S, Nohara T, Iwamoto M, Sumiyoshi K, Kimura K, Takahashi Y, et al. Tau expression and efficacy of paclitaxel treatment in metastatic breast cancer. Cancer Chemother Pharmacol. 2009;64(2):341–346. pmid:19039589
  88. 88. Fountzilas G, Kotoula V, Pectasides D, Kouvatseas G, Timotheadou E, Bobos M, et al. Ixabepilone administered weekly or every three weeks in HER2-negative metastatic breast cancer patients; a randomized non-comparative phase II trial. PLoS ONE. 2013;8(7):e69256. pmid:23935969
  89. 89. Mihaly Z, Kormos M, Lanczky A, Dank M, Budczies J, Szasz MA, et al. A meta-analysis of gene expression-based biomarkers predicting outcome after tamoxifen treatment in breast cancer. Breast Cancer Res Treat. 2013;140(2):219–232. pmid:23836010
  90. 90. Fletcher GC, Patel S, Tyson K, Adam PJ, Schenker M, Loader JA, et al. hAG-2 and hAG-3, human homologues of genes involved in differentiation, are associated with oestrogen receptor-positive breast tumours and interact with metastasis gene C4.4a and dystroglycan. Br J Cancer. 2003;88(4):579–585. pmid:12592373
  91. 91. Persson S, Rosenquist M, Knoblach B, Khosravi-Far R, Sommarin M, Michalak M. Diversity of the protein disulfide isomerase family: identification of breast tumor induced Hag2 and Hag3 as novel members of the protein family. Mol Phylogenet Evol. 2005;36(3):734–740. pmid:15935701
  92. 92. Wang K, Deng QT, Liao N, Zhang GC, Liu YH, Xu FP, et al. Tau expression correlated with breast cancer sensitivity to taxanes-based neoadjuvant chemotherapy. Tumour Biol. 2013;34(1):33–38. pmid:22976542
  93. 93. Kim SJ, Kang H, Chang HL, Jung YC, Sim H, Lee KS, et al. Promoter hypomethylation of the N-acetyltransferase 1 gene in breast cancer. Oncol Rep. 2008;19:663–668. pmid:18288399
  94. 94. Sim E, Lack N, Wang CJ, Long H, Westwood I, Fullam E, et al. Arylamine N-acetyltransferases: structural and functional implications of polymorphisms. Toxicology. 2008;254(3):170–183. pmid:18852012
  95. 95. Sim E, Walters K, Boukouvala S. Arylamine N-acetyltransferases: from structure to function. Drug Metab Rev. 2008;40(3):479–510. pmid:18642144
  96. 96. Kabbage M, Trimeche M, Nasr H, Hammann P, Kuhn L, Hamrita B, et al. Expression of the molecular chaperone aB-crystallin in infiltrating ductal breast carcinomas and the significance thereof: an immunohistochemical and proteomics-based strategy. Tumour Biol. 2012;33(6):2279–2288. pmid:22972503
  97. 97. Ruan Q, Han S, Jiang WG, Boulton ME, Chen ZJ, Law BK, et al. AlphaB-crystallin, an effector of unfolded protein response, confers anti-VEGF resistance to breast cancer via maintenance of intracrine VEGF in endothelial cells. Mol Cancer Res. 2011;9(12):1632–1643. pmid:21984182
  98. 98. Campbell-Lloyd AJ, Mundy J, Deva R, Lampe G, Hawley C, Boyle G, et al. Is alpha-B crystallin an independent marker for prognosis in lung cancer? Heart Lung Circ. 2013;22(9):759–766. pmid:23582651
  99. 99. Cortesi L, Barchetti A, De Matteis E, Rossi E, Della Casa L, Marcheselli L, et al. Identification of protein clusters predictive of response to chemotherapy in breast cancer patients. J Proteome Res. 2009;8(11):4916–4933. pmid:19739612
  100. 100. van de Rijn M, Perou CM, Tibshirani R, Haas P, Kallioniemi O, Kononen J, et al. Expression of Cytokeratins 17 and 5 Identifies a Group of Breast Carcinomas with Poor Clinical Outcome. Am J Pathol. 2002;161(6):1991–1996. Available from: http://www.ncbi.nlm.nih.gov/pmc/articles/PMC1850928/. pmid:12466114
  101. 101. Gorski JJ, James CR, Quinn JE, Stewart GE, Staunton KC, Buckley NE, et al. BRCA1 transcriptionally regulates genes associated with the basal-like phenotype in breast cancer. Breast Cancer Res Treat. 2010;122(3):721–731. pmid:19882246
  102. 102. Liu Q, Li JG, Zheng XY, Jin F, Dong HT. Expression of CD133, PAX2, ESA, and GPR30 in invasive ductal breast carcinomas. Chin Med J. 2009;122(22):2763–2769. pmid:19951611
  103. 103. Coradini D, Fornili M, Ambrogi F, Boracchi P, Biganzoli E. TP53 mutation, epithelial-mesenchymal transition, and stemlike features in breast cancer subtypes. J Biomed Biotechnol. 2012;2012:254085. pmid:22899882
  104. 104. Bertheau P, Turpin E, Rickman DS, Espie M, de Reynies A, Feugeas J, et al. Exquisite Sensitivity of TP53 Mutant and Basal Breast Cancers to a Dose-Dense Epirubicin-Cyclophosphamide Regimen. PLoS Med. 2007;4(3):585–594.
  105. 105. Liu N, Yu Q, Liu TJ, Gebreamlak EP, Wang SL, Zhang RJ, et al. P-cadherin expression and basal-like subtype in breast cancers. Med Oncol. 2012;29(4):2606–2612. pmid:22467076
  106. 106. Nadal R, Ortega FG, Salido M, Lorente JA, Rodriguez-Rivera M, Delgado-Rodriguez M, et al. CD133 expression in circulating tumor cells from breast cancer patients: potential role in resistance to chemotherapy. Int J Cancer. 2013;133(10):2398–2407. pmid:23661576
  107. 107. Qian Y, Shen L, Cheng L, Wu Z, Yao H. B7–H4 expression in various tumors determined using a novel developed monoclonal antibody. Clin Exp Med. 2011;11(3):163–170. pmid:21190056
  108. 108. Suh WK, Wang S, Duncan GS, Miyazaki Y, Cates E, Walker T, et al. Generation and characterization of B7–H4/B7S1/B7x-deficient mice. Mol Cell Biol. 2006;26(17):6403–6411. pmid:16914726
  109. 109. Salceda S, Tang T, Kmet M, Munteanu A, Ghosh M, Macina R, et al. The immunomodulatory protein B7–H4 is overexpressed in breast and ovarian cancers and promotes epithelial cell transformation. Exp Cell Res. 2005;306(1):128–141. pmid:15878339
  110. 110. Heinonen H, Nieminen A, Saarela M, Kallioniemi A, Klefstrom J, Hautaniemi S, et al. Deciphering downstream gene targets of PI3K/mTOR/p70S6K pathway in breast cancer. BMC Genomics. 2008;9:348. pmid:18652687
  111. 111. Tringler B, Zhuo S, Pilkington G, Torkko KC, Singh M, Lucia MS, et al. B7–h4 is highly expressed in ductal and lobular breast cancer. Clin Cancer Res. 2005;11(5):1842–1848. pmid:15756008
  112. 112. Zhang W, Peng G, Lin SY, Zhang P. DNA damage response is suppressed by the high cyclin-dependent kinase 1 activity in mitotic mammalian cells. J Biol Chem. 2011;286(41):35899–35905. pmid:21878640
  113. 113. Kim SJ, Nakayama S, Miyoshi Y, Taguchi T, Tamaki Y, Matsushima T, et al. Determination of the specific activity of CDK1 and CDK2 as a novel prognostic indicator for early breast cancer. Ann Oncol. 2008;19(1):68–72. pmid:17956886
  114. 114. Kim SJ, Nakayama S, Shimazu K, Tamaki Y, Akazawa K, Tsukamoto F, et al. Recurrence risk score based on the specific activity of CDK1 and CDK2 predicts response to neoadjuvant paclitaxel followed by 5-fluorouracil, epirubicin and cyclophosphamide in breast cancers. Ann Oncol. 2012;23(4):891–897. pmid:21821547
  115. 115. Torikoshi Y, Gohda K, Davis ML, Symmans WF, Pusztai L, Kazansky A, et al. Novel functional assay for spindle-assembly checkpoint by cyclin-dependent kinase activity to predict taxane chemosensitivity in breast tumor patient. J Cancer. 2013;4(9):697–702. pmid:24312139
  116. 116. Xia Q, Cai Y, Peng R, Wu G, Shi Y, Jiang W. The CDK1 inhibitor RO3306 improves the response of BRCA-pro fi cient breast cancer cells to PARP inhibition. Int J Oncol. 2014;44(3):735–744. pmid:24378347
  117. 117. Yu YN, Yip GW, Tan PH, Thike AA, Matsumoto K, Tsujimoto M, et al. Y-box binding protein 1 is up-regulated in proliferative breast cancer and its inhibition deregulates the cell cycle. Int J Oncol. 2010;37(2):483–492. pmid:20596676
  118. 118. Kim C, Tang G, Pogue-Geile KL, Costantino JP, Baehner FL, Baker J, et al. Estrogen receptor (ESR1) mRNA expression and benefit from tamoxifen in the treatment and prevention of estrogen receptor-positive breast cancer. J Clin Oncol. 2011;29(31):4160–4167. pmid:21947828
  119. 119. Stossi F, Madak-Erdogan Z, Katzenellenbogen BS. Macrophage-elicited loss of estrogen receptor-alpha in breast cancer cells via involvement of MAPK and c-Jun at the ESR1 genomic locus. Oncogene. 2012;31(14):1825–1834. pmid:21860415
  120. 120. Aguilar H, Sole X, Bonifaci N, Serra-Musach J, Islam A, Lopez-Bigas N, et al. Biological reprogramming in acquired resistance to endocrine therapy of breast cancer. Oncogene. 2010;29(45):6071–6083. pmid:20711236
  121. 121. Dunbier AK, Anderson H, Ghazoui Z, Lopez-Knowles E, Pancholi S, Ribas R, et al. ESR1 is co-expressed with closely adjacent uncharacterised genes spanning a breast cancer susceptibility locus at 6q25.1. PLoS Genet. 2011;7(4):e1001382. pmid:21552322
  122. 122. Vargas AC, McCart Reed AE, Waddell N, Lane A, Reid LE, Smart CE, et al. Gene expression profiling of tumour epithelial and stromal compartments during breast cancer progression. Breast Cancer Res Treat. 2012;135(1):153–165. pmid:22718308
  123. 123. Gauger KJ, Chenausky KL, Murray ME, Schneider SS. SFRP1 reduction results in an increased sensitivity to TGF-beta signaling. BMC Cancer. 2011;11:59. pmid:21303533
  124. 124. Mukherjee N, Bhattacharya N, Alam N, Roy A, Roychoudhury S, Panda CK. Subtype-specific alterations of the Wnt signaling pathway in breast cancer: clinical and prognostic significance. Cancer Sci. 2012;103(2):210–220. pmid:22026417
  125. 125. Gostner JM, Fong D, Wrulich OA, Lehne F, Zitt M, Hermann M, et al. Effects of EpCAM overexpression on human breast cancer cell lines. BMC Cancer. 2011;11:45. pmid:21281469
  126. 126. Matsuda Y, Schlange T, Oakeley EJ, Boulay A, Hynes NE. WNT signaling enhances breast cancer cell motility and blockade of the WNT pathway by sFRP1 suppresses MDA-MB-231 xenograft growth. Breast Cancer Res. 2009;11(3):R32. pmid:19473496
  127. 127. Suzuki H, Toyota M, Carraway H, Gabrielson E, Ohmura T, Fujikane T, et al. Frequent epigenetic inactivation of Wnt antagonist genes in breast cancer. Br J Cancer. 2008;98(6):1147–1156. pmid:18283316
  128. 128. Dahl E, Wiesmann F, Woenckhaus M, Stoehr R, Wild PJ, Veeck J, et al. Frequent loss of SFRP1 expression in multiple human solid tumours: association with aberrant promoter methylation in renal cell carcinoma. Oncogene. 2007;26(38):5680–5691. pmid:17353908
  129. 129. Shulewitz M, Soloviev I, Wu T, Koeppen H, Polakis P, Sakanaka C. Repressor roles for TCF-4 and Sfrp1 in Wnt signaling in breast cancer. Oncogene. 2006;25(31):4361–4369. pmid:16532032
  130. 130. Ugolini F, Charafe-Jauffret E, Bardou VJ, Geneix J, Adelaide J, Labat-Moleur F, et al. WNT pathway and mammary carcinogenesis: loss of expression of candidate tumor suppressor gene SFRP1 in most invasive carcinomas except of the medullary type. Oncogene. 2001;20(41):5810–5817. pmid:11593386
  131. 131. Dumont N, Crawford YG, Sigaroudinia M, Nagrani SS, Wilson MB, Buehring GC, et al. Human mammary cancer progression model recapitulates methylation events associated with breast premalignancy. Breast Cancer Res. 2009;11(6):R87. pmid:19995452
  132. 132. Park SY, Kwon HJ, Choi Y, Lee HE, Kim SW, Kim JH, et al. Distinct patterns of promoter CpG island methylation of breast cancer subtypes are associated with stem cell phenotypes. Mod Pathol. 2012;25(2):185–196. pmid:22037257
  133. 133. Cooper SJ, von Roemeling CA, Kang KH, Marlow LA, Grebe SK, Menefee ME, et al. Reexpression of tumor suppressor, sFRP1, leads to antitumor synergy of combined HDAC and methyltransferase inhibitors in chemoresistant cancers. Mol Cancer Ther. 2012;11(10):2105–2115. pmid:22826467
  134. 134. Browne EP, Punska EC, Lenington S, Otis CN, Anderton DL, Arcaro KF. Increased promoter methylation in exfoliated breast epithelial cells in women with a previous breast biopsy. Epigenetics. 2011;6(12):1425–1435. pmid:22139572
  135. 135. Yang ZQ, Liu G, Bollig-Fischer A, Haddad R, Tarca AL, Ethier SP. Methylation-associated silencing of SFRP1 with an 8p11–12 amplification inhibits canonical and non-canonical WNT pathways in breast cancers. Int J Cancer. 2009;125(7):1613–1621. pmid:19569235
  136. 136. Martin-Manso G, Calzada MJ, Chuman Y, Sipes JM, Xavier CP, Wolf V, et al. sFRP-1 binds via its netrin-related motif to the N-module of thrombospondin-1 and blocks thrombospondin-1 stimulation of MDA-MB-231 breast carcinoma cell adhesion and migration. Arch Biochem Biophys. 2011;509(2):147–156. pmid:21402050
  137. 137. Veeck J, Niederacher D, An H, Klopocki E, Wiesmann F, Betz B, et al. Aberrant methylation of the Wnt antagonist SFRP1 in breast cancer is associated with unfavourable prognosis. Oncogene. 2006;25(24):3479–3488. pmid:16449975
  138. 138. Gauger KJ, Schneider SS. Tumour supressor secreted frizzled related protein 1 regulates p53-mediated apoptosis. Cell Biol Int. 2014;38(1):124–130. pmid:24038862
  139. 139. Klopocki E, Kristiansen G, Wild PJ, Klaman I, Castanos-Velez E, Singer G, et al. Loss of SFRP1 is associated with breast cancer progression and poor prognosis in early stage tumors. Int J Oncol. 2004;25(3):641–649. pmid:15289865
  140. 140. Sano H, Wada S, Eguchi H, Osaki A, Saeki T, Nishiyama M. Quantitative prediction of tumor response to neoadjuvant chemotherapy in breast cancer: novel marker genes and prediction model using the expression levels. Breast Cancer. 2012;19(1):37–45. pmid:21437666
  141. 141. Cimino D, Fuso L, Sfiligoi C, Biglia N, Ponzone R, Maggiorotto F, et al. Identification of new genes associated with breast cancer progression by gene expression analysis of predefined sets of neoplastic tissues. Int J Cancer. 2008;123(6):1327–1338. pmid:18561318
  142. 142. Yamamura J, Miyoshi Y, Tamaki Y, Taguchi T, Iwao K, Monden M, et al. mRNA expression level of estrogen-inducible gene, alpha 1-antichymotrypsin, is a predictor of early tumor recurrence in patients with invasive breast cancers. Cancer Sci. 2004;95(11):887–892. pmid:15546506
  143. 143. Miller WR, Larionov A. Changes in expression of oestrogen regulated and proliferation genes with neoadjuvant treatment highlight heterogeneity of clinical resistance to the aromatase inhibitor, letrozole. Breast Cancer Res. 2010;12(4):R52. pmid:20646288
  144. 144. Moy I, Todorovic V, Dubash AD, Coon JS, Parker JB, Buranapramest M, et al. Estrogen-dependent sushi domain containing 3 regulates cytoskeleton organization and migration in breast cancer cells. Oncogene. 2014;.
  145. 145. Sasaki Y, Koyama R, Maruyama R, Hirano T, Tamura M, Sugisaka J, et al. CLCA2, a target of the p53 family, negatively regulates cancer cell migration and invasion. Cancer Biol Ther. 2012;13(14):1512–1521. pmid:22990203
  146. 146. Walia V, Ding M, Kumar S, Nie D, Premkumar LS, Elble RC. hCLCA2 Is a p53-Inducible Inhibitor of Breast Cancer Cell Proliferation. Cancer Res. 2009;69(16):6624–6632. pmid:19654313
  147. 147. Walia V, Yu Y, Cao D, Sun M, McLean JR, Hollier BG, et al. Loss of breast epithelial marker hCLCA2 promotes epithelial-to-mesenchymal transition and indicates higher risk of metastasis. Oncogene. 2012;31(17):2237–2246. pmid:21909135
  148. 148. Li X, Cowell JK, Sossey-Alaoui K. CLCA2 tumour suppressor gene in 1p31 is epigenetically regulated in breast cancer. Oncogene. 2004;23(7):1474–1480. pmid:14973555
  149. 149. Waluk DP, Schultz N, Hunt MC. Identification of glycine N-acyltransferase-like 2 (GLYATL2) as a transferase that produces N-acyl glycines in humans. FASEB J. 2010;24(8):2795–2803. pmid:20305126
  150. 150. Waluk DP, Sucharski F, Sipos L, Silberring J, Hunt MC. Reversible lysine acetylation regulates activity of human glycine N-acyltransferase-like 2 (hGLYATL2): implications for production of glycine-conjugated signaling molecules. J Biol Chem. 2012;287(20):16158–16167. pmid:22408254
  151. 151. Liu L, Liu Z, Qu S, Zheng Z, Liu Y, Xie X, et al. Small breast epithelial mucin tumor tissue expression is associated with increased risk of recurrence and death in triple-negative breast cancer patients. Diagn Pathol. 2013;8:71. pmid:23635316
  152. 152. Liu ZZ, Xie XD, Qu SX, Zheng ZD, Wang YK. Small breast epithelial mucin (SBEM) has the potential to be a marker for predicting hematogenous micrometastasis and response to neoadjuvant chemotherapy in breast cancer. Clin Experim Metast. 2010;27(4):251–259.
  153. 153. Valladares-Ayerbes M, Iglesias-Diaz P, Diaz-Prado S, Ayude D, Medina V, Haz M, et al. Diagnostic accuracy of small breast epithelial mucin mRNA as a marker for bone marrow micrometastasis in breast cancer: a pilot study. J Cancer Res Clin Oncol. 2009;135(9):1185–1195. pmid:19221791
  154. 154. Weigelt B, Verduijn P, Bosma AJ, Rutgers EJ, Peterse HL, van’t Veer LJ. Detection of metastases in sentinel lymph nodes of breast cancer patients by multiple mRNA markers. Br J Cancer. 2004;90(8):1531–1537. pmid:15083181
  155. 155. Skliris GP, Hube F, Gheorghiu I, Mutawe MM, Penner C, Watson PH, et al. Expression of small breast epithelial mucin (SBEM) protein in tissue microarrays (TMAs) of primary invasive breast cancers. Histopathology. 2008;52(3):355–369. pmid:18269587
  156. 156. Miksicek RJ, Myal Y, Watson PH, Walker C, Murphy LC, Leygue E. Identification of a Novel Breast- and Salivary Gland-specific, Mucin-like Gene Strongly Expressed in Normal and Tumor Human Mammary Epithelium. Cancer Res. 2002;62:2736–2740. pmid:12019145
  157. 157. Lopez FJ, Cuadros M, Cano C, Concha A, Blanco A. Biomedical application of fuzzy association rules for identifying breast cancer biomarkers. Med Biol Eng Comput. 2012;50(9):981–990. pmid:22622817
  158. 158. Zvelebil M, Oliemuller E, Gao Q, Wansbury O, Mackay A, Kendrick H, et al. Embryonic mammary signature subsets are activated in Brca1-/- and basal-like breast cancers. Breast Cancer Res. 2013;15(2):R25. pmid:23506684
  159. 159. Bolanz KA, Hediger MA, Landowski CP. The role of TRPV6 in breast carcinogenesis. Mol Cancer Ther. 2008;7(2):271–279. pmid:18245667
  160. 160. Bowen CV, DeBay D, Ewart HS, Gallant P, Gormley S, Ilenchuk TT, et al. In vivo detection of human TRPV6-rich tumors with anti-cancer peptides derived from soricidin. PLoS ONE. 2013;8(3):e58866. pmid:23554944
  161. 161. Peters AA, Simpson PT, Bassett JJ, Lee JM, Da Silva L, Reid LE, et al. Calcium channel TRPV6 as a potential therapeutic target in estrogen receptor-negative breast cancer. Mol Cancer Ther. 2012;11(10):2158–2168. pmid:22807578
  162. 162. Kim SY, Yang D, Myeong J, Ha K, Kim SH, Park EJ, et al. Regulation of calcium influx and signaling pathway in cancer cells via TRPV6-Numb1 interaction. Cell Calcium. 2013;53(2):102–111. pmid:23140583
  163. 163. Landowski CP, Bolanz KA, Suzuki Y, Hediger MA. Chemical inhibitors of the calcium entry channel TRPV6. Pharm Res. 2011;28(2):322–330. pmid:21057859
  164. 164. Dhennin-Duthille I, Gautier M, Faouzi M, Guilbert A, Brevet M, Vaudry D, et al. High expression of transient receptor potential channels in human breast cancer epithelial cells and tissues: correlation with pathological parameters. Cell Physiol Biochem. 2011;28(5):813–822. pmid:22178934
  165. 165. Xing P, Li Jg, Jin F, Zhao Tt, Liu Q, Dong Ht, et al. Clinical and Biological Significance of Hepsin Overexpression in Breast Cancer. J Investig Med. 2011;59(5):803–810. pmid:21383634
  166. 166. Delassus GS, Cho H, Park J, Eliceiri GL. New pathway links from cancer-progression determinants to gene expression of matrix metalloproteinases in breast cancer cells. J Cell Physiol. 2008;217(3):739–744. pmid:18651563
  167. 167. Delassus GS, Cho H, Eliceiri GL. New signaling pathways from cancer progression modulators to mRNA expression of matrix metalloproteinases in breast cancer cells. J Cell Physiol. 2011;226(12):3378–3384. pmid:21344390
  168. 168. Kwon YJ, Hurst DR, Steg AD, Yuan K, Vaidya KS, Welch DR, et al. Gli1 enhances migration and invasion via up-regulation of MMP-11 and promotes metastasis in ERalpha negative breast cancer cell lines. Clin Experim Metast. 2011;28(5):437–449.
  169. 169. Min KW, Kim DH, Do SI, Pyo JS, Kim K, Chae S, et al. Diagnostic and Prognostic Relevance of MMP-11 Expression in the Stromal Fibroblast-Like Cells Adjacent to Invasive Ductal Carcinoma of the Breast. Ann Surg Oncol. 2013;20(3):433–442.
  170. 170. Kasper G, Reule M, Tschirschmann M, Dankert N, Stout-Weider K, Lauster R, et al. Stromelysin-3 over-expression enhances tumourigenesis in MCF-7 and MDA-MB-231 breast cancer cell lines: involvement of the IGF-1 signalling pathway. BMC Cancer. 2007;7:12. pmid:17233884
  171. 171. Cheng CW, Yu JC, Wang HW, Huang CS, Shieh JC, Fu YP, et al. The clinical implications of MMP-11 and CK-20 expression in human breast cancer. Clin Chim Acta. 2010;411(3–4):234–241. pmid:19914229
  172. 172. Eiseler T, Doppler H, Yan IK, Goodison S, Storz P. Protein kinase D1 regulates matrix metalloproteinase expression and inhibits breast cancer cell invasion. Breast Cancer Res. 2009;11(1):R13. pmid:19243594
  173. 173. Garcia MF, Gonzalez-Reyes S, Gonzalez LO, Junquera S, Berdize N, Del Casar JM, et al. Comparative study of the expression of metalloproteases and their inhibitors in different localizations within primary tumours and in metastatic lymph nodes of breast cancer. Int J Exp Pathol. 2010;91(4):324–334. pmid:20412339
  174. 174. Eiro N, Fernandez-Garcia B, Gonzalez LO, Vizoso FJ. Cytokines related to MMP-11 expression by inflammatory cells and breast cancer metastasis. Oncoimmunology. 2013;2(5):e24010. pmid:23762791
  175. 175. Tan J, Buache E, Alpy F, Daguenet E, Tomasetto CL, Ren GS, et al. Stromal matrix metalloproteinase-11 is involved in the mammary gland postnatal development. Oncogene. 2013;.
  176. 176. Hegedus L, Cho H, Xie X, Eliceiri GL. Additional MDA-MB-231 breast cancer cell matrix metalloproteinases promote invasiveness. J Cell Physiol. 2008;216(2):480–485. pmid:18286480
  177. 177. Ueki T, Park JH, Nishidate T, Kijima K, Hirata K, Nakamura Y, et al. Ubiquitination and downregulation of BRCA1 by ubiquitin-conjugating enzyme E2T overexpression in human breast cancer cells. Cancer Res. 2009;69(22):8752–8760. pmid:19887602
  178. 178. Li Y, Wu J, Zhang W, Zhang N, Guo H. Identification of serum CCL15 in hepatocellular carcinoma. Br J Cancer. 2013;108(1):99–106. pmid:23321514
  179. 179. Itatani Y, Kawada K, Fujishita T, Kakizaki F, Hirai H, Matsumoto T, et al. Loss of SMAD4 From Colorectal Cancer Cells Promotes CCL15 Expression to Recruit CCR1+ Myeloid Cells and Facilitate Liver Metastasis. Gastroenterology. 2013;145(5):1064–1075.e11. pmid:23891973
  180. 180. Ellsworth RE, Seebach J, Field LA, Heckman C, Kane J, Hooke JA, et al. A gene expression signature that defines breast cancer metastases. Clin Experim Metast. 2009;26(3):205–213.
  181. 181. Kim H, Watkinson J, Varadan V, Anastassiou D. Multi-cancer computational analysis reveals invasion-associated variant of desmoplastic reaction involving INHBA, THBS2 and COL11A1. BMC Med Genomics. 2010;3:51. pmid:21047417
  182. 182. van Zalen S, Nijenhuis M, Jonkman MF, Pas HH. Two major 5’-untranslated regions for type XVII collagen mRNA. J Dermatol Sci. 2006;43(1):11–19. pmid:16580182
  183. 183. Wang J, He Q, Shao YG, Ji M. Chemokines fluctuate in the progression of primary breast cancer. Eur Rev Med Pharmacol Sci. 2013;17:596–608. pmid:23543442
  184. 184. Bandyopadhyay S, Zhan R, Chaudhuri A, Watabe M, Pai SK, Hirota S, et al. Interaction of KAI1 on tumor cells with DARC on vascular endothelium leads to metastasis suppression. Nat Med. 2006;12(8):933–938. pmid:16862154
  185. 185. Zeng XH, Ou ZL, Yu KD, Feng LY, Yin WJ, Li J, et al. Coexpression of atypical chemokine binders (ACBs) in breast cancer predicts better outcomes. Breast Cancer Res Treat. 2011;125(3):715–727. pmid:20369284
  186. 186. Liu XF, Li LF, Ou ZL, Shen R, Shao ZM. Correlation between Duffy blood group phenotype and breast cancer incidence. BMC Cancer. 2012;12:374. pmid:22928984
  187. 187. Wu J, Liu S, Liu G, Dombkowski A, Abrams J, Martin-Trevino R, et al. Identification and functional analysis of 9p24 amplified genes in human breast cancer. Oncogene. 2012;31(3):333–341. pmid:21666724
  188. 188. Ni M, Chen Y, Lim E, Wimberly H, Bailey ST, Imai Y, et al. Targeting androgen receptor in estrogen receptor-negative breast cancer. Cancer Cell. 2011;20(1):119–131. pmid:21741601
  189. 189. Augello MA, Hickey TE, Knudsen KE. FOXA1: master of steroid receptor function in cancer. EMBO J. 2011;30(19):3885–3894. pmid:21934649
  190. 190. Robinson JL, Macarthur S, Ross-Innes CS, Tilley WD, Neal DE, Mills IG, et al. Androgen receptor driven transcription in molecular apocrine breast cancer is mediated by FoxA1. EMBO J. 2011;30(15):3019–3027. pmid:21701558
  191. 191. Bernardo GM, Keri RA. FOXA1: a transcription factor with parallel functions in development and cancer. Biosci Rep. 2012;32(2):113–130. pmid:22115363
  192. 192. Fu X, Huang C, Schiff R. More on FOX News: FOXA1 on the horizon of estrogen receptor function and endocrine response. Breast Cancer Res. 2011;13(2):307. pmid:21575280
  193. 193. Meyer KB, Carroll JS. FOXA1 and breast cancer risk. Nat Genet. 2012;44(11):1176–1177. pmid:23104063
  194. 194. Cowper-Sal lari R, Zhang X, Wright JB, Bailey SD, Cole MD, Eeckhoute J, et al. Breast cancer risk-associated SNPs modulate the affinity of chromatin for FOXA1 and alter gene expression. Nat Genet. 2012;44(11):1191–1198. pmid:23001124
  195. 195. Katika MR, Hurtado A. A functional link between FOXA1 and breast cancer SNPs. Breast Cancer Res. 2013;15(1):303. pmid:23427833
  196. 196. Badve S, Turbin D, Thorat MA, Morimiya A, Nielsen TO, Perou CM, et al. FOXA1 expression in breast cancer–correlation with luminal subtype A and survival. Clin Cancer Res. 2007;13(15 Pt 1):4415–4421. pmid:17671124
  197. 197. Mehta RJ, Jain RK, Leung S, Choo J, Nielsen T, Huntsman D, et al. FOXA1 is an independent prognostic marker for ER-positive breast cancer. Breast Cancer Res Treat. 2012;131(3):881–890. pmid:21503684
  198. 198. Habashy HO, Powe DG, Rakha EA, Ball G, Paish C, Gee J, et al. Forkhead-box A1 (FOXA1) expression in breast cancer and its prognostic significance. Europ J Cancer. 2008;44(11):1541–1551.
  199. 199. Albergaria A, Paredes J, Sousa B, Milanezi F, Carneiro V, Bastos J, et al. Expression of FOXA1 and GATA-3 in breast cancer: the prognostic significance in hormone receptor-negative tumours. Breast Cancer Res. 2009;11(3):R40. pmid:19549328
  200. 200. Hurtado A, Holmes KA, Ross-Innes CS, Schmidt D, Carroll JS. FOXA1 is a key determinant of estrogen receptor function and endocrine response. Nat Genet. 2011;43(1):27–33. pmid:21151129
  201. 201. Naderi A, Meyer M, Dowhan DH. Cross-regulation between FOXA1 and ErbB2 signaling in estrogen receptor-negative breast cancer. Neoplasia. 2012;14(4):283–296. pmid:22577344
  202. 202. Kong SL, Li G, Loh SL, Sung WK, Liu ET. Cellular reprogramming by the conjoint action of ERalpha, FOXA1, and GATA3 to a ligand-inducible growth state. Mol Syst Biol. 2011;7:526. pmid:21878914
  203. 203. Magnani L, Lupien M. Chromatin and epigenetic determinants of estrogen receptor alpha (ESR1) signaling. Mol Cell Endocrinol. 2014;382(1):633–641. pmid:23684889
  204. 204. Lasa A, Garcia A, Alonso C, Millet P, Cornet M, Ramon y Cajal T, et al. Molecular detection of peripheral blood breast cancer mRNA transcripts as a surrogate biomarker for circulating tumor cells. PLoS ONE. 2013;8(9):e74079. pmid:24058517
  205. 205. Lacroix M. Significance, detection and markers of disseminated breast cancer cells. Endocr Relat Cancer. 2006;13:1033–1067. pmid:17158753
  206. 206. Ahmed AR, Griffiths AB, Tilby MT, Westley BR, May FE. TFF3 is a normal breast epithelial protein and is associated with differentiated phenotype in early breast cancer but predisposes to invasion and metastasis in advanced disease. Am J Pathol. 2012;180(3):904–916. pmid:22341453
  207. 207. Fenne IS, Helland T, Flageng MH, Dankel SN, Mellgren G, Sagen JV. Downregulation of steroid receptor coactivator-2 modulates estrogen-responsive genes and stimulates proliferation of mcf-7 breast cancer cells. PLoS ONE. 2013;8(7):e70096. pmid:23936147
  208. 208. Roll JD, Rivenbark AG, Sandhu R, Parker JS, Jones WD, Carey LA, et al. Dysregulation of the epigenome in triple-negative breast cancers: basal-like and claudin-low breast cancers express aberrant DNA hypermethylation. Exp Mol Pathol. 2013;95(3):276–287. pmid:24045095
  209. 209. Sandhu R, Rivenbark AG, Mackler RM, Livasy CA, Coleman WB. Dysregulation of microRNA expression drives aberrant DNA hypermethylation in basal-like breast cancer. Int J Oncol. 2014;44(2):563–572. pmid:24297604
  210. 210. Wang J, Ray PS, Sim MS, Zhou XZ, Lu KP, Lee AV, et al. FOXC1 regulates the functions of human basal-like breast cancer cells by activating NF-kappaB signaling. Oncogene. 2012;31(45):4798–4802. pmid:22249250
  211. 211. Sizemore ST, Keri RA. The forkhead box transcription factor FOXC1 promotes breast cancer invasion by inducing matrix metalloprotease 7 (MMP7) expression. J Biol Chem. 2012;287(29):24631–24640. pmid:22645147
  212. 212. Tkocz D, Crawford NT, Buckley NE, Berry FB, Kennedy RD, Gorski JJ, et al. BRCA1 and GATA3 corepress FOXC1 to inhibit the pathogenesis of basal-like breast cancers. Oncogene. 2012;31(32):3667–3678. pmid:22120723
  213. 213. Andres SA, Brock GN, Wittliff JL. Interrogating differences in expression of targeted gene sets to predict breast cancer outcome. BMC Cancer. 2013;13(326):1–18.
  214. 214. Symmans WF, Fiterman DJ, Anderson SK, Ayers M, Rouzier R, Dunmire V, et al. A single-gene biomarker identifies breast cancers associated with immature cell type and short duration of prior breastfeeding. Endocr Relat Cancer. 2005;12(4):1059–1069. pmid:16322343
  215. 215. Bell JL, Wachter K, Muhleck B, Pazaitis N, Kohn M, Lederer M, et al. Insulin-like growth factor 2 mRNA-binding proteins (IGF2BPs): post-transcriptional drivers of cancer progression? Cell Mol Life Sci. 2013;70(15):2657–2675. pmid:23069990
  216. 216. Fadare O, Liang SX, Crispens MA, Jones HW, Khabele D, Gwin K, et al. Expression of the oncofetal protein IGF2BP3 in endometrial clear cell carcinoma: assessment of frequency and significance. Hum Pathol. 2013;44(8):1508–1515. pmid:23465280
  217. 217. Lochhead P, Imamura Y, Morikawa T, Kuchiba A, Yamauchi M, Liao X, et al. Insulin-like growth factor 2 messenger RNA binding protein 3 (IGF2BP3) is a marker of unfavourable prognosis in colorectal cancer. Europ J Cancer. 2012;48(18):3405–3413.
  218. 218. Samanta S, Pursell B, Mercurio AM. IMP3 protein promotes chemoresistance in breast cancer cells by regulating breast cancer resistance protein (ABCG2) expression. J Biol Chem. 2013;288(18):12569–12573. pmid:23539627
  219. 219. Samanta S, Sharma VM, Khan A, Mercurio AM. Regulation of IMP3 by EGFR signaling and repression by ERbeta: implications for triple-negative breast cancer. Oncogene. 2012;31(44):4689–4697. pmid:22266872
  220. 220. Won JR, Gao D, Chow C, Cheng J, Lau SY, Ellis MJ, et al. A survey of immunohistochemical biomarkers for basal-like breast cancer against a gene expression profile gold standard. Mod Pathol. 2013;26(11):1438–1450. pmid:23702728
  221. 221. Walter O, Prasad M, Lu S, Quinlan RM, Edmiston KL, Khan A. IMP3 is a novel biomarker for triple negative invasive mammary carcinoma associated with a more aggressive phenotype. Hum Pathol. 2009;40(11):1528–1533. pmid:19695680
  222. 222. Bu D, Lewis CM, Sarode V, Chen M, Ma X, Lazorwitz AM, et al. Identification of Breast Cancer DNA Methylation Markers Optimized for Fine-Needle Aspiration Samples. Cancer Epidemiol Biomarkers Prev. 2013;22(12):2212–2221. pmid:24089458
  223. 223. Martens JW, Nimmrich I, Koenig T, Look MP, Harbeck N, Model F, et al. Association of DNA methylation of phosphoserine aminotransferase with response to endocrine therapy in patients with recurrent breast cancer. Cancer Res. 2005;65(10):4101–4117. pmid:15899800
  224. 224. Lan J, Zhao J, Liu Y. Molecular cloning, sequence characterization, polymorphism and association analysis of porcine ROPN1 gene. Mol Biol Rep. 2012;39(3):2739–2743. pmid:21667248
  225. 225. Berretta R, Moscato P. Cancer Biomarker Discovery: The Entropic Hallmark. PLoS ONE. 2010;5(8):e12262. pmid:20805891
  226. 226. de Kruijf EM, Bastiaannet E, Rubertá F, de Craen AJM, Kuppen PJK, Smit VTHBM, et al. Comparison of frequencies and prognostic effect of molecular subtypes between young and elderly breast cancer patients. Molecular Oncology. 2014;8(5):1014–1025. Available from: http://www.sciencedirect.com/science/article/pii/S1574789114000775. pmid:24767310
  227. 227. Nguyen PL, Taghian AG, Katz MS, Niemierko A, Abi Raad RF, Boon WL, et al. Breast cancer subtype approximated by estrogen receptor, progesterone receptor, and HER-2 is associated with local and distant recurrence after breast-conserving therapy. J Clin Oncol. 2008;26(14):2373–2378. pmid:18413639
  228. 228. Polyak K. Heterogeneity in breast cancer. J Clin Invest. 2011;121(10):3786–3788. pmid:21965334
  229. 229. Peppercorn J, Perou CM, Carey LA. Molecular subtypes in breast cancer evaluation and management: divide and conquer. Cancer Invest. 2008;26(1):1–10. pmid:18181038
  230. 230. Mackay A, Weigelt B, Grigoriadis A, Kreike B, Natrajan R, A’Hern R, et al. Microarray-based class discovery for molecular classification of breast cancer: analysis of interobserver agreement. J Natl Cancer Inst. 2011;103(8):662–673. pmid:21421860