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

A Bioinformatics Filtering Strategy for Identifying Radiation Response Biomarker Candidates

  • Jung Hun Oh,

    Affiliation Department of Medical Physics, Memorial Sloan-Kettering Cancer Center, New York, New York, United States of America

  • Harry P. Wong,

    Affiliation Department of Infectious Diseases, Washington University School of Medicine, St. Louis, Missouri, United States of America

  • Xiaowei Wang,

    Affiliation Department of Radiation Oncology, Washington University School of Medicine, St. Louis, Missouri, United States of America

  • Joseph O. Deasy

    deasyj@mskcc.org

    Affiliation Department of Medical Physics, Memorial Sloan-Kettering Cancer Center, New York, New York, United States of America

Abstract

The number of biomarker candidates is often much larger than the number of clinical patient data points available, which motivates the use of a rational candidate variable filtering methodology. The goal of this paper is to apply such a bioinformatics filtering process to isolate a modest number (<10) of key interacting genes and their associated single nucleotide polymorphisms involved in radiation response, and to ultimately serve as a basis for using clinical datasets to identify new biomarkers. In step 1, we surveyed the literature on genetic and protein correlates to radiation response, in vivo or in vitro, across cellular, animal, and human studies. In step 2, we analyzed two publicly available microarray datasets and identified genes in which mRNA expression changed in response to radiation. Combining results from Step 1 and Step 2, we identified 20 genes that were common to all three sources. As a final step, a curated database of protein interactions was used to generate the most statistically reliable protein interaction network among any subset of the 20 genes resulting from Steps 1 and 2, resulting in identification of a small, tightly interacting network with 7 out of 20 input genes. We further ranked the genes in terms of likely importance, based on their location within the network using a graph-based scoring function. The resulting core interacting network provides an attractive set of genes likely to be important to radiation response.

Introduction

In the ‘omics’ era, the number of biomarker candidates potentially available for statistical testing is often much larger than the number of patient data points. This presents a fundamental problem in biomarker research: the number of candidate genetic or epigenetic markers often overwhelms the inherent statistical power available in a clinical dataset, which usually has tens or hundreds of patient cases available rather than thousands. This statistical mismatch is typically becoming worse as more of the intracellular complexity of molecular machinery is identified. At one extreme, a genome-wide association study (GWAS) examining the correlations of millions of tag single-nucleotide polymorphisms (SNPs) to cancer treatment outcome may require a very high, and biologically unlikely, odds ratio given the number of multiple comparisons, to reach statistical significance. At the other extreme, it is clear that investigators cannot a priori identify the most important biomarker genes or SNPs for testing. These unsatisfying extreme cases motivated our search for a middle strategy that would objectively identify a modest number of promising SNPs/proteins, etc. as a cohort for testing against a given dataset. Because clinical datasets for a given endpoint are commonly of modest size (tens or hundreds, not thousands, of patients), we searched for key protein interaction networks that result in less than approximately a hundred candidate SNPs. Our methodology, of course, could be adopted to throw a wider net if much larger datasets become available. Our endpoint of interest is late toxicity following radiation therapy for cancer. Many cancer patients who receive radiation therapy suffer from acute or late side effects; the risk for experiencing these side effects is expected to have a genetic component [1]. Numerous genes participate in a cascade of events in response to radiation and the resulting DNA damage in a complex signal transduction network [2].

Recently, many studies have focused on finding radio-responsive genes at the whole genome level with gene expression microarrays. Rieger and Chu used oligonucleotide microarrays to develop a genome-wide portrait of transcriptional response to ionizing radiation (IR) and ultraviolet (UV) radiation in cell lines collected from 15 healthy individuals [3]. In another study [1] using samples extracted from cancer patients with acute radiation toxicity, Rieger et al. showed that toxicity after radiation therapy (radiotherapy) could be associated with abnormal transcriptional responses to DNA. Jen and Cheung [2] assessed transcriptional levels of genes in lymphoblastoid cells at various time points with 3 Gy and 10 Gy of ex vivo IR exposure. Following 10 Gy of IR exposure, more genes were induced, suggesting that a higher radiation dose causes a more complex response. A high percentage of significant genes were involved in cell cycle, cell death, DNA repair, DNA metabolism, and RNA processing. Eschrich et al. [4] analyzed microarray gene expression data derived from 48 human cancer cell lines and generated an interaction network using MetaCore software (GeneGo, Encinitas, CA) with the top 500 genes identified by linear regression analysis. Subsequently, based on 10 hub genes obtained from the network, they modeled radiosensitivity (survival fraction at 2 Gy) using a linear regression method.

Normal tissue toxicity after radiotherapy may partially be attributable to specific genetic mutations. In an effort to identify candidate polymorphisms at the SNP level involved in the cellular response to irradiation in breast and prostate cancers, Popanda et al. [5] surveyed many published studies that show associations of SNPs in candidate genes with acute or late side effects of radiotherapy. Andreassen and Alsner [6] summarized studies published on genetic variation in normal tissue toxicity and proposed a model of allelic architecture that illustrates relative risk for genetic variants associated with normal tissue radiosensitivity.

In this study, we attempted to define an objective method for identifying key radiosensitivity genes likely to have a significant impact on clinical outcome following radiotherapy. We elected to construct a staged filter. The first step was a comprehensive literature review of radiosensitivity-related genes. These genes were then further delimited to genes responding to IR in an analysis of publicly available microarray gene expression datasets. We further focused the search on interacting networks, based on the hypothesis that good biomarkers are likely to be embedded in important pathways or networks involving multiple genes known to be important to the endpoint in question [7]. This last step may potentially add new, previously unreported targets, based on curated pathway libraries.

Materials and Methods

In summary, we used a multi-component filtering process: (1) genes associated with radiation response in the literature and (2) genes associated with radiation response in two microarray mRNA datasets. Overlapping genes from these three sources were fed into a curated protein interaction network system (MetaCore) to identify key interacting networks. The most important network was taken as our target set.

Literature Review of Radiosensitivity-related Genes

We attempted a complete literature review of all genes implicated in radiation response. Published papers were searched by using PubMed and Scopus search engines in 2010 and by following citations within the identified papers. The search strategy was based on a combination of the following search keywords: “SNPs, polymorphisms, or microsatellites” and “irradiation, radiation, or radiotherapy” and “morbidity, radiosensitivity, normal tissue, toxicity, or complications” and “siRNA, knockdown, or knockout”. Papers referred to in the original search returns, or referring to the original papers at a later date were also reviewed. This resulted in an in-depth review of around 200 published papers, and a list of 221 genes implicated in radiation response.

Microarray Gene Expression Datasets

To identify significant radio-responsive genes based on microarray gene expression profiling, we searched for all relevant, publically available microarray datasets, resulting in locating two datasets. We analyzed GSE1977 and GSE23393, downloaded from the publicly available Gene Expression Omnibus (GEO) database (http://www.ncbi.nlm.nih.gov/geo/). In GSE1977, lymphoblastoid cell lines obtained from 15 healthy individuals were established by immortalization of peripheral blood B-lymphocytes [3]. The response of numerous genes was measured by mock treatment, UV, and X-ray exposures. Cells were exposed to 5 Gy radiation doses and harvested for RNA 4 hours later. In our work, the differential between mock and X-ray cases was used. In contrast, in GSE23393 [8], blood was gathered from eight radiotherapy patients (at our institution): eight samples were collected immediately before irradiation and another eight samples were collected at 4 hours after total body irradiation with 1.25 Gy X-rays.

Preprocessing for Identification of Significant Genes

Before the microarray datasets were analyzed, gene expression values were log-base-2 transformed, followed by quantile normalization across all samples [9]. Microarray gene expression values from two different conditions (before and after exposure) were compared using a two-tailed t-test to identify differentially expressed genes (radio-responsive genes). To estimate the likelihood of identifying significant genes by chance, we computed permutation-based p-values using 10,000 permutations. Then, using Storey’s method, the false discovery rate (FDR) and q-value for each gene were calculated [10]. Significance Analysis of Microarrays (SAM) and t-test are widely used for indentifying differentially expressed genes in the analysis of microarray data [11]. We chose a permutation t-test with an assumption that the permutation t-test and SAM could yield a set of similar significant genes, as recommended by Chen et al. [11]. In this analysis, we did not use a fold change cutoff in order to avoid losing some important genes, a problem described by Larsson et al. [12].

Pathway and Process Analysis

Significant genes were identified both in the literature review and the analysis of two microarray datasets (GSE1977 and GSE23393). These genes were then entered into a manually curated pathway analysis database (MetaCore™, GeneGo, Inc., Carlsbad, California). The commercial pathway analysis system, MetaCore, computes p-values for overrepresented pathways and processes. MetaCore is based on a comprehensive manually curated attempt to capture protein interactions as networks. We used MetaCore to attempt to find the most probable interaction pathways among a set of genes uploaded by the user. Several algorithms are available to do this; we used the “Analyze network” option. If necessary, MetaCore adds appropriate genes to complete a network.

Gene Ontology Analysis

A further analysis of the resulting significant genes was performed using the Gene Ontology (GO) database (www.geneontology.org), in which genes are annotated with known molecular functions, biological processes, and cellular component locations.

Gene Ranking

In our previous work to identify blood-based protein biomarkers to predict radiation-induced pneumonitis [7], we proposed a graph-based scoring function to rank proteins in a protein–protein interaction network. The network consisted of candidate proteins we identified in mass spectrometry analysis and four previously identified (‘regularization’) biomarker proteins. Using the proposed method, we attempted to measure a ‘functional distance’ between each candidate protein and the four regularization proteins, based on the hypothesis that some proteins relevant to a specific disease exist in close proximity, in a network sense. In the current study, we modified that algorithm such that within a given protein–protein interaction network for a biological process, we estimate the functional distance between each protein and all the remaining proteins in the network, since all the proteins in the network are more likely to be related to one another and act together in the biological process.

To rank biomarkers, we modelled each protein–protein interaction network as a directed graph, G = (V, E), where V consists of a set of nodes (proteins) and E is the set of possible edges (protein–protein interactions) between pairs of nodes. Let A and B be two proteins in a network. We assume that there are two concepts of distance between A and B: a geometrical distance that is defined in terms of the number of nodes in the shortest path between A and B, as well as a virtual distance that is defined in terms of the number of publications that verify the interactions along the shortest path. Intuitively, as the number of intermediate nodes between A and B increases, the geometrical distance increases and the two proteins are less likely to be correlated. In contrast, considering virtual distance, we expect that as the number of references demonstrating a relationship between two proteins increases, they are more likely to be related. In other words, the number of references is proportional to relatedness while the number of nodes is inversely proportional.

Using a power law, we calculate two scores from A to B: a reference score (rs) and a node score (ns) as follows:(1)(2)where r and n are the total number of references and nodes in the shortest path from A to B. We suppose that the influence of the number of nodes is greater than that of the number of references. Therefore, as the number of intermediate nodes between any two given nodes increases, the relationship between the two nodes becomes much less likely. The score capturing the path from A to B is defined as the summation of two different scores:(3)Likewise, we also estimate a score from B to A, Then, the final score, between A and B, is defined as the maximal value among and :

(4)We suppose that the final score of a protein is computed by the summation of all scores between the protein and all the remaining proteins in the network. Hence, the final score of a protein A is defined by:(5)

To estimate the number of references and nodes, we employed two methods. For the number of references, we used a function in the MetaCore software that provides the number of references between two connected proteins in a network. For the number of nodes, we used the Floyd-Warshall algorithm that was originally designed to find the shortest paths between all pairs of nodes based on dynamic programming [13]. To apply this algorithm to our problem of estimating the number of nodes, we modified the original Floyd-Warshall algorithm such that an equal weight of 1 was assigned to all connected edges in a network. As a result, the modified algorithm generated a matrix that represents the number of nodes on the all-pairs shortest-paths in a given protein–protein interaction network.

Results

Identification of Significant Biomarkers via Literature Review

Based on the literature review, several types of biomarkers, including genes, proteins, kinases, ligands, and protein complexes were identified. To unify the biomarker terms differently used across studies, we converted all the biomarkers into their corresponding gene symbols. As a result, 221 unique genes and 4 protein complexes (DNA-PK, HSP70, MRN(95), RAS) were identified from around 200 papers that studied radiation response-related biomarkers [4], [14][185]. Table 1 displays the 221 unique genes and their corresponding GO processes, including DNA repair, cell proliferation/cycle, apoptosis, RNA processing, and response to stress. It is well known that ionizing radiation causes DNA damage that activates the p53 pathway through ATM [186]. Genes that are involved in cell cycle, such as CDKN1A, GADD45A, MDM2, and CCNG1, are known to be dependent on p53 [2]. Also, other cell cycle-related genes including CCNB1 and CDC20 were identified. Among cell cycle or proliferation genes, TOB1, BTG2, and CDKN1A are anti-proliferative/check-point related [3]. Several genes (XPC, DDB2, PCNA, ERCC4, and NBN) are involved in DNA repair. Two major pathways to repair IR-induced DNA double-strand breaks are homologous recombination (HR; genes include XRCC2, XRCC3, MRE11A, RAD50, NBN, BRCA1, and BRCA2) and non-homologous end joining (NHEJ; genes include LIG4, XRCC4, XRCC5, XRCC6, and DNA-PK) [3]. Some genes, including FAS, BBC3, and TNF, are involved in apoptosis [187]. BCL2 and DDR1 are anti-apoptotic.

thumbnail
Table 1. Radio-responsive biomarkers identified by literature review and their biological processes.

https://doi.org/10.1371/journal.pone.0038870.t001

For biological process and pathway analysis, the 221 unique genes were uploaded into the MetaCore. Figure 1 illustrates a direct interaction network generated with these genes. As shown, numerous genes are strongly connected to one another, suggesting that interacting genes are more likely to play related roles. Table 2 shows the top ten GeneGo pathways, GeneGo processes, and GO processes. As can be seen in the table, the most highly ranked pathways and processes are associated with DNA damage and repair, cell cycle, and apoptosis.

thumbnail
Figure 1. Direct protein-protein interaction network.

A network representation that illustrates the complexity of direct connections among genes identified via literature review.

https://doi.org/10.1371/journal.pone.0038870.g001

thumbnail
Table 2. The top ten GeneGo pathways/processes and GO processes resulting from genes identified via literature review.

https://doi.org/10.1371/journal.pone.0038870.t002

Identification of Significant Genes via Microarray Dataset Analysis

To identify significant changes in gene expression values between the two groups (before and after irradiation) in two microarray datasets, a t-test with 10,000 permutations was performed. To estimate p-values, we counted the number of permutations for each gene whose t-scores are greater than or equal to the t-score calculated with observed values. Then, the number of permutations passed the criterion was divided by the total number of permutations [188]. With an FDR of 20%, 631 probes (corresponding to 550 unique genes) were significantly identified for GSE1977. Figure 2 shows a normal quantile plot of t-scores for GSE1977. Data points of genes that are farther away from the black diagonal line are considered to be differentially expressed. Figure 3 displays a volcano plot that depicts the –log10 of q-values against log2 of fold changes for all genes. The majority of genes with an FDR of 20% changed 1.2-fold or higher. For GSE23393, with an FDR of 20%, 224 probes (corresponding to 184 unique genes) were identified (Figure S1 and Figure S2).

thumbnail
Figure 2. A normal quantile plot of t-scores for GSE1977.

Significant genes have red circles.

https://doi.org/10.1371/journal.pone.0038870.g002

thumbnail
Figure 3. Significant gene detection.

A volcano plot that depicts the –log10 of q-values against log2 of fold changes for all genes in GSE1977.

https://doi.org/10.1371/journal.pone.0038870.g003

Overlapping Genes

To delimit our potential biomarker set, we investigated which genes are commonly or uniquely found among the set of genes identified by our literature review and two sets of genes identified in the analysis of the two gene microarray datasets, as summarized in Figure 4. The rationale is that those are genes likely to be key to an active response, but unlikely to be false positives due to the literature review. Twenty genes were commonly identified among the three different analyses (literature review and two microarray datasets), as shown in Table 3. We further analyzed pathways and biological processes associated with the 20 genes. Table 4 shows the top ten GeneGo pathways generated by the MetaCore software (Table S1). Not surprisingly, even with the 20 genes, DNA damage/repair and apoptosis-related pathways were highly ranked.

thumbnail
Figure 4. Comparison of significant genes among three sources.

A Venn diagram depicting the number of shared and unique genes among a set of genes identified by literature review and two sets of genes identified in the analysis of two gene microarray datasets.

https://doi.org/10.1371/journal.pone.0038870.g004

thumbnail
Table 3. Twenty genes commonly identified by literature review and analysis of two microarray datasets.

https://doi.org/10.1371/journal.pone.0038870.t003

thumbnail
Table 4. The top ten GeneGo pathways generated by MetaCore when the 20 overlapping genes were used.

https://doi.org/10.1371/journal.pone.0038870.t004

thumbnail
Figure 5. The most probable interaction network when 20 genes were entered into MetaCore software.

The resulting interacting network uses only 7 genes. Red, green, and gray lines indicate inhibitory, stimulatory, and unspecified interactions, respectively.

https://doi.org/10.1371/journal.pone.0038870.g005

thumbnail
Table 5. The results of the proposed scoring function test applied to the network in Figure 5.

https://doi.org/10.1371/journal.pone.0038870.t005

Gene Ranking and Identification of a Core Radio-response Network

Figure 5 shows the most probable/robust single interaction network when the 20 overlapping genes were entered into the MetaCore software. Of the 20 input genes, seven genes appeared in this core radio-response network. We applied our graph-based scoring function to this network and the results are summarized in Table 5. MYC was ranked first with a score of 113.74, which had a high p-value in GSE23393 and a statistically significant p-value, yet still relatively high compared to other genes, in GSE1977. As a hub gene, MYC had the highest number of edges (n = 12) that seem to contribute to the score. Overall p-values in GSE23393 (in situ IR) are higher than those of GSE1977 (ex vivo IR). Intuitively, as the number of edges increases, the score seems to increase. However, it should be noted that although GADD45A has 9 edges, it obtained a higher score than PPM1D, which has 11 edges. This is attributed to the fact that when we calculate the score for a gene, our scoring function takes into account all network interactions and the number of references on the interactions in the network. Interestingly, CDKN1A obtained a relatively high score of 100.13, considering only 3 edges and substantially low p-values (0.00027 in GSE1977 and 0.00367 in GSE23393).

Discussion

We have demonstrated an unbiased bioinformatics filtering methodology to objectively identify a core network of key interacting genes that are important to radiation response. We hypothesized that, by combining several different types of datasets, we are increasingly likely to identify interacting genes that are particularly important to radiation response. We also hypothesize that these genes are therefore attractive candidates for biomarker testing. For example, the 7 key genes contain 89 relevant SNPs in our radiation therapy cancer dataset and we are in the process of testing late toxicity with the dataset. We make no claim that the network shown in Figure 5 dominates radiation response and do not expect that to be the case. Nevertheless, this network seems to be highly relevant to radiation response: among the 7 genes, 5 and 4 genes are involved in cell cycle control and apoptosis, respectively. More detailed information is shown in Table S2. Five of these genes, including MYC, BBC3, GADD45A, CDKN1A, and XPC belong to a list of 34 radio-responsive genes observed by Tusher et al. [187]. Moreover, this network is consistent with (though slightly different from) the programmed cell death network reported by Moussay et al. [189].

Figure 4 shows the number of genes commonly or uniquely identified among three different studies (literature review and analysis of two microarray datasets). Interestingly, relatively few genes overlapped among the three analyses. Literature coverage is expected to be incomplete regarding coverage of radiosensitivity genes. Microarray analysis is subject to high false positive and false negative rates [190]. Another possible reason for the small number of overlapped genes is the widely differing irradiation conditions and doses. Despite this, the biological processes and pathways generated from the 20 overlapping genes were similar to those generated from the whole literature review.

We further analyzed the 20 genes, uploading these genes into the MetaCore software. In the network of the most probable biological process shown in Figure 5, only seven out of 20 genes appeared in the network. Additional genes were automatically added to the network by MetaCore, including AKT1, RELA, BCL2L1, PTEN, CDK1, and XIAP. Note, however, that these genes were also members of the list generated by our radiation response literature review, suggesting some consistency between these sources. This also suggests a potential ability to find novel biomarker candidates through the network mapping/ranking process, though that did not occur in this case.

The graph-based scoring function proposed in our previous study [7] was modified and applied to the network shown in Figure 5. In some studies, researchers tend to regard genes with high degrees of connectivity (hub genes) as significant in an interaction network, while neglecting others [4]. While this is rational, finding hub genes based on edge connectivity considers only direct interactions between genes whereas our proposed approach takes into account all interactions in a network (that is, the entire graph structure) and the number of published references on the interactions. To measure the closeness between two proteins (say A and B), we employed two scores; a node score and a reference score. In a protein interaction network, it is obvious that as the number of internal nodes between A and B increases, these two proteins are less likely to be related with each other. In contrast, the reference score is a score calculated using the number of papers that studied on an interaction between two proteins, which can be important evidence that there is an actual relationship between the two proteins. As can be seen in Table S3, MYC was first ranked using a total score. However, BBC3 and PPM1D were first ranked using a reference score and a node score, respectively. CDKN1A, PLK3, and XPC obtained somewhat high scores considering their connectivity, suggesting that they could play important roles in this core network. We believe that the use of both scores could be more effective for ranking proteins in a protein interaction network.

Future work will test SNPs identified in this network against toxicity resulting from radiation therapy. As the number of patients available for SNP analyses increases, it may be rational to expand the number of candidate SNPs to several hundreds or more. The general methodology may be applied in many genetic/protein biomarker studies with limited patient data.

Supporting Information

Figure S1.

A normal quantile plot of t-scores for GSE23393 after 10,000 permutations.

https://doi.org/10.1371/journal.pone.0038870.s001

(TIF)

Figure S2.

Significant gene detection. A volcano plot that depicts the –log10 of q-values against log2 of fold changes for all genes in GSE23393.

https://doi.org/10.1371/journal.pone.0038870.s002

(TIF)

Table S1.

The top ten GeneGo pathways/processes and GO processes generated by the MetaCore software when 20 overlapped genes were used.

https://doi.org/10.1371/journal.pone.0038870.s003

(DOC)

Table S2.

Biological processes for the seven genes shown in Table 5.

https://doi.org/10.1371/journal.pone.0038870.s004

(DOC)

Table S3.

Scores obtained using the graph-based scoring function.

https://doi.org/10.1371/journal.pone.0038870.s005

(DOC)

Author Contributions

Conceived and designed the experiments: JHO XW JOD. Performed the experiments: JHO HPW. Analyzed the data: JHO. Wrote the paper: JHO JOD.

References

  1. 1. Rieger KE, Hong WJ, Tusher VG, Tang J, Tibshirani R (2004) Toxicity from radiation therapy associated with abnormal transcriptional responses to DNA damage. Proc Natl Acad Sci U S A 101: 6635.6640
  2. 2. Jen KY, Cheung VG (2003) Transcriptional response of lymphoblastoid cells to ionizing radiation. Genome Res 13: 2092.2100
  3. 3. Rieger KE, Chu G (2004) Portrait of transcriptional responses to ultraviolet and ionizing radiation in human cells. Nucleic Acids Res 32: 4786.4803
  4. 4. Eschrich S, Zhang H, Zhao H, Boulware D, Lee JH (2009) Systems biology modeling of the radiation sensitivity network: a biomarker discovery platform. Int J Radiat Oncol Biol Phys 75: 497.505
  5. 5. Popanda O, Marquardt JU, Chang-Claude J, Schmezer P (2009) Genetic variation in normal tissue toxicity induced by ionizing radiation. Mutat Res 667: 58.69
  6. 6. Andreassen CN, Alsner J (2009) Genetic variants and normal tissue toxicity after radiotherapy: a systematic review. Radiother Oncol 92: 299.309
  7. 7. Oh JH, Craft JM, Townsend R, Deasy JO, Bradley JD (2011) A bioinformatics approach for biomarker identification in radiation-induced lung inflammation from limited proteomics data. J Proteome Res 10: 1406.1415
  8. 8. Templin T, Paul S, Amundson SA, Young EF, Barker CA (2011) Radiation-induced micro-RNA expression changes in peripheral blood cells of radiotherapy patients. Int J Radiat Oncol Biol Phys 80: 549.557
  9. 9. Dexter TJ, Sims D, Mitsopoulos C, Mackay A, Grigoriadis A (2010) Genomic distance entrained clustering and regression modelling highlights interacting genomic regions contributing to proliferation in breast cancer. BMC Syst Biol 4: 127.
  10. 10. Storey JD, Tibshirani R (2003) Statistical significance for genomewide studies. Proc Natl Acad Sci U S A 100: 9440.9445
  11. 11. Chen T, Guo L, Zhang L, Shi L, Fang H (2006) Gene expression profiles distinguish the carcinogenic effects of aristolochic acid in target (kidney) and non-target (liver) tissues in rats. BMC Bioinformatics 7: S20.
  12. 12. Larsson O, Wahlestedt C, Timmons JA (2005) Considerations when using the significance analysis of microarrays (SAM) algorithm. BMC Bioinformatics 6: 129.
  13. 13. Brigl B, Strübing A, Wendt T, Winter A (2006) Modeling interdependencies between information processes and communication paths in hospitals. Methods Inf Med 45: 216.224
  14. 14. Isomura M, Oya N, Tachiiri S, Kaneyasu Y, Nishimura Y (2008) IL12RB2 and ABCA1 genes are associated with susceptibility to radiation dermatitis. Clin Cancer Res 14: 6683.6689
  15. 15. Chiani F, Iannone C, Negri R, Paoletti D, D'Antonio M (2009) Radiation Genes: a database devoted to microarrays screenings revealing transcriptome alterations induced by ionizing radiation in mammalian cells. Database (Oxford) 2009: bap007.
  16. 16. Turtoi A, Schneeweiss FH (2009) Effect of (211)At alpha-particle irradiation on expression of selected radiation responsive genes in human lymphocytes. Int J Radiat Biol 85: 403.412
  17. 17. Valdagni R, Rancati T, Ghilotti M, Cozzarini C, Vavassori V (2009) To bleed or not to bleed. A prediction based on individual gene profiling combined with dose-volume histogram shapes in prostate cancer patients undergoing three-dimensional conformal radiation therapy. Int J Radiat Oncol Biol Phys 74: 1431.1440
  18. 18. Toulany M, Kehlbach R, Florczak U, Sak A, Wang S (2008) Targeting of AKT1 enhances radiation toxicity of human tumor cells by inhibiting DNA-PKcs-dependent DNA double-strand break repair. Mol Cancer Ther 7: 1772.1781
  19. 19. Suga T, Ishikawa A, Kohda M, Otsuka Y, Yamada S (2007) Haplotype-based analysis of genes associated with risk of adverse skin reactions after radiotherapy in breast cancer patients. Int J Radiat Oncol Biol Phys 69: 685.693
  20. 20. Suga T, Iwakawa M, Tsuji H, Ishikawa H, Oda E (2008) Influence of multiple genetic polymorphisms on genitourinary morbidity after carbon ion radiotherapy for prostate cancer. Int J Radiat Oncol Biol Phys 72: 808.813
  21. 21. Xu QY, Gao Y, Liu Y, Yang WZ, Xu XY (2008) Identification of differential gene expression profiles of radioresistant lung cancer cell line established by fractionated ionizing radiation in vitro. Chin Med J (Engl) 121: 1830.1837
  22. 22. Chang-Claude J, Popanda O, Tan XL, Kropp S, Helmbold I (2005) Association between polymorphisms in the DNA repair genes, XRCC1, APE1, and XPD and acute side effects of radiotherapy in breast cancer patients. Clin Cancer Res 11: 4802.4809
  23. 23. Sak SC, Harnden P, Johnston CF, Paul AB, Kiltie AE (2005) APE1 and XRCC1 protein expression levels predict cancer-specific survival following radical radiotherapy in bladder cancer. Clin Cancer Res 11: 6205.6211
  24. 24. Xiang DB, Chen ZT, Wang D, Li MX, Xie JY (2008) Chimeric adenoviral vector Ad5/F35-mediated APE1 siRNA enhances sensitivity of human colorectal cancer cells to radiotherapy in vitro and in vivo. Cancer Gene Ther 15: 625.635
  25. 25. Andreassen CN, Alsner J, Overgaard M, Sørensen FB, Overgaard J (2006) Risk of radiation-induced subcutaneous fibrosis in relation to single nucleotide polymorphisms in TGFB1, SOD2, XRCC1, XRCC3, APEX and ATM–a study based on DNA from formalin fixed paraffin embedded tissue samples. Int J Radiat Biol 82: 577.586
  26. 26. Andreassen CN, Alsner J, Overgaard J, Herskind C, Haviland J (2005) TGFB1 polymorphisms are associated with risk of late normal tissue complications in the breast after radiotherapy for early breast cancer. Radiother Oncol 75: 18.21
  27. 27. Higuchi Y, Nelson GA, Vazquez M, Laskowitz DT, Slater JM (2002) Apolipoprotein E expression and behavioral toxicity of high charge, high energy (HZE) particle radiation. J Radiat Res 43:
  28. 28. Kool J, Hamdi M, Cornelissen-Steijger P, van der Eb AJ, Terleth C (2003) Induction of ATF3 by ionizing radiation is mediated via a signaling pathway that includes ATM, Nibrin1, stress-induced MAPkinases and ATF-2. Oncogene 22: 4235.4242
  29. 29. Borgmann K, Röper B, El-Awady R, Brackrock S, Bigalke M (2002) Indicators of late normal tissue response after radiotherapy for head and neck cancer: fibroblasts, lymphocytes, genetics, DNA repair, and chromosome aberrations. Radiother Oncol 64: 141.152
  30. 30. Azria D, Ozsahin M, Kramar A, Peters S, Atencio DP (2008) Single nucleotide polymorphisms, apoptosis, and the development of severe late adverse effects after radiotherapy. Clin Cancer Res 14: 6284.6288
  31. 31. Moore J, Stock RG, Cesaretti JA, Stone NN, Li W (2007) Single nucleotide polymorphisms as predictors for development of erectile dysfunction in African-American men treated with radiotherapy for prostate cancer. Int J Radiat Oncol Biol Phys 69: S7.
  32. 32. Appleby JM, Barber JB, Levine E, Varley JM, Taylor AM (1997) Absence of mutations in the ATM gene in breast cancer patients with severe responses to radiotherapy. Br J Cancer 76: 1546.1549
  33. 33. Iannuzzi CM, Atencio DP, Green S, Stock RG, Rosenstein BS (2002) ATM mutations in female breast cancer patients predict for an increase in radiation-induced late effects. Int J Radiat Oncol Biol Phys 52: 606.613
  34. 34. Hall EJ, Schiff PB, Hanks GE, Brenner DJ, Russo J (1998) A preliminary report: frequency of A-T heterozygotes among prostate cancer patients with severe late responses to radiation therapy. Cancer J Sci Am 4: 385.389
  35. 35. Pugh TJ, Keyes M, Barclay L, Delaney A, Krzywinski M (2009) Sequence variant discovery in DNA repair genes from radiosensitive and radiotolerant prostate brachytherapy patients. Clin Cancer Res 15: 5008.5016
  36. 36. Zschenker O, Raabe A, Boeckelmann IK, Borstelmann S, Szymczak S (2010) Association of single nucleotide polymorphisms in ATM, GSTP1, SOD2, TGFB1, XPD and XRCC1 with clinical and cellular radiosensitivity. Radiother Oncol 97: 26.32
  37. 37. Hendry JH, Hoyes KP, Wadeson P, Roberts SA (2002) Role of damage-sensing/processing genes in the radiation response of haemopoietic in vitro colony-forming cells. Int J Radiat Biol 78: 559.566
  38. 38. Mao JH, Wu D, DelRosario R, Castellanos A, Balmain A (2008) Atm heterozygosity does not increase tumor susceptibility to ionizing radiation alone or in a p53 heterozygous background. Oncogene 27: 6596.6600
  39. 39. Worgul BV, Smilenov L, Brenner DJ, Junk A, Zhou W (2002) Atm heterozygous mice are more sensitive to radiation-induced cataracts than are their wild-type counterparts. Proc Natl Acad Sci U S A 99: 9836.9839
  40. 40. Clarke RA, Fang ZH, Marr PJ, Lee CS, Kearsley JH (2002) ATM induction insufficiency in a radiosensitive breast-cancer patient. Australas Radiol 46: 329.335
  41. 41. Mirzayans R, Severin D, Murray D (2006) Relationship between DNA double-strand break rejoining and cell survival after exposure to ionizing radiation in human fibroblast strains with differing ATM/p53 status: implications for evaluation of clinical radiosensitivity. Int J Radiat Oncol Biol Phys 66: 1498.1505
  42. 42. Paterson MC, Anderson AK, Smith BP, Smith PJ (1979) Enhanced radiosensitivity of cultured fibroblasts from ataxia telangiectasia heterozygotes manifested by defective colony-forming ability and reduced DNA repair replication after hypoxic gamma-irradiation. Cancer Res 39: 3725.3734
  43. 43. Fernet M, Hall J (2004) Genetic biomarkers of therapeutic radiation sensitivity. DNA Repair (Amst) 3: 1237.1243
  44. 44. Chorna IV, Datsyuk LO, Stoika RS (2005) Expression of Bax, Bad and Bcl-2 proteins under x-radiation effect towards human breast carcinoma MCF-7 cells and their doxorubicin-resistant derivatives. Exp Oncol 27: 196.201
  45. 45. Moretti L, Attia A, Kim KW, Lu B (2007) Crosstalk between Bak/Bax and mTOR signaling regulates radiation-induced autophagy. Autophagy 3: 142.144
  46. 46. Haffty BG, Glazer PM (2003) Molecular markers in clinical radiation oncology. Oncogene 22: 5915.5925
  47. 47. Burns TF, Bernhard EJ, El-Deiry WS (2001) Tissue specific expression of p53 target genes suggests a key role for KILLER/DR5 in p53-dependent apoptosis in vivo. Oncogene 20: 4601.4612
  48. 48. Badie C, Dziwura S, Raffy C, Tsigani T, Alsbeih G (2008) Aberrant CDKN1A transcriptional response associates with abnormal sensitivity to radiation treatment. Br J Cancer 98: 1845.1851
  49. 49. Kabacik S, Mackay A, Tamber N, Manning G, Finnon P (2011) Gene expression following ionising radiation: identification of biomarkers for dose estimation and prediction of individual response. Int J Radiat Biol 87: 115.129
  50. 50. Paul S, Amundson SA (2008) Development of gene expression signatures for practical radiation biodosimetry. Int J Radiat Oncol Biol Phys 71: 1236.1244
  51. 51. Kim DW, Seo SW, Cho SK, Chang SS, Lee HW (2007) Targeting of cell survival genes using small interfering RNAs (siRNAs) enhances radiosensitivity of Grade II chondrosarcoma cells. J Orthop Res 25: 820.828
  52. 52. Guo WF, Lin RX, Huang J, Zhou Z, Yang J (2005) Identification of differentially expressed genes contributing to radioresistance in lung cancer cells using microarray analysis. Radiat Res 164: 27.35
  53. 53. Guan HT, Xue XH, Dai ZJ, Wang XJ, Li A (2006) Down-regulation of survivin expression by small interfering RNA induces pancreatic cancer cell apoptosis and enhances its radiosensitivity. World J Gastroenterol 12: 2901.2907
  54. 54. Gaffney DK, Brohet RM, Lewis CM, Holden JA, Buys SS (1998) Response to radiation therapy and prognosis in breast cancer patients with BRCA1 and BRCA2 mutations. Radiother Oncol 47: 129.136
  55. 55. Shanley S, McReynolds K, Ardern-Jones A, Ahern R, Fernando I (2006) Late toxicity is not increased in BRCA1/BRCA2 mutation carriers undergoing breast radiotherapy in the United Kingdom. Clin Cancer Res 12: 7025.7032
  56. 56. Pierce LJ, Strawderman M, Narod SA, Oliviotto I, Eisen A (2000) Effect of radiotherapy after breast-conserving treatment in women with breast cancer and germline BRCA1/2 mutations. J Clin Oncol 18: 3360.3369
  57. 57. Leong T, Whitty J, Keilar M, Mifsud S, Ramsay J (2000) Mutation analysis of BRCA1 and BRCA2 cancer predisposition genes in radiation hypersensitive cancer patients. Int J Radiat Oncol Biol Phys 48: 959.965
  58. 58. Abbott DW, Thompson ME, Robinson-Benion C, Tomlinson G, Jensen RA (1999) BRCA1 expression restores radiation resistance in BRCA1-defective cancer cells through enhancement of transcription-coupled DNA repair. J Biol Chem 274: 18808.18812
  59. 59. Ernestos B, Nikolaos P, Koulis G, Eleni R, Konstantinos B (2010) Increased chromosomal radiosensitivity in women carrying BRCA1/BRCA2 mutations assessed with the G2 assay. Int J Radiat Oncol Biol Phys 76: 1199.1205
  60. 60. Higgins GS, Prevo R, Lee YF, Helleday T, Muschel RJ (2010) A small interfering RNA screen of genes involved in DNA repair identifies tumor-specific radiosensitization by POLQ knockdown. Cancer Res 70: 2984.2993
  61. 61. Zhou T, Chou JW, Simpson DA, Zhou Y, Mullen TE (2006) Profiles of global gene expression in ionizing-radiation-damaged human diploid fibroblasts reveal synchronization behind the G1 checkpoint in a G0-like state of quiescence. Environ Health Perspect 114: 553.559
  62. 62. Kuptsova N, Chang-Claude J, Kropp S, Helmbold I, Schmezer P (2008) Genetic predictors of long-term toxicities after radiation therapy for breast cancer. Int J Cancer 122: 1333.1339
  63. 63. Hehlgans S, Eke I, Storch K, Haase M, Baretton GB (2009) Caveolin-1 mediated radioresistance of 3D grown pancreatic cancer cells. Radiother Oncol 92: 362.370
  64. 64. Hixon JA, Blazar BR, Anver MR, Wiltrout RH, Murphy WJ (2001) Antibodies to CD40 induce a lethal cytokine cascade after syngeneic bone marrow transplantation. Biol Blood Marrow Transplant 7: 136.143
  65. 65. Wischhusen J, Jung G, Radovanovic I, Beier C, Steinbach JP (2002) Identification of CD70-mediated apoptosis of immune effector cells as a novel immune escape pathway of human glioblastoma. Cancer Res 62: 2592.2599
  66. 66. Taniguchi K, Momiyama N, Ueda M, Matsuyama R, Mori R (2008) Targeting of CDC20 via small interfering RNA causes enhancement of the cytotoxicity of chemoradiation. Anticancer Res 28: 1559.1563
  67. 67. Snyder AR, Morgan WF (2004) Gene expression profiling after irradiation: clues to understanding acute and persistent responses? Cancer Metastasis Rev 23: 259.268
  68. 68. Correa CR, Cheung VG (2004) Genetic variation in radiation-induced expression phenotypes. Am J Hum Genet 75: 885.890
  69. 69. Moussavi-Harami F, Mollano A, Martin JA, Ayoob A, Domann FE (2006) Intrinsic radiation resistance in human chondrosarcoma cells. Biochem Biophys Res Commun 346: 379.385
  70. 70. Lu HR, Wang X, Wang Y (2006) A stronger DNA damage-induced G2 checkpoint due to over-activated CHK1 in the absence of PARP-1. Cell Cycle 5: 2364.2370
  71. 71. Xiao Z, Xue J, Sowin TJ, Rosenberg SH, Zhang H (2005) A novel mechanism of checkpoint abrogation conferred by Chk1 downregulation. Oncogene 24: 1403.1411
  72. 72. Kim JS, Chang JW, Yun HS, Yang KM, Hong EH (2010) Chloride intracellular channel 1 identified using proteomic analysis plays an important role in the radiosensitivity of HEp-2 cells via reactive oxygen species production. Proteomics 10: 2589.2604
  73. 73. Liu L, Zou JJ, Luo HS, Wu DH (2009) [Effect of protein kinase CK2 gene silencing on radiosensitization in human nasopharyngeal carcinoma cells]. Nan Fang Yi Ke Da Xue Xue Bao 29: 1551.1553
  74. 74. Damaraju S, Murray D, Dufour J, Carandang D, Myrehaug S (2006) Association of DNA repair and steroid metabolism gene polymorphisms with clinical late toxicity in patients treated with conformal radiotherapy for prostate cancer. Clin Cancer Res 12: 2545.2554
  75. 75. Kim KW, Moretti L, Mitchell LR, Jung DK, Lu B (2010) Endoplasmic reticulum stress mediates radiation-induced autophagy by perk-eIF2alpha in caspase-3/7-deficient cells. Oncogene 29: 3241.3251
  76. 76. Shi HS, Ren DL, Cao XM, Huang L, Jiang Y (2008) [Small interfering RNA targeting C-erbB-2 gene increases the radiosensitivity of lung adenocarcinoma cells in vitro]. Nan Fang Yi Ke Da Xue Xue Bao 28: 1977.1980
  77. 77. Pietras RJ, Poen JC, Gallardo D, Wongvipat PN, Lee HJ (1999) Monoclonal antibody to HER-2/neureceptor modulates repair of radiation-induced DNA damage and enhances radiosensitivity of human breast cancer cells overexpressing this oncogene. Cancer Res 59: 1347.1355
  78. 78. Xu W, Debeb BG, Lacerda L, Woodward WA (2010) Potential targets for improving radiosensitivity of breast tumor-initiating cells. Anticancer Agents Med Chem 10: 152.156
  79. 79. Kornguth DG, Garden AS, Zheng Y, Dahlstrom KR, Wei Q (2005) Gastrostomy in oropharyngeal cancer patients with ERCC4 (XPF) germline variants. Int J Radiat Oncol Biol Phys 62: 665.671
  80. 80. Reap EA, Roof K, Maynor K, Borrero M, Booker J (1997) Radiation and stress-induced apoptosis: a role for Fas/Fas ligand interactions. Proc Natl Acad Sci U S A 94: 5750.5755
  81. 81. Peña LA, Fuks Z, Kolesnick RN (2000) Radiation-induced apoptosis of endothelial cells in the murine central nervous system: protection by fibroblast growth factor and sphingomyelinase deficiency. Cancer Res 60: 321.327
  82. 82. Chang JT, Chan SH, Lin CY, Lin TY, Wang HM (2007) Differentially expressed genes in radioresistant nasopharyngeal cancer cells: gp96 and GDF15. Mol Cancer Ther 6: 2271.2279
  83. 83. Cao Y, Fu YL, Yu M, Yue PB, Ge CH (2009) Human augmenter of liver regeneration is important for hepatoma cell viability and resistance to radiation-induced oxidative stress. Free Radic Biol Med 47: 1057.1066
  84. 84. Barahmani N, Carpentieri S, Li XN, Wang T, Cao Y (2009) Glutathione S-transferase M1 and T1 polymorphisms may predict adverse effects after therapy in children with medulloblastoma. Neuro Oncol 11: 292.300
  85. 85. Edvardsen H, Kristensen VN, Grenaker Alnaes GI, Bøhn M, Erikstein B (2007) Germline glutathione S-transferase variants in breast cancer: relation to diagnosis and cutaneous long-term adverse effects after two fractionation patterns of radiotherapy. Int J Radiat Oncol Biol Phys 67: 1163.1171
  86. 86. Ambrosone CB, Tian C, Ahn J, Kropp S, Helmbold I (2006) Genetic predictors of acute toxicities related to radiation therapy following lumpectomy for breast cancer: a case-series study. Breast Cancer Res 8: R40.
  87. 87. Munshi A, Kurland JF, Nishikawa T, Tanaka T, Hobbs ML (2005) Histone deacetylase inhibitors radiosensitize human melanoma cells by suppressing DNA repair activity. Clin Cancer Res 11: 4912.4922
  88. 88. Yoshida K, Morita T (2004) Control of radiosensitivity of F9 mouse teratocarcinoma cells by regulation of histone H2AX gene expression using a tetracycline turn-off system. Cancer Res 64: 4131.4136
  89. 89. Zhang B, Wang Y, Pang X (2010) Enhanced radiosensitivity of EC109 cells by inhibition of HDAC1 expression. Med Oncol.
  90. 90. Bekker-Jensen S, Rendtlew Danielsen J, Fugger K, Gromova I, Nerstedt A (2010) HERC2 coordinates ubiquitin-dependent assembly of DNA repair factors on damaged chromosomes. Nat Cell Biol 12: 80–86; sup 81–12.
  91. 91. Kassem HSh, Sangar V, Cowan R, Clarke N, Margison GP (2002) A potential role of heat shock proteins and nicotinamide N-methyl transferase in predicting response to radiation in bladder cancer. Int J Cancer 101: 454.460
  92. 92. Hadchity E, Aloy MT, Paulin C, Armandy E, Watkin E (2009) Heat shock protein 27 as a new therapeutic target for radiation sensitization of head and neck squamous cell carcinoma. Mol Ther 17: 1387.1394
  93. 93. Wang X, Hu B, Weiss RS, Wang Y (2006) The effect of Hus1 on ionizing radiation sensitivity is associated with homologous recombination repair but is independent of nonhomologous end-joining. Oncogene 25: 1980.1983
  94. 94. Ishigami T, Uzawa K, Fushimi K, Saito K, Kato Y (2008) Inhibition of ICAM2 induces radiosensitization in oral squamous cell carcinoma cells. Br J Cancer 98: 1357.1365
  95. 95. Ding KK, Shang ZF, Hao C, Xu QZ, Shen JJ (2009) Induced expression of the IER5 gene by gamma-ray irradiation and its involvement in cell cycle checkpoint control and survival. Radiat Environ Biophys 48: 205.213
  96. 96. Yavari K, Taghikhani M, Maragheh MG, Mesbah-Namin SA, Babaei MH (2010) SiRNA-mediated IGF-1R inhibition sensitizes human colon cancer SW480 cells to radiation. Acta Oncol 49: 70.75
  97. 97. Tan W, Huang W, Zhong Q, Schwarzenberger P (2006) IL-17 receptor knockout mice have enhanced myelotoxicity and impaired hemopoietic recovery following gamma irradiation. J Immunol 176: 6186.6193
  98. 98. Eke I, Leonhardt F, Storch K, Hehlgans S, Cordes N (2009) The small molecule inhibitor QLT0267 Radiosensitizes squamous cell carcinoma cells of the head and neck. PLoS One 4: e6434.
  99. 99. Brunner TB, Cengel KA, Hahn SM, Wu J, Fraker DL (2005) Pancreatic cancer cell radiation survival and prenyltransferase inhibition: the role of K-Ras. Cancer Res 65: 8433.8441
  100. 100. Riballo E, Doherty AJ, Dai Y, Stiff T, Oettinger MA (2001) Cellular and biochemical impact of a mutation in DNA ligase IV conferring clinical radiosensitivity. J Biol Chem 276: 31124.31132
  101. 101. Fogarty GB, Muddle R, Sprung CN, Chen W, Duffy D (2010) Unexpectedly severe acute radiotherapy side effects are associated with single nucleotide polymorphisms of the melanocortin-1 receptor. Int J Radiat Oncol Biol Phys 77: 1486.1492
  102. 102. Guoan X, Hanning W, Kaiyun C, Hao L (2010) Adenovirus-mediated siRNA targeting Mcl-1 gene increases radiosensitivity of pancreatic carcinoma cells in vitro and in vivo. Surgery 147: 553.561
  103. 103. Guo W, Ahmed KM, Hui Y, Guo G, Li JJ (2007) siRNA-mediated MDM2 inhibition sensitizes human lung cancer A549 cells to radiation. Int J Oncol 30: 1447.1452
  104. 104. Chetty C, Bhoopathi P, Rao JS, Lakka SS (2009) Inhibition of matrix metalloproteinase-2 enhances radiosensitivity by abrogating radiation-induced FoxM1-mediated G2/M arrest in A549 lung cancer cells. Int J Cancer 124: 2468.2477
  105. 105. Ahn GO, Brown JM (2008) Matrix metalloproteinase-9 is required for tumor vasculogenesis but not for angiogenesis: role of bone marrow-derived myelomonocytic cells. Cancer Cell 13: 193.205
  106. 106. Ahn J, Ambrosone CB, Kanetsky PA, Tian C, Lehman TA (2006) Polymorphisms in genes related to oxidative stress (CAT, MnSOD, MPO, and eNOS) and acute toxicities from radiation therapy following lumpectomy for breast cancer. Clin Cancer Res 12: 7063.7070
  107. 107. Cecchin E, Agostini M, Pucciarelli S, De Paoli A, Canzonieri V (2011) Tumor response is predicted by patient genetic profile in rectal cancer patients treated with neo-adjuvant chemo-radiotherapy. Pharmacogenomics J 11: 214.226
  108. 108. Sheen JH, Dickson RB (2002) Overexpression of c-Myc alters G(1)/S arrest following ionizing radiation. Mol Cell Biol 22: 1819.1833
  109. 109. Popanda O, Tan XL, Ambrosone CB, Kropp S, Helmbold I (2006) Genetic polymorphisms in the DNA double-strand break repair genes XRCC3, XRCC2, and NBS1 are not associated with acute side effects of radiotherapy in breast cancer patients. Cancer Epidemiol Biomarkers Prev 15: 1048.1050
  110. 110. Das A, Boldogh I, Lee JW, Harrigan JA, Hegde ML (2007) The human Werner syndrome protein stimulates repair of oxidative DNA base damage by the DNA glycosylase NEIL1. J Biol Chem 282: 26591.26602
  111. 111. Mi J, Guo C, Brautigan DL, Larner JM (2007) Protein phosphatase-1alpha regulates centrosome splitting through Nek2. Cancer Res 67: 1082.1089
  112. 112. Wang Y, Meng A, Lang H, Brown SA, Konopa JL (2004) Activation of nuclear factor kappaB In vivo selectively protects the murine small intestine against ionizing radiation-induced damage. Cancer Res 64: 6240.6246
  113. 113. Li S, Kuhne WW, Kulharya A, Hudson FZ, Ha K (2009) Involvement of p54(nrb), a PSF partner protein, in DNA double-strand break repair and radioresistance. Nucleic Acids Res 37: 6746.6753
  114. 114. Park S, Ahn JY, Lim MJ, Kim MH, Yun YS (2010) Sustained expression of NADPH oxidase 4 by p38 MAPK-Akt signaling potentiates radiation-induced differentiation of lung fibroblasts. J Mol Med (Berl) 88: 807.816
  115. 115. Sterpone S, Cornetta T, Padua L, Mastellone V, Giammarino D (2010) DNA repair capacity and acute radiotherapy adverse effects in Italian breast cancer patients. Mutat Res 684: 43.48
  116. 116. Zhang B, Wang Y, Liu K, Yang X, Song M (2008) Adenovirus-mediated transfer of siRNA against peroxiredoxin I enhances the radiosensitivity of human intestinal cancer. Biochem Pharmacol 75: 660.667
  117. 117. Fernet M, Ponette V, Deniaud-Alexandre E, Ménissier-De Murcia J, De Murcia G (2000) Poly(ADP-ribose) polymerase, a major determinant of early cell response to ionizing radiation. Int J Radiat Biol 76: 1621.1629
  118. 118. Shan B, Xu J, Zhuo Y, Morris CA, Morris GF (2003) Induction of p53-dependent activation of the human proliferating cell nuclear antigen gene in chromatin by ionizing radiation. J Biol Chem 278: 44009.44017
  119. 119. Hamilton J, Grawenda AM, Bernhard EJ (2009) Phosphatase inhibition and cell survival after DNA damage induced by radiation. Cancer Biol Ther 8: 1577.1586
  120. 120. Prevo R, Deutsch E, Sampson O, Diplexcito J, Cengel K (2008) Class I PI3 kinase inhibition by the pyridinylfuranopyrimidine inhibitor PI-103 enhances tumor radiosensitivity. Cancer Res 68: 5915.5923
  121. 121. Vens C, Dahmen-Mooren E, Verwijs-Janssen M, Blyweert W, Graversen L (2002) The role of DNA polymerase beta in determining sensitivity to ionizing radiation in human tumor cells. Nucleic Acids Res 30: 2995.3004
  122. 122. Rossi M, Demidov ON, Anderson CW, Appella E, Mazur SJ (2008) Induction of PPM1D following DNA-damaging treatments through a conserved p53 response element coincides with a shift in the use of transcription initiation sites. Nucleic Acids Res 36: 7168.7180
  123. 123. Park JJ, Chang HW, Jeong EJ, Roh JL, Choi SH (2009) Peroxiredoxin IV protects cells from radiation-induced apoptosis in head-and-neck squamous cell carcinoma. Int J Radiat Oncol Biol Phys 73: 1196.1202
  124. 124. Abbaszadeh F, Clingen PH, Arlett CF, Plowman PN, Bourton EC (2010) A novel splice variant of the DNA-PKcs gene is associated with clinical and cellular radiosensitivity in a patient with xeroderma pigmentosum. J Med Genet 47: 176.181
  125. 125. Peng Y, Zhang Q, Nagasawa H, Okayasu R, Liber HL (2002) Silencing expression of the catalytic subunit of DNA-dependent protein kinase by small interfering RNA sensitizes human cells for radiation-induced chromosome damage, cell killing, and mutation. Cancer Res 62: 6400.6404
  126. 126. Ogawa K, Boucher Y, Kashiwagi S, Fukumura D, Chen D (2007) Influence of tumor cell and stroma sensitivity on tumor response to radiation. Cancer Res 67: 4016.4021
  127. 127. Chiou SH, Kao CL, Chen YW, Chien CS, Hung SC (2008) Identification of CD133-positive radioresistant cells in atypical teratoid/rhabdoid tumor. PLoS One 3: e2090.
  128. 128. Pazzaglia S, Tanori M, Mancuso M, Rebessi S, Leonardi S (2006) Linking DNA damage to medulloblastoma tumorigenesis in patched heterozygous knockout mice. Oncogene 25: 1165.1173
  129. 129. Lee C, Kim JS, Waldman T (2004) PTEN gene targeting reveals a radiation-induced size checkpoint in human cancer cells. Cancer Res 64: 6906.6914
  130. 130. Chen WS, Yu YC, Lee YJ, Chen JH, Hsu HY (2010) Depletion of securin induces senescence after irradiation and enhances radiosensitivity in human cancer cells regardless of functional p53 expression. Int J Radiat Oncol Biol Phys 77: 566.574
  131. 131. Severin DM, Leong T, Cassidy B, Elsaleh H, Peters L (2001) Novel DNA sequence variants in the hHR21 DNA repair gene in radiosensitive cancer patients. Int J Radiat Oncol Biol Phys 50: 1323.1331
  132. 132. Djuzenova C, Mühl B, Schakowski R, Oppitz U, Flentje M (2004) Normal expression of DNA repair proteins, hMre11, Rad50 and Rad51 but protracted formation of Rad50 containing foci in X-irradiated skin fibroblasts from radiosensitive cancer patients. Br J Cancer 90: 2356.2363
  133. 133. Welsh JW, Ellsworth RK, Kumar R, Fjerstad K, Martinez J (2009) Rad51 protein expression and survival in patients with glioblastoma multiforme. Int J Radiat Oncol Biol Phys 74: 1251.1255
  134. 134. Raaphorst GP, Leblanc M, Li LF (2005) A comparison of response to cisplatin, radiation and combined treatment for cells deficient in recombination repair pathways. Anticancer Res 25: 53.58
  135. 135. Singhal SS, Roth C, Leake K, Singhal J, Yadav S (2009) Regression of prostate cancer xenografts by RLIP76 depletion. Biochem Pharmacol 77: 1074.1083
  136. 136. Singhal SS, Sehrawat A, Sahu M, Singhal P, Vatsyayan R (2010) Rlip76 transports sunitinib and sorafenib and mediates drug resistance in kidney cancer. Int J Cancer 126: 1327.1338
  137. 137. Kuo ML, Hwang HS, Sosnay PR, Kunugi KA, Kinsella TJ (2003) Overexpression of the R2 subunit of ribonucleotide reductase in human nasopharyngeal cancer cells reduces radiosensitivity. Cancer J 9: 277.285
  138. 138. Yokomakura N, Natsugoe S, Okumura H, Ikeda R, Uchikado Y (2007) Improvement in radiosensitivity using small interfering RNA targeting p53R2 in esophageal squamous cell carcinoma. Oncol Rep 18: 561.567
  139. 139. Jia L, Yang J, Hao X, Zheng M, He H (2010) Validation of SAG/RBX2/ROC2 E3 ubiquitin ligase as an anticancer and radiosensitizing target. Clin Cancer Res 16: 814.824
  140. 140. Chung HJ, Yoon SI, Shin SH, Koh YA, Lee SJ (2006) p53-Mediated enhancement of radiosensitivity by selenophosphate synthetase 1 overexpression. J Cell Physiol 209: 131.141
  141. 141. Milliat F, Sabourin JC, Tarlet G, Holler V, Deutsch E (2008) Essential role of plasminogen activator inhibitor type-1 in radiation enteropathy. Am J Pathol 172: 691.701
  142. 142. Chang CJ, Hsu CC, Yung MC, Chen KY, Tzao C (2009) Enhanced radiosensitivity and radiation-induced apoptosis in glioma CD133-positive cells by knockdown of SirT1 expression. Biochem Biophys Res Commun 380: 236.242
  143. 143. Peter Y, Rotman G, Lotem J, Elson A, Shiloh Y (2001) Elevated Cu/Zn-SOD exacerbates radiation sensitivity and hematopoietic abnormalities of Atm-deficient mice. EMBO J 20: 1538.1546
  144. 144. Burri RJ, Stock RG, Cesaretti JA, Atencio DP, Peters S (2008) Association of single nucleotide polymorphisms in SOD2, XRCC1 and XRCC3 with susceptibility for the development of adverse effects resulting from radiotherapy for prostate cancer. Radiat Res 170: 49.59
  145. 145. Andreassen CN, Alsner J, Overgaard M, Overgaard J (2003) Prediction of normal tissue radiosensitivity from polymorphisms in candidate genes. Radiother Oncol 69: 127.135
  146. 146. Dittmann K, Mayer C, Kehlbach R, Rodemann HP (2008) Radiation-induced caveolin-1 associated EGFR internalization is linked with nuclear EGFR transport and activation of DNA-PK. Mol Cancer 7: 69.
  147. 147. Hui Z, Tretiakova M, Zhang Z, Li Y, Wang X (2009) Radiosensitization by inhibiting STAT1 in renal cell carcinoma. Int J Radiat Oncol Biol Phys 73: 288.295
  148. 148. Bonner JA, Trummell HQ, Willey CD, Plants BA, Raisch KP (2009) Inhibition of STAT-3 results in radiosensitization of human squamous cell carcinoma. Radiother Oncol 92: 339.344
  149. 149. Li X, Wang H, Lu X, Di B (2010) STAT3 blockade with shRNA enhances radiosensitivity in Hep-2 human laryngeal squamous carcinoma cells. Oncol Rep 23: 345.353
  150. 150. Giotopoulos G, Symonds RP, Foweraker K, Griffin M, Peat I (2007) The late radiotherapy normal tissue injury phenotypes of telangiectasia, fibrosis and atrophy in breast cancer patients have distinct genotype-dependent causes. Br J Cancer 96: 1001.1007
  151. 151. Meyer A, Dörk T, Bogdanova N, Brinkhaus MJ, Wiese B (2009) TGFB1 gene polymorphism Leu10Pro (c.29T>C), prostate cancer incidence and quality of life in patients treated with brachytherapy. World J Urol 27: 371.377
  152. 152. Yuan X, Liao Z, Liu Z, Wang LE, Tucker SL (2009) Single nucleotide polymorphism at rs1982073:T869C of the TGFbeta 1 gene is associated with the risk of radiation pneumonitis in patients with non-small-cell lung cancer treated with definitive radiotherapy. J Clin Oncol 27: 3370.3378
  153. 153. Kim dR, Laurence B, Jan VM, Wilfried dN, Hubert T (2010) Association of TGFbeta1 polymorphisms involved in radiation toxicity with TGFbeta1 secretion in vitro. Cytokine 50: 37.41
  154. 154. Alsbeih G, Al-Harbi N, Al-Hadyan K, El-Sebaie M, Al-Rajhi N (2010) Association between normal tissue complications after radiotherapy and polymorphic variations in TGFB1 and XRCC1 genes. Radiat Res 173: 505.511
  155. 155. Zhang M, Qian J, Xing X, Kong FM, Zhao L (2008) Inhibition of the tumor necrosis factor-alpha pathway is radioprotective for the lung. Clin Cancer Res 14: 1868.1876
  156. 156. Shankar S, Singh TR, Chen X, Thakkar H, Firnin J (2004) The sequential treatment with ionizing radiation followed by TRAIL/Apo-2L reduces tumor growth and induces apoptosis of breast tumor xenografts in nude mice. Int J Oncol 24: 1133.1140
  157. 157. Jiao Y, Ge CM, Meng QH, Cao JP, Tong J (2007) Adenovirus-mediated expression of Tob1 sensitizes breast cancer cells to ionizing radiation. Acta Pharmacol Sin 28: 1628.1636
  158. 158. Terry SY, Riches AC, Bryant PE (2009) Suppression of topoisomerase IIalpha expression and function in human cells decreases chromosomal radiosensitivity. Mutat Res 663: 40.45
  159. 159. Thurfjell N, Coates PJ, Vojtesek B, Benham-Motlagh P, Eisold M (2005) Endogenous p63 acts as a survival factor for tumour cells of SCCHN origin. Int J Mol Med 16: 1065.1070
  160. 160. Firat E, Tsurumi C, Gaedicke S, Huai J, Niedermann G (2009) Tripeptidyl peptidase II plays a role in the radiation response of selected primary cell types but not based on nuclear translocation and p53 stabilization. Cancer Res 69: 3325.3331
  161. 161. Zheng M, Morgan-Lappe SE, Yang J, Bockbrader KM, Pamarthy D (2008) Growth inhibition and radiosensitization of glioblastoma and lung cancer cells by small interfering RNA silencing of tumor necrosis factor receptor-associated factor 2. Cancer Res 68: 7570.7578
  162. 162. Demizu Y, Sasaki R, Trachootham D, Pelicano H, Colacino JA (2008) Alterations of cellular redox state during NNK-induced malignant transformation and resistance to radiation. Antioxid Redox Signal 10: 951.961
  163. 163. Javvadi P, Hertan L, Kosoff R, Datta T, Kolev J (2010) Thioredoxin reductase-1 mediates curcumin-induced radiosensitization of squamous carcinoma cells. Cancer Res 70: 1941.1950
  164. 164. Li XL, Meng QH, Fan SJ (2009) Adenovirus-mediated expression of UHRF1 reduces the radiosensitivity of cervical cancer HeLa cells to gamma-irradiation. Acta Pharmacol Sin 30: 458.466
  165. 165. Yan J, Kim YS, Yang XP, Li LP, Liao G (2007) The ubiquitin-interacting motif containing protein RAP80 interacts with BRCA1 and functions in DNA damage repair response. Cancer Res 67: 6647.6656
  166. 166. Langsenlehner T, Kapp KS, Langsenlehner U (2008) TGFB1 single-nucleotide polymorphisms are associated with adverse quality of life in prostate cancer patients treated with radiotherapy. In regard to Peters, et al. (Int J Radiat Oncol Biol Phys 2008;70: 752–759). Int J Radiat Oncol Biol Phys 71: 960; author reply 960–961.
  167. 167. Clark AJ, Chan DC, Chen MY, Fillmore H, Dos Santos WG (2007) Down-regulation of Wilms' tumor 1 expression in glioblastoma cells increases radiosensitivity independently of p53. J Neurooncol 83: 163.172
  168. 168. Giagkousiklidis S, Vellanki SH, Debatin KM, Fulda S (2007) Sensitization of pancreatic carcinoma cells for gamma-irradiation-induced apoptosis by XIAP inhibition. Oncogene 26: 7006.7016
  169. 169. Wang R, Li B, Wang X, Lin F, Gao P (2009) Inhibiting XIAP expression by RNAi to inhibit proliferation and enhance radiosensitivity in laryngeal cancer cell line. Auris Nasus Larynx 36: 332.339
  170. 170. Wiebalk K, Schmezer P, Kropp S, Chang-Claude J, Celebi O (2007) In vitro radiation-induced expression of XPC mRNA as a possible biomarker for developing adverse reactions during radiotherapy. Int J Cancer 121: 2340.2345
  171. 171. Moullan N, Cox DG, Angèle S, Romestaing P, Gérard JP (2003) Polymorphisms in the DNA repair gene XRCC1, breast cancer risk, and response to radiotherapy. Cancer Epidemiol Biomarkers Prev 12: 1168.1174
  172. 172. De Ruyck K, Wilding CS, Van Eijkeren M, Morthier R, Tawn EJ (2005) Microsatellite polymorphisms in DNA repair genes XRCC1, XRCC3 and XRCC5 in patients with gynecological tumors: association with late clinical radiosensitivity and cancer incidence. Radiat Res 164: 237.244
  173. 173. Brem R, Cox DG, Chapot B, Moullan N, Romestaing P (2006) The XRCC1–77T->C variant: haplotypes, breast cancer risk, response to radiotherapy and the cellular response to DNA damage. Carcinogenesis 27: 2469.2474
  174. 174. Price EA, Bourne SL, Radbourne R, Lawton PA, Lamerdin J (1997) Rare microsatellite polymorphisms in the DNA repair genes XRCC1, XRCC3 and XRCC5 associated with cancer in patients of varying radiosensitivity. Somat Cell Mol Genet 23: 237.247
  175. 175. De Ruyck K, Van Eijkeren M, Claes K, Morthier R, De Paepe A (2005) Radiation-induced damage to normal tissues after radiotherapy in patients treated for gynecologic tumors: association with single nucleotide polymorphisms in XRCC1, XRCC3, and OGG1 genes and in vitro chromosomal radiosensitivity in lymphocytes. Int J Radiat Oncol Biol Phys 62: 1140.1149
  176. 176. Werbrouck J, De Ruyck K, Duprez F, Veldeman L, Claes K (2009) Acute normal tissue reactions in head-and-neck cancer patients treated with IMRT: influence of dose and association with genetic polymorphisms in DNA DSB repair genes. Int J Radiat Oncol Biol Phys 73: 1187.1195
  177. 177. Wachsberger PR, Li WH, Guo M, Chen D, Cheong N (1999) Rejoining of DNA double-strand breaks in Ku80-deficient mouse fibroblasts. Radiat Res 151: 398.407
  178. 178. Vandersickel V, Mancini M, Slabbert J, Marras E, Thierens H (2010) The radiosensitizing effect of Ku70/80 knockdown in MCF10A cells irradiated with X-rays and p(66)+Be(40) neutrons. Radiat Oncol 5: 30.
  179. 179. Negroni A, Stronati L, Grollino MG, Barattini P, Gumiero D (2008) Radioresistance in a tumour cell line correlates with radiation inducible Ku 70/80 end-binding activity. Int J Radiat Biol 84: 265.276
  180. 180. Vandersickel V, Mancini M, Marras E, Willems P, Slabbert J (2010) Lentivirus-mediated RNA interference of Ku70 to enhance radiosensitivity of human mammary epithelial cells. Int J Radiat Biol 86: 114.124
  181. 181. Urano M, He F, Minami A, Ling CC, Li GC (2010) Response to multiple radiation doses of human colorectal carcinoma cells infected with recombinant adenovirus containing dominant-negative Ku70 fragment. Int J Radiat Oncol Biol Phys 77: 877.885
  182. 182. Sakata K, Someya M, Matsumoto Y, Hareyama M (2007) Ability to repair DNA double-strand breaks related to cancer susceptibility and radiosensitivity. Radiat Med 25: 433.438
  183. 183. Du XL, Jiang T, Wen ZQ, Gao R, Cui M (2009) Silencing of heat shock protein 70 expression enhances radiotherapy efficacy and inhibits cell invasion in endometrial cancer cell line. Croat Med J 50: 143.150
  184. 184. O'Malley BW, Li D, Carney J, Rhee J, Suntharalingam M (2003) Molecular disruption of the MRN(95) complex induces radiation sensitivity in head and neck cancer. Laryngoscope 113: 1588.1594
  185. 185. Cengel KA, McKenna WG (2005) Molecular targets for altering radiosensitivity: lessons from Ras as a pre-clinical and clinical model. Crit Rev Oncol Hematol 55: 103.116
  186. 186. Pawlik TM, Keyomarsi K (2004) Role of cell cycle in mediating sensitivity to radiotherapy. Int J Radiat Oncol Biol Phys 59: 928.942
  187. 187. Tusher VG, Tibshirani R, Chu G (2001) Significance analysis of microarrays applied to the ionizing radiation response. Proc Natl Acad Sci U S A 98: 5116.5121
  188. 188. Merbl Y, Kirschner MW (2009) Large-scale detection of ubiquitination substrates using cell extracts and protein microarrays. Proc Natl Acad Sci U S A 106: 2543.2548
  189. 189. Moussay E, Palissot V, Vallar L, Poirel HA, Wenner T (2010) Determination of genes and microRNAs involved in the resistance to fludarabine in vivo in chronic lymphocytic leukemia. Mol Cancer 9: 115.
  190. 190. Blalock EM, Chen KC, Stromberg AJ, Norris CM, Kadish I (2005) Harnessing the power of gene microarrays for the study of brain aging and Alzheimer's disease: statistical reliability and functional correlation. Ageing Res Rev 4: 481.512