Radiomics: Images Are More than Pictures, They Are Data

Robert J Gillies, Paul E Kinahan, Hedvig Hricak, Robert J Gillies, Paul E Kinahan, Hedvig Hricak

Abstract

In the past decade, the field of medical image analysis has grown exponentially, with an increased number of pattern recognition tools and an increase in data set sizes. These advances have facilitated the development of processes for high-throughput extraction of quantitative features that result in the conversion of images into mineable data and the subsequent analysis of these data for decision support; this practice is termed radiomics. This is in contrast to the traditional practice of treating medical images as pictures intended solely for visual interpretation. Radiomic data contain first-, second-, and higher-order statistics. These data are combined with other patient data and are mined with sophisticated bioinformatics tools to develop models that may potentially improve diagnostic, prognostic, and predictive accuracy. Because radiomics analyses are intended to be conducted with standard of care images, it is conceivable that conversion of digital images to mineable data will eventually become routine practice. This report describes the process of radiomics, its challenges, and its potential power to facilitate better clinical decision making, particularly in the care of patients with cancer.

Figures

Figure 1:
Figure 1:
Flowchart shows the process of radiomics and the use of radiomics in decision support. Patient work-up requires information from disparate sources to be combined into a coherent model to describe where the lesion is, what it is, and what it is doing. Radiomics begins with acquisition of high-quality images. From these images, a region of interest (ROI) that contains either the whole tumor or subregions (ie, habitats) within the tumor can be identified. These are segmented with operator edits and are eventually rendered in three dimensions (3D). Quantitative features are extracted from these rendered volumes to generate a report, which is placed in a database along with other data, such as clinical and genomic data. These data are then mined to develop diagnostic, predictive, or prognostic models for outcomes of interest.
Figure 2:
Figure 2:
Habitats in a patient with glioblastoma multiforme. Habitats were defined by combining unenhanced and contrast-enhanced T1-weighted, 120-msec echo time T2-weighted, and fluid-attenuated inversion recovery (FLAIR) images. Data from each acquisition were sorted into low and high values with automated histogram analyses, yielding a potential for eight different combinations. In practice, only four distinct combinations were observed. They correspond to the red (low T1, high T2 and FLAIR), yellow (low T1 and T2, high FLAIR), blue (High T1 and FLAIR, low T2), and green (high T1, low FLAIR and T2) areas. Notably, while the identities of individual voxels were determined mathematically, they spatially clustered into contiguous regions reflecting different physiologic microenvironments. (Image courtesy of R. A. Gatenby.)
Figure 3:
Figure 3:
Covariance matrix of radiomic features. A total of 219 features were extracted from each non–small cell lung cancer tumor in 235 patients. Across all tumors, each feature was individually compared with all other features by using regression analysis, thereby generating correlation coefficients (R2). Individual features were then clustered and plotted along both axes, and R2 is shown as a heat map, with areas of high correlation (R2 > 0.95) shown in red. Thus, each of the red squares along the diagonal contains a group of features that are highly correlated with one another and are thus redundant. For data analysis, one feature was chosen to be representative of each of these groups. The representative feature chosen was the one that had the highest natural biologic range (interpatient variability) across the entire patient data set, with the explicit assumption that features that show the highest interpatient variability will be the most informative. (Image courtesy of Y. Balagurunathan.)
Figure 4:
Figure 4:
Application of texture analysis to T2-weighted MR images and ADC maps of prostate cancer. A lesion in the transition zone is barely discernible on the T2-weighted image (top left) and has higher conspicuity on the ADC map (top right). Texture features were computed on a per-voxel basis (using a 5 × 5 × 1 pixel window) from manually segmented regions of interest identifying the normal peripheral zone (outlined in blue) and cancer (outlined in red). From the computed texture features, a machine learning method was applied to distinguish between normal and cancerous structures and to stratify the Gleason patterns. Heat map images show clear differences between healthy tissue and cancer and depict intratumoral heterogeneity that may be useful in assessing tumor aggressiveness and informing fused MR imaging–ultrasonography biopsy.
Figure 5:
Figure 5:
Attenuation gradients of lung CT images. Data are representative of patients with (left) and without (right) recurrence after lobectomy. Although differences in texture were visible on the CT images (top), the color-coded attenuation maps (bottom) more dramatically show intratumoral complexity. Maps were generated by separating the attenuation into quartiles, with hotter colors representing higher attenuation.
Figure 6:
Figure 6:
Application of texture analysis to CT images of bladder cancer. On original contrast-enhanced CT image of bladder cancer (top left), a high-attenuation lesion is clearly visible, and there is some evidence of intratumoral heterogeneity. However, when the texture features of energy (top right), entropy (bottom left), and homogeneity (bottom right) are displayed over the source image, intratumoral heterogeneity can be readily appreciated. Other studies have shown that higher intratumoral heterogeneity is associated with a worse prognosis.
Figure 7:
Figure 7:
Application of radiomics to FDG-avid lymph nodes on PET and CT images in a patient with metastatic breast cancer. Left: Standard PET image shows there is little evidence of intranodal heterogeneity. Right: CT image shows calculation and display of the Haralick co-occurrence statistics with a 9 × 9 × 9 voxel matrix and clearly reveals some areas with lower co-occurrence (red), which have higher regional heterogeneity and would therefore be considered more suspicious for cancer. The results were used to select lymph nodes for image-informed biopsy.

References

    1. Institute of Medicine I . Evolution of Translational Omics. Lessons learned and the path forward. Washington, DC: National Academies Press, 2012.
    1. Giger ML. Update on the potential of computer-aided diagnosis for breast cancer. Future Oncol 2010;6(1):1–4.
    1. Doi K. Computer-aided diagnosis in medical imaging: historical review, current status and future potential. Comput Med Imaging Graph 2007;31(4-5):198–211.
    1. Guo Z, Shu Y, Zhou H, Zhang W, Wang H. Radiogenomics helps to achieve personalized therapy by evaluating patient responses to radiation treatment. Carcinogenesis 2015;36(3):307–317.
    1. West CM, Barnett GC. Genetics and genomics of radiotherapy toxicity: towards prediction. Genome Med 2011;3(8):52.
    1. Wibmer A, Hricak H, Gondo T, et al. . Haralick texture analysis of prostate MRI: utility for differentiating non-cancerous prostate from prostate cancer and differentiating prostate cancers with different Gleason scores. Eur Radiol 2015;25(10):2840–2850.
    1. Coroller TP, Grossmann P, Hou Y, et al. . CT-based radiomic signature predicts distant metastasis in lung adenocarcinoma. Radiother Oncol 2015;114(3):345–350.
    1. Committee on the Review of Omics-Based Tests for Predicting Patient Outcomes in Clinical Trials, Board on Health Care Services, Board on Health Sciences Policy, Institute of Medicine . Evolution of Translational Omics: Lessons Learned and the Path Forward. Micheel CM, Nass SJ, Omenn GS, eds. Washington, DC: National Academies Press.
    1. Bates SE. It’s all about the test: the complexity of companion diagnostic co-development in personalized medicine. Clin Cancer Res 2014;20(6):1418.
    1. Collins FS, Varmus H. A new initiative on precision medicine. N Engl J Med 2015;372(9):793–795.
    1. Office of the Press Secretary . Fact sheet: President Obama’s precision medicine initiative. The White House Web site. . Updated January 30, 2015. Accessed April 4, 2015.
    1. Siegel R, Ma J, Zou Z, Jemal A. Cancer statistics, 2014. CA Cancer J Clin 2014;64(1):9–29.
    1. Yachida S, Jones S, Bozic I, et al. . Distant metastasis occurs late during the genetic evolution of pancreatic cancer. Nature 2010;467(7319):1114–1117.
    1. Gerlinger M, Rowan AJ, Horswell S, et al. . Intratumor heterogeneity and branched evolution revealed by multiregion sequencing. N Engl J Med 2012;366(10):883–892.
    1. Sottoriva A, Spiteri I, Piccirillo SG, et al. . Intratumor heterogeneity in human glioblastoma reflects cancer evolutionary dynamics. Proc Natl Acad Sci U S A 2013;110(10):4009–4014.
    1. Sequist LV, Martins RG, Spigel D, et al. . First-line gefitinib in patients with advanced non-small-cell lung cancer harboring somatic EGFR mutations. J Clin Oncol 2008;26(15):2442–2449.
    1. Burstein HJ, Harris LN, Marcom PK, et al. . Trastuzumab and vinorelbine as first-line therapy for HER2-overexpressing metastatic breast cancer: multicenter phase II trial with clinical outcomes, analysis of serum tumor markers as predictive factors, and cardiac surveillance algorithm. J Clin Oncol 2003;21(15):2889–2895.
    1. Wagle N, Emery C, Berger MF, et al. . Dissecting therapeutic resistance to RAF inhibition in melanoma by tumor genomic profiling. J Clin Oncol 2011;29(22):3085–3096.
    1. Sequist LV, Waltman BA, Dias-Santagata D, et al. . Genotypic and histological evolution of lung cancers acquiring resistance to EGFR inhibitors. Sci Transl Med 2011;3(75):75ra26.
    1. Clauson J, Hsieh YC, Acharya S, Rademaker AW, Morrow M. Results of the Lynn Sage Second-Opinion Program for local therapy in patients with breast carcinoma: changes in management and determinants of where care is delivered. Cancer 2002;94(4):889–894.
    1. Nguyen PL, Schultz D, Renshaw AA, et al. . The impact of pathology review on treatment recommendations for patients with adenocarcinoma of the prostate. Urol Oncol 2004;22(4):295–299.
    1. Staradub VL, Messenger KA, Hao N, Wiley EL, Morrow M. Changes in breast cancer therapy because of pathology second opinions. Ann Surg Oncol 2002;9(10):982–987.
    1. Kronz JD, Westra WH, Epstein JI. Mandatory second opinion surgical pathology at a large referral hospital. Cancer 1999;86(11):2426–2435.
    1. Kumar V, Gu Y, Basu S, et al. . Radiomics: the process and the challenges. Magn Reson Imaging 2012;30(9):1234–1248.
    1. Lambin P, Rios-Velazquez E, Leijenaar R, et al. . Radiomics: extracting more information from medical images using advanced feature analysis. Eur J Cancer 2012;48(4):441–446.
    1. Clarke LP, Sriram RD, Schilling LB. Imaging as a Biomarker: Standards for Change Measurements in Therapy workshop summary. Acad Radiol 2008;15(4):501–530.
    1. Shankar LK, Hoffman JM, Bacharach S, et al. . Consensus recommendations for the use of 18F-FDG PET as an indicator of therapeutic response in patients in National Cancer Institute Trials. J Nucl Med 2006;47(6):1059–1066.
    1. Clarke LP, Nordstrom RJ, Zhang H, et al. . The Quantitative Imaging Network: NCI’s Historical Perspective and Planned Goals. Transl Oncol 2014;7(1):1–4.
    1. Buckler AJ, Bresolin L, Dunnick NR, Sullivan DC; for the Group. A collaborative enterprise for multi-stakeholder participation in the advancement of quantitative imaging. Radiology 2011;258(3):906–914.
    1. Kinahan P, Wahl R, Shao L, Frank R, Perlman E. The QIBA profile for quantitative FDG-PET/CT oncology imaging [abstr]. J Nucl Med 2014;55(Suppl 1):1520.
    1. FDG-PET/CT Technical Committee . FDG-PET/CT as an Imaging Biomarker Measuring Response to Cancer Therapy, Quantitative Imaging Biomarkers Alliance, Version 1.05, Publicly Reviewed Version. RSNA Web site. . Published December 11, 2013. Accessed May 20, 2015.
    1. van Persijn van Meerten EL, Gelderblom H, Bloem JL. RECIST revised: implications for the radiologist: a review article on the modified RECIST guideline. Eur Radiol 2010;20(6):1456–1467.
    1. Burnside ES, Sickles EA, Bassett LW, et al. . The ACR BI-RADS experience: learning from history. J Am Coll Radiol 2009;6(12):851–860.
    1. Bomers JG, Barentsz JO. Standardization of multiparametric prostate MR imaging using PI-RADS. BioMed Res Int 2014;2014:431680.
    1. Kazerooni EA, Armstrong MR, Amorosa JK, et al. . ACR CT accreditation program and the lung cancer screening program designation. J Am Coll Radiol 2015;12(1):38–42.
    1. Gatenby RA, Grove O, Gillies RJ. Quantitative imaging in cancer evolution and ecology. Radiology 2013;269(1):8–15.
    1. Zhou M, Hall L, Goldgof D, et al. . Radiologically defined ecological dynamics and clinical outcomes in glioblastoma multiforme: preliminary results. Transl Oncol 2014;7(1):5–13.
    1. Rios Velazquez E, Aerts HJ, Gu Y, et al. . A semiautomatic CT-based ensemble segmentation of lung tumors: comparison with oncologists’ delineations and with the surgical specimen. Radiother Oncol 2012;105(2):167–173.
    1. van Dam IE, van Sörnsen de Koste JR, Hanna GG, Muirhead R, Slotman BJ, Senan S. Improving target delineation on 4-dimensional CT scans in stage I NSCLC using a deformable registration tool. Radiother Oncol 2010;96(1):67–72.
    1. Segal E, Sirlin CB, Ooi C, et al. . Decoding global gene expression programs in liver cancer by noninvasive imaging. Nat Biotechnol 2007;25(6):675–680.
    1. Haralick RM, Shanmugam K, Dinstein I. Textural features for image classification. IEEE Trans Syst Man Cybern 1973;SMC-3(6):610–621.
    1. Davnall F, Yip CS, Ljungqvist G, et al. . Assessment of tumor heterogeneity: an emerging imaging tool for clinical practice? Insights Imaging 2012;3(6):573–589.
    1. O’Connor JP, Rose CJ, Waterton JC, Carano RA, Parker GJ, Jackson A. Imaging intratumor heterogeneity: role in therapy response, resistance, and clinical outcome. Clin Cancer Res 2015;21(2):249–257.
    1. Rose CJ, Mills SJ, O’Connor JP, et al. . Quantifying spatial heterogeneity in dynamic contrast-enhanced MRI parameter maps. Magn Reson Med 2009;62(2):488–499.
    1. Larkin TJ, Canuto HC, Kettunen MI, et al. . Analysis of image heterogeneity using 2D Minkowski functionals detects tumor responses to treatment. Magn Reson Med 2014;71(1):402–410.
    1. Grossmann P, Grove O, El-Hachem N, et al. . Identification of molecular phenotypes in lung cancer by integrating radiomics and genomics. Sci Transl Med (in press).
    1. Aerts HJ, Velazquez ER, Leijenaar RT, et al. . Decoding tumour phenotype by noninvasive imaging using a quantitative radiomics approach. Nat Commun 2014;5:4006.
    1. Pickup L, Talwar A, Stalin S, et al. . Lung nodule classification using learnt texture features on a single patient population [abstr]. In: Radiological Society of North America Scientific Assembly and Annual Meeting Program. Oak Brook, Ill: Radiological Society of North America, 2015; SSM06.
    1. Schabath MB, Massion PP, Thompson ZJ, et al. . Behavior of incidence lung cancers in the CT arm of the National Lung Screening Trial. JAMA (in press).
    1. Balagurunathan Y, Kumar V, Gu Y, et al. . Test-retest reproducibility analysis of lung CT image features. J Digit Imaging 2014;27(6):805–823.
    1. Lin LI. A concordance correlation coefficient to evaluate reproducibility. Biometrics 1989;45(1):255–268.
    1. Carrasco JL, King TS, Chinchilli VM. The concordance correlation coefficient for repeated measures estimated by variance components. J Biopharm Stat 2009;19(1):90–105.
    1. Fehr D, Veeraraghavan H, Wibmer A, et al. . Automatic classification of prostate cancer Gleason scores from multi-parametric magnetic resonance images. Proc Natl Acad Sci USA (in press).
    1. Diehn M, Nardini C, Wang DS, et al. . Identification of noninvasive imaging surrogates for brain tumor gene-expression modules. Proc Natl Acad Sci U S A 2008;105(13):5213–5218.
    1. Kuo MD, Gollub J, Sirlin CB, Ooi C, Chen X. Radiogenomic analysis to identify imaging phenotypes associated with drug response gene expression programs in hepatocellular carcinoma. J Vasc Interv Radiol 2007;18(7):821–831.
    1. Grove O, Berglund AE, Schabath MB, et al. . Quantitative computed tomographic descriptors associate tumor shape complexity and intratumor heterogeneity with prognosis in lung adenocarcinoma. PLoS One 2015;10(3):e0118261.
    1. Gevaert O, Mitchell LA, Achrol AS, et al. . Glioblastoma multiforme: exploratory radiogenomic analysis by using quantitative image features. Radiology 2014;273(1):168–174.
    1. Vignati A, Mazzetti S, Giannini V, et al. . Texture features on T2-weighted magnetic resonance imaging: new potential biomarkers for prostate cancer aggressiveness. Phys Med Biol 2015;60(7):2685–2701.
    1. Teruel JRHM, Heldahl MG, Goa PE, et al. . Dynamic contrast-enhanced MRI texture analysis for pretreatment prediction of clinical and pathological response to neoadjuvant chemotherapy in patients with locally advanced breast cancer. NMR Biomed 2014;27(8):887–896.
    1. Klaeser B, Wiskirchen J, Wartenberg J, et al. . PET/CT-guided biopsies of metabolically active bone lesions: applications and clinical impact. Eur J Nucl Med Mol Imaging 2010;37(11):2027–2036.
    1. Tatli S, Gerbaudo VH, Mamede M, Tuncali K, Shyn PB, Silverman SG. Abdominal masses sampled at PET/CT-guided percutaneous biopsy: initial experience with registration of prior PET/CT images. Radiology 2010;256(1):305–311.
    1. Simon RM, Paik S, Hayes DF. Use of archived specimens in evaluation of prognostic and predictive biomarkers. J Natl Cancer Inst 2009;101(21):1446–1452.
    1. Hayes DF, Bast RC, Desch CE, et al. . Tumor marker utility grading system: a framework to evaluate clinical utility of tumor markers. J Natl Cancer Inst 1996;88(20):1456–1466.
    1. Chalmers I, Glasziou P. Avoidable waste in the production and reporting of research evidence. Lancet 2009;374(9683):86–89.
    1. Simera I, Moher D, Hirst A, Hoey J, Schulz KF, Altman DG. Transparent and accurate reporting increases reliability, utility, and impact of your research: reporting guidelines and the EQUATOR Network. BMC Med 2010;8:24.
    1. Prinz F, Schlange T, Asadullah K. Believe it or not: how much can we rely on published data on potential drug targets? Nat Rev Drug Discov 2011;10(9):712.
    1. Begley CG, Ellis LM. Drug development: Raise standards for preclinical cancer research. Nature 2012;483(7391):531–533.
    1. Journals unite for reproducibility. Nature 2014;515(7525):7.
    1. McNutt M. Journals unite for reproducibility. Science 2014;346(6210):679.
    1. Landis SC, Amara SG, Asadullah K, et al. . A call for transparent reporting to optimize the predictive value of preclinical research. Nature 2012;490(7419):187–191.
    1. Drawing Causal Inference from Big Data . Arthur M. Sackler Colloquia of the National Academy of Sciences. . Published April 2, 2015. Accessed May 5, 2015.
    1. Nelson B. Data sharing: empty archives. Nature 2009;461(7261):160–163.
    1. Devidas M, London WB, Anderson JR. The use of central laboratories and remote electronic data capture to risk-adjust therapy for pediatric acute lymphoblastic leukemia and neuroblastoma. Semin Oncol 2010;37(1):53–59.
    1. Heilbrun ME. Evaluating RadLex and real world radiology reporting: are we there yet? Acad Radiol 2013;20(11):1327–1328.
    1. Kahn CE, Jr. Annotation of figures from the biomedical imaging literature: a comparative analysis of RadLex and other standardized vocabularies. Acad Radiol 2014;21(3):384–392.
    1. Mabotuwana T, Lee MC, Cohen-Solal EV, Chang P. Mapping institution-specific study descriptions to RadLex Playbook entries. J Digit Imaging 2014;27(3):321–330.
    1. Clark K, Vendt B, Smith K, et al. . The Cancer Imaging Archive (TCIA): maintaining and operating a public information repository. J Digit Imaging 2013;26(6):1045–1057.

Source: PubMed

3
Předplatit