Open access peer-reviewed chapter

Role of Biomarkers in Hepatocellular Carcinoma and Their Disease Progression

Written By

S.S. Haque, Ravi Bhushan Raman and Mehboobus Salam

Submitted: 13 May 2022 Reviewed: 14 June 2022 Published: 19 August 2022

DOI: 10.5772/intechopen.105856

From the Edited Volume

Liver Cancer - Genesis, Progression and Metastasis

Edited by Mark Feitelson and Alla Arzumanyan

Chapter metrics overview

131 Chapter Downloads

View Full Metrics

Abstract

Hepatocellular carcinoma (HCC) is one of the third leading and common lethal cancers worldwide. Early detection of tumorigenesis of hepatocellular carcinoma is through ultrasonography, computerized tomography (CT) scans, and magnetic resonance imaging (MRI) scans; however, these methods are not up to the mark, so a search for an efficient biomarker for early diagnosis and treatment of hepatocarcinogenesis is important. Proteomic and genomic approaches aid to develop new promising biomarkers for the diagnosis of HCC at the early stages. These biomarkers not only help in prognosis but also provide better therapeutic intervention against HCC. Among the different biomarker candidates, liquid biopsy [including circulating tumor cells (CTCs) and circulating tumor DNA (ctDNA)] has recently emerged as a noninvasive detection technique for the characterization of circulating cells, providing a strong basis and early diagnosis for the individualized treatment of patients. This review provides the current understanding of HCC biomarkers that predict the risk of HCC recurrence.

Keywords

  • hepatocellular carcinoma
  • biomarkers
  • diagnosis
  • prognosis

1. Introduction

Hepatocellular carcinoma (HCC) is the fourth most common type of primary liver malignancy and is prevalent in almost most parts of the world [1]. Hepatitis B virus or hepatitis C virus is the most important contributing factor [2]. Liver biopsy is invasive but still the gold standard for determining the presence of the tumor. However, in some cases, preoperative biopsy of the tumor samples does not accurately predict microvascular invasion when compared with the final specimen examination after liver resection [3, 4]. Due to the risk of needle tract tumor seeding and other limitations, preoperative biopsy is not currently recommended for routine HCC evaluation. Biomarkers are biological indicator molecules of physiological and disease states. The detection of HCC in the blood is simple, noninvasive, reliable, and the best choice for screening of HCC. In the coming year, newly identified biomarkers will be more thoroughly evaluated for the diagnosis of HCC (Figure 1). In this review, some promising biomarkers are discussed.

Figure 1.

Biomarker for HCC.

As a clinical biomarker, D-dimer is commonly used in HCC because portal vein thrombosis (PVT) is one of the important complications of HCC. D-dimer is the end product of fibrin degradation and is produced by fibrinogen, so D-dimer is a marker of endogenous fibrinolysis and its levels increase with fibrinolysis [5, 6]. Prealbumin (PA) or transthyretin is a homotetrameric protein that has been recently identified as a biomarker by the liver [7]. In contrast to albumin, prealbumin has a short biological half-life and reflects recent status [8]. Preoperative prealbumin level is useful for predicting long-term outcomes in patients undergoing liver resection for HCC.

The protein biomarker alpha-fetoprotein (AFP) is an important traditional diagnostic biomarker for hepatocellular carcinoma (HCC) for the past several decades. It is a 67-kDa glycoprotein that is produced by the liver in early fetal life when hepatocytes have not yet matured, and in HCC because of the dedifferentiation of hepatocytes to a more fetal pattern of gene expression. Serum AFP levels of more than 500 ng/ml are considered to have diagnostic significance; however, these values are reported only in a small percentage of patients with HCC. Many studies have reported that elevated serum AFP levels are correlated with increased risks of HCC in individuals infected with the hepatitis C virus (HCV) [9, 10]. In the practical aspects, the serum AFP concentrations do not correlate well with the prognostic values of HCC, such as the size of the tumor, stages, or disease progression, and ethnic variability may also be reported. Furthermore, in some cases of HCC, AFP elevations are not apparent at all and lack apparent discriminating power [11]. In HCC, the liver is damaged by one or more preexisting pathologic conditions, including cirrhosis and chronic hepatitis resulting from HBV or HCV infection [12], which also exhibited high serum AFP levels [13]. Three types of glycoforms of AFP have been identified so far, which are based on the difference in their binding affinities for Lens culinaris agglutinin (LCA). AFPL1 is a nonbinding fraction and AFP-L2 is a weakly binding fraction of total AFP. AFP-L3 is associated effectively with LCA [14], which is why it is a more specific biomarker for HCC. The AFP-L3 fraction has significantly increased the sensitivity for the early detection of HCC. However, AFP-based diagnostic approaches are still far from satisfactory results.

Advertisement

2. Serum autoantibodies

At an early stage of carcinogenesis, a small amount of tumor antigens can be produced by tumor cells that ultimately generate autoantibodies. These autoantibodies are stable in blood circulation and remain elevated for a long time [15]. IgG-L3% HCC-derived immunoglobulin G (IgG) and any type of unwanted glycosylations cause carcinogenesis. The fraction of Lens culinaris agglutinin-binding IgG (IgG-L3) among total serum IgG (IgG-L3%) increases with tumor load compared to healthy volunteers and asymptomatic HBV carriers.

Advertisement

3. Emerging serum protein biomarkers

Although traditional biomarkers have certain diagnostic values for HCC, none of them have been suitable in clinical practice, but there are emerging serum biomarkers. For example, aldo-keto reductase family 1 member B10 (AKR1B10) is a novel secretory protein that is associated with lung, breast, and colorectal cancers but induced in hepatocellular carcinoma (HCC), and this protein is located on chromosome 7q33 [16, 17] and is one of the important diagnostic and prognostic biomarkers for HCC [18, 19, 20]. AKR1B10 protein is an oncogenic protein that induces tumor development and progression by the removal of cytotoxic carbonyl compounds [21, 22, 23]. Dickkopf-1 (DKK1) is a 266-amino acid (35-kDa) secreted glycoprotein and antagonist of the Wnt/beta-catenin pathway signaling pathway discovered in 1998, which is expressed in a variety of human tumors. It is one of the impaired signaling pathways in hepatocellular carcinoma (HCC). Its function seems to be contradictory in the process of tumorigenesis, acting either as an oncogenic promoter of metastasis or as a tumor suppressor. In many tumors, high expression of Dkk1 may promote tumor metastasis. However, Dkk1 can inhibit tumor invasion and metastasis [24, 25, 26] and plays an important role in regulating human HCC cell migration, invasion, and tumor growth [27]. Therefore, novel genetic biomarkers for early diagnosis and detection are needed.

With the advent of cellular and molecular techniques for a better understanding of tumor biology, the role of biomarkers related to early detection has attracted a great deal of research interest resulting in the discovery and utilization of several novel markers in this disease. Liquid biopsy is one of the potential and noninvasive procedures that have attracted much attention to identify tumor markers in peripheral blood for diagnosis, monitoring, and prognosis of cancer, and overcoming tissue biopsy limitations. In this procedure, the sampling and analysis of biological samples, such as blood, urine, saliva, or stool, are done where nucleic acids originating from all or part of the body can be found including circulating tumor DNA (ctDNA) or RNA, exosomes, and circulating tumor cells (CTC). Most of the studies are carried out for mutations of ctDNA to detect minimal residual disease, as diagnostic markers or response to therapy. For the mapping of genetic alterations in ctDNA either next-generation sequencing (NGS) or digital PCR are used. Circulating tumoral cells and tumoral cell-free nucleic acids in peripheral blood could signal the presence of micrometastasis, and their utility has been explored in HCC diagnosis and prognosis.

Some other circulating RNAs have been explored, but none of them have been widely recognized as valuable markers of HCC recurrence, probably because none of them are specific for HCC [28]. In cancer biology, different types of circulating cellular elements have been identified as tumor markers [29, 30]. One of them is circulating tumor cells (CTC), which consist of obtaining a sample in a convenient and minimally invasive manner at multiple time points over the course of the disease.

Advertisement

4. Circulating tumor cells (CTCs)

Circulating tumor cells (CTCs) are defined as the primary solid tumor cells shed into blood, bone marrow or lymphatic vessels, or other healthy organs. These cells have a strong potential for distant metastasis, circulating through the bloodstream, and traveling to different tissues or organs of the body [31]. This type of tumor development process occurs at every stage. Australian doctor Thomas R. Ashworth was the first to discover CTCs in 1869 in the blood of a breast cancer patient [32]. CTCs are one of the useful markers for early diagnosis and monitoring of disease relapse. Their concentration is very low in blood, which limited the studies on CTC. Metastatic cells struggle to survive in the bloodstream and less than 0.01% of CTCs introduced into the circulation survive to produce metastases [33]. In recent years, with the advent of new technology, the separation and enrichment of CTCs have been greatly improved. In cancer, CTCs have received great attention in trying to estimate the future course in patients with breast cancer, colon cancer, and prostate cancer.

CTC analysis might provide personalized and effective strategies for clinicians and researchers because CTCs are sensitive biomarkers that enable early diagnosis, real-time monitoring, and molecular characterization to facilitate the implementation of precision medicine. In addition, the different study shows powerful evidence for the potential clinical value of the CTC assay [34]. However, numerous obstructions must be overcome before CTC analysis can be applied in the clinic. The CTC detection methods mentioned above have their own advantages and drawbacks. It is one of the important tasks to establish a highly sensitive and specific method that provides the full spectrum of CTCs. Therefore, the standardized method for CTC evaluation includes sample preparation, enrichment, and detection. In addition, lots of studies are single-center case-control research, with limited sample sizes. Validation is sometimes difficult or impossible to achieve. A multicenter prospective study with sufficient sample size and long follow-up is needed for CTC detection methodologies, the detection method must be uniform, and large samples can provide powerful validation for accurate analysis and standard evaluation of the final data. Although CTC detection is currently used in research, technological advancement will make it feasible in clinical practice in the near future.

Advertisement

5. Circulating tumor DNA (ctDNA)

Circulating tumor DNA (ctDNA) has attracted extensive attention for its wide application in cancer research, and it consists of mutant DNAs or tumor-derived fragmented DNA released into the circulation by tumor cells and constitute part of circulating cell-free DNA (cfDNA) in different types of cancer [35], cfDNA levels vary substantially from <0.01% to >60% of alleles in circulation [36, 37]. ctDNA carries the genetic information of the tumor, and it is highly specific and can detect at very low concentrations, making it suitable for early diagnosis, treatment, and monitoring of tumor progression. ctDNA is a liquid biopsy to profile the genome of a tumor more comprehensively than conventional sampling methods. Thus, it is an important tool to provide information for guiding targeted therapy [38], unveiling drug resistance [39], and monitoring treatment response [40]. Analysis of ctDNA helps in the efficient evaluation of disease status and early detection of recurrence, providing an average of 10–12 months’ lead time for detection of metastatic recurrence compared to traditional modalities [41]. After resection, detectable ctDNA could identify cancer patients at high risk of recurrence, [42] while dynamic ctDNA changes can predict clinical relapse [43]. ctDNA provides information about specific tumor genes, such as DNA methylation, point mutations, copy number variations (CNVs), and chromosomal rearrangements. It also provides a unique opportunity for serially monitoring tumor genomes in a noninvasive, convenient, and accurate manner. Two different changes are monitored during the detection of ctDNA—quantitative changes and qualitative changes. The first detection method measures the quantity of ctDNA in circulation, and the second detects tumor-specific genetic aberrations. Many studies have carried out quantitative changes in cfDNA in the blood of HCC patients and show that elevated levels of cfDNA may represent an important complementary tool with potential clinical applications for detection, screening, treatment monitoring, and predicting metastatic potential [44, 45, 46, 47, 48, 49, 50, 51].

Cell-free nucleic acids (cfNA) were first reported in human peripheral blood by Mandel and Metais in 1948 [52]. However, their studies did not recognize until 30 years later with the discovery of higher concentrations of cell-free DNA (cfDNA) in serum and plasma from cancer patients compared to healthy individuals. Currently, cfDNA is supposed to be released by normal cells at an average concentration of 30 ng/ml (0–100 ng/ml) into peripheral blood at the physiological level [53]. The concentration of cfDNA was accompanied by a decrease in DNase activity because cfDNA is degraded by peripheral blood deoxyribonuclease activity. Normal cells in peripheral circulation can also release cfDNA, and this reduces ctDNA concentrations [54]. cfDNA lysis occurs secondary to the clotting process of blood cells in collection tubes; thus, several studies have found significantly high cfDNA concentrations in serum than in plasma [55, 56]. Similarly, the blood specimen collected improperly or mechanical shearing leads to the destruction of the blood cells, causing the release of cfDNA into plasma [57]. Until recently, many researchers preferred plasma fraction over that in serum for cfDNA analysis [58]. Although DNA in the plasma is least contaminated with blood cells, the amount of DNA in plasma is more or less affected due to the time interval between analysis and blood collection [59].

A large number of hypermethylated genes, such as developing brain homeobox protein 2 (DBX2) [60], G-protein-coupled bile acid receptor (TGR5) [61], metallothionein 1M (MT1M), metallothionein 1G (MT1G) [62] and I the cyclin kinase inhibitor (NK4A) [63], were detected as cfDNA from HCC patients and were identified as biomarkers of vascular invasion. In the process of HCC diagnosis, high degree of methylation at multiple genes has been shown to play an important role. In addition, to improve the diagnostic efficiency, the combined detection of the methylation status of multiple genes may be effective [64].

The presence of cell-free DNA in plasma/serum has been used to reveal tumor-associated biomarkers, such as the increased plentiful of cell-free DNA (cfDNA) in cancer patients or the presence of epigenetic or specific genetic alterations, which have been discovered in numerous types of cancers, including HCC. Many studies have reported that the cfDNA is a source of HCC biomarkers in the diagnostic and prognosis of HCC in clinical settings also; HCC-specific biomarkers should be validated to determine their association with HCC recurrence. Finally, different micro (mi) RNA signatures in liver tissue have been associated with HCC [65, 66]. However, the necessity of obtaining liver tissue samples limits their application preoperatively, and circulating miRNAs are at present being explored. Several circulating miRNAs have been reported as important biomarkers for HCC diagnosis [67], prognosis, and vascular invasion [68, 69]. To date, there are no data about the association of miRNAs with HCC recurrence, and future studies are needed to explore the utility of these promising biomarkers.

Advertisement

6. Circulating microRNAs

Circulating microRNAs (miRNAs) were first proposed as potential cancer biomarkers in 2008 [70]. Blood miRNAs, which circulate in a highly stable, cell-free form, show promise as novel potential biomarkers for early detection of HCC. A number of evidence indicate that these noncoding nucleotide sequences are resistant to RNase degradation, repeated freeze-thaw cycles, boiling as well as acid/base treatment [71, 72]. The stability of circulating miRNAs has attracted attention from clinical researchers, who want to investigate their diagnostic utility for a wide range of diseases including HCC.

Under extreme conditions, their stability and their abundance in sera made circulating miRNAs as promising biomarkers for cancer pathologies [73, 74]. Serum miRNAs are not digested by ribonuclease because they are encapsulated in protein complexes or in membranous microvesicles that transport them in the circulatory system [75, 76]. The circulating miRNAs are stable when the samples are stored at −80°C [77]. Despite the valuable investigation of extracellular miRNAs, the use of miRNAs as biomarkers of cancer is still regarded as a “work in progress” and mostly confined to research programs [78]. Continuing technological advancements, however, like second-generation sequencing, as well as a perceptive of the pathobiological role of miRNAs, emphasize their future promise as clinical biomarkers [79].

Advertisement

7. Circular RNAs

Circular RNAs (circRNAs) are covalently closed, single-stranded, and stable RNA molecules [80] that have been evaluated for the diagnosis of various cancers. They contribute importantly to gastric cancer [81, 82], breast cancer [83], lung cancer [84, 85], pancreatic cancer [86], and HCC [87]. In a multicenter study, three circRNAs (hsa_circ_0000976, hsa_circ_0007750, and hsa_circ_0139897) were successfully validated in the plasma of the hepatitis B virus-related HCC patients, and their plasma levels positively correlated with HCC relapse, while after hepatectomy their levels decreased [87].

Advertisement

8. Conclusion

In the present scenario, no single or biomarker combination is predictable enough to diagnose a HCC lesion without confirmatory histological or radiological data. None of the new tumor markers excel the conventional ones in such a way that it has been widely endorsed in clinical practice. The liquid biopsy is one of the critical parts of precision medicine, and it is likely to be useful in the near future. However, future research should develop useful HCC biomarkers for monitoring treatment activity, detecting early resistance to treatment, and identifying patients who would more likely benefit from treatment.

We expect that identifying novel cost-effective and high-efficient biomarkers for the early diagnosis of HCC will be promising.

References

  1. 1. Njei B, Rotman Y, Ditah I, Lim JK. Emerging trends in hepatocellular carcinoma incidence and mortality. Hepatology. 2015;61:191-199
  2. 2. Global Burden of Disease Liver Cancer Collaboration, Akinyemiju T, Abera S, Ahmed M, Alam N, Alemayohu MA, et al. Choi JAMA Oncology. 2017;3(12):1683-1691. DOI: 10.1001/jamaoncol.2017
  3. 3. Pawlik TM, Gleisner AL, Anders RA, Assumpcao L, Maley W, Choti MA. Preoperative assessment of hepatocellular carcinoma tumor grade using needle biopsy: Implications for transplant eligibility. Annals of Surgery. 2007;245:435-442
  4. 4. Court CM, Harlander-Locke MP, Markovic D, French SW, Naini BV, Lu DS, et al. Determination of hepatocellular carcinoma grade by needle biopsy is unreliable for liver transplant candidate selection. Liver Transplantation. 2017;23:1123-1132
  5. 5. Uccello M, Malaguarnera G, Pelligra EM, Biondi A, Basile F, Motta M. Lipoprotein(a) as a potential marker of residual liver function in hepatocellular carcinoma. Indian Journal of Medical and Paediatric Oncology. 2011;32(2):71-75
  6. 6. Malaguarnera G, Gagliano C, Giordano M, et al. Homocysteine serum levels in diabetic patients with non proliferative, proliferative and without retinopathy. Biomed Research International. 2014;2014:1-4e191497
  7. 7. Tripodi A. D-dimer testing in laboratory practice. Clinical Chemistry. 2011;57(9):1256-1262
  8. 8. Kumada T et al. Predictive value of tumor markers for hepatocarcinogenesis in patients with hepatitis C virus. Journal of Gastroenterology. 2011;46:536-544
  9. 9. Chen DS, Sung JL, Sheu JC. Serum α-fetoprotein in the early stage of human hepatocellular carcinoma. Gastroenterology. 1984;86(6):1404-1409
  10. 10. Saffroy R et al. New perspectives and strategy research biomarkers for hepatocellular carcinoma. Clinical Chemistry and Laboratory Medicine. 2007;45:1169-1179
  11. 11. Fattovich G et al. Hepatocellular carcinoma in cirrhosis: Incidence and risk factors. Gastroenterology. 2004;127(suppl. 1):S35-S50
  12. 12. Lok AS, Lai CL. Alpha-fetoprotein monitoring in Chinese patients with chronic hepatitis B virus infection: Role in the early detection of hepatocellular carcinoma. Hepatology. 1989;9:110-115
  13. 13. Li J, Han X, Yu X, Xu Z, Yang G, Liu B, et al. Clinical applications of liquid biopsy as prognostic and predictive biomarkers in hepatocellular carcinoma: Circulating tumor cells and circulating tumor DNA. Journal of Experimental and Clinical Cancer Research. 2018;37:213
  14. 14. Schuler F, Dolken G. Detection and monitoring of minimal residual disease by quantitative real-time PCR. Clinica Chimica Acta. 2006;363:147-156.7
  15. 15. Tan HT, Low J, Lim SG, Chung MC. Serum autoantibodies as biomarkers for early cancer detection. The FEBS Journal. 2009;276:6880-6904. DOI: 10.1111/j.1742-4658.2009.07396.x
  16. 16. Wang NY, Wang C, Li W, Wang GJ, Cui GZ, He H, et al. Prognostic value of serum AFP, AFP-L3, and GP73 in monitoring short-term treatment response and recurrence of hepatocellular carcinoma after radiofrequency ablation. Asian Pacific Journal of Cancer Prevention. 2014;15:1539-1544. DOI: 10.7314/APJCP.2014.15.4.1539
  17. 17. Xia Y, Yan ZL, Xi T, Wang K, Li J, Shi LH, et al. A case control study of correlation between preoperative serum AFP and recurrence of hepatocellular carcinoma after curative hepatectomy. Hepato-Gastroenterology. 2012;59:2248-2254. DOI: 10.5754/hge11978
  18. 18. Bi X, Yan T, Zhao H, Zhao J, Li Z, Huang Z, et al. Correlation of alpha fetoprotein with the prognosis of hepatocellular carcinoma after hepatectomy in an ethnic Chinese population. Zhonghua Yi Xue Za Zhi. 2014;94:2645-2649. DOI: 10.3760/cma.j.issn.0376-2491.2014. 34.002
  19. 19. Singal AG, Chan V, Getachew Y, Guerrero R, Reisch JS, Cuthbert JA. Predictors of liver transplant eligibility for patients with hepatocellular carcinoma in a safety net hospital. Digestive Diseases and Sciences. 2012;57:580-586. DOI: 10.1007/s10620-011-1904-7
  20. 20. Eurich D, Boas-Knoop S, Morawietz L, Neuhaus R, Somasundaram R, Ruehl M, et al. Association of mannose-binding lectin-2 gene polymorphism with the development of hepatitis C-induced hepatocellular carcinoma. Liver International. 2011;31:1006-1012. DOI: 10.1111/j.1478-3231.2011.02522.x
  21. 21. Zhong L, Liu Z, Yan R, Johnson S, Zhao Y, Fang X, et al. Aldo-Keto reductase family 1 B10 protein detoxifies dietary and lipid-derived alpha, beta-unsaturated carbonyls at physiological levels. Biochemical and Biophysical Research Communications. 2009;387(2):245-250. DOI: 10.1016/j.bbrc.2009.06.123
  22. 22. Shen Y, Zhong L, Johnson S, Cao D. Human aldo-keto reductases 1B1 and 1B10: A comparative study on their enzyme activity toward electrophilic carbonyl compounds. Chemico-Biological Interactions. 2011;191(1-3):192-198. DOI: 10.1016/j.cbi.2011.02.004
  23. 23. Yan R, Zu X, Ma J, Liu Z, Adeyanju M, Cao D. Aldo-keto reductase family 1 B10 gene silencing results in growth inhibition of colorectal cancer cells: Implication for cancer intervention. International Journal of Cancer. 2007;121(10):2301-2306. DOI: 10.1002/ijc.22933
  24. 24. Liu C, Xiao GQ , Yan LN, Li B, Jiang L, Wen TF, et al. Value of alpha-fetoprotein in association with clinicopathological features of hepatocellular carcinoma. World Journal of Gastroenterology. 2013;19:1811-1819. DOI: 10.3748/wjg.v19.i11.1811
  25. 25. Ilikhan SU, Bilici M, Sahin H, Akca AS, Can M, Oz II, et al. Assessment of the correlation between serum prolidase and alpha-fetoprotein levels in patients with hepatocellular carcinoma. World Journal of Gastroenterology. 2015;21:6999-7007. DOI: 10.3748/wjg.v21.i22.6999
  26. 26. Abbasi A, Bhutto AR, Butt N, Munir SM. Corelation of serum alpha fetoprotein and tumor size in hepatocellular carcinoma. The Journal of the Pakistan Medical Association. 2012;62:33-36
  27. 27. Tang H, Tang XY, Liu M, Li X. Targeting alpha-fetoprotein represses the proliferation of hepatoma cells via regulation of the cell cycle. Clinica Chimica Acta. 2008;394:81-88. DOI: 10.1016/j.cca.2008.04.012
  28. 28. Dardaei L, Shahsavani R, Ghavamzadeh A, Behmanesh M, Aslankoohi E, Alimoghaddam K, et al. The detection of disseminated tumor cells in bone marrow and peripheral blood of gastric cancer patients by multimarker (CEA, CK20, TFF1 and MUC2) quantitative real-time PCR. Clinical Biochemistry. 2011;44:325-330
  29. 29. Chaffer CL, Weinberg RA. A perspective on cancer cell metastasis. Science. 2011;331:1559-1564
  30. 30. Ashworth T. A case of cancer in which cells similar to those in the tumours were seen in the blood after death. The Australasian Medical Journal. 1869;14:146-149
  31. 31. Ma J, Yan R, Zu X, Cheng JM, Rao K, Liao DF, et al. Aldo-keto reductase family 1 B10 affects fatty acid synthesis by regulating the stability of acetylCoA carboxylase-alpha in breast cancer cells. Journal of Biological Chemistry. 2008;283:3418-3423. DOI: 10.1074/jbc.M707650200
  32. 32. Ashworth TR. A case of cancer in which cells similar to those in the tumours were seen in the blood after death. The Australasian Medical Journal. 1869;14:146-147
  33. 33. Tremblay PL, Huot J, Auger FA. Mechanisms by which E-selectin regulates diapedesis of colon cancer cells under flow conditions. Cancer Research. 2008;68(13):5167-5176. DOI: 10.1158/0008-5472.CAN-08-1229
  34. 34. Alunni-Fabbroni M, Sandri MT. Circulating tumour cells in clinical practice: Methods of detection and possible characterization. Methods. 2010;50:289-297
  35. 35. Han C, Gao L, Bai H, Dou X. Identification of a role for serum aldoketo reductase family 1 member B10 in early detection of hepatocellular carcinoma. Oncology Letters. 2018;16:7123-7130. DOI: 10.3892/ol.2018.9547
  36. 36. Zhu R, Xiao J, Luo D, Dong M, Sun T, Jin J. Serum AKR1B10 predicts the risk of hepatocellular carcinoma—A retrospective single-center study. Gastroenterología y Hepatología. 2019;42:614-621. DOI: 10.1016/j.gastrohep.2019.06.007
  37. 37. Ye X, Li C, Zu X, Lin M, Liu Q , Liu J, et al. A large-scale multicenter study validates AKR1B10 as a new prevalent serum marker for detection of hepatocellular carcinoma. Hepatology. 2019;69:2489-2501. DOI: 10.1002/hep.30519
  38. 38. Sato N, Yamabuki T, Takano A, Koinuma J, Aragaki M, Masuda K, et al. Wnt inhibitor Dickkopf-1 as a target for passive cancer immunotherapy. Cancer Research. 2010;70:5326-5336. DOI: 10.1158/0008-5472.CAN-09-3879
  39. 39. Rachner TD, Thiele S, Gobel A, Browne A, Fuessel S, Erdmann K, et al. High serum levels of Dickkopf-1 are associated with a poor prognosis in prostate cancer patients. BMC Cancer. 2014;14:649. DOI: 10.1186/1471-2407-14-649
  40. 40. Peng YH, Xu YW, Guo H, Huang LS, Tan HZ, Hong CQ , et al. Combined detection of serum Dickkopf-1 and its autoantibodies to diagnose esophageal squamous cell carcinoma. Cancer Medicine. 2016;5:1388-1396. DOI: 10.1002/cam4.702
  41. 41. Tung EK, Mak CK, Fatima S, Lo RC, Zhao H, Zhang C, et al. Clinicopathological and prognostic significance of serum and tissue Dickkopf-1 levels in human hepatocellular carcinoma. Liver International. 2011;31:1494-1504. DOI: 10.1111/j.1478-3231.2011.02597.x
  42. 42. Yin CQ , Yuan CH, Qu Z, Guan Q , Chen H, Wang FB. Liquid biopsy of hepatocellular carcinoma: Circulating tumor-derived biomarkers. Disease Markers. 2016;2016:1427849
  43. 43. Miller MC, Doyle GV, Terstappen LW. Significance of circulating tumor cells detected by the cell search system in patients with metastatic breast colorectal and prostate cancer. Journal of Oncology. 2010;2010:617421
  44. 44. Sun C, Liao W, Deng Z, Li E, Feng Q , Lei J, et al. The diagnostic value of assays for circulating tumor cells in hepatocellular carcinoma: A meta-analysis. Medicine. 2017;96(29):e7513. DOI: 10.1097/MD.0000000000007513
  45. 45. Bettegowda C, Sausen M, Leary RJ, Kinde I, Wang Y, Agrawal N, et al. Detection of circulating tumor DNA in early- and late-stage human malignancies. Science Translational Medicine. 2014;6:224ra224
  46. 46. Leung F, Kulasingam V, Diamandis EP, Hoon DS, Kinzler K, Pantel K, et al. Circulating tumor DNA as a cancer biomarker: Fact or fiction? Clinical Chemistry. 2016;62:1054-1060. DOI: 10.1373/clinchem.2016.260331
  47. 47. Wong SQ, Raleigh JM, Callahan J, Vergara IA, Ftouni S, Hatzimihalis A, et al. Circulating tumor DNA analysis and functional imaging provide complementary approaches for comprehensive disease monitoring in metastatic melanoma. JCO Precision Oncology. 2017;1:1-14. DOI: 10.1200/PO.16.00009
  48. 48. Luke JJ, Oxnard GR, Paweletz CP, Camidge DR, Heymach JV, Solit DB, et al. Cell free DNAWG. Realizing the potential of plasma genotyping in an age of genotype-directed therapies. Journal of the National Cancer Institute. 2014;106(8):dju214
  49. 49. Murtaza M, Dawson SJ, Tsui DW, Gale D, Forshew T, Piskorz AM, et al. Non-invasive analysis of acquired resistance to cancer therapy by sequencing of plasma DNA. Nature. 2013;497:108-112
  50. 50. Montagut C, Siravegna G, Bardelli A. Liquid biopsies to evaluate early therapeutic response in colorectal cancer. Annals of Oncology. 2015;26:1525-1527
  51. 51. Reinert T, Scholer LV, Thomsen R, Tobiasen H, Vang S, Nordentoft I, et al. Analysis of circulating tumour DNA to monitor disease burden following colorectal cancer surgery. Gut. 2016;65:625-634
  52. 52. Sausen M, Phallen J, Adleff V, Jones S, Leary RJ, Barrett MT, et al. Clinical implications of genomic alterations in the tumour and circulation of pancreatic cancer patients. Nature Communications. 2015;6:7686. DOI: 10.1038/ncomms8686
  53. 53. Garcia-Murillas I, Schiavon G, Weigelt B, Ng C, Hrebien S, Cutts RJ, et al. Mutation tracking in circulating tumor DNA predicts relapse in early breast cancer. Science Translational Medicine. 2015;7:302ra133
  54. 54. Mandel P, Métais P. Les acides nucleiques du plasma sanguin chez l’homme. Comptes Rendus de l’Académie des Sciences de Paris. 1948;142:241-243
  55. 55. Sorenson GD, Pribish DM, Valone FH, Memoli VA, Bzik DJ, Yao S-L. Soluble normal and mutated DNA sequences from single-copy genes in human blood. Cancer Epidemiology, Biomarkers and Prevention. 1994;3(1):67-71
  56. 56. Fleischhacker M, Schmidt B. Circulating nucleic acids (CNAs) and cancer—A survey. Biochimica et Biophysica Acta. 2007;1775(1):181-232
  57. 57. Ren N, Qin LX, Tu H, Liu YK, Zhang BH, Tang ZY. The prognostic value of circulating plasma DNA level and its allelic imbalance on chromosome 8p in patients with hepatocellular carcinoma. Journal of Cancer Research and Clinical Oncology. 2006;132(6):399-407
  58. 58. Iizuka N, Sakaida I, Moribe T, Fujita N, Miura T, Stark M, et al. Elevated levels of circulating cell-free DNA in the blood of patients with hepatitis C virus-associated hepatocellular carcinoma. Anticancer Research. 2006;26(6C):4713-4719
  59. 59. Yang YJ, Chen H, Huang P, Li CH, Dong ZH, Hou YL. Quantification of plasma hTERT DNA in hepatocellular carcinoma patients by quantitative fluorescent polymerase chain reaction. Clinical and Investigative Medicine. 2011;34(4):E238
  60. 60. Piciocchi M, Cardin R, Vitale A, Vanin V, Giacomin A, Pozzan C, et al. Circulating free DNA in the progression of liver damage to hepatocellular carcinoma. Hepatology International. 2013;7(4):1050-1057
  61. 61. Huang A, Zhang X, Zhou SL, Cao Y, Huang XW, Fan J, et al. Plasma circulating cell-free DNA integrity as a promising biomarker for diagnosis and surveillance in patients with hepatocellular carcinoma. Journal of Cancer. 2016;7(13):1798-1803
  62. 62. Yan L, Chen Y, Zhou J, Zhao H, Zhang H, Wang G. Diagnostic value of circulating cell-free DNA levels for hepatocellular carcinoma. International Journal of Infectious Diseases. 2017;67:92-97
  63. 63. Bettegowda C, Sausen M, Leary RJ, Kinde I, Wang Y, Agrawal N, et al. Detection of circulating tumor DNA in early and late-stage human malignancies. Science Translational Medicine. 2014;6(224):224ra224
  64. 64. Diehl F, Schmidt K, Choti MA, Romans K, Goodman S, Li M, et al. Circulating mutant DNA to assess tumor dynamics. Nature Medicine. 2008;14(9):985-990
  65. 65. Tamkovich SN, Cherepanova AV, Kolesnikova EV, Rykova EY, Pyshnyi DV, Vlassov VV, et al. Circulating DNA and DNase activity in human blood. Annals of the New York Academy of Sciences. 2006;1075:191-196
  66. 66. Lee TH, Montalvo L, Chrebtow V, Busch MP. Quantitation of genomic DNA in plasma and serum samples: Higher concentrations of genomic DNA found in serum than in plasma. Transfusion. 2001;41(2):276-282
  67. 67. Jung M, Klotzek S, Lewandowski M, Fleischhacker M, Jung K. Changes in concentration of DNA in serum and plasma during storage of blood samples. Clinical Chemistry. 2003;49(6 Pt 1):1028-1029
  68. 68. Volik S, Alcaide M, Morin RD, Collins C. Cell-free DNA (cfDNA): Clinical significance and utility in cancer shaped by emerging technologies. Molecular Cancer Research. 2016;14(10):898-908
  69. 69. Vallee A, Marcq M, Bizieux A, Kouri CE, Lacroix H, Bennouna J, et al. Plasma is a better source of tumor-derived circulating cell-free DNA than serum for the detection of EGFR alterations in lung tumor patients. Lung Cancer. 2013;82(2):373-374
  70. 70. van Dessel LF, Beije N, Helmijr JC, Vitale SR, Kraan J, Look MP, et al. Application of circulating tumor DNA in prospective clinical oncology trials—Standardization of preanalytical conditions. Molecular Oncology. 2017;11(3):295-304
  71. 71. Liese J, Peveling-Oberhag J, Doering C, Schnitzbauer AA, Herrmann E, Zangos S, et al. A possible role of microRNAs as predictive markers for the recurrence of hepatocellular carcinoma after liver transplantation. Transplant International. 2016;29:369-380
  72. 72. Han ZB, Zhong L, Teng MJ, Fan JW, Tang HM, Wu JY, et al. Identification of recurrence-related microRNAs in hepatocellular carcinoma following liver transplantation. Molecular Oncology. 2012;6:445-457
  73. 73. Bharali D, Jebur HB, Baishya D, Kumar S, Sarma MP, Masroor M, et al. Expression analysis of serum microRNA-34a and microRNA-183 in hepatocellular carcinoma. Asian Pacific Journal of Cancer Prevention. 2018;19:2561-2568
  74. 74. Cho HJ, Kim SS, Nam JS, Kim JK, Lee JH, Kim B, et al. Low levels of circulating microRNA-26a/29a as poor prognostic markers in patients with hepatocellular carcinoma who underwent curative treatment. Clinics and Research in Hepatology and Gastroenterology. 2017;41:181-189
  75. 75. Zhang J, Lin H, Wang XY, Zhang DQ , Chen JX, Zhuang Y, et al. Predictive value of microRNA-143 in evaluating the prognosis of patients with hepatocellular carcinoma. Cancer Biomarkers. 2017;19:257-262
  76. 76. Zhang P, Wen X, Gu F, Deng X, Li J, Dong J, et al. Methylation profiling of serum DNA from hepatocellular carcinoma patients using an Infinium human methylation 450 BeadChip. Hepatology International. 2013;7(3):893-900
  77. 77. Chen LL. The biogenesis and emerging roles of circular RNAs. Nature Reviews. Molecular Cell Biology. 2016;17:205-211. DOI: 10.1038/nrm.2015.32
  78. 78. Sun H, Tang W, Rong D, Jin H, Fu K, Zhang W, et al. Hsa_circ_0000520, a potential new circular RNA biomarker, is involved in gastric carcinoma. Cancer Biomarkers. 2018;21:299-306. DOI: 10.3233/CBM-170379
  79. 79. Li P, Chen S, Chen H, Mo X, Li T, Shao Y, et al. Using circular RNA as a novel type of biomarker in the screening of gastric cancer. Clinica Chimica Acta. 2015;444:132-136. DOI: 10.1016/j.cca.2015.02.018
  80. 80. Li T, Shao Y, Fu L, Xie Y, Zhu L, Sun W, et al. Plasma circular RNA profiling of patients with gastric cancer and their droplet digital RT-PCR detection. Journal of Molecular Medicine. 2018;96:85-96. DOI: 10.1007/s00109-017-1600-y
  81. 81. Yin WB, Yan MG, Fang X, Guo JJ, Xiong W, Zhang RP. Circulating circular RNA hsa_circ_0001785 acts as a diagnostic biomarker for breast cancer detection. Clinica Chimica Acta. 2018;487:363-368. DOI: 10.1016/j.cca.2017.10.011
  82. 82. Zhu X, Wang X, Wei S, Chen Y, Chen Y, Fan X, et al. hsa_circ_0013958: A circular RNA and potential novel biomarker for lung adenocarcinoma. FEBS Journal. 2017;284:2170-2182. DOI: 10.1111/febs.14132
  83. 83. Hang D, Zhou J, Qin N, Zhou W, Ma H, Jin G, et al. A novel plasma circular RNA circFARSA is a potential biomarker for non-small cell lung cancer. Cancer Medicine. 2018;7:2783-2791. DOI: 10.1002/cam4.1514
  84. 84. Yang F, Liu DY, Guo JT, Ge N, Zhu P, Liu X, et al. Circular RNA circLDLRAD3 as a biomarker in diagnosis of pancreatic cancer. World Journal of Gastroenterology. 2017;23:8345-8354. DOI: 10.3748/wjg.v23.i47.8345
  85. 85. Zhang X, Xu Y, Qian Z, Zheng W, Wu Q , Chen Y, et al. circRNA_104075 stimulates YAP-dependent tumorigenesis through the regulation of HNF4a and may serve as a diagnostic marker in hepatocellular carcinoma. Cell Death and Disease. 2018;9:1091. DOI: 10.1038/s41419-018-1132-6
  86. 86. Li Z, Zhou Y, Yang G, He S, Qiu X, Zhang L, et al. Using circular RNA SMARCA5 as a potential novel biomarker for hepatocellular carcinoma. Clinica Chimica Acta. 2019;492:37-44. DOI: 10.1016/j.cca.2019.02.001
  87. 87. Yu J, Ding WB, Wang MC, Guo XG, Xu J, Xu QG, et al. Plasma circular RNA panel to diagnose hepatitis B virus-related hepatocellular carcinoma: A large-scale, multicenter study. International Journal of Cancer. 2019;146:1754-1763. DOI: 10.1002/ijc.32647

Written By

S.S. Haque, Ravi Bhushan Raman and Mehboobus Salam

Submitted: 13 May 2022 Reviewed: 14 June 2022 Published: 19 August 2022