Next Article in Journal
Centralised RECIST Assessment and Clinical Outcomes with Lenvatinib Monotherapy in Recurrent and Metastatic Adenoid Cystic Carcinoma
Next Article in Special Issue
Immediate Sample Fixation Increases Circulating Tumour Cell (CTC) Capture and Preserves Phenotype in Head and Neck Squamous Cell Carcinoma: Towards a Standardised Approach to Microfluidic CTC Biomarker Discovery
Previous Article in Journal
Recurrent Laryngeal Nerve Preservation Strategies in Pediatric Thyroid Oncology: Continuous vs. Intermittent Nerve Monitoring
Previous Article in Special Issue
The Emerging Role of Circulating Tumor DNA in the Management of Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Liquid Biopsy in Hepatocellular Carcinoma: Opportunities and Challenges for Immunotherapy

1
Department of Laboratory Medicine Karolinska Institutet, 14152 Stockholm, Sweden
2
Division of Oncology, Department of Medical and Surgical Sciences for Children & Adults, University-Hospital of Modena and Reggio Emilia, 41100 Modena, Italy
3
Unit of Gastroenterology and Hepatology, Department of Medicine/Huddinge, Karolinska Institutet, Department of Upper GI Diseases, Karolinska University Hospital, 14186 Stockholm, Sweden
4
Department of Pathology, Winthrop P. Rockefeller Cancer Institute, University of Arkansas for Medical Sciences, Little Rock, AR 72205, USA
*
Authors to whom correspondence should be addressed.
Submission received: 30 July 2021 / Revised: 21 August 2021 / Accepted: 23 August 2021 / Published: 27 August 2021
(This article belongs to the Special Issue Circulating Tumor Cells: From the Laboratory to the Cancer Clinic)

Abstract

:

Simple Summary

Hepatocellular carcinoma (HCC) causes many deaths worldwide, and current treatments have limitations. Immunotherapies have shown the most promising clinical outcomes for advanced HCC. However, there are many patients with HCC who still respond poorly to these treatments. Circulating biomarkers that can easily be obtained through blood sampling are promising in predicting treatment responses, since they are minimally invasive and enable us to constantly monitor disease progression. The aim of this review is to discuss the most promising types of blood-based biomarkers for the diagnosis and prognosis of HCC, with the focus on circulating tumor cells and circulating tumor DNA. We also discuss technologies for detecting these biomarkers, as well as their clinical applications for immunotherapies in HCC. We conclude that despite their encouraging results to accurately predict responses to immunotherapies, more and larger clinical studies are still necessary, in order to improve the precision of biomarkers, which are used in the treatment decision for patients with HCC.

Abstract

Hepatocellular carcinoma (HCC) is one of the deadliest cancer types worldwide. HCC is often diagnosed at a late stage when the therapeutic options are very limited. However, even at the earlier stages, the best treatment is liver transplantation, surgical resection or ablation. Surgical resection and ablation may carry a high risk of tumor recurrence. The recent introduction of immunotherapies resulted in clinical responses for a subgroup of patients, but there were still no effective predictive markers for response to immunotherapy or for recurrence after surgical therapy. The identification of biomarkers that could correlate and predict response or recurrence would require close monitoring of the patients throughout and after the completion of treatment. However, this would not be performed efficiently by repeated and invasive tissue biopsies. A better approach would be to use liquid biopsies including circulating tumor DNA (ctDNA), circulating RNA (e.g., microRNAs), circulating tumor cells (CTC) and extracellular vesicles (EVs) (e.g., exosomes) for disease monitoring in a non-invasive manner. In this review, we discuss the currently available technology that can enable the use of liquid biopsy as a diagnostic and prognostic tool. Moreover, we discuss the opportunities and challenges of the clinical application of liquid biopsy for immunotherapy of HCC.

1. Introduction

It has been estimated that more than one million deaths will be attributed to hepatocellular carcinoma (HCC) by 2030 [1], making it one of the deadliest cancer types worldwide. Depending on the stage of HCC, the treatment options can vary. When diagnosed at an early stage, the standard of care options include resection, local ablation or liver transplantation, but the risk of tumor recurrence still remains high [2]. When diagnosed at an intermediate stage, treatment options are limited to transarterial chemoembolization, whereas systemic therapies, such as the multi-kinase inhibitors sorafenib or lenvatinib, have, until recently, been the treatment of choice for late-stage tumors [3,4]. For advanced HCC, as for a few other solid cancers, immunotherapy is one of the most promising and novel treatment approaches. A number of ongoing clinical trials have been reported [2,5] in which various immunotherapies, such as immune checkpoint inhibitors (ICIs), are utilized for the treatment of HCC, either alone or in combination with targeted and/or systemic therapies [2,5].
Despite the great clinical benefit that immunotherapies have offered, there are still many patients who do not respond or respond poorly to this type of treatment. In particular, only about 15–20% of advanced HCC patients respond to ICIs [6]. The reasons for the unsatisfactory clinical responses are not clear. One major area of research is indeed focusing on the identification of biomarkers that can better predict tumor responses to the immunotherapy, in order to improve the clinical outcomes and cover a broader number of cancer patients. Various biomarkers have been shown to predict responses to ICIs including tumoral PD-L1 expression [7] and tumor mutational burden (TMB) [8]. A higher TMB, based on genomic profiling of various tumor biopsies, may reflect a higher likelihood for response to ICIs [9], whereas PD-L1 expression can positively correlate with better responses to anti-PD-L1 therapy [10,11]. In addition, gene expression analysis on HCC adjacent tumor tissues has been able to identify signatures correlated with improved survival [12]. However, tissue biopsies require invasive tumor sampling, therefore making it harder to collect multiple samples and comprehensively track tumor genomic changes throughout the treatment [13,14,15]. In particular, HCC is a heterogeneous and molecularly complex cancer type, and conventional tissue biopsies are not able to fully reflect its heterogeneity and thus accurately predict therapy efficacy [16]. In addition, unlike other solid tumors, tissue biopsies for HCC are not frequently available, since diagnosis mainly relies on imaging [17]. Additionally, at a late stage, when the lesions are unresectable, a liver biopsy is usually not recommended for advanced HCC [18], while there is a risk of extrahepatic tumor spread along the needle track [19].
Liquid biopsy, where only a blood sample is taken to analyze circulating tumor cells (CTC) [20] or circulating tumor DNA (ctDNA) [21], can overcome these issues due to its minimally invasive nature. Additionally, it can be used to monitor the disease status systematically [22]. Alpha-fetoprotein (AFP) is one of the first liquid biopsy biomarkers used for the early diagnosis of HCC [23]. However, concerns about its sensitivity and high levels of AFP in non-HCC patients highlight the need to identify more sensitive and reliable biomarkers, which can be used alone or in combination with AFP [16].
In this review, we will describe the currently available technology that can enable the use of liquid biopsy, with the focus on CTCs and ctDNA as diagnostic and prognostic tools. Furthermore, we will discuss the opportunities and challenges of the clinical application of liquid biopsy for immunotherapy of HCC.

2. Liquid Biopsy in HCC

Liquid biopsy refers to all the non-solid biologic materials used for the diagnosis and monitoring of HCC and is mainly based on the detection of ctDNA [24,25,26], circulating RNAs (e.g., microRNAs) [27,28,29], CTCs [20,30,31] and extracellular vesicles (EVs) (e.g., exosomes) [32,33,34] (Figure 1). In the following paragraphs, we will focus on two of the most promising liquid biopsy types in HCC, ctDNA and CTCs.

2.1. ctDNA in HCC

ctDNA can arise in the bloodstream of cancer patients as a result of tumor cell apoptosis or necrosis [35]. As ctDNA represents the total tumor genome, its role in determining clinical outcomes gains more and more attention, especially in cases of advanced and unresectable HCC in which surgical or other invasive procedures, including tissue biopsy, are not recommended [18]. ctDNA contains cancer-associated molecular characteristics, such as mutational signatures [36], epigenetic changes [37,38] and cancer-derived viral sequences [39], which allow its discrimination from total normal circulating cellular free DNA [40,41,42]. Therefore, it could significantly contribute to the improvement in sensitivity for the current diagnostic tools, such as AFP, whose sensitivity remains at an average of 50% among HCC cases [16]. In a study including 42 patients with unresectable primary liver cancer, ctDNA could correlate more closely with the tumor load and could predict treatment efficacy with higher sensitivity, compared to AFP or imaging [18]. Another pilot study showed that ultra-deep targeted sequencing of cell-free DNA (cfDNA) could confidently detect somatic mutations, which were previously identified in tissue biopsies and were frequently found in HCC patients, highlighting the benefits of cfDNA-derived mutation sequencing [43]. Similarly, in a larger cohort study enrolling 121 advanced HCC patients, mutation profiling of ctDNA revealed mutations in the most frequent HCC-associated driver oncogenes and tumor suppressors, including the TERT promoter, TP53, CTNNB1, PTEN, AXIN1, ARID2, KMT2D and TSC2. This technique was able to reveal predictive mutational signatures associated with responses to systemic therapy with tyrosine kinase inhibitors (TKIs) [36]. Other studies based on detection and mutational analysis of ctDNA also showed promising results [40,44,45,46,47].
Despite the proven valuable role of ctDNA as a tumor biomarker, it still has some limitations, including the low levels of detection in the early stages, which makes it challenging for the early diagnosis of HCC. Another limitation is the lack of standardized procedures to prepare samples and analyze data [16]. Lastly, this approach is limited by the uncertain ability to capture spatial tumor heterogeneity, which can reflect clonal differences within or across cancer metastatic sites [48]. This implies that combinational and/or multi-parametric approaches may be needed, in order to increase the sensitivity and specificity of ctDNA as a biomarker for HCC.

2.2. CTCs in HCC

CTCs are also emerging as a promising biomarker for the prediction of HCC treatment efficacies. CTCs arise in the circulation after detachment from primary or metastatic tumor lesions [49]. They differ from other types of cancer biomarkers as they represent viable tumor cells circulating in the patient’s bloodstream. Therefore, CTCs can also provide comprehensive genetic information about tumor heterogeneity and drug sensitivity [20]. CTCs have been approved by the FDA as diagnostic markers for specific epithelial cancers [16]. However, their diagnostic role in HCC still requires further studies. A widely known CTC biomarker is the epithelial cell adhesion molecule (EpCAM) [49], a pan-cancer biomarker which has also been observed in HCC patients [50]. Several studies highlighted the role of EpCAM+ CTCs in predicting HCC recurrence after surgery, as well as their associations with disease progression, vascular invasion and overall survival [51]. Detection of EpCAM-positive CTCs with co-existing T regulatory cells (CD4+/CD25+/Foxp3+) indicated HCC recurrence [52]. In this study, patients with high CTC/Tregs levels had a significantly higher risk of developing postoperative HCC recurrence than those with low CTC and Treg levels. In addition, other subtypes of CTCs were also explored, including the presence of mesenchymal CTCs, which were also associated with a higher risk of tumor recurrence in HCC patients [53,54].
Despite the highly promising role of CTCs as a biomarker for HCC [20], it remains challenging to detect HCC CTCs early and accurately because of the lack of specific markers. Another limitation is that the frequencies of CTCs are usually low in the circulation, especially at the early stages [16,20]. Thus, combinational strategies may be needed, in order to improve the prognostic and diagnostic value of HCC. We will discuss this in detail in Section 4.

3. Technology Platforms for Isolation and Detection

3.1. The Detection of CTCs

(i) 
Biophysics-based approach
The first approach utilizes the difference in biophysical properties between CTCs and normal blood cells, such as size and density. For example, the ISET (isolation by size of epithelial tumor cells) technique utilizes a polycarbonate membrane with calibrated, 8 mm-diameter, cylindrical pores to filter out rare CTCs from the blood [55]. After filtering, each membrane is allowed to dry and then stained with H&E or antibodies. In addition, fluorescence in situ hybridization (FISH) and PCR-based genetic analyses can be applied to ISET-isolated cells. The ISET technique has been used to isolate CTCs from HCC patients [55,56].
Another approach is based on the difference in density. Ficoll-Paque was originally developed to isolate peripheral blood mononuclear cells (PBMCs) from other blood components. Ficoll-Paque is placed at the bottom of a conical tube, and then the blood is layered above Ficoll-Paque. After centrifugation, PBMCs are located in a layer between the plasma and Ficoll-Paque. The same approach can be used to enrich CTCs from other blood cells [57]. By following the same process, CTCs are enriched in the PBMC layer. In general, biophysics-based approaches can easily enrich CTCs, but the purity is far less than the antibody-based approach. Therefore, additional steps are required to detect CTCs, such as antibody staining or PCR-based approaches [57,58,59] (Figure 2Ai).
(ii) 
Antibody-based approach
The most common approach is using an antibody to detect the cell surface marker EpCAM on CTCs. EpCAM is over-expressed on tumor cells, but it can be expressed on some normal epithelial cells [60]. CellSearch is the first FDA-cleared test to provide CTC-related information to clinicians [61,62]. In a CellSearch system, CTCs in a plasma sample are enriched by EpCAM antibody-labeled magnetic beads. These cells are further stained with fluorescence-labeled CD45, cytokeratin-8 (CK-8), CK-18 and CK-19 antibodies. Cells with CD45(−), CK-8(+), CK-18(+) and CK-19(+ or −) are considered as CTCs. CellSearch has been cleared for use as a diagnostic test for patients with metastatic breast, prostate or colorectal cancer, in conjunction with other diagnostic tests. This system has also been used for several research studies in HCC [31,63,64]. In addition to CellSearch, other commonly used techniques, such as MACS (magnetic-activated cell sorting), have also been used to enrich antibody-labeled CTCs [65,66,67] (Figure 2Aii).

3.2. The Detection of ctDNA

(i) 
Quantitative PCR-based approach
The quantitative PCR-based approach utilizes pre-designed PCR primer/probe sets to detect known mutations, usually hotspot mutations, in the plasma DNA. In recent years, scientists have employed droplet digital PCR (ddPCR) techniques to further improve the precision [68]. ddPCR is very similar to quantitative PCR, except that one or a few DNA templates are compartmentalized by small droplets. Therefore, DNA templates can be amplified independently without interference, in order to achieve a better precision. In addition, ddPCR overcomes issues of normalization to housekeeping since it is based on absolute quantification of sample fractionalizing and statistics correction for multiple target molecules identified per droplet [69]. In an HCC study, ddPCR could detect hotspot mutations in plasma samples from 48 HCC patients [70]. In this study, ddPCR assays were performed to target TP53 (c.747G > T), CTNNB1 (c.121A > G, c.133T > C) and TERT (c.1-124C > T) using wild-type and mutant probes. A total of 56.3% of the HCC patients in the study were found to have at least one of these mutations in ctDNA (Figure 2Bi).
(ii) 
NGS-based approach
The NGS (next-generation sequencing) technique can be utilized to detect tumor DNA from patients’ plasma, by either detecting tumor mutations or chromosome abnormalities. The advantage of this approach is that it can detect tumor DNA in plasma without prior knowledge or pre-defined hotspot sites. However, because the vast majority of plasma DNA comes from normal cells, it can be technically challenging to conduct this.
Tumor cells usually have some chromosomal alterations, including copy number changes, amplifications, deletions and translocations, which can be detected by low-coverage whole-genome sequencing, followed by bioinformatics analysis [71]. The sensitivity and specificity of this approach largely rely on the depths of sequencing and also the design of bioinformatic pipelines. Additionally, the nature of tumor cells, such as the magnitude of chromosomal alterations, can also directly affect the sensitivity of this approach. This approach has been used in many cancer types, including HCC. For example, shortened DNA associated with copy number aberrations was detected in 84% of HCC patients in one report [72].
A more sensitive approach is to detect tumor mutations directly by NGS. For instance, an approach to detect tumor mutations from plasma, called CancerSEEK, has been developed [21]. Plasma samples were collected from 1005 patients with nonmetastatic, clinically detected common cancers, including HCC. DNA materials were purified from plasma and then amplified by 61 PCR primer pairs targeting regions of interest from 16 genes, including TP53, KRAS, CTNNB1, PIK3CA, APC, EGFR, PTEN, FGFR2, CDKN2A, GNAS, PPP2R1A, AKT1, HRAS, BRAF, FBXW7 and NRAS. PCR products were deep sequenced, and hotspot mutations were identified by bioinformatic analysis. The median sensitivity of CancerSEEK was 70% among eight cancer types, and 97% for HCC. The specificity for CancerSEEK was greater than 99%.
Although NGS approaches could detect ctDNA in patients with some tumor burdens, the sensitivity was not sufficient to detect residual diseases for patients with nonmetastatic cancer. A targeted digital sequencing (TARDIS) approach was developed to improve the sensitivity [73]. Tumor mutations from each individual breast cancer patient were identified by whole-exome sequencing using their tumor biopsy specimens. Approximately 18 patient-specific mutations were selected for PCR amplification for each patient. Importantly, a random nucleotide sequence, also known as a unique molecular identifier (UMI), was added to each single-stranded DNA template prior to the PCR amplification. The PCR products were deep sequenced, and the UMI and fragment size were used to remove potential errors introduced during the PCR amplification. This approach could improve the sensitivity up to 100-fold, compared to other approaches. As a result, ctDNA was detected in 100% of patients with early and locally advanced breast cancer prior to treatments, and in 12 out of 13 patients with invasive or in situ residual diseases after treatments (Figure 2Bii).

4. Liquid Biopsy as a Diagnostic and Prognostic Tool

Historically, serum AFP and diagnostic imaging have been the primary diagnostic modalities used for HCC [74]. Elevated levels of AFP have been associated with increased tumor size and portal vein thrombosis, as well as an increased risk of recurrence after liver transplantation [51,75]. However, the role of AFP as a biomarker has a limited diagnostic value because of the low sensitivity in HCC, at 50% [51]. Alternatively, other markers such as the AFP lectin fraction (AFP-L3) and des-y-carboxy prothrombin (DCP) have been shown to improve the diagnostic performances when used in combination with AFP [76]. Besides these, Glypican 3 (GPC3) [77], cytokeratin 19 (CK19) [78], Golgi protein 73 (GP73) [79], osteopontin [80], squamous cell carcinoma antigen (SCCA) [81] and annexin A2 [82] have all been shown to have diagnostic and prognostic roles in HCC as well, but they have not been validated sufficiently for routine clinical use [20].
Therefore, research on biomarker combinations has been performed in order to provide more accurate and valuable information for a future individualized HCC diagnosis and/or prognosis assessment [83]. In this context, liquid biopsy has been explored as a way to monitor cancer prognosis and diagnosis in a non-invasive manner. This technology has shown promising results in early diagnosis [84], detection of minimal residual disease [85] and decision making for systemic therapies of different types of cancers, including HCC [8,43,86].
Among all liquid biopsy analytes, ctDNA plays an important role in HCC prognosis [17]. ctDNA maintains the same genomic signatures that are present in the matching tumor tissue, allowing for the quantitative and qualitative evaluation of the mutation burden in body fluids [87]. In this way, ctDNA has been considered as a good biomarker and can be utilized in disease monitoring. The data of ctDNA include quantitative changes, such as differences in the concentration of ctDNA, as well as qualitative changes, such as gene mutations, DNA copy number variations and DNA methylation [16]. Indicatively, studies based on the detection of somatic single-nucleotide mutations and methylation changes in ctDNA could closely correlate with tumor burden over time in HCC patients and could be used to predict recurrence after surgery [17,88,89]. Initially, it was shown that p15 and p16 methylations were positive in the plasma/serum of 92% of HCC patients [90]. In another study, Ras association domain family 1A(RASSF1A) promoter hypermethylation was detected in 90% of cases, with an overall predictive accuracy of 77.5%, compared to healthy controls [91]. In addition to RASSF1A, two abnormally methylated genes (APC and COX2) and one miRNA (miR-203) were combined to establish a predictive model by which nearly 75% of HCC patients were detected, who could not be diagnosed with AFP [92].
As ctDNA represents only a very small proportion of cell-free DNA, very sensitive and reliable detection methods are required. Levels of ctDNA are measured mainly by real-time PCR (RT-PCR) [93], while digital PCR (dPCR) [94] or sequencing methods are used for the detection of point mutations [95]. In addition to TERT and TP53 mutations as the prognostic factors of poor survival [45,47], other mutations have been shown to have prognostic values for HCC. MLH1 mutation was specifically associated with lower survival [1], whereas mutations of genes from the PI3K/mTOR pathway were shown to be the predictors of non-responders to TKI treatments for patients with advanced HCC [86].
A number of studies have also shown the prognostic values of circulating miRNAs in HCC. Lower survival rates were associated with patients with low levels of miR-1, miR-122, miR-26a, miR-29a and miR-223-3p [96,97,98,99] or high levels of miR-155, miR-96 and miR-193-5p [100,101]. Furthermore, six additional miRNAs were identified as prognostic factors. Low levels of miR-424-5p or miR- 101-3p and high levels of miR-128, miR-139-5p, miR-382-5p and miR410 were associated with lower survival rates in HCC patients [17]. Alternatively, miRNAs have been studied in association with EVs [32,33,34]. In a cohort of 59 HCC patients, a correlation was found between tumor recurrences after liver transplantation and a high level of exosomal miR-718 [102]. Additionally, high levels of exosomal miR-665 or low levels of exosomal miR-638 and miR-320a were identified as predictors of poor survival [103,104,105].
Another cornerstone of liquid biopsy is the isolation and detection of CTCs, which have been described as a useful tool for the prognostication of HCC [106]. As introduced above, EpCAM-positive CTC cells have been intensively investigated in HCC [50,51]. However, since CTCs can lose their epithelial phenotype through epithelial-to-mesenchymal transition (EMT) in order to survive and metastasize [49], EpCAM cannot always be considered an optimal biomarker to detect HCC. Alternatively, other phenotypic markers have been explored, such as the hepatocyte-specific asialoglycoprotein receptor (ASGPR) [67], and the hepatocyte paraffin 1 [107], or incorporation of several markers simultaneously, as it has extensively been reviewed elsewhere [20]. Most recently, in a prospective study of 80 HCC patients, a multimarker assay combining cell surface markers EpCAM, ASGPR and GPC3 was able to detect HCC CTCs in 97% of the patients with high accuracy. Moreover, a phenotypic variant subset of CTCs was associated with aggressive disease progression and underlying metastasis, therefore highlighting the important implications of CTCs in treatment selection [108]. Another study showed that the detection of phosphorylated ERK (pERK) and pAkt in CTCs could predict the response to sorafenib efficacy in advanced HCC patients, similarly to tumor tissue biopsy [59].

5. Clinical Applications for Immunotherapy in HCC

5.1. Immunotherapy in HCC

After years of sorafenib predominance and desolated perspectives, the skyline of systemic therapies for unresectable advanced HCC has considerably grown in the last decade. Not only more angiogenesis- and proliferation pathway-directed targeted therapies are available (TKIs; monoclonal antibodies), covering the therapeutic scenario from the first- to the third-line of treatment [109], but also ICIs are now well-established active agents and are gaining growing attention in the context of liver cancer [5]. Immuno-oncology (IO) represents a major breakthrough in this context, leading to a significant increase in median overall survival (OS) and to the possibility of long-term survival [110].
Nivolumab and pembrolizumab showed promising anti-tumoral activity with a 20% and 17% objective response rate (ORR), respectively, in patients who were refractory or intolerant to sorafenib in the phase 1/2 Checkmate-040 [6] and phase 2 Keynote-224 [111] clinical trials, respectively. Based on these results, the FDA granted an accelerated approval to the two anti-PD-1 antibodies for the second-line treatment of HCC. Notwithstanding these relevant and exciting results, the phase 3 trials testing nivolumab in the first line (Checkmate-459) [112] and pembrolizumab in the second line (Keynote-240) [113] of treatment failed to meet the protocol-defined statistical significance threshold of their primary endpoints. However, a clinically meaningful improvement in the overall response rate and complete responses were seen, confirming that ICIs are active and stressing their role for the treatment of liver cancer. Importantly, the safety profile was favorable and consistent with that of the primary analysis, supporting a favorable risk-to-benefit ratio. More recently, both the FDA and European Medicine Agency (EMA) approved the combination of atezolizumab (anti-PD-L1) and bevacizumab (anti-Vascular Endothelial Growth Factor-A monoclonal antibody) as a first-line systemic option for unresectable advanced-stage HCC, based on the striking results of the phase 3 trial IMbrave150 [114,115]. This study reported a 27% ORR and showed clinically meaningful and statistically significant improved outcomes for the experimental arm (vs. sorafenib) in terms of OS (NR (not reached) vs. 13.2 months, HR 0.58) and progression-free survival (6.8 vs. 4.3 months, HR 0.59) [4]. The treatment benefit was also meaningful after 12 months of additional follow-up, showing a median OS of 19.2 months with atezolizumab and bevacizumab vs. 13.4 months with sorafenib (HR 0.66, p = 0.0009), and a 29.8% ORR for the experimental arm, consistent with the primary analysis. Providing the longest survival ever reached in a first-line phase 3 trial in advanced HCC, this combination represents a practice-changing first-line treatment for HCC patients [116]. Following this stream, novel immuno-oncology-based combinations (ICI + ICI, ICI + TKI) are currently under development with the promise of improving the therapeutic management not only of advanced HCC but also of the early (in neo-adjuvant and adjuvant settings) and intermediate stages of the disease [110].
A common finding, that all trials, to date, have definitely revealed, is a significant heterogeneity in the magnitude of the tumor response to IO drugs and in the duration of the clinical response [117]. Despite a clear clinical benefit of immunotherapies in HCC, ICI showed efficacy only in a minority of HCC patients. A deeper knowledge of the dynamic interplay between all the components of the HCC ecosystem (tumor and immune cells, stromal cells, endothelial cells and nonmalignant cells) [1] is needed to dissect the mechanisms behind the clinical response to ICI and could result in an higher efficacy rate in the future.
Today, treatment allocation is not guided by any particular tumor characteristic, since no biomarker can effectively predict the response to a particular drug; the only exceptions, at present, are AFP and ramucirumab (anti-Vascular Endothelial Growth Factor Receptor-2 monoclonal antibody) [118,119]. Therefore, a crucial, urgent and still unmet need in the context of HCC research is the availability of robust and validated predictive biomarkers than can help in identifying the best candidates for a personalized IO approach [5,110,117,120]. Predictive biomarkers for the IO response and/or IO resistance might be extremely useful in order to optimize patient selection, spare toxicity to patients unlikely to respond and improve the design of clinical trials in the upcoming years [4,117].
The expression of PD-L1 on tumor cells, assessed by immunohistochemistry (IHC) from formalin-fixed paraffin-embedded (FFPE) tissue sections, is the only approved biomarker commonly used in the clinical routine to identify subgroups of patients with a higher chance of benefit from ICI [120,121]. This biomarker, however, is far from being perfect and does not confidently predict the response to cancer treatment. A few reasons can be indicated for this poor predictive value, for example, the discrepancy in PD-L1 IHC assessment in terms of the positivity cutoff, which is mostly related to the use of different detection antibodies or the choice of cell type to be stained [122]. Moreover, the unavailability of tissue or the low percentage of tumor cells in the biopsy makes this test not always feasible [120]. Despite the IHC measure, a proportion of PD-L1-“negative” patients will still respond and a proportion of PD-L1-“positive” patients will fail, showing how complex and still not fully unraveled the interplay is between cancer and the immune system [120,123]. In addition, tumor heterogeneity as a result of tumor evolution and subsequent clonal mutational differences [15] or due to treatment-induced resistant sub-clones [124] may further challenge disease monitoring and choice of therapy. These issues have prompted research to discover other markers, exploring not only tissue but also blood samples of cancer patients [121].

5.2. Liquid Biopsy for Immunotherapy in HCC

The race towards the identification of immunotherapy predictive biomarkers is at the forefront of research in HCC. Among the biomarkers of interest, there are TMB and mutational signatures identified from ctDNA, and PD-L1 expression detected on CTCs [125]. TMB and PD-L1 expression are considered good predictors in several cancers, but the evidence in liver cancer has not been as established thus far [117]. In Table 1, we summarize the most recent literature in the field, and we highlight the key findings for each study.
A proof-of-concept study was able to show that changes in the ctDNA levels could significantly correlate with tumor size in cancer patients treated with anti-PD1 drugs and be a valuable prognostic factor of progression-free and overall survival [126]. Another recent emerging predictor of immunotherapy efficacy is considered to be the TMB which is defined as the total number of somatic non-synonymous mutations per mega-base identified in tumor tissue or circulating tumor DNA [2,127]. In one study, targeted gene analysis of ctDNA showed high consistency in the levels of TMB between tissue and blood samples which were higher in TP53-mutated patients with advanced liver cancer, indicating that ctDNA analysis could be a better option to evaluate TMB prior to immunotherapy in cases of advanced primary liver cancers where tissue biopsy was not recommended [18]. However, in another prospective study of advanced HCC, mutational analysis of ctDNA could not be associated with response to ICI therapy but only to systemic treatment with TKIs [36], which implied that more and larger cohort clinical studies would be required in order to elucidate the potentials of mutational ctDNA analysis in determining immunotherapy efficacy.
In a small subset of patients enrolled in the phase 1b clinical trial receiving atezolizumab plus bevacizumab, higher levels of ctDNA at baseline were associated with an increased baseline tumor burden (p < 0.03), and ctDNA turned negative in 70%, 27%, 9% and 0% of patients achieving a complete response, partial response, stable disease and disease progression, respectively. Moreover, undetectable ctDNA levels during treatment were linked to a longer progression-free survival (PFS) [128], suggesting a role for the non-invasive ctDNA analysis in the prediction of immune response.
It has also been shown that a hyper-mutated ctDNA phenotype, the liquid counterpart of the tissue-based high level of TMB, was associated with a favorable outcome in a cohort of 69 cancer patients with different histologies, including three HCC patients, treated with different immune checkpoint inhibitors. In particular, the overall response rate, PFS and OS in the high-alteration groups, defined as variants of unknown significance (VUS) > 3 or total alterations ≥ 6, were significantly higher than in the low-alteration groups, defined as VUS ≤ 3 or total alterations < 6 (45% vs. 15% for high and low alterations, respectively, p = 0.014) [122].
Along with predictive biomarkers of response, the identification of predictive biomarkers of resistance is also compelling and might have relevant implications when designing future clinical trials. Among the oncogenic pathways that have been linked to potential tumor immunotherapy resistance, the Wnt/b-catenin signaling pathway is one of the earliest, with evidence coming from genomic, tissue and mouse model studies [129,130,131,132]. In this context, a preliminary study demonstrated that liquid biopsy was potentially able to detect Wnt/b-catenin-activating mutations in HCC [117]. However, another study could not confirm this hypothesis and postulated that the detection of Wnt/b-catenin pathway-activating mutations might not be sufficient to identify advanced HCC patients with primary resistance to ICI [119]. CTNNB1 is one of the genes involved in the Wnt/b-catenin pathway, and its mutations, such as p.T41A, are among the most prevalent genetic alterations in HCC [133]. Oversoe et al. evaluated the presence of the CTNNB1 p.T41A mutation comparing tumor tissue DNA and ctDNA and found that liquid biopsy managed to reveal mutations that were not detected in single tumor biopsies, thus increasing the detection rate of the CTNNB1 mutation in HCC patients and suggesting that ctDNA could empower the perspective of a tailored treatment strategy [134].
Table 1. Summary of the key findings in the most recent literature.
Table 1. Summary of the key findings in the most recent literature.
Type of Biomarker AnalyzedKey FindingsReference
Changes in the ctDNA levelsCould significantly correlate with tumor size in cancer patients treated with anti-PD1 drugs and be a valuable prognostic factor of progression-free and overall survival.[126]
Targeted gene analysis of ctDNACan be a better option to evaluate TMB prior to immunotherapy in cases of advanced primary liver cancers when tissue biopsy is not recommended.[18]
Mutational analysis of ctDNACould not be associated with response to ICI therapy but only to systemic treatment.[36]
Levels of ctDNA at baselineHigher levels of ctDNA at baseline were associated with an increased baseline tumor burden, and ctDNA turned negative in 70%, 27%, 9% and 0% of patients achieving a complete response, partial response, stable disease and disease progression, respectively.[128]
Undetectable ctDNA levels during treatment were linked to a longer progression-free survival.
Hyper-mutated ctDNA phenotypeIs associated with a favorable outcome in a cohort of 69 cancer patients with different histologies, including three HCC patients, treated with different immune checkpoint inhibitors.[122]
Overall response rate, PFS and OS in high-alteration groups were significantly higher than in low-alteration groups.
Detection of Wnt/b-catenin-activating mutationsWnt/b-catenin-activating mutations in HCC linked to potential tumor immunotherapy resistance in several studies.[129,130,131,132]
Detection of Wnt/b-catenin-activating mutationsDemonstration that liquid biopsy is potentially able to detect Wnt/b-catenin-activating mutations in HCC.[117]
Detection of Wnt/b-catenin-activating mutationsDetection of Wnt/b-catenin pathway-activating mutations might not be sufficient to identify advanced HCC patients with primary resistance to ICI.[119]
Targeted mutational analysis of CTNNB1 p.T41A mutationctDNA liquid biopsy managed to reveal mutations that were not detected in single tumor biopsies, thus increasing the detection rate of CTNNB1 mutation in HCC patients.[134]
PD-L1 expression on CTCsBiomarker to assess ICI-based immunotherapy efficacy of advanced solid tumors.[135]
CTCs expressing PD-L1PD-L1-positive CTCs are mainly found in advanced stages of disease, and they represent an independent prognostic factor for overall survival.
6 out of 10 patients receiving anti-PD-1 treatment had PD-L1-positive CTCs at baseline, and of these, 5 out of 6 had a favorable treatment response.
4 out of 10 patients receiving anti-PD-1 treatment did not have PD-L1+ CTCs and were non-responders.
[136]
Apart from mutational signatures and TMB in connection to ctDNA, CTCs expressing PD-L1 have been suggested as another promising biomarker to assess the ICI-based immunotherapy efficacy of advanced solid tumors [135]. Winograd and colleagues analyzed the expression of PD-L1 on CTCs and found that PD-L1-positive CTCs are mainly found in the advanced stages of disease and that they represent an independent prognostic factor for overal survival. Moreover, of 10 patients receiving anti-PD-1 treatment, 6 had PD-L1-positive CTCs at baseline, and of these, 5 had a favorable treatment response, whereas 4 patients did not have PD-L1+ CTCs and were non-responders [136]. They definitely unraveled the prognostic and potentially predictive value of CTCs in the context of HCC. Clearly, as pointed out by Hofman and coauthors, together with technical issues, some relevant questions could also be asked: e.g., (1) Is PD-L1 expression in CTCs correlated with matched tissue samples? (2) Is this expression homogeneous on all CTCs or is it restricted to a subpopulation of CTCs? (3) What is the implication of PD-L1 expression in circulating immune cells when associated with CTCs? [120]
Undoubtedly, liquid biopsy-based biomarkers for immunotherapy in HCC require further rigorous testing and validation in carefully and well-designed clinical trials where their performance can be evaluated and their clinical implication can be measured. Hopefully, liquid biopsy will be integrated in the daily clinical management of HCC in the near future.
Notably, it is likely that immune-genomic biomarkers may further enhance immunotherapy through novel additional checkpoint inhibitors, but also neoantigen vaccine or adoptive cell transfer approaches [5,137]. This can certainly be of paramount importance to magnify the immunotherapy treatment response and implement precision immune-oncology.

6. Limitations and Future Perspectives

Immunotherapy represents the latest and most promising clinical development for the treatment of advanced HCC. However, despite the great benefits immunotherapies have brought, there is still a significant number of cancer patients who do not respond or respond poorly to these new approaches [6]. Therefore, there is an imperative need to increase the reliability and enrich the repertoire of currently available biomarkers, in order to more accurately predict therapy efficacy and to improve response rates, compared to existing therapies. In particular, the attractive types of biomarkers are the ones that can be monitored without invasive procedures. Liquid biopsies, from which it is possible to isolate ctDNA and CTCs circulating in the blood stream of cancer patients, have shown promising data as prognostic and diagnostic tools for HCC while, at the same time, also allowing sequential and real-time monitoring of disease status in a minimally invasive manner [16,22,138]. This is especially important for patients with advanced and unresectable HCC, when surgical and other invasive procedures are not an option [18]. Interestingly, in some cases, ctDNA has been shown to be superior in identifying mutational signatures that could not be traced in single tumor biopsies [134] or that could correlate more closely with the tumor load and predict treatment efficacy with higher sensitivity, compared to AFP or imaging in patients with unresectable liver cancer [18]. These data prove the predictive value of analyzing ctDNA, which can add to the existing and well-established diagnostic tools for HCC. However, other studies have not been able to confirm the association of tumoral signature mutations with resistance to ICIs in HCC based on their ctDNA analysis [36]. These conflicting data highlight that larger and more comprehensive clinical studies are required, in order to obtain widely applicable and consistent results, which, at the moment, are missing in the field of liquid biopsy for HCC. CTCs are another important source of biomarkers, and many studies point out their role in the prognosis of HCC [51]. Importantly, CTCs have also been used to prognosticate responses to ICIs through the expression of PD-L1 in patients with advanced liver cancer [136]. However, low levels of CTCs in the early stages of HCC and the lack of standardized procedures make it challenging to integrate CTC techniques in the clinical practice for HCC diagnosis [51].
In this context, factors associated with the individual’s genetic background, the tumor microenvironment and interactions with the host immune system may additionally challenge the selection and evaluation of biomarkers able to predict tumor responses. In order to overcome these issues, integration of multiple biomarkers rather than single analytes as well as combinational approaches based on genomic and proteomic analyses will most probably be able to improve the precision of personalized treatments. Here, the implementation of novel NGS technologies and artificial intelligence might be of great importance to identify genomic and immunologic signatures predictive of treatment responses, as it has encouragingly been shown in recent studies [73,139,140,141].
Up to the present, ICIs have been the most widely used immune-based approaches in the clinical management of advanced HCC [110]. Therefore, the majority of cancer immunotherapy predictive biomarkers described thus far refer mainly to responses to ICIs. Other immunotherapy modalities such as CAR-T cell therapies and neoantigen vaccines are currently being tested in ongoing clinical studies for HCC, and the preliminary results are awaited with great interest [2,5].

7. Conclusions

Although liquid biopsy biomarkers for HCC are not well evaluated, compared to other malignancies, there are many studies highlighting their prognostic and diagnostic value for clinical management. Here, we focus on ctDNA and CTCs, two emerging liquid biopsy biomarkers, and their role in HCC immunotherapy. In the near future, the identification and further validation of novel and existing biomarkers, as well as incorporation of high-throughput technologies, will remain of paramount importance in order to improve precision in treatment decision making in the field of immuno-oncology, especially for HCC immunotherapy.

Author Contributions

Conceptualization, A.P. and P.M.; writing—original draft preparation, P.M., D.N.S., G.R. and Y.-C.L.; writing—review and editing, P.M., D.N.S., G.R., M.C., P.S. and Y.-C.L. All authors have read and agreed to the published version of the manuscript.

Funding

P.M., D.N.S., M.C. and A.P. were supported by the Sjöberg Foundation (2020-01-07-04), Center for Innovative Medicine (CIMED FoUI-954541), Svenska Läkaresällskapet (SLS-935150), The Swedish Cancer Society (200769Pj01H) and the Ruth and Richard Julins Foundation (2020-00263). Y.-C.L. was supported by funding from the Winthrop P. Rockefeller Cancer Institute and Translational Research Institute KL2 Award (TR003107 and TR003108). P.S. was supported by the Stockholm County Council (K2017-4579 and ALF LS 2018-1157), Center for Innovative Medicine (CIMED 20180889) and The Swedish Cancer Society (170690).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kim, E.; Viatour, P. Hepatocellular carcinoma: Old friends and new tricks. Exp. Mol. Med. 2020, 52, 1898–1907. [Google Scholar] [CrossRef]
  2. Lai, E.; Astara, G.; Ziranu, P.; Pretta, A.; Migliari, M.; Dubois, M.; Donisi, C.; Mariani, S.; Liscia, N.; Impera, V.; et al. Introducing immunotherapy for advanced hepatocellular carcinoma patients: Too early or too fast? Crit. Rev. Oncol. Hematol. 2021, 157, 103167. [Google Scholar] [CrossRef]
  3. Forner, A.; Reig, M.; Bruix, J. Hepatocellular carcinoma. Lancet 2018, 391, 1301–1314. [Google Scholar] [CrossRef]
  4. Finn, R.S.; Qin, S.; Ikeda, M.; Galle, P.R.; Ducreux, M.; Kim, T.-Y.; Kudo, M.; Breder, V.; Merle, P.; Kaseb, A.O.; et al. Atezolizumab plus Bevacizumab in Unresectable Hepatocellular Carcinoma. N. Engl. J. Med. 2020, 382, 1894–1905. [Google Scholar] [CrossRef] [PubMed]
  5. Sangro, B.; Sarobe, P.; Hervás-Stubbs, S.; Melero, I. Advances in immunotherapy for hepatocellular carcinoma. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 525–543. [Google Scholar] [CrossRef] [PubMed]
  6. El-Khoueiry, A.B.; Sangro, B.; Yau, T.; Crocenzi, T.S.; Kudo, M.; Hsu, C.; Kim, T.-Y.; Choo, S.-P.; Trojan, J.; Welling, T.H.; et al. Nivolumab in patients with advanced hepatocellular carcinoma (CheckMate 040): An open-label, non-comparative, phase 1/2 dose escalation and expansion trial. Lancet 2017, 389, 2492–2502. [Google Scholar] [CrossRef]
  7. Davis, A.A.; Patel, V.G. The role of PD-L1 expression as a predictive biomarker: An analysis of all US Food and Drug Administration (FDA) approvals of immune checkpoint inhibitors. J. Immunother. Cancer 2019, 7, 1–8. [Google Scholar] [CrossRef]
  8. Lee, E.Y.; Kulkarni, R.P. Circulating biomarkers predictive of tumor response to cancer immunotherapy. Expert Rev. Mol. Diagn. 2019, 19, 895–904. [Google Scholar] [CrossRef] [Green Version]
  9. Goodman, A.M.; Kato, S.; Bazhenova, L.; Patel, S.P.; Frampton, G.M.; Miller, V.; Stephens, P.J.; Daniels, G.A.; Kurzrock, R. Tumor Mutational Burden as an Independent Predictor of Response to Immunotherapy in Diverse Cancers. Mol. Cancer Ther. 2017, 16, 2598–2608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Kowanetz, M.; Zou, W.; Gettinger, S.N.; Koeppen, H.; Kockx, M.; Schmid, P.; Kadel, E.E.; Wistuba, I., 3rd; Chaft, J.; Rizvi, N.A.; et al. Differential regulation of PD-L1 expression by immune and tumor cells in NSCLC and the response to treatment with atezolizumab (anti-PD-L1). Proc. Natl. Acad. Sci. USA 2018, 115, e10119–e10126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Herbst, R.S.; Baas, P.; Kim, D.-W.; Felip, E.; Perez-Gracia, J.L.; Han, J.-Y.; Molina, J.; Kim, J.-H.; Arvis, C.D.; Ahn, M.-J.; et al. Pembrolizumab versus docetaxel for previously treated, PD-L1-positive, advanced non-small-cell lung cancer (KEYNOTE-010): A randomised controlled trial. Lancet 2015, 387, 1540–1550. [Google Scholar] [CrossRef]
  12. Hoshida, Y.; Villanueva, A.; Kobayashi, M.; Peix, J.; Chiang, D.; Camargo, A.; Gupta, S.; Moore, J.; Wrobel, M.J.; Lerner, J.; et al. Gene Expression in Fixed Tissues and Outcome in Hepatocellular Carcinoma. N. Engl. J. Med. 2008, 359, 1995–2004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Quandt, D.; Zucht, H.D.; Amann, A.; Wulf-Goldenberg, A.; Borrebaeck, C.; Cannarile, M.; Lambrechts, D.; Oberacher, H.; Garrett, J.; Nayak, T.; et al. Implementing liquid biopsies into clinical decision making for cancer immunotherapy. Oncotarget 2017, 8, 48507–48520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Cheng, J.; Li, M.; Lv, Y. Sublethal heat treatment promotes epithelial-mesenchymal transition and enhances the malignant potential of hepatocellular carcinoma. Hepatology 2013, 59, 1650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Jamal-Hanjani, M.; Quezada, S.; Larkin, J.; Swanton, C. Translational Implications of Tumor Heterogeneity. Clin. Cancer Res. 2015, 21, 1258–1266. [Google Scholar] [CrossRef] [Green Version]
  16. Mocan, T.; Simao, A.L.; Castro, R.E.; Rodrigues, C.M.P.; Slomka, A.; Wang, B.; Strassburg, C.; Wohler, A.; Willms, A.G.; Kornek, M. Liquid Biopsies in Hepatocellular Carcinoma: Are We Winning? J. Clin. Med. 2020, 9, 1541. [Google Scholar] [CrossRef] [PubMed]
  17. Labgaa, I.; Villanueva, A.; Dormond, O.; Demartines, N.; Melloul, E. The Role of Liquid Biopsy in Hepatocellular Carcinoma Prognostication. Cancers 2021, 13, 659. [Google Scholar] [CrossRef]
  18. Zhao, W.; Qiu, L.; Liu, H.; Xu, Y.; Zhan, M.; Zhang, W.; Xin, Y.; He, X.; Yang, X.; Bai, J.; et al. Circulating tumor DNA as a potential prognostic and predictive biomarker during interventional therapy of unresectable primary liver cancer. J. Gastrointest. Oncol. 2020, 11, 1065–1077. [Google Scholar] [CrossRef]
  19. Silva, M.A.; Hegab, B.; Hyde, C.; Guo, B.; Buckels, J.A.C.; Mirza, D.F. Needle track seeding following biopsy of liver lesions in the diagnosis of hepatocellular cancer: A systematic review and meta-analysis. Gut 2008, 57, 1592–1596. [Google Scholar] [CrossRef]
  20. Chen, F.; Zhong, Z.; Tan, H.-Y.; Wang, N.; Feng, Y. The Significance of Circulating Tumor Cells in Patients with Hepatocellular Carcinoma: Real-Time Monitoring and Moving Targets for Cancer Therapy. Cancers 2020, 12, 1734. [Google Scholar] [CrossRef]
  21. Cohen, J.D.; Li, L.; Wang, Y.; Thoburn, C.; Afsari, B.; Danilova, L.; Douville, C.; Javed, A.A.; Wong, F.; Mattox, A.; et al. Detection and localization of surgically resectable cancers with a multi-analyte blood test. Science 2018, 359, 926–930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Li, J.; Han, X.; Yu, X.; Xu, Z.; Yang, G.; Liu, B.; Xiu, P. Clinical applications of liquid biopsy as prognostic and predictive biomarkers in hepatocellular carcinoma: Circulating tumor cells and circulating tumor DNA. J. Exp. Clin. Cancer Res. 2018, 37, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Ryder, S.D. Guidelines for the diagnosis and treatment of hepatocellular carcinoma (HCC) in adults. Gut 2003, 52, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Alunni-Fabbroni, M.; Rönsch, K.; Huber, T.; Cyran, C.C.; Seidensticker, M.; Mayerle, J.; Pech, M.; Basu, B.; Verslype, C.; Benckert, J.; et al. Circulating DNA as prognostic biomarker in patients with advanced hepatocellular carcinoma: A translational exploratory study from the SORAMIC trial. J. Transl. Med. 2019, 17, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Huang, A.; Zhang, X.; Zhou, S.-L.; Cao, Y.; Huang, X.-W.; Fan, J.; Yang, X.-R.; Zhou, J. Plasma Circulating Cell-free DNA Integrity as a Promising Biomarker for Diagnosis and Surveillance in Patients with Hepatocellular Carcinoma. J. Cancer 2016, 7, 1798–1803. [Google Scholar] [CrossRef] [Green Version]
  26. Xiong, Y.; Xie, C.-R.; Zhang, S.; Chen, J.; Yin, Z.-Y. Detection of a novel panel of somatic mutations in plasma cell-free DNA and its diagnostic value in hepatocellular carcinoma. Cancer Manag. Res. 2019, 11, 5745–5756. [Google Scholar] [CrossRef]
  27. El-Tawdi, A.H.F.; Matboli, M.; Shehata, H.H.; Tash, F.; El-Khazragy, N.; Azazy, A.E.-S.M.; Abdel-Rahman, O. Evaluation of Circulatory RNA-Based Biomarker Panel in Hepatocellular Carcinoma. Mol. Diagn. Ther. 2016, 20, 265–277. [Google Scholar] [CrossRef]
  28. Moshiri, F.; Salvi, A.; Gramantieri, L.; Sangiovanni, A.; Guerriero, P.; De Petro, G.; Bassi, C.; Lupini, L.; Sattari, A.; Cheung, D.; et al. Circulating miR-106b-3p, miR-101-3p and miR-1246 as diagnostic biomarkers of hepatocellular carcinoma. Oncotarget 2018, 9, 15350–15364. [Google Scholar] [CrossRef]
  29. Yamamoto, Y.; Kondo, S.; Matsuzaki, J.; Esaki, M.; Okusaka, T.; Shimada, K.; Murakami, Y.; Enomoto, M.; Tamori, A.; Kato, K.; et al. Highly Sensitive Circulating MicroRNA Panel for Accurate Detection of Hepatocellular Carcinoma in Patients With Liver Disease. Hepatol. Commun. 2019, 4, 284–297. [Google Scholar] [CrossRef]
  30. Kelley, R.K.; Magbanua, M.J.M.; Butler, T.M.; Collisson, E.A.; Hwang, J.; Sidiropoulos, N.; Evason, K.; McWhirter, R.M.; Hameed, B.; Wayne, E.M.; et al. Circulating tumor cells in hepatocellular carcinoma: A pilot study of detection, enumeration, and next-generation sequencing in cases and controls. BMC Cancer 2015, 15, 1–11. [Google Scholar] [CrossRef]
  31. Von Felden, J.; Schulze, K.; Krech, T.; Ewald, F.; Nashan, B.; Pantel, K.; Lohse, A.W.; Riethdorf, S.; Wege, H. Circulating tumor cells as liquid biomarker for high HCC recurrence risk after curative liver resection. Oncotarget 2017, 8, 89978–89987. [Google Scholar] [CrossRef] [Green Version]
  32. Julich-Haertel, H.; Urban, S.K.; Krawczyk, M.; Willms, A.; Jankowski, K.; Patkowski, W.; Kruk, B.; Krasnodębski, M.; Ligocka, J.; Schwab, R.; et al. Cancer-associated circulating large extracellular vesicles in cholangiocarcinoma and hepatocellular carcinoma. J. Hepatol. 2017, 67, 282–292. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, Y.; Zhang, C.; Zhang, P.; Guo, G.; Jiang, T.; Zhao, X.; Jiang, J.; Huang, X.; Tong, H.; Tian, Y. Serum exosomal microRNAs combined with alpha-fetoprotein as diagnostic markers of hepatocellular carcinoma. Cancer Med. 2018, 7, 1670–1679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Tucci, M.; Passarelli, A.; Mannavola, F.; Stucci, L.S.; Ascierto, P.A.; Capone, M.; Madonna, G.; Lopalco, P.; Silvestris, F. Serum exosomes as predictors of clinical response to ipilimumab in metastatic melanoma. OncoImmunology 2017, 7, e1387706. [Google Scholar] [CrossRef]
  35. Chakrabarti, S.; Xie, H.; Urrutia, R.; Mahipal, A. The Promise of Circulating Tumor DNA (ctDNA) in the Management of Early-Stage Colon Cancer: A Critical Review. Cancers 2020, 12, 2808. [Google Scholar] [CrossRef] [PubMed]
  36. von Felden, J.; Craig, A.J.; Garcia-Lezana, T.; Labgaa, I.; Haber, P.K.; D’Avola, D.; Asgharpour, A.; Dieterich, D.; Bonaccorso, A.; Torres-Martin, M.; et al. Mutations in circulating tumor DNA predict primary resistance to systemic therapies in advanced hepatocellular carcinoma. Oncogene 2020, 40, 140–151. [Google Scholar] [CrossRef] [PubMed]
  37. Cai, J.; Chen, L.; Zhang, Z.; Zhang, X.; Lu, X.; Liu, W.; Shi, G.; Ge, Y.; Gao, P.; Yang, Y.; et al. Genome-wide mapping of 5-hydroxymethylcytosines in circulating cell-free DNA as a non-invasive approach for early detection of hepatocellular carcinoma. Gut 2019, 68, 2195–2205. [Google Scholar] [CrossRef] [Green Version]
  38. Xu, R.-H.; Wei, W.; Krawczyk, M.; Wang, W.; Luo, H.; Flagg, K.; Yi, S.; Shi, W.; Quan, Q.; Li, K.; et al. Circulating tumour DNA methylation markers for diagnosis and prognosis of hepatocellular carcinoma. Nat. Mater. 2017, 16, 1155–1161. [Google Scholar] [CrossRef]
  39. Kaseb, A.O.; Sánchez, N.S.; Sen, S.; Kelley, R.K.; Tan, B.R.; Bocobo, A.G.; Lim, K.H.; Abdel-Wahab, R.; Uemura, M.; Pestana, R.C.; et al. Molecular Profiling of Hepatocellular Carcinoma Using Circulating Cell-Free DNA. Clin. Cancer Res. 2019, 25, 6107–6118. [Google Scholar] [CrossRef] [Green Version]
  40. Liao, W.; Yang, H.; Xu, H.; Wang, Y.; Ge, P.; Ren, J.; Xu, W.; Lü, X.; Sang, X.; Zhong, S.; et al. Noninvasive detection of tumor-associated mutations from circulating cell-free DNA in hepatocellular carcinoma patients by targeted deep sequencing. Oncotarget 2016, 7, 40481–40490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Chan, K.C.A.; Lai, P.B.-S.; Mok, T.; Chan, H.L.Y.; Ding, C.; Yeung, S.W.; Lo, Y.M.D. Quantitative Analysis of Circulating Methylated DNA as a Biomarker for Hepatocellular Carcinoma. Clin. Chem. 2008, 54, 1528–1536. [Google Scholar] [CrossRef] [PubMed]
  42. Tran, N.H.; Kisiel, J.; Roberts, L.R. Using cell-free DNA for HCC surveillance and prognosis. JHEP Rep. 2021, 3. [Google Scholar] [CrossRef]
  43. Labgaa, I.; Villacorta-Martin, C.; D’Avola, D.; Craig, A.J.; Von Felden, J.; Filho, S.M.; Sia, D.; Stueck, A.; Ward, S.C.; Fiel, M.I.; et al. A pilot study of ultra-deep targeted sequencing of plasma DNA identifies driver mutations in hepatocellular carcinoma. Oncogene 2018, 37, 3740–3752. [Google Scholar] [CrossRef]
  44. Jiao, J.; Watt, G.P.; Stevenson, H.; Calderone, T.L.; Fisher-Hoch, S.P.; Ye, Y.; Wu, X.; Vierling, J.M.; Beretta, L. Telomerase reverse transcriptase mutations in plasma DNA in patients with hepatocellular carcinoma or cirrhosis: Prevalence and risk factors. Hepatol. Commun. 2018, 2, 718–731. [Google Scholar] [CrossRef] [PubMed]
  45. Oversoe, S.K.; Clement, M.S.; Pedersen, M.H.; Weber, B.; Aagaard, N.K.; Villadsen, G.E.; Grønbæk, H.; Hamilton-Dutoit, S.J.; Sorensen, B.S.; Kelsen, J. TERT promoter mutated circulating tumor DNA as a biomarker for prognosis in hepatocellular carcinoma. Scand. J. Gastroenterol. 2020, 55, 1433–1440. [Google Scholar] [CrossRef]
  46. Hirai, M.; Kinugasa, H.; Nouso, K.; Yamamoto, S.; Terasawa, H.; Onishi, Y.; Oyama, A.; Adachi, T.; Wada, N.; Sakata, M.; et al. Prediction of the prognosis of advanced hepatocellular carcinoma by TERT promoter mutations in circulating tumor DNA. J. Gastroenterol. Hepatol. 2020, 36, 1118–1125. [Google Scholar] [CrossRef]
  47. Shen, T.; Li, S.; Wang, J.; Zhang, T.; Zhang, S.; Chen, H.; Xiao, Q.; Ren, W.; Liu, C.; Peng, B.; et al. TP53 R249S mutation detected in circulating tumour DNA is associated with Prognosis of hepatocellular carcinoma patients with or without hepatectomy. Liver Int. 2020, 40, 2834–2847. [Google Scholar] [CrossRef]
  48. David, W.; Cescon, S.V.B.; Steven, M. Chan and Lillian, L. Siu1 Circulating tumor DNA and liquid biopsy in oncology. Nat. Cncer 2020, 1, 276–290. [Google Scholar]
  49. Plaks, V.; Koopman, C.D.; Werb, Z. Cancer. Circulating tumor cells. Science 2013, 341, 1186–1188. [Google Scholar] [CrossRef]
  50. Guo, W.; Yang, X.R.; Sun, Y.F.; Shen, M.N.; Ma, X.L.; Wu, J.; Zhang, C.Y.; Zhou, Y.; Xu, Y.; Hu, B.; et al. Clinical significance of EpCAM mRNA-positive circulating tumor cells in hepatocellular carcinoma by an optimized negative enrichment and qRT-PCR-based platform. Clin. Cancer Res. 2014, 20, 4794–4805. [Google Scholar] [CrossRef] [Green Version]
  51. Ahn, J.C.; Teng, P.-C.; Chen, P.; Posadas, E.; Tseng, H.; Lu, S.C.; Yang, J.D. Detection of Circulating Tumor Cells and Their Implications as a Biomarker for Diagnosis, Prognostication, and Therapeutic Monitoring in Hepatocellular Carcinoma. Hepatology 2020, 73, 422–436. [Google Scholar] [CrossRef]
  52. Zhou, Y.; Wang, B.; Wu, J.; Zhang, C.; Zhou, Y.; Yang, X.; Zhou, J.; Guo, W.; Fan, J. Association of preoperative EpCAM Circulating Tumor Cells and peripheral Treg cell levels with early recurrence of hepatocellular carcinoma following radical hepatic resection. BMC Cancer 2016, 16, 506. [Google Scholar] [CrossRef] [Green Version]
  53. Wang, Z.; Luo, L.; Cheng, Y.; Ze-Sheng, J.; Peng, B.; Gao, Y.; Jiang, Z.-S.; Pan, M. Correlation Between Postoperative Early Recurrence of Hepatocellular Carcinoma and Mesenchymal Circulating Tumor Cells in Peripheral Blood. J. Gastrointest. Surg. 2017, 22, 633–639. [Google Scholar] [CrossRef] [Green Version]
  54. Qi, L.-N.; Xiang, B.-D.; Wu, F.-X.; Ye, J.-Z.; Zhong, J.-H.; Wang, Y.-Y.; Chen, Y.-Y.; Chen, Z.-S.; Ma, L.; Chen, J.; et al. Circulating Tumor Cells Undergoing EMT Provide a Metric for Diagnosis and Prognosis of Patients with Hepatocellular Carcinoma. Cancer Res. 2018, 78, 4731–4744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Vona, G.; Sabile, A.; Louha, M.; Sitruk, V.; Romana, S.; Schutze, K.; Capron, F.; Franco, D.; Pazzagli, M.; Vekemans, M.; et al. Isolation by size of epithelial tumor cells: A new method for the immunomorphological and molecular characterization of circulatingtumor cells. Am. J. Pathol. 2000, 156, 57–63. [Google Scholar] [CrossRef]
  56. Vona, G.; Estepa, L.; Beroud, C.; Damotte, D.; Capron, F.; Nalpas, B.; Mineur, A.; Franco, M.; Lacour, B.; Pol, S.; et al. Impact of cytomorphological detection of circulating tumor cells in patients with liver cancer. Hepatology 2004, 39, 792–797. [Google Scholar] [CrossRef] [PubMed]
  57. Weitz, J.; Kienle, P.; Lacroix, J.; Willeke, F.; Benner, A.; Lehnert, T.; Herfarth, C.; Doeberitz, M.V.K. Dissemination of tumor cells in patients undergoing surgery for colorectal cancer. Clin. Cancer Res. 1998, 4, 343–348. [Google Scholar]
  58. Liu, H.-Y.; Qian, H.-H.; Zhang, X.-F.; Li, J.; Yang, X.; Sun, B.; Ma, J.-Y.; Chen, L.; Yin, Z.-F. Improved method increases sensitivity for circulating hepatocellular carcinoma cells. World J. Gastroenterol. 2015, 21, 2918–2925. [Google Scholar] [CrossRef] [PubMed]
  59. Li, J.; Shi, L.; Zhang, X.; Sun, B.; Yang, Y.; Ge, N.; Liu, H.; Yang, X.; Chen, L.; Qian, H.; et al. pERK/pAkt phenotyping in circulating tumor cells as a biomarker for sorafenib efficacy in patients with advanced hepatocellular carcinoma. Oncotarget 2015, 7, 2646–2659. [Google Scholar] [CrossRef] [Green Version]
  60. Macdonald, J.; Henri, J.; Roy, K.; Hays, E.; Bauer, M.; Veedu, R.N.; Pouliot, N.; Shigdar, S. EpCAM Immunotherapy versus Specific Targeted Delivery of Drugs. Cancers 2018, 10, 19. [Google Scholar] [CrossRef] [Green Version]
  61. Cristofanilli, M.; Budd, G.T.; Ellis, M.J.; Stopeck, A.; Matera, J.; Miller, M.C.; Reuben, J.M.; Doyle, G.V.; Allard, W.J.; Terstappen, L.W.; et al. Circulating tumor cells, disease progression, and survival in metastatic breast cancer. N. Engl. J. Med. 2004, 351, 781–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Allard, W.J.; Matera, J.; Miller, M.C.; Repollet, M.; Connelly, M.C.; Rao, C.; Tibbe, A.G.J.; Uhr, J.W.; Terstappen, L. Tumor Cells Circulate in the Peripheral Blood of All Major Carcinomas but not in Healthy Subjects or Patients with Nonmalignant Diseases. Clin. Cancer Res. 2004, 10, 6897–6904. [Google Scholar] [CrossRef] [Green Version]
  63. Sun, Y.-F.; Xu, Y.; Yang, X.-R.; Guo, W.; Zhang, X.; Qiu, S.-J.; Shi, R.-Y.; Hu, B.; Zhou, J.; Fan, J. Circulating stem cell-like epithelial cell adhesion molecule-positive tumor cells indicate poor prognosis of hepatocellular carcinoma after curative resection. Hepatology 2012, 57, 1458–1468. [Google Scholar] [CrossRef] [PubMed]
  64. Yu, J.-J.; Xiao, W.; Dong, S.-L.; Liang, H.-F.; Zhang, Z.-W.; Zhang, B.-X.; Huang, Z.-Y.; Chen, Y.-F.; Zhang, W.-G.; Luo, H.-P.; et al. Effect of surgical liver resection on circulating tumor cells in patients with hepatocellular carcinoma. BMC Cancer 2018, 18, 1–9. [Google Scholar] [CrossRef] [PubMed]
  65. Xu, W.; Cao, L.; Chen, L.; Li, J.; Zhang, X.-F.; Qian, H.-H.; Kang, X.-Y.; Zhang, Y.; Liao, J.; Shi, L.-H.; et al. Isolation of Circulating Tumor Cells in Patients with Hepatocellular Carcinoma Using a Novel Cell Separation Strategy. Clin. Cancer Res. 2011, 17, 3783–3793. [Google Scholar] [CrossRef] [Green Version]
  66. Li, Y.-M.; Xu, S.-C.; Li, J.; Han, K.-Q.; Pi, H.-F.; Zheng, L.; Zuo, G.-H.; Huang, X.-B.; Li, H.-Y.; Zhao, H.-Z.; et al. Epithelial–mesenchymal transition markers expressed in circulating tumor cells in hepatocellular carcinoma patients with different stages of disease. Cell Death Dis. 2013, 4, e831. [Google Scholar] [CrossRef]
  67. Li, J.; Chen, L.; Zhang, X.; Zhang, Y.; Liu, H.; Sun, B.; Zhao, L.; Ge, N.; Qian, H.; Yang, Y.; et al. Detection of Circulating Tumor Cells in Hepatocellular Carcinoma Using Antibodies against Asialoglycoprotein Receptor, Carbamoyl Phosphate Synthetase 1 and Pan-Cytokeratin. PLoS ONE 2014, 9, e96185. [Google Scholar] [CrossRef] [Green Version]
  68. Taylor, S.C.; Laperriere, G.; Germain, H. Droplet Digital PCR versus qPCR for gene expression analysis with low abundant targets: From variable nonsense to publication quality data. Sci. Rep. 2017, 7, 1–8. [Google Scholar] [CrossRef] [Green Version]
  69. Mao, X.; Liu, C.; Tong, H.; Chen, Y.; Liu, K. Principles of digital PCR and its applications in current obstetrical and gynecological diseases. Am. J. Transl. Res. 2019, 11, 7209–7222. [Google Scholar]
  70. Huang, A.; Zhang, X.; Zhou, S.-L.; Cao, Y.; Huang, X.-W.; Fan, J.; Yang, X.-R.; Zhou, J. Detecting Circulating Tumor DNA in Hepatocellular Carcinoma Patients Using Droplet Digital PCR Is Feasible and Reflects Intratumoral Heterogeneity. J. Cancer 2016, 7, 1907–1914. [Google Scholar] [CrossRef] [Green Version]
  71. Leary, R.J.; Sausen, M.; Kinde, I.; Papadopoulos, N.; Carpten, J.D.; Craig, D.; O’Shaughnessy, J.; Kinzler, K.W.; Parmigiani, G.; Vogelstein, B.; et al. Detection of Chromosomal Alterations in the Circulation of Cancer Patients with Whole-Genome Sequencing. Sci. Transl. Med. 2012, 4, 162ra154. [Google Scholar] [CrossRef] [Green Version]
  72. Jiang, P.; Chan, C.W.M.; Chan, K.C.A.; Cheng, S.H.; Wong, J.; Wong, V.W.-S.; Wong, G.L.-H.; Chan, S.; Mok, T.; Chan, H.L.Y.; et al. Lengthening and shortening of plasma DNA in hepatocellular carcinoma patients. Proc. Natl. Acad. Sci. USA 2015, 112, E1317–E1325. [Google Scholar] [CrossRef] [Green Version]
  73. McDonald, B.R.; Contente-Cuomo, T.; Sammut, S.-J.; Odenheimer-Bergman, A.; Ernst, B.; Perdigones, N.; Chin, S.-F.; Farooq, M.; Mejia, R.; Cronin, P.A.; et al. Personalized circulating tumor DNA analysis to detect residual disease after neoadjuvant therapy in breast cancer. Sci. Transl. Med. 2019, 11, eaax7392. [Google Scholar] [CrossRef]
  74. Marques, H.P.; da Silva, S.G.; De Martin, E.; Agopian, V.G.; Martins, P.N. Emerging biomarkers in HCC patients: Current status. Int. J. Surg. 2020, 82, 70–76. [Google Scholar] [CrossRef]
  75. Agopian, V.G.; Harlander-Locke, M.P.; Markovic, D.; Zarrinpar, A.; Kaldas, F.M.; Cheng, E.Y.; Yersiz, H.; Farmer, D.G.; Hiatt, J.R.; Busuttil, R.W. Evaluation of Patients with Hepatocellular Carcinomas That Do Not Produce α-Fetoprotein. JAMA Surg. 2017, 152, 55–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Johnson, P.J.; Pirrie, S.J.; Cox, T.; Berhane, S.; Teng, M.; Palmer, D.; Morse, J.; Hull, D.; Patman, G.; Kagebayashi, C.; et al. The Detection of Hepatocellular Carcinoma Using a Prospectively Developed and Validated Model Based on Serological Biomarkers. Cancer Epidemiol. Biomark. Prev. 2013, 23, 144–153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Hippo, Y.; Watanabe, K.; Watanabe, A.; Midorikawa, Y.; Yamamoto, S.; Ihara, S.; Tokita, S.; Iwanari, H.; Ito, Y.; Nakano, K.; et al. Identification of Soluble NH2-Terminal Fragment of Glypican-3 as a Serological Marker for Early-Stage Hepatocellular Carcinoma. Cancer Res. 2004, 64, 2418–2423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Feng, J.; Zhu, R.; Chang, C.; Yu, L.; Cao, F.; Zhu, G.; Chen, F.; Xia, H.; Lv, F.; Zhang, S.; et al. CK19 and Glypican 3 Expression Profiling in the Prognostic Indication for Patients with HCC after Surgical Resection. PLoS ONE 2016, 11, e0151501. [Google Scholar] [CrossRef]
  79. Marrero, J.A.; Romano, P.R.; Nikolaeva, O.; Steel, L.; Mehta, A.; Fimmel, C.J.; Comunale, M.A.; D’Amelio, A.; Lok, A.S.; Block, T.M. GP73, a resident Golgi glycoprotein, is a novel serum marker for hepatocellular carcinoma. J. Hepatol. 2005, 43, 1007–1012. [Google Scholar] [CrossRef]
  80. Wan, H.-G.; Xu, H.; Gu, Y.-M.; Wang, H.; Xu, W.; Zu, M.-H. Comparison osteopontin vs AFP for the diagnosis of HCC: A meta-analysis. Clin. Res. Hepatol. Gastroenterol. 2014, 38, 706–714. [Google Scholar] [CrossRef]
  81. Pozzan, C.; Cardin, R.; Piciocchi, M.; Cazzagon, N.; Maddalo, G.; Vanin, V.; Giacomin, A.; Pontisso, P.; Cillo, U.; Farinati, F. Diagnostic and prognostic role of SCCA-IgM serum levels in hepatocellular carcinoma (HCC). J. Gastroenterol. Hepatol. 2014, 29, 1637–1644. [Google Scholar] [CrossRef]
  82. Zhang, H.; Yao, M.; Wu, W.; Qiu, L.; Sai, W.; Yang, J.; Zheng, W.; Huang, J.; Yao, D. Up-regulation of annexin A2 expression predicates advanced clinicopathological features and poor prognosis in hepatocellular carcinoma. Tumor Biol. 2015, 36, 9373–9383. [Google Scholar] [CrossRef]
  83. de Lima, L.T.F.; Broszczak, D.; Zhang, X.; Bridle, K.; Crawford, D.; Punyadeera, C. The use of minimally invasive biomarkers for the diagnosis and prognosis of hepatocellular carcinoma. Biochim. Biophys. Acta Bioenerg. 2020, 1874, 188451. [Google Scholar] [CrossRef]
  84. Bettegowda, C.; Sausen, M.; Leary, R.J.; Kinde, I.; Wang, Y.; Agrawal, N.; Bartlett, B.; Wang, H.; Luber, B.; Alani, R.M.; et al. Detection of Circulating Tumor DNA in Early- and Late-Stage Human Malignancies. Sci. Transl. Med. 2014, 6, 224ra24. [Google Scholar] [CrossRef] [Green Version]
  85. Tie, J.; Wang, Y.; Tomasetti, C.; Li, L.; Springer, S.; Kinde, I.; Silliman, N.; Tacey, M.; Wong, H.-L.; Christie, M.; et al. Circulating tumor DNA analysis detects minimal residual disease and predicts recurrence in patients with stage II colon cancer. Sci. Transl. Med. 2016, 8, 346ra92. [Google Scholar] [CrossRef] [Green Version]
  86. Von Felden, J.; Garcia-Lezana, T.; Schulze, K.; Losic, B.; Villanueva, A. Liquid biopsy in the clinical management of hepatocellular carcinoma. Gut 2020, 69, 2025–2034. [Google Scholar] [CrossRef] [PubMed]
  87. Vymetalkova, V.; Cervena, K.; Bartu, L.; Vodicka, P. Circulating Cell-Free DNA and Colorectal Cancer: A Systematic Review. Int. J. Mol. Sci. 2018, 19, 3356. [Google Scholar] [CrossRef] [Green Version]
  88. Ono, A.; Fujimoto, A.; Yamamoto, Y.; Akamatsu, S.; Hiraga, N.; Imamura, M.; Kawaoka, T.; Tsuge, M.; Abe, H.; Hayes, C.N.; et al. Circulating Tumor DNA Analysis for Liver Cancers and Its Usefulness as a Liquid Biopsy. Cell. Mol. Gastroenterol. Hepatol. 2015, 1, 516–534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Cai, Z.; Chen, G.; Zeng, Y.; Dong, X.; Li, Z.; Huang, Y.; Xin, F.; Qiu, L.; Xu, H.; Zhang, W.; et al. Comprehensive Liquid Profiling of Circulating Tumor DNA and Protein Biomarkers in Long-Term Follow-Up Patients with Hepatocellular Carcinoma. Clin. Cancer Res. 2019, 25, 5284–5294. [Google Scholar] [CrossRef]
  90. Wong, I.H.; Lo, Y.M.; Yeo, W.; Lau, W.Y.; Johnson, P.J. Frequent p15 promoter methylation in tumor and peripheral blood from hepatocellular carcinoma patients. Clin. Cancer Res. 2000, 6, 3516–3521. [Google Scholar] [PubMed]
  91. Yeo, W.; Wong, N.; Wong, W.-L.; Lai, P.B.-S.; Zhong, S.; Johnson, P.J. High frequency of promoter hypermethylation of RASSF1A in tumor and plasma of patients with hepatocellular carcinoma. Liver Int. 2005, 25, 266–272. [Google Scholar] [CrossRef]
  92. Lu, C.-Y.; Chen, S.-Y.; Peng, H.-L.; Kan, P.-Y.; Chang, W.-C.; Yen, C.-J. Cell-free methylation markers with diagnostic and prognostic potential in hepatocellular carcinoma. Oncotarget 2016, 8, 6406–6418. [Google Scholar] [CrossRef] [Green Version]
  93. Iizuka, N.; Sakaida, I.; Moribe, T.; Fujita, N.; Miura, T.; Stark, M.; Tamatsukuri, S.; Ishitsuka, H.; Uchida, K.; Terai, S.; et al. Elevated levels of circulating cell-free DNA in the blood of patients with hepatitis C virus-associated hepatocellular carcinoma. Anticancer. Res. 2007, 26, 4713–4719. [Google Scholar]
  94. Hindson, B.J.; Ness, K.D.; Masquelier, D.A.; Belgrader, P.; Heredia, N.J.; Makarewicz, A.J.; Bright, I.J.; Lucero, M.Y.; Hiddessen, A.L.; Legler, T.C.; et al. High-throughput droplet digital PCR system for absolute quantitation of DNA copy number. Anal. Chem. 2011, 83, 8604–8610. [Google Scholar] [CrossRef] [PubMed]
  95. Chan, K.C.A.; Jiang, P.; Chan, C.W.M.; Sun, K.; Wong, J.; Hui, E.P.; Chan, S.; Chan, W.C.; Hui, D.; Ng, S.S.M.; et al. Noninvasive detection of cancer-associated genome-wide hypomethylation and copy number aberrations by plasma DNA bisulfite sequencing. Proc. Natl. Acad. Sci. USA 2013, 110, 18761–18768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Köberle, V.; Kronenberger, B.; Pleli, T.; Trojan, J.; Imelmann, E.; Peveling-Oberhag, J.; Welker, M.-W.; Elhendawy, M.; Zeuzem, S.; Piiper, A.; et al. Serum microRNA-1 and microRNA-122 are prognostic markers in patients with hepatocellular carcinoma. Eur. J. Cancer 2013, 49, 3442–3449. [Google Scholar] [CrossRef]
  97. Xu, Y.; Bu, X.; Dai, C.; Shang, C. High serum microRNA-122 level is independently associated with higher overall survival rate in hepatocellular carcinoma patients. Tumor Biol. 2015, 36, 4773–4776. [Google Scholar] [CrossRef] [PubMed]
  98. Cho, H.J.; Kim, S.S.; Nam, J.S.; Kim, J.K.; Lee, J.H.; Kim, B.; Wang, H.J.; Kim, B.W.; Lee, J.D.; Kang, D.Y.; et al. Low levels of circulating microRNA-26a/29a as poor prognostic markers in patients with hepatocellular carcinoma who underwent curative treatment. Clin. Res. Hepatol. Gastroenterol. 2017, 41, 181–189. [Google Scholar] [CrossRef]
  99. Pratedrat, P.; Chuaypen, N.; Nimsamer, P.; Payungporn, S.; Pinjaroen, N.; Sirichindakul, B.; Tangkijvanich, P. Diagnostic and prognostic roles of circulating miRNA-223-3p in hepatitis B virus–related hepatocellular carcinoma. PLoS ONE 2020, 15, e0232211. [Google Scholar] [CrossRef] [Green Version]
  100. Ning, S.; Liu, H.; Gao, B.; Wei, W.; Yang, A.; Li, J.; Zhang, J. miR-155, miR-96 and miR-99a as potential diagnostic and prognostic tools for the clinical management of hepatocellular carcinoma. Oncol. Lett. 2019, 18, 3381–3387. [Google Scholar] [CrossRef]
  101. Loosen, S.H.; Wirtz, T.H.; Roy, S.; Vucur, M.; Castoldi, M.; Schneider, A.T.; Koppe, C.; Ulmer, T.F.; Roeth, A.A.; Bednarsch, J.; et al. Circulating levels of microRNA193a-5p predict outcome in early stage hepatocellular carcinoma. PLoS ONE 2020, 15, e0239386. [Google Scholar] [CrossRef]
  102. Sugimachi, K.; Matsumura, T.; Hirata, H.; Uchi, R.; Ueda, M.; Ueo, H.; Shinden, Y.; Iguchi, T.; Eguchi, H.; Shirabe, K.; et al. Identification of a bona fide microRNA biomarker in serum exosomes that predicts hepatocellular carcinoma recurrence after liver transplantation. Br. J. Cancer 2015, 112, 532–538. [Google Scholar] [CrossRef]
  103. Qu, Z.; Wu, J.; Wu, J.; Ji, A.; Qiang, G.; Jiang, Y.; Jiang, C.; Ding, Y. Exosomal miR-665 as a novel minimally invasive biomarker for hepatocellular carcinoma diagnosis and prognosis. Oncotarget 2017, 8, 80666–80678. [Google Scholar] [CrossRef] [Green Version]
  104. Shi, M.; Jiang, Y.; Yang, L.; Yan, S.; Wang, Y.; Lu, X. Decreased levels of serum exosomal miR-638 predict poor prognosis in hepatocellular carcinoma. J. Cell. Biochem. 2017, 119, 4711–4716. [Google Scholar] [CrossRef] [PubMed]
  105. Hao, X.; Xin, R.; Dong, W. Decreased serum exosomal miR-320a expression is an unfavorable prognostic factor in patients with hepatocellular carcinoma. J. Int. Med. Res. 2020, 48. [Google Scholar] [CrossRef]
  106. Ye, Q.; Ling, S.; Zheng, S.; Xu, X. Liquid biopsy in hepatocellular carcinoma: Circulating tumor cells and circulating tumor DNA. Mol. Cancer 2019, 18, 1–13. [Google Scholar] [CrossRef] [PubMed]
  107. Lamps, L.W.; Folpe, A.L. The Diagnostic Value of Hepatocyte Paraffin Antibody 1 in Differentiating Hepatocellular Neoplasms from Nonhepatic Tumors: A Review. Adv. Anat. Pathol. 2003, 10, 39–43. [Google Scholar] [CrossRef]
  108. Court, C.M.; Hou, S.; Winograd, P.; Segel, N.H.; Li, Q.W.; Zhu, Y.; Sadeghi, S.; Finn, R.S.; Ganapathy, E.; Song, M.; et al. A novel multimarker assay for the phenotypic profiling of circulating tumor cells in hepatocellular carcinoma. Liver Transplant. 2018, 24, 946–960. [Google Scholar] [CrossRef] [PubMed]
  109. Schipilliti, F.M.; Garajová, I.; Rovesti, G.; Balsano, R.; Piacentini, F.; Dominici, M.; Gelsomino, F. The Growing Skyline of Advanced Hepatocellular Carcinoma Treatment: A Review. Pharmaceuticals 2021, 14, 43. [Google Scholar] [CrossRef]
  110. Merle, P. The New Immuno-Oncology-Based Therapies and Their Perspectives in Hepatocellular Carcinoma. Cancers 2021, 13, 238. [Google Scholar] [CrossRef] [PubMed]
  111. Zhu, A.X.; Finn, R.S.; Edeline, J.; Cattan, S.; Ogasawara, S.; Palmer, D.; Verslype, C.; Zagonel, V.; Fartoux, L.; Vogel, A.; et al. Pembrolizumab in patients with advanced hepatocellular carcinoma previously treated with sorafenib (KEYNOTE-224): A non-randomised, open-label phase 2 trial. Lancet Oncol. 2018, 19, 940–952. [Google Scholar] [CrossRef]
  112. Yau, T.; Hsu, C.; Kim, T.-Y.; Choo, S.-P.; Kang, Y.-K.; Hou, M.-M.; Numata, K.; Yeo, W.; Chopra, A.; Ikeda, M.; et al. Nivolumab in advanced hepatocellular carcinoma: Sorafenib-experienced Asian cohort analysis. J. Hepatol. 2019, 71, 543–552. [Google Scholar] [CrossRef] [Green Version]
  113. Finn, R.S.; Ryoo, B.-Y.; Merle, P.; Kudo, M.; Bouattour, M.; Lim, H.Y.; Breder, V.; Edeline, J.; Chao, Y.; Ogasawara, S.; et al. Pembrolizumab as Second-Line Therapy in Patients With Advanced Hepatocellular Carcinoma in KEYNOTE-240: A Randomized, Double-Blind, Phase III Trial. J. Clin. Oncol. 2020, 38, 193–202. [Google Scholar] [CrossRef] [PubMed]
  114. FDA. FDA Approves Atezolizumab Plus Bevacizumab for Unresectable Hepatocellular Carcinoma. Available online: https://www.fda.gov/drugs/drug-approvals-and-databases/fda-approves-atezolizumab-plus-bevacizumab-unresectable-hepatocellular-carcinoma (accessed on 16 May 2021).
  115. ESMO. Available online: https://www.esmo.org/oncology-news/ema-recommends-extension-of-indications-for-atezolizumab3 (accessed on 19 May 2021).
  116. Finn, R.S.; Qin, S.; Ikeda, M.; Galle, P.R.; Ducreux, M.; Kim, T.-Y.; Lim, H.Y.; Kudo, M.; Breder, V.V.; Merle, P.; et al. IMbrave150: Updated overall survival (OS) data from a global, randomized, open-label phase III study of atezolizumab (atezo) + bevacizumab (bev) versus sorafenib (sor) in patients (pts) with unresectable hepatocellular carcinoma (HCC). J. Clin. Oncol. 2021, 39, 267. [Google Scholar] [CrossRef]
  117. Kwee, S.A.; Tiirikainen, M. Beta-catenin activation and immunotherapy resistance in hepatocellular carcinoma: Mechanisms and biomarkers. Hepatoma Res. 2021, 2021. [Google Scholar] [CrossRef] [PubMed]
  118. Zhu, A.X.; Kang, Y.K.; Yen, C.J.; Finn, R.S.; Galle, P.R.; Llovet, J.M.; Assenat, E.; Brandi, G.; Pracht, M.; Lim, H.Y.; et al. Ramucirumab after sorafenib in patients with advanced hepatocellular carcinoma and increased α-fetoprotein concentrations (REACH-2): A randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol. 2019, 20, 282–296. [Google Scholar] [CrossRef]
  119. von Felden, J.; Villanueva, A. Role of Molecular Biomarkers in Liver Transplantation for Hepatocellular Carcinoma. Liver Transpl. 2020, 26, 823–831. [Google Scholar] [CrossRef]
  120. Hofman, P.; Heeke, S.; Alix-Panabières, C.; Pantel, K. Liquid biopsy in the era of immuno-oncology: Is it ready for prime-time use for cancer patients? Ann. Oncol. 2019, 30, 1448–1459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. De Mattos-Arruda, L.; Siravegna, G. How to use liquid biopsies to treat patients with cancer. ESMO Open 2021, 6, 100060. [Google Scholar] [CrossRef]
  122. Khagi, Y.; Kurzrock, R.; Patel, S.P. Next generation predictive biomarkers for immune checkpoint inhibition. Cancer Metastas. Rev. 2016, 36, 179–190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Patel, S.P.; Kurzrock, R. PD-L1 Expression as a Predictive Biomarker in Cancer Immunotherapy. Mol. Cancer Ther. 2015, 14, 847–856. [Google Scholar] [CrossRef] [Green Version]
  124. McGranahan, N.; Swanton, C. Biological and Therapeutic Impact of Intratumor Heterogeneity in Cancer Evolution. Cancer Cell 2015, 27, 15–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Armstrong, P.P.S.; Aiwu, R.H. Immunotherapy and immunotherapy biomarkers for hepatocellular carcinoma. Hepatoma Res. 2021, 7, 18. [Google Scholar]
  126. Cabel, L.; Riva, F.; Servois, V.; Livartowski, A.; Daniel, C.; Rampanou, A.; Lantz, O.; Romano, E.; Milder, M.; Buecher, B.; et al. Circulating tumor DNA changes for early monitoring of anti-PD1 immunotherapy: A proof-of-concept study. Ann. Oncol. 2017, 28, 1996–2001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Chan, T.; Yarchoan, M.; Jaffee, E.; Swanton, C.; Quezada, S.; Stenzinger, A.; Peters, S. Development of tumor mutation burden as an immunotherapy biomarker: Utility for the oncology clinic. Ann. Oncol. 2018, 30, 44–56. [Google Scholar] [CrossRef]
  128. Hsu, C.-H.; Lu, S.; Abbas, A.; Guan, Y.; Zhu, A.X.; Aleshin, A.; Lee, K.-H.; Lee, M.S.; Mahipal, A.; Ryoo, B.-Y.; et al. Longitudinal and personalized detection of circulating tumor DNA (ctDNA) for monitoring efficacy of atezolizumab plus bevacizumab in patients with unresectable hepatocellular carcinoma (HCC). J. Clin. Oncol. 2020, 38, 3531. [Google Scholar] [CrossRef]
  129. Luke, J.J.; Bao, R.; Sweis, R.F.; Spranger, S.; Gajewski, T.F. WNT/β-catenin Pathway Activation Correlates with Immune Exclusion across Human Cancers. Clin. Cancer Res. 2019, 25, 3074–3083. [Google Scholar] [CrossRef] [Green Version]
  130. Sia, D.; Jiao, Y.; Martinez-Quetglas, I.; Kuchuk, O.; Villacorta-Martin, C.; de Moura, M.C.; Putra, J.; Campreciós, G.; Bassaganyas, L.; Akers, N.; et al. Identification of an Immune-specific Class of Hepatocellular Carcinoma, Based on Molecular Features. Gastroenterology 2017, 153, 812–826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Harding, J.J.; Nandakumar, S.; Armenia, J.; Khalil, D.N.; Albano, M.; Ly, M.; Shia, J.; Hechtman, J.; Kundra, R.; El Dika, I.; et al. Prospective Genotyping of Hepatocellular Carcinoma: Clinical Implications of Next-Generation Sequencing for Matching Patients to Targeted and Immune Therapies. Clin. Cancer Res. 2018, 25, 2116–2126. [Google Scholar] [CrossRef] [Green Version]
  132. De Galarreta, M.R.; Bresnahan, E.; Molina-Sánchez, P.; Lindblad, K.E.; Maier, B.; Sia, D.; Puigvehi, M.; Miguela, V.; Casanova-Acebes, M.; Dhainaut, M.; et al. β-Catenin Activation Promotes Immune Escape and Resistance to Anti–PD-1 Therapy in Hepatocellular Carcinoma. Cancer Discov. 2019, 9, 1124–1141. [Google Scholar] [CrossRef]
  133. Gao, C.; Wang, Y.; Broaddus, R.; Sun, L.; Xue, F.; Zhang, W. Exon 3 mutations of CTNNB1 drive tumorigenesis: A review. Oncotarget 2017, 9, 5492–5508. [Google Scholar] [CrossRef] [Green Version]
  134. Oversoe, S.K.; Clement, M.S.; Weber, B.; Grønbæk, H.; Hamilton-Dutoit, S.J.; Sorensen, B.S.; Kelsen, J. Combining tissue and circulating tumor DNA increases the detection rate of a CTNNB1 mutation in hepatocellular carcinoma. BMC Cancer 2021, 21, 1–7. [Google Scholar] [CrossRef] [PubMed]
  135. Yue, C.; Jiang, Y.; Li, P.; Wang, Y.; Xue, J.; Li, N.; Li, D.; Wang, R.; Dang, Y.; Hu, Z.; et al. Dynamic change of PD-L1 expression on circulating tumor cells in advanced solid tumor patients undergoing PD-1 blockade therapy. OncoImmunology 2018, 7, e1438111. [Google Scholar] [CrossRef] [PubMed]
  136. Winograd, P.; Hou, S.; Court, C.M.; Lee, Y.; Chen, P.; Zhu, Y.; Sadeghi, S.; Finn, R.S.; Teng, P.; Wang, J.J.; et al. Hepatocellular Carcinoma–Circulating Tumor Cells Expressing PD-L1 Are Prognostic and Potentially Associated with Response to Checkpoint Inhibitors. Hepatol. Commun. 2020, 4, 1527–1540. [Google Scholar] [CrossRef] [PubMed]
  137. De Mattos-Arruda, L.; Vazquez, M.; Finotello, F.; Lepore, R.; Porta, E.; Hundal, J.; Amengual-Rigo, P.; Ng, C.; Valencia, A.; Carrillo, J.; et al. Neoantigen prediction and computational perspectives towards clinical benefit: Recommendations from the ESMO Precision Medicine Working Group. Ann. Oncol. 2020, 31, 978–990. [Google Scholar] [CrossRef]
  138. Chen, V.L.; Xu, D.; Wicha, M.S.; Lok, A.S.; Parikh, N.D. Utility of Liquid Biopsy Analysis in Detection of Hepatocellular Carcinoma, Determination of Prognosis, and Disease Monitoring: A Systematic Review. Clin. Gastroenterol. Hepatol. 2020, 18, 2879–2902.e9. [Google Scholar] [CrossRef]
  139. Dhar, M.; Wong, J.; Che, J.; Matsumoto, M.; Grogan, T.; Elashoff, D.; Garon, E.B.; Goldman, J.W.; Christen, E.S.; Di Carlo, D.; et al. Evaluation of PD-L1 expression on vortex-isolated circulating tumor cells in metastatic lung cancer. Sci. Rep. 2018, 8, 1–10. [Google Scholar] [CrossRef]
  140. Krieg, C.; Nowicka, M.; Guglietta, S.; Schindler, S.; Hartmann, F.J.; Weber, L.M.; Dummer, R.; Robinson, M.D.; Levesque, M.P.; Becher, B. High-dimensional single-cell analysis predicts response to anti-PD-1 immunotherapy. Nat. Med. 2018, 24, 144–153. [Google Scholar] [CrossRef]
  141. Masucci, G.V.; Cesano, A.; Hawtin, R.; Janetzki, S.; Zhang, J.; Kirsch, I.; Dobbin, K.K.; Alvarez, J.; Robbins, P.B.; Selvan, S.R.; et al. Validation of biomarkers to predict response to immunotherapy in cancer: Volume I—Pre-analytical and analytical validation. J. Immunother. Cancer 2016, 4, 76. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Liquid biopsy in hepatocellular carcinoma (HCC). Illustration of liquid biopsy biomarkers investigated in HCC, including circulating nucleic acids, circulating tumor DNA (ctDNA), circulating RNA (cRNA)/microRNAs (miRNA), extracellular vesicles (EVs)/exosomes and circulating tumor cells (CTC). Created with BioRender.com accessed on 11 August 2021.
Figure 1. Liquid biopsy in hepatocellular carcinoma (HCC). Illustration of liquid biopsy biomarkers investigated in HCC, including circulating nucleic acids, circulating tumor DNA (ctDNA), circulating RNA (cRNA)/microRNAs (miRNA), extracellular vesicles (EVs)/exosomes and circulating tumor cells (CTC). Created with BioRender.com accessed on 11 August 2021.
Cancers 13 04334 g001
Figure 2. Main technology platforms utilized for the detection of CTCs and ctDNA in HCC, together with advantages and limitations for each approach. (A) The detection of CTCs has been based either on biophysics-based properties, including size for filtration and density by centrifugation (i), or antibody-based approaches targeting known CTC expression markers, such as the epithelial cell adhesion molecule (EpCAM), cytokeratin markers (CK) and asialoglycoprotein receptor (ASGPR). Antibodies enriched with magnetic beads can be subsequently magnetically separated to isolate CTCs (ii). (B) For the detection of ctDNA, various quantitative PCR-based approaches have been utilized, such as droplet digital PCR (ddPCR), to improve the precision of mutational analysis of ctDNA (i). The next-generation sequencing (NGS)-based approach represents another alternative platform for ctDNA detection. A unique advantage of NGS is that it can detect tumor DNA mutations in plasma without prior knowledge of pre-defined hotspot mutation sites. Apart from DNA mutations, chromosomal alterations can also be detected by this approach (ii). Created with BioRender.com accessed on 11 August 2021.
Figure 2. Main technology platforms utilized for the detection of CTCs and ctDNA in HCC, together with advantages and limitations for each approach. (A) The detection of CTCs has been based either on biophysics-based properties, including size for filtration and density by centrifugation (i), or antibody-based approaches targeting known CTC expression markers, such as the epithelial cell adhesion molecule (EpCAM), cytokeratin markers (CK) and asialoglycoprotein receptor (ASGPR). Antibodies enriched with magnetic beads can be subsequently magnetically separated to isolate CTCs (ii). (B) For the detection of ctDNA, various quantitative PCR-based approaches have been utilized, such as droplet digital PCR (ddPCR), to improve the precision of mutational analysis of ctDNA (i). The next-generation sequencing (NGS)-based approach represents another alternative platform for ctDNA detection. A unique advantage of NGS is that it can detect tumor DNA mutations in plasma without prior knowledge of pre-defined hotspot mutation sites. Apart from DNA mutations, chromosomal alterations can also be detected by this approach (ii). Created with BioRender.com accessed on 11 August 2021.
Cancers 13 04334 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Maravelia, P.; Silva, D.N.; Rovesti, G.; Chrobok, M.; Stål, P.; Lu, Y.-C.; Pasetto, A. Liquid Biopsy in Hepatocellular Carcinoma: Opportunities and Challenges for Immunotherapy. Cancers 2021, 13, 4334. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers13174334

AMA Style

Maravelia P, Silva DN, Rovesti G, Chrobok M, Stål P, Lu Y-C, Pasetto A. Liquid Biopsy in Hepatocellular Carcinoma: Opportunities and Challenges for Immunotherapy. Cancers. 2021; 13(17):4334. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers13174334

Chicago/Turabian Style

Maravelia, Panagiota, Daniela Nascimento Silva, Giulia Rovesti, Michael Chrobok, Per Stål, Yong-Chen Lu, and Anna Pasetto. 2021. "Liquid Biopsy in Hepatocellular Carcinoma: Opportunities and Challenges for Immunotherapy" Cancers 13, no. 17: 4334. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers13174334

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop