Skip to main content
Top

24-02-2017 | Liquid biopsy | Review | Article

Liquid biopsies come of age: Towards implementation of circulating tumor DNA

Authors: Jonathan C. M. Wan, Charles Massie, Javier Garcia-Corbacho, Florent Mouliere, James D. Brenton, Carlos Caldas, Simon Pacey, Richard Baird, Nitzan Rosenfeld

Abstract

Improvements in genomic and molecular methods are expanding the range of potential applications for circulating tumour DNA (ctDNA), both in a research setting and as a 'liquid biopsy' for cancer management. Proof-of-principle studies have demonstrated the translational potential of ctDNA for prognostication, molecular profiling and monitoring. The field is now in an exciting transitional period in which ctDNA analysis is beginning to be applied clinically, although there is still much to learn about the biology of cell-free DNA. This is an opportune time to appraise potential approaches to ctDNA analysis, and to consider their applications in personalized oncology and in cancer research.

Nat Rev Cancer 2017; 17: 223–238. doi: 10.1038/nrc.2017.7

© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.

Key points

  • Cell-free DNA (cfDNA) is released predominantly by cell death into the bloodstream, although active secretion may have a role. Since the discovery of fetal cfDNA in the maternal circulation, cfDNA analysis has been rapidly implemented in clinical practice for noninvasive prenatal testing.
  • Mutations were first detected in cfDNA more than two decades ago, and interest in circulating tumour DNA (ctDNA) as a noninvasive cancer diagnostic has increased considerably with the development of molecular methods that permit the sensitive detection and monitoring of multiple classes of mutation.
  • ctDNA may have utility at almost every stage of the management of patients with cancer, including diagnosis, minimally invasive molecular profiling, treatment monitoring, the detection of residual disease and the identification of resistance mutations. ctDNA analysis may be broadly considered as a tool both for quantitative analysis of disease burden and for genomic analysis.
  • The identification of ctDNA in individuals before a cancer diagnosis, and in presymptomatic individuals, suggests the possibility of ctDNA analysis as a tool for earlier detection and screening. Noninvasive cancer classification or subtyping has also emerged as a possibility, although for early detection, both technical and biological factors introduce challenges to the detection of mutant DNA in plasma and its interpretation.
  • Monitoring multiple mutations in parallel can enhance the sensitivity of ctDNA detection, can be used to assess the clonal evolution of patients' disease and may identify resistance mutations before clinical progression is observed.
  • ctDNA analysis is beginning to transition from the research setting into the clinic. The US Food and Drug Administration and the European Medicines Agency have approved ctDNA tests for specific indications in the absence of evaluable tumour tissue. Analysis of gene panels in plasma has now become available as a potential clinical tool. Large studies are under way to establish the overall performance and clinical utility of such assays when a tumour biopsy is not available for analysis.
  • Potential applications of ctDNA have been demonstrated by a number of proof-of-principle studies. Prospective clinical trials are beginning to assess the clinical utility of ctDNA analysis for molecular profiling and disease monitoring. The increasing acceptance of ctDNA is enabling the field to move from exploratory ctDNA studies towards clinical trials in which ctDNA guides decision-making.
  • In order to fully exploit the potential utility of liquid biopsies, it is essential that the biology of cfDNA and ctDNA is explored further. The mechanisms of release and degradation, and the factors that affect the representation of ctDNA in plasma, are poorly understood. The nature of ctDNA will be clarified through both large, well-annotated clinical studies and in vivo studies in which variables can be controlled.

Main

The presence of fragments of cell-free nucleic acids in human blood was first described in 1948 by Mandel and Métais1. The origins and characteristics of cell-free DNA (cfDNA) were studied intermittently in subsequent decades2. In healthy individuals, cfDNA concentrations tend to range between 1 and 10 ng ml−1 in plasma3,4. Raised cfDNA levels were first reported in the serum of patients with cancer in 1977 (Ref. 5); cfDNA concentrations can also be raised in other physiological conditions or clinical scenarios, such as acute trauma6, cerebral infarction7, exercise8, transplantation9 and infection10. The identification of fetal DNA sequences in maternal plasma by Lo et al.11 in 1997 has led to multiple applications of cfDNA in prenatal medicine, including sex determination12, the identification of monogenic disorders13 and noninvasive prenatal testing (NIPT) for aneuploidies such as Down syndrome (trisomy 21). NIPT was first demonstrated in 2007 by Lo et al.14 and has moved rapidly into widespread clinical use15,16.

In 1989, Stroun and colleagues2,17 reported that at least some cfDNA in the plasma of patients with cancer originates from cancer cells. In 1991, Sidransky et al.18 showed that DNA from urinary sediments (cell pellets) from patients with invasive bladder cancer carried mutations in TP53(which encodes p53), and this set the stage for the use of genomic analysis methods in liquid biopsy applications. KRAS mutations that matched the mutations in colorectal19, pancreatic20 or lung21,22 cancers were soon found in stool or sputum samples. In 1994, mutated KRASsequences were first reported to be detected in the plasma cfDNA of patients with pancreatic cancer by PCR with allele-specific primers23. For each patient, the KRAS mutation found in the plasma was identical to that found in the patient's tumour, thereby confirming that the mutant DNA fragments in the plasma were of tumour origin. Mutations in cfDNA are highly specific markers for cancer, and this gave rise to the term 'circulating tumour DNA' (ctDNA).

In the following decades, ctDNA was explored as a prognostic or predictive marker24,25, and for cancer detection26. Such studies confirmed the potential of ctDNA, although the levels of ctDNA in different clinical contexts were not yet accurately defined. These studies nonetheless could demonstrate potential clinical applications, for example, the detection of KRAS mutations in plasma as a potential prognostic factor in colorectal cancer27. The introduction of a digital PCR (dPCR) method in 1999 by Vogelstein and Kinzler28 enabled the accurate identification and absolute quantification of rare mutant fragments. A modification of this technique using beads in emulsions29 and flow cytometry enabled the quantification of the mutant allele fraction of cancer mutations in the plasma of patients with various stages of colorectal cancer30. Diehl, Diaz et al.31 then showed in 2008 that ctDNA is a highly specific marker of tumour dynamics and may be able to indicate residual disease. In parallel, allele-specific PCR and other methods were devised and tested for their ability to identify epidermal growth factor receptor (EGFR) mutations in serum or plasma of patients with lung cancer, following the elucidation of the role of such mutations in predicting response to treatment with molecularly targeted inhibitors25,32,33.

The development of next-generation sequencing-based technologies has facilitated the interrogation of the genome at a broader scale than previously possible. In 2012, deep sequencing of multiple genes in cfDNA was demonstrated using panels of tagged amplicons, which allowed the identification of mutations directly in the plasma of patients with cancer and enabled the monitoring of multiple tumour-specific mutations in a single assay34. This method was subsequently applied to monitor ctDNA in a cohort of patients with metastatic breast cancer35. Shortly thereafter, whole- genome sequencing (WGS) of plasma cfDNA was shown to identify tumour-derived chromosomal aberrations36, focal amplifications37 and gene rearrangements38, and hybrid-capture sequencing was introduced as a noninvasive method of analysing the evolving genomic profile of mutations across the entire exome in cancer39.

There is a clear clinical need for novel diagnostic and molecular tools in oncology (Box 1). Conventional sampling methods such as needle biopsies are subject to procedural complications in up to one in six biopsies40, difficulty in obtaining sufficient material of adequate quality for genomic profiling (reported failure rates range from <10% to >30% of cases)41,42 and sampling biases that arise from genetic heterogeneity43,44,45,46,47. The detection and monitoring of disease often relies on body fluid-based markers that often lack specificity48, and on imaging, which exposes patients to ionizing radiation49 and has limited resolution (in terms of both time and space). Recent advances in ctDNA research highlight the potential applications of liquid biopsies at each stage of patient management (Fig. 1a). These potential applications arise primarily from two types of information that are obtainable through ctDNA analysis: quantification of disease burden, and genomic analysis of cancer (Fig. 1b). These may be combined and/or leveraged through serial sampling in order to monitor disease burden and clonal evolution.

Box 1: The clinical need for liquid biopsies in oncology

There are multiple areas of oncology in which novel diagnostics may have utility and produce clinical benefit, including:

  • Cancer diagnosis: earlier diagnosis of cancer would enable treatment to be initiated sooner, and curative surgery may be carried out if the tumour is diagnosed at an early stage. For symptomatic patients, sensitive and specific cancer detection may speed up the time to diagnosis and treatment. At a population level, screening for markers of disease could enable early intervention in presymptomatic individuals.
  • Prognosis, residual disease and risk of relapse: the assessment of risk of progression is essential to select the extent or aggressiveness of treatment. After treatment with curative intent, the identification of patients with residual disease who are at high risk of relapse may be used to stratify patients to adjuvant therapy. Effective stratification would also spare low-risk patients from overtreatment.
  • Treatment selection: the introduction of a wide array of novel molecularly targeted and immunotherapy agents necessitates improved tools for molecular profiling of patients and for treatment stratification. At present, tumour biopsies are the standard for obtaining tumour DNA; however, these cannot always be obtained and their interpretation may be confounded by intratumour heterogeneity43,44,45,46,47, which could lead to false-negative results and suboptimal therapy selection.
  • Monitoring disease burden: treatment monitoring, which is currently performed using imaging or molecular methods, may identify response or progression, enabling clinicians and patients to adapt therapy accordingly. Current methods have limited accuracy, or are associated with a logistical burden or radiation exposure49. The ideal monitoring assay should be repeatable serially over time, with minimal risk to patients, and should provide an accurate read-out of tumour burden.

Figure 1: Applications of circulating tumour DNA analysis during the course of disease management.

The increasing availability and reliability of techniques for PCR and high-throughput sequencing are facilitating novel high-sensitivity applications for ctDNA, the generation of large clinical data sets, and a better understanding of the origins of both cfDNA and ctDNA. This Review highlights and explores recent advances in the field, and discusses the implications for oncology.

cfDNA and ctDNA biology

Characteristics of cfDNA and ctDNA

cfDNA is thought to be released from cells mostly through apoptosis and necrosis, and possibly also active secretion2,50,51,52,53. Outside the blood circulation, cfDNA has been detected in various body fluids, including urine54,55,56,57,58, cerebrospinal fluid59,60,61,62, pleural fluid63 and saliva64. Genetic and epigenetic modifications of cfDNA molecules reflect the genome or epigenome of the cell of origin65,66,67 (Fig. 2). Methylation analysis has revealed that the majority of cfDNA in plasma is released from haematopoietic cells in healthy individuals66,67,68. These cells have been suggested to be the source of cfDNA release following intense exercise69. Observational studies have determined the half-life of cfDNA in the circulation as between 16 minutes and 2.5 hours31,70,71,72, which enables ctDNA analysis to be considered as a 'real-time' snapshot of disease burden. Other observational studies indicate that cfDNA is cleared from the circulation via nuclease action71,73 and renal excretion into the urine54,58,74. cfDNA uptake by the liver and spleen, followed by degradation by macrophages, may also contribute30,75. The stability of individual fragments in the circulation may be increased through association with cell membranes, extracellular vesicles or proteins2.

Figure 2: Origins and range of alterations in cell-free DNA.

Nearly two decades ago, the modal size of cfDNA was determined using gel electrophoresis as ∼180 bp, indicating that cfDNA was likely to be associated with nucleosomes76. Sequencing-based approaches have since refined this measurement by identifying a prominent peak at 166 bp77,78, which corresponds to the length of DNA wrapped around a nucleosome (∼147 bp) plus linker DNA associated with histone H1. Fragment size traces of cfDNA show a 10 bp ladder pattern53,74, which is ostensibly caused by nucleases cleaving the DNA strand at periodically exposed sites with each turn of the DNA double helix. The fragmentation patterns of cfDNA differ between plasma and urine74, and this is potentially related to relatively higher nuclease activity in urine79.

ctDNA molecules are shorter than non-mutant cfDNA molecules in plasma, as demonstrated by PCR4,80 and sequencing67,81. Animal xenograft experiments78,80,81,82 provide an elegant means by which to interrogate ctDNA, as any human DNA sequences must have originated from the tumour xenograft. The modal length of ctDNA fragments has been measured in a rat xenograft model as between 134 and 144 bp81, although the cause of this shortening is not clear. Shortening of fragments is also observed in fetal cfDNA relative to maternal cfDNA77 and in non-haematopoetically derived relative to haematopoietically derived cfDNA fragments in patients who have received transplants83,84. Differences in nucleosome wrapping or nuclease action between haematopoietic cells, which contribute most to the cfDNA pool, and other tissues may have a role. Long cfDNA fragments (>1,000 bp) have been identified in healthy individuals using long-read sequencing techniques85 and may be released into the circulation in association with exosomes50,51 or by tumour cells via necrosis52. Current extraction methods often poorly recover these long fragments86,87 (Box 2). Commonly used library preparation methods introduce further biases: single-stranded DNA library preparation88 can recover DNA fragments with damaged ends, and when applied to cfDNA65,89 this approach identified a large proportion of fragments shorter than 100 bases. Diverse extraction and sequencing methods may therefore yield complementary data. Combining those data with histological analysis of corresponding tissue samples could provide new insights into the biological determinants of cfDNA fragmentation and the biological origins of cfDNA.

Box 2: Pre-analytical considerations

In low-burden disease, or certain cancer types, the concentration of circulating tumour DNA (ctDNA) molecules may be low, and any loss of sampled material could reduce the sensitivity of molecular profiling. For quantitative applications, the reproducibility of measurement is essential to achieve a robust result, and so the following pre-analytical factors should be considered:

  • Samples should be collected in tubes containing an anticoagulant that is compatible with PCR; ethylenediaminetetraacetic acid (EDTA) is preferred. PCR is inhibited by plasma from heparinized blood232, although some studies have been able to utilize such samples32.
  • It is important that the first centrifugation of the blood is done within a few hours of the blood draw in order to remove blood cells that may lyse and release germline DNA, which would dilute ctDNA68,233,234,235,236. Tubes containing fixative agents may stabilize cells and prevent lysis for several days at room temperature234,236,237,238,239, including during shipping239,240.
  • Following centrifugation, buffy coat DNA from the same tubes can be used as a source of germline DNA, although this may contain small or trace amounts of ctDNA.
  • From a blood draw, plasma is preferred over serum for ctDNA analysis235. Serum also contains ctDNA241, but blood cell lysis during the preparation of serum samples could release DNA from non-cancerous cells, which would dilute any ctDNA signal. Other body fluids or cytological specimens may be used, and may contain a higher amount or concentration of tumour DNA than that in plasma, depending on tumour proximity (Fig. 3c).
  • Cell-free DNA (cfDNA) extraction may be carried out using affinity column-based, magnetic bead-based, polymer-based and phenol–chloroform-based methods, or by filtration. Different methods show variation in their ability to recover particular fragment sizes86,87, and this could have implications for ctDNA detection, given the differences in size between cfDNA and ctDNA.

Physiological and pathological considerations

cfDNA has been proposed as a ligand for Toll-like receptor 9 (TLR9)90,91, which is a sensor of exogenous DNA fragments that is found primarily in tissues rich in immune cells. In mice, obesity-related adipocyte degeneration was shown to release cfDNA, which contributed to macrophage accumulation via TLR9 activation and led to adipose tissue inflammation and insulin resistance91. Another study has suggested that cfDNA may inhibit pro-apoptotic caspases via TLR9-dependent signalling90, which could imply a potential immunomodulatory role for cfDNA.

In vitro experiments suggest that cfDNA may be internalized by cells2,92,93, which raises the possibility that cfDNA molecules could mediate the horizontal transfer of genes or DNA. One report showed that NIH-3T3 mouse cells that were in contact with samples of plasma from patients with KRAS-mutant colorectal cancers underwent in vitrotransformation, despite being separated to avoid tumour cell contamination94. Another study demonstrated the integration of ctDNA into the nuclear DNA of recipient cells and suggested that this may occur through non-homologous end-joining92. A similar phenomenon has been observed for mitochondrial DNA95. Overall, it is clear that there is a lot to learn about the biology of cfDNA and ctDNA, and improved knowledge could have an important impact on their potential applications in oncology.

Approaches to ctDNA analysis

The analysis of ctDNA ranges in scale from single mutations to whole-genome analyses (Fig. 3; see Table 1 for a comparison of selected techniques). Appropriately designed assays for individual mutations can achieve high sensitivity using a simple workflow. Allele-specific PCR methods96,97 have been applied since the mid-2000s for detection of hot-spot mutations in serum and plasma25, and some assays are available as kits that are approved for clinical use98,99, but have limited analytical sensitivity. dPCR assays on microfluidic platforms are quantitative and highly sensitive, and are used extensively to quantify ctDNA levels29,30,56,100,101,102,103. Improved detection at selected loci has been demonstrated by methods such as single-base extension104 or enrichment for mutant alleles by electrophoretic methods105,106, nuclease activity107 or modified PCR58,108,109. The multiplexing capacity of such assays, which rely on differential binding affinities of mutant and wild-type alleles, and for the most part require primers or probes that are specific to a defined mutation or targeted locus, is limited. These are, therefore, generally suited to investigating a small number of mutations and are often applied to analysis of cancer hot-spot mutations. If samples need to be split into multiple reactions, this increases sampling error and may impair the overall performance of an assay for very low copy numbers of mutant DNA.

Figure 3: Current and future paradigms for the sensitive detection of circulating tumour DNA.

In order to interrogate a larger number of loci, targeted sequencingusing PCR amplicons or hybrid capture have been used34,39,110,111. Regions for sequencing may range from individual exons of interest (kilobases) to the entire exome (∼50 Mb). Current off-the-shelf panels for gene sequencing can detect mutations with an allele fraction greater than 1%45,112. By reducing the background error rates of sequencing34,113— for example, by molecular barcoding (Fig. 3b) or running multiple replicates (Fig. 4) — ctDNA can be detected at allele fractions below 0.1%114,115,116,117 (Table 1). Amplicon-based assays that have been optimized for the purpose of ctDNA analysis can target dozens to hundreds of amplicons across multiple kilobases with high sensitivity34,115,117. Hybrid capture-based approaches can increase the genomic region studied to dozens or hundreds of kilobases110,111,114,118. The sensitivity of ctDNA detection can be further enhanced, even with limited amounts of input material, by using multiplexed patient-specific panels in combination with targeted sequencing methods34,114 (Fig. 4).

Figure 4: Leveraging multiple mutations to detect low-burden disease and overcome sampling noise.


Amplifications and deletions may be identified through low-depth (∼0.1× coverage) sequencing of the whole genome, and subsequent comparison of the relative number of sequencing reads in equally sized genomic regions across a sample, or in samples versus controls38. Such shallow WGS (sWGS) has been used to detect fetal aneuploidies14, and it can also be used to detect cancer-specific copy number alterations36,38,119. sWGS has a limit of detection of a mutant allele fraction of between 5% and 10% (Table 1), and so has limited sensitivity for profiling earlier-stage disease. If molecular profiling of a small number of recurrent copy number alterations is desired, higher sensitivity may be achieved through targeted sequencing of single-nucleotide polymorphisms, which may detect copy number alterations as low as 0.5%120.

The limit of detection for assays will vary on the basis of whether the individual's disease status and tumour mutations have already been characterized. Tumour burden in plasma has often been assessed by quantifying mutations (or other alterations) that were previously identified in the patient's tumour sample30,31. For mutation calling across a panel of genes or hot spots, the risk of false positives increases with the size of the panel owing to multiple hypothesis testing, and filters need to be applied to increase specificity; however, their application erodes sensitivity for rare variants. Prior knowledge of the mutation profile (for example, from tumour sequencing data) enables the detection of known patient-specific mutations above the background error rate, as opposed to calling mutations de novo34. Thus, sequencing-based assays can be used as sensitive and quantitative tools for ctDNA measurement and monitoring, in addition to their use for mutation profiling34,35,110 (Fig. 1b).

ctDNA can be quantified using various metrics, such as mutant allele concentration (that is, copies per millilitre) or mutant allele fraction121. Each of these metrics would be affected in a different way by analytical, pre- analytical and physiological characteristics. For example, metabolic changes in the rate of cfDNA turnover would affect the concentration of mutant alleles more than the mutant allele fraction, whereas pre-analytical factors affecting the release of germline DNA from blood cells would reduce the mutant allele fraction to a relatively greater extent. The analysis of ctDNA (both fraction and concentration)121, as well as total cfDNA and cfDNA fragmentation3,122, could therefore provide complementary information, and may have advantages in different applications or in combination.

ctDNA detection across cancer stages

ctDNA relates to stage and prognosis

The concentration of ctDNA in plasma has been shown to correlate with tumour size78,123 and stage124. A study of 640 patients with various types and stages of cancer found a 100-fold increase in the median ctDNA concentration in patients with stage IV disease compared with those who had stage I disease124. Quantification of individual tumour mutations in each patient revealed that patients with stage I disease had fewer than 10 copies per 5 ml of plasma. In sharp contrast, patients with advanced prostate, ovarian or colorectal cancer had a median concentration of 100–1,000 copies per 5 ml of plasma. ctDNA levels vary greatly even between patients with the same type and stage of disease. This variability in ctDNA concentration is partially explained by differences in the extent of metastatic spread or disease burden. In a recent report that compared ctDNA levels with tumour volume assessed by imaging in patients with relapsed high-grade serous ovarian cancer, ctDNA levels and disease volume were significantly correlated121. The mutant alleles in plasma increased in fraction by approximately 0.08%, and in concentration by 6 mutant copies per millilitre of plasma, for every cubic centimetre of disease121. Notwithstanding these correlations, substantial variation in ctDNA concentration may arise from interindividual differences. For example, poor tumour vascularization could hinder ctDNA release into the bloodstream or, conversely, could promote ctDNA release by producing hypoxia and cell death. Histological differences could foreseeably influence both the rate and type of cell death. Patients with primary brain tumours have very low levels of plasma ctDNA, with a median concentration for individual mutations of less than 10 copies per 5 ml of plasma124, whereas the proportion of tumour DNA in cerebrospinal fluid was found to be significantly higher60,61,124. Although not directly proved, the blood–brain barrier has been suggested to impede the movement of cfDNA fragments into the circulation60,61,124.

The relationship between ctDNA levels and cancer stage suggests prognostic utility for ctDNA. Patients with detectable ctDNA have been shown to have worse survival outcomes than those without27,125,126,127,128,129. In one of the earliest examples in the field, the 2-year overall survival rate for patients with colorectal cancer who had detectable ctDNA was 48%, as opposed to 100% for patients without27. In patients with detectable ctDNA, it has been found to be a significantly better prognostic predictor than are commonly used tumour markers35,121; specifically, an increasing concentration of ctDNA correlates with poorer clinical and radiological outcomes35,121,124,130,131. For example, in patients with metastatic breast cancer, a significant inverse correlation was shown between ctDNA concentration and overall survival up to 2,000 copies per millilitre, with a uniformly poor prognosis above this level35. In addition to ctDNA levels, mutational patterns identified in ctDNA (Fig. 1b) can help to group patients into molecular subtypes with different prognoses132.

Earlier diagnosis of disease

Diagnosing cancer at an earlier stage, particularly before metastatic spread, may allow earlier intervention and could improve survival133. A number of studies have demonstrated the potential for noninvasive early diagnosis. Mutations have been detected in saliva and plasma from individuals up to 2 years before cancer diagnosis21,134, and there have been reports of incidental presymptomatic detection of cancers in pregnant women who underwent NIPT67,135,136, as WGS can identify copy number alterations of both fetal and tumour origin. Screening in asymptomatic populations introduces the risks of over-diagnosis and false positives; implementation could therefore be explored in stages, and a first step could involve the use of ctDNA for earlier diagnosis of disease in symptomatic individuals, who at present may undergo lengthy investigative procedures.

In a survey across several cancer types, ctDNA was detected in 82% of patients with stage IV disease, and this fell to 47% for patients with stage I disease124. The method applied by the investigators was benchmarked as being able to detect one copy of an individual cancer mutation per 5 ml of plasma124. Using a sequencing gene panel that targeted a median of four mutations per patient, ctDNA was detected in 50% of patients with stage I non-small-cell lung cancer (NSCLC)110. Targeting known tumour mutations in plasma using droplet dPCR (ddPCR) assays in early-stage breast cancer showed a sensitivity of 93.3%137. A sWGS method adapted from an NIPT assay was recently shown to detect 6 of 16 (37.5%) cases of early ovarian cancer138, although this approach may not perform as well in other cancer types that have fewer copy number alterations. Together, these studies outline the possibility, and the challenge, of ctDNA detection in early-stage disease.

If analysis is performed on a few millilitres of plasma that contain only a few thousand copies of the genome, increasing the analytical sensitivity beyond the range of ∼1 in 1,000 may not produce any sensitivity benefit, as it becomes increasingly likely at low allele fractions that the mutation of interest may not be found owing to sampling noise. One approach may be to collect greater volumes of plasma (and more cfDNA) through methods such as plasmapheresis or implanted devices containing materials that bind cfDNA; similar approaches have been tested for enhancing the yield of circulating tumour cells (CTCs)139. Given these challenges, it is important that pre-analytical factors relating to the collection, processing and extraction of cfDNA are optimized (Box 2). Alternatively, for some cancer types, other minimally invasive samples may have a higher tumour DNA content than plasma: for example, urine for bladder cancers56; stool for colorectal cancers19; cytological specimens such as cervical smears140 or uterine lavage141 for gynaecological cancers, or oesophageal brushings142 for oesophageal cancers (Fig. 3c). For cancers with a viral aetiology — for example, nasopharyngeal carcinoma or cervical cancer — detection of the cancer-associated viral DNA that may be present in body fluids in many more copies than tumour DNA can enhance the identification of individuals with early-stage disease, or those with premalignant lesions and a high risk of cancer143,144.

Technical and biological advances facilitate enhanced mutation detection: the finding that ctDNA is shorter than cfDNA4,80,81 suggests the utility of experimental or in silico size selection (Fig. 3d). If the input amount of plasma or DNA is limiting, assays that interrogate multiple mutations in the same reaction (through broad genomic coverage or the use of patient-specific multiplexed panels) may produce a higher overall sensitivity for detecting the presence of any ctDNA34,35,110,112,114 than could be achieved by a focused analysis of individual loci (Fig. 4). Taken to a (currently impractical) extreme, ultra-deep sequencing of the entire genome could enable sensitive detection of cancer even from small volumes of plasma (or other fluids) in the future36.

Although technical advances may improve the sensitivity of ctDNA analysis further, biological and genomic factors may eventually become limiting. For confident de novo cancer detection using ctDNA, detected alterations should have a high positive predictive value for cancer. However, mutations known to be associated with cancer (for example, those in TP53, KRAS and Notch pathway genes) have been found at low levels in skin biopsies from healthy individuals145. If non- tumorigenic clones were to increase to a sufficient size and release mutated cfDNA, they could introduce biological noise. Clonal haematopoiesis with leukaemia- associated mutations has been observed in 10% of individuals older than 65 years of age, although the absolute risk of conversion into haematological cancer is 1%146. In addition, genomic alterations known to be associated with cancer have been found in plasma from healthy individuals114,118,134,147. Clinical outcomes for apparently healthy individuals in whom mutant DNA is detected in plasma should be characterized in order to understand the biological and clinical implications of such findings.

Cancer localization

Tissue-of-origin information from liquid biopsies might be able to aid cancer localization, for example, for cancers of unknown primary148. Methylation and nucleosome occupancy patterns in cfDNA have been found to encode tissue-specific and cell-specific information65,66,67,149: in one case, a pregnant woman had chromosomal abnormalities that were detectable by NIPT, and so tissue-specific methylation signals in the plasma were studied, enabling the quantification of the relative contributions of each tissue to the total cfDNA pool67. This revealed an increased contribution of cfDNA from B lymphocytes, which was consistent with a diagnosis of follicular lymphoma67. Although it may be possible in the future to deduce the tissue type of origin of a cancer in this manner, it remains to be seen whether the site or sites of metastatic spread can be determined through a measurable increase in the tissue-specific cfDNA signal.

Noninvasive molecular profiling

Analysis of heterogeneity

The extent of genetic heterogeneity has been confirmed in recent years, as multiregional sequencing studies have demonstrated clear differences between the mutation profiles of different tumour regions in the same patient150,151, and those of different specimens from primary and metastatic sites152. Although the potentially confounding effects of heterogeneity are recognized, it is often neither feasible nor desirable to perform multiple tumour biopsies on patients to try to account for this. Analysis of an individual biopsy might not accurately reflect the genomic architecture of a patient's cancer, and this could introduce bias to the selection and efficacy of personalized medicines. Furthermore, in a recent study of patients with lung cancer who were treated with an EGFR inhibitor, the tumour EGFRT790M allele fraction correlated with the degree of tumour shrinkage153, which suggests that the current paradigm of treatment selection based on mutation presence or absence alone may be suboptimal.

Liquid biopsies sample ctDNA released from multiple tumour regions, and may thereby reflect both intratumour heterogeneity44,46,61 and spatially separated disease foci47,154,155,156. Although individual tumour biopsies from different tumour regions may differ in their mutation profiles owing to intratumour heterogeneity157,158, ctDNA analysis has detected mutations that have been missed in corresponding tissue samples44,47,159,160. Multiregion tumour sequencing data show that stem mutations (which are shared by all tumour regions) have a higher allele fraction in plasma than do private mutations46,47. Therefore, stem mutations would provide the most reliable method of tracking tumour burden in plasma. Alternatively, tracking a large set of mutations may compensate for potential biases that may be introduced by heterogeneous or private mutations.

Hot-spot mutations and gene panels

By comparing mutation detection in plasma against that in matched tumour samples, the sensitivity of ctDNA analysis has been estimated in retrospective studies as between 65% and 98%100,102,122,156,159,161–163. For profiling specific loci — for example, in order to stratify patients to receive matched molecular therapies — international studies have begun to demonstrate that large-scale testing is feasible and can be standardized, although the use of assays with limited analytical sensitivity resulted in low detection rates of ctDNA164–166. Using assays developed specifically to detect low levels of ctDNA3, a blinded prospective study demonstrated sensitivities of 92% and 100% (and concordance rates of 96% and 100%) for KRAS and BRAF mutations, respectively, in metastatic colorectal cancer122.

Considerable attention has been devoted to the analysis of EGFRmutations in patients with NSCLC33,167, as it is often challenging to obtain tissue biopsies from these patients to help to inform treatment40,41. A meta-analysis of 27 selected studies published between 2007 and 2015 that used a variety of methods and comprised nearly 4,000 patients, found a pooled sensitivity of 60% and specificity of 94% for detection of EGFR mutations in plasma or serum33. In a phase IV study of the EGFR inhibitor gefitinib, mutation status was compared between tumour and plasma samples from 652 patients. The sensitivity and specificity for detecting mutations in plasma were determined as 65.7% and 99.8%, respectively166,168; these measurements were probably affected by the limited analytical sensitivity of the allele-specific PCR method used98, as the version of the kit that was used in that study had limits of detection (at ≥95% analytical sensitivity) of 1.64% and 1.26% for EGFR deletions and L858R mutations, respectively166. Low rates of concordance of EGFRT790M status were also observed in a recent phase III trial of osimertinib169, in which tissue testing was compared with plasma using an allele- specific PCR assay for ctDNA analysis that has a limit of detection (with ≥95% analytical sensitivity) of 100 copies of EGFRT790M per millilitre of plasma99. Using methods with greater analytical sensitivity, higher concordance rates can be obtained159,172. The EGFRT790M mutation confers resistance to gefitinib and erlotinib and frequently emerges after initial treatment with EGFR tyrosine kinase inhibitors32,170,171. In plasma, the sensitivity for detecting this mutation was lower than for EGFRL858R and EGFREx19del mutations (70% versus ≥90%), which occur earlier in the development of disease159,172. Initial data suggest that this may be due to the heterogeneous presence of resistance mutations at disease relapse159.

In retrospective analyses, despite limited concordance rates of EGFRmutation status between plasma and tumour samples observed in some studies, response rates for patients who were plasma positive for mutations in EGFR were similar to the response rates of patients who were tissue positive159,166,168,169,172. Data showing the response of patients treated solely on the basis of ctDNA analysis are starting to emerge: patients who were treated with osimertinib on the basis of the detection of EGFRT790M in plasma showed response rates that were similar to those of patients treated on the basis of tissue analysis173. Interestingly, objective responses were also seen in patients with very low allele fractions of mutant EGFRT790M in plasma (<0.5%)173.

At present, the European Medicines Agency (EMA) and the US Food and Drug Administration (FDA) approve the use of information from ctDNA analysis to help to select patients with EGFR-mutant NSCLC for gefitinib (EMA)174, erlotinib (FDA)99 or osimertinib (EMA and FDA)99,175 therapy in the event that a tumour sample is not evaluable. This could offer a pragmatic solution that provides molecular profiling information for patients while avoiding repeat biopsies for some individuals. Current recommendations99,175 state that if liquid biopsies are carried out in advance of a tumour biopsy, ctDNA detection may abrogate the need for tissue biopsy, but if ctDNA analysis is negative, a tissue biopsy may still provide valuable genomic information.

Molecular profiling using ctDNA may have particular utility for stratifying patients in 'basket trials', which enrol patients independently of tumour histology, or 'umbrella trials', which assign patients to multiple investigational drugs or treatment options176. For example, a 54-gene panel detected ctDNA in 58% of patients across multiple cancer types177. Of the patients with alterations, 71.4% had at least one mutation that was actionable by an FDA-approved drug177. This panel is being used to test the feasibility of matching patients who have different metastatic cancer types to targeted therapies in a prospective clinical trial178. In another study presented at the 2016 Molecular Analysis for Personalised Therapy meeting, a 34-gene panel identified mutations in 79% of 174 patients with NSCLC, and this allowed 28 patients (17%) to receive personalized treatment on the basis of ctDNA molecular profiling179. Personalized therapy selection presents challenges: even if mutations are successfully detected using ctDNA in patients, an efficacious molecularly targeted agent may not exist. However, data from a prospective clinical trial presented at the 2016 Molecular Analysis for Personalised Therapy meeting demonstrated that selecting therapies on the basis of genomic analysis can improve outcomes for patients with cancer, even when patients with well- established actionable targets (for which approved drugs are available) were excluded180.

Improvements in the analytical sensitivity of molecular profiling tools could further increase detection and concordance rates, or allow for sensitive multiplexed analysis (Table 1), although biological factors and heterogeneity may reduce sensitivity in some cancer types and stages124,159,181. The utility of ctDNA should, therefore, be assessed for different clinical indications. However, benchmarking ctDNA against individual tumour biopsies may be confounded by sampling error, as rare private mutations may be sampled in the biopsy but ctDNA carrying such mutations may be present at levels that are too low to be detectable.

Structural variants

Copy number alterations can be detected in cfDNA using WGS36,37,182,183, amplicon-based117,119,184 and hybrid-capture approaches39,110,162,185. In patients with hepatocellular carcinoma, WGS was able to identify amplifications and deletions in plasma that matched those identified in tumour tissue37,186. Heterogeneous copy number changes were also identified in a patient with synchronous breast and ovarian cancers, as copy number changes unique to each cancer were detected in plasma37. In a study of 80 patients with prostate cancer, androgen receptor (AR) copy number gain before abiraterone therapy predicted a worse overall survival, thus identifying patients with primary resistance187. For patients with advanced disease, sWGS may provide a relatively cost-effective measure of ctDNA levels that is applicable across cancer types. This approach may have utility as a sample screening step in a ctDNA analysis workflow119, such that high-burden patient samples are triaged for exome sequencing39.

Chromosomal rearrangements in plasma can be identified through both WGS36 and hybrid-capture sequencing approaches38,188, although the latter may be more economical owing to the depth of coverage needed to confidently identify a rearrangement. In one study of patients with prostate cancer, sWGS was able to detect a deletion on chromosome 21 in five patients, although higher-depth hybrid-capture sequencing was necessary to identify a rearrangement between exon 1 of TMPRSS2(which encodes transmembrane protease serine 2) and exon 3 of ERG38.

Longitudinal monitoring

Monitoring response

The short half-life of cfDNA31,70,71,72, as well as the ease and reduced risk of repeating liquid biopsies relative to imaging49 or tissue biopsies40, enables liquid biopsies to be used for real-time monitoring of cancer burden in response to therapy. Studies monitoring patients during treatment have shown that ctDNA dynamics correlate with treatment response31,34,35,110,129 and may identify response earlier than does clinical detection35,189,190. In patients with breast cancer, ctDNA showed the greatest range in concentration and provided the earliest measure of response to chemotherapies, as well as the earliest indication of impending relapse when compared with imaging and other blood-based cancer markers, such as CTCs and cancer antigen 15-3 (CA15-3; also known as MUC1)35. In relapsed ovarian cancer, pretreatment ctDNA levels and the extent of ctDNA decrease after chemotherapy initiation were significantly associated with time to progression, and were more informative than levels of CA125 (also known as MUC16)121.

A recent study suggested that an early spike in ctDNA levels (specifically, an increase in the allele fractions of BRAF mutations) in the first week following the initiation of immunotherapy in patients with melanoma may predict response191. This may reflect a transient increase in cell death. If these data are confirmed, sampling at early time points could be applied in the clinic as well as in drug development. However, the presence or timing of such spikes in cell death would probably vary on the basis of the pharmacological properties of the treatments used and the biological responses to them. An early spike was not observed a few days after the initiation of treatment with chemotherapy in patients with colorectal cancer161 or treatment with an EGFR inhibitor in patients with NSCLC190. If the analysis of plasma immediately after the start of therapy could reliably detect the destruction of sensitive cancer cells, this raises an exciting possibility that the existence of resistant subclones could be identified very rapidly by detecting the differential early dynamics of mutations. In the context of immunotherapy, liquid biopsies may provide the opportunity to monitor both ctDNA and the response of the immune system through, for example, the analysis of cfDNA released from distinct T cell clones192.

Minimal residual disease and recurrence monitoring

Following surgery or treatment with curative intent, detection of ctDNA may signal the presence of minimal residual disease (MRD) even in the absence of any other clinical evidence of disease, and the presence of ctDNA could identify patients who may be at higher risk of relapse. Stratification of patients into high-risk and low-risk groups would enable adjuvant therapy to be given to patients who are likely to benefit most, while sparing low-risk patients from unnecessary comorbidities and the risk of adverse events. In a prospective study of 230 patients with early-stage colorectal cancer, the assessment of ctDNA at the first follow-up visit after surgical resection indicated that recurrence-free survival at 3 years was 0% for the ctDNA-positive group and 90% for the ctDNA-negative group193. In a separate study of 55 patients with early-stage breast cancer, assessment of ctDNA showed that the detection of ctDNA at first follow-up could also indicate poor prognosis194. Furthermore, stratification based on mutation detection across serial samples improved the prediction of relapse, and this and other studies have observed an interval of 7.9–11.0 months between ctDNA detection and clinical relapse194–196, which is similar to that identified in the metastatic setting35. With more sensitive approaches (Fig. 4), earlier identification of clinical relapse should be possible. Combining monitoring for known mutations with molecular profiling assays could enable the identification of potential targets for early therapeutic intervention; ideally, such intervention would, in future, prevent or postpone overt relapse.

Patient-specific DNA rearrangements identified from sequencing tumour samples can be used to design assays to track tumour burden in plasma196,197,198,199. Curative surgery could provide an excellent opportunity to obtain tumour DNA that can be sequenced to guide the design of assays for postoperative monitoring. Patient-specific rearrangements may be more readily detected in ctDNA at high sensitivity197,198, as rearrangements are less confounded by background noise than mutations are. One of the challenges of individualized panel design is that sequencing an individual tumour biopsy may not sample every mutation in heterogeneous disease; therefore, sequencing of matched body fluid and tumour samples may be desirable for comprehensive mutation profiling. In future, if tumour sequencing becomes routine, monitoring disease using patient-specific panels could become viable, although the routine implementation of such assays may face challenges related to validation, regulation and quality assurance of individual assays.

Clonal evolution and resistance

As discussed above, a rising or falling ctDNA concentration may provide an indication of the effect of treatment on overall tumour burden. If multiple tumour mutations are interrogated, then the relative change between each may provide insight into the molecular evolution of a patient's cancer39,200,201. Ratios between the levels of different mutations in plasma can indicate heterogeneity and may be informative for the prediction of patient response to treatments that target particular alterations159,162. Liquid biopsies have been shown to contain ctDNA from multiple tumour sites44,46,47,61, and their analysis can have a faster turnaround time38,102,188 and might be less prone to biases150,151,202 resulting from the analysis of individual tumour biopsies.

Studies demonstrate that ctDNA can be used to monitor clonal evolution and identify mechanisms of resistance to treatment112,160,162,200,203,204. Serial ctDNA analysis in patients with colorectal cancer demonstrated the positive selection of mutant KRASclones during EGFR blockade, and a later decline in the mutant KRASclones upon the withdrawal of anti-EGFR therapy160,205. In patients with NSCLC undergoing treatment with EGFR inhibitors, resistance-conferring mutations emerged in plasma ctDNA ahead of clinical progression170,171. Exome sequencing of plasma DNA may identify resistance mechanisms in patients with various cancer types39, although the sensitivity of exome sequencing currently limits its application to patients with advanced cancer in whom ctDNA levels are high (>5% mutant allele fraction). The design of patient-specific mutation panels34,47,114,206 could be a more cost-effective alternative for high-sensitivity monitoring, although these panels may miss subsequent de novo events, unless appropriately designed. Serial sWGS analysis also demonstrates highly dynamic copy number adaptations in response to selection pressures, with a mean interval of 26.4 weeks between new amplifications207.

Serial liquid biopsies may have particular utility for adaptive or reactive therapy, in which resistance mutations are prospectively identified and therapy is adapted in real time (Fig. 5). In the clinical research setting, noninvasive monitoring could facilitate clinical trials that prospectively identify efficacious treatment regimens or drug combinations, and identify resistance mechanisms to novel therapies. In addition, in vitroor in vivo experiments carried out in parallel may provide further insight into cancer biology. For example, experiments in a colorectal cancer cell line that were carried out in parallel with ctDNA analysis showed that resistance mutations may arise both from the selection of pre-existing minor clones and through ongoing mutagenesis200. Another study investigating resistance to a pan-tropomyosin-related kinase (TRK; also known as NTRK) inhibitor in colorectal cancer demonstrated that the simultaneous analysis of patient-derived xenografts and liquid biopsies may characterize resistance more comprehensively than the analysis of plasma alone199.

Figure 5: Adaptive or reactive treatment paradigms using liquid biopsies.


Future directions

Proof-of-concept studies provide an excellent starting point for larger prospective studies of the clinical utility of ctDNA and demonstrate that ctDNA may be a useful research tool for drug development, and for the study of intratumour heterogeneity and clonal evolution. In the future, randomized trials comparing ctDNA- guided decision-making against the standard of care would be definitive, and the EMA have outlined good-practice guidelines for the design of such trials208. Trials to test the clinical utility of ctDNA analysis for treatment monitoring are now being carried out209,210. In one trial, patients with NSCLC receiving erlotinib are being prospectively monitored, and if resistance mutations emerge in plasma then additional scans to search for signs of disease progression will be carried out209. Another clinical trial aims to demonstrate the efficacy of targeting mutations identified in plasma from patients with advanced breast cancer211, and the findings of this study could support the future use of plasma- only mutation profiling and treatment stratification. Together, these studies highlight that the field is moving from exploratory ctDNA studies towards clinical trials in which ctDNA is guiding decision-making.

A better understanding of the origin and biology of cfDNA and ctDNA would aid the implementation of liquid biopsies2. The relative contributions of apoptosis, necrosis and active release, particularly at different time points during treatment, should be explored. Our limited understanding of the release and clearance mechanisms of cfDNA hampers the interpretation of current studies. Studies of the dynamics and reproducibility of ctDNA measurement in the absence of intervention will become increasingly important as we aim to interpret ctDNA signal in response to treatment. It is also not clear whether all tumour subclones contribute proportionately to the total ctDNA pool or whether their representation in the bloodstream is biased by other biological factors, such as tumour vascularity or metabolic activity. In vivo cellular barcoding experiments212 and autopsy studies47 could elucidate the contribution of individual subclones, and histological studies may clarify the factors that modulate ctDNA release. The difference in size between cfDNA and ctDNA fragments4,67,80,81 suggests that optimizing processing and extraction methods (as well as downstream assays) for the recovery of selected fragment sizes may provide further improvement to overall performance.

Although ctDNA analysis can have increased sensitivity and specificity compared with the analysis of other circulating biomarkers35, taking a multi-marker approach may offer a more comprehensive insight into a patient's disease3,122,213. For example, the total cfDNA concentration correlates with disease status3,5 and is associated with prognosis214. Epigenetic analysis of cfDNA may identify cancer gene hypermethylation27,155 or the cell type that gives rise to cfDNA fragments65,66,67, and may provide information about the tumour microenvironment, which usually lacks somatic mutations. Other circulating nucleic acids such as mRNA and microRNA can provide additional layers of information215. Targeting multiple types of nucleic acid that have independent mechanisms of release (for example, through the co- isolation of exosomal RNA and cfDNA216) may increase the sensitivity of detecting MRD. Actively released nucleic acids may be preferred for the detection of mutations in subclones that are resistant to therapy, whereas fragments that arise from dying cells following the initiation of therapy may identify treatment-responsive subclones. Next, although it may be possible to infer gene expression patterns from cfDNA217, sequencing RNA within exosomes218, CTCs219 or platelets220 could provide more direct evidence. Analysis of cfDNA in plasma alongside other fluids such as urine or cerebrospinal fluid can provide complementary information. We further echo the suggestion made by Gormally et al.221 a decade ago that the characterization of proteins associated with cfDNA may provide a rich source of information about an individual's disease and about the biology of cfDNA.

The clinical uptake of liquid biopsies will depend on the practical advantages for patients and clinicians, the infrastructure required and its cost-effectiveness. Tissue biopsies will continue to have a key role in cancer management, particularly for the histological diagnosis and classification of cancers. At present, specialized laboratories handle CTC and ctDNA samples222, although in future, hospital laboratories may carry out analysis locally if appropriate processes can be established223.

Point-of-care devices that have clinically meaningful sensitivities for the identification of individual hot-spot mutations are starting to be used for tissue and plasma samples224,225. The feasibility of single-molecule (third-generation) sequencing of maternal plasma DNA was first demonstrated in 2015 (Ref. 85), and subsequently it was shown that structural variants in cell line DNA can be detected226. The portability of such technologies was demonstrated by the real-time genomic surveillance in the field during the Ebola virus disease epidemic227. At present, such platforms are limited by a high error rate226, which makes single- nucleotide variant and indel detection challenging. Another challenge is that of sequencing short DNA fragments, which requires optimized library preparation methods228. Sequencing capacity is also limited (currently to ∼150 Mb)229, although this is likely to increase in the near future, and specific amplicons can be targeted through real-time selective sequencing230. These studies support the possibility of molecular profiling at the point of care, especially if blood plasma can be analysed without the relatively cumbersome and time-consuming step of DNA purification86,231.

The initial approvals by the EMA and FDA for mutation detection in plasma as a companion diagnostic99,174,175, and emerging ctDNA-guided clinical trials209−211, represent key milestones towards the implementation of liquid biopsies in personalized oncology. Improving technologies are enabling an ever-wider scope for noninvasive molecular analysis of cancer, providing information that opens new avenues for genomic research and may aid in clinical decisions. In order to fully exploit the potential utility of liquid biopsies, it is essential that the biology of ctDNA is explored further. Thus far, liquid biopsies have demonstrated potential utility across a range of applications, and they are beginning to be used for patient benefit.

Acknowledgements

The authors would like to thank all members of the Rosenfeld laboratory, in particular C. G. Smith, D. Gale, K. M. Patel and W. N. Cooper for expert advice and proofreading of this manuscript. They would like to thank D. Lo (The Chinese University of Hong Kong) for providing the image used in Figure 3a. The authors would also like to acknowledge the support of the University of Cambridge (UK), Cancer Research UK (grant numbers A11906, A20240 and A15601; to N.R. and J.D.B.), the European Research Council (ERC) under the European Union's Seventh Framework Programme (FP/2007-2013) ERC Grant Agreement number 337905 (to N.R.), Hutchison Whampoa Limited (Hong Kong; to N.R.), AstraZeneca (UK; to R.B. and S.P.), the Cambridge Experimental Cancer Medicine Centre (to R.B. and S.P.), and the National Institute for Health Research Biomedical Research Centre (UK; to R.B. and S.P.). J.G.-C. acknowledges clinical fellowship support from the Spanish Society of Medical Oncology (SEOM).

Author information

Richard Baird & Nitzan Rosenfeld

These authors contributed equally to this work.

Affiliations

Cancer Research UK Cambridge Institute, Li Ka Shing Centre, Robinson Way, Cambridge CB2 0RE, UK.
Jonathan C. M. Wan, Charles Massie, Florent Mouliere, James D. Brenton, Carlos Caldas & Nitzan Rosenfeld

Cancer Research UK Cambridge Centre, Cambridge CB2 0RE, UK.
Jonathan C. M. Wan, Charles Massie, Florent Mouliere, James D. Brenton, Carlos Caldas, Simon Pacey, Richard Baird & Nitzan Rosenfeld

Clinical Trials Unit, Clinic Institute of Haematological and Oncological Diseases, Hospital Clinic de Barcelona, IDIBAPs, Carrer de Villarroel, 170 Barcelona 08036, Spain.
Javier Garcia-Corbacho

Department of Oncology, University of Cambridge Hutchison–MRC Research Centre, Box 197, Cambridge Biomedical Campus, Cambridge CB2 0XZ, UK.
Carlos Caldas, Simon Pacey & Richard Baird

Competing interests

N.R. is the Chief Scientific Officer of Inivata (Cambridge, UK, and Research Triangle Park, North Carolina, USA). N.R. and J.D.B. are co-founders and shareholders of Inivata. N.R. and F.M. are co-inventors of patent applications that describe methods for the analysis of DNA fragments and applications of circulating tumour DNA.

Corresponding author

Correspondence to Nitzan Rosenfeld.

Glossary

Liquid biopsy

Analysis of tumour material (for example, cells or nucleic acids) obtained in a minimally invasive or noninvasive manner through the sampling of blood or other body fluids.

Digital PCR

(dPCR). An assay in which many microlitre- or nanolitre-scale PCR reactions are run in parallel within physically separated reaction chambers or as droplets in an emulsion (droplet dPCR (ddPCR)). By partitioning molecules into hundreds or up to millions of reactions, rare mutant molecules can be accurately identified and quantified.

Mutant allele fraction

The proportion of mutant DNA fragments at a given locus.

Hybrid-capture sequencing

DNA sequencing of kilobases to megabases of the genome, in which the DNA to be sequenced is selected using complementary oligonucleotide baits that hybridize to the target DNA. The DNA is then captured in solution, commonly through binding to magnetic beads.

Targeted sequencing

Massively parallel (next-generation) sequencing that uses methods such as PCR amplification or hybrid capture to focus on a subset of the genome, which can range from few genes or mutation loci to large proportions of the genome, such as the entire exome. Smaller panels yield a higher sequencing depth at lower costs than do larger panels.

Molecular barcoding

The addition of unique molecular sequences to each molecule when creating a sequencing library, so that reads originating from the same molecule can be identified and the consensus taken, correcting for some PCR or sequencing errors.

Limit of detection

The threshold below which mutations cannot be confidently discriminated from background noise; for sequencing-based approaches, this is often determined by technical artefacts such as PCR or sequencing errors.

Mutant allele concentration

The number of mutant DNA fragments at a given locus per unit volume.

Stem mutations

Mutations that occur early in a cancer's development and are present in all cells.

Private mutations

Mutations that are present only in a specific region of a tumour, or in a subset of cells, owing to intratumour heterogeneity.

Rights and permissions

To obtain permission to re-use content from this article visit RightsLink.

Literature
  1. Mandel, P. & Métais, P. Les acides nucléiques du plasma sanguin chez l'homme. C. R. Seances Soc. Biol. Fil. 142, 241–243 (in French) (1948).
  2. Thierry, A. R., El Messaoudi, S., Gahan, P. B., Anker, P. & Stroun, M.Origins, structures, and functions of circulating DNA in oncology. Cancer Metastasis Rev. 35, 347–376 (2016).
  3. Mouliere, F., El Messaoudi, S., Pang, D., Dritschilo, A. & Thierry, A. R. Multi-marker analysis of circulating cell-free DNA toward personalized medicine for colorectal cancer. Mol. Oncol. 8, 927–941 (2014).
  4. Mouliere, F. et al. High fragmentation characterizes tumour-derived circulating DNA. PLoS ONE 6, e23418 (2011). This study specifically characterizes ctDNA fragmentation patterns and the biological properties of ctDNA using an animal model system and clinical samples.
  5. Leon, S. A., Shapiro, B., Sklaroff, D. M. & Yaros, M. J. Free DNA in the serum of cancer patients and the effect of therapy. Cancer Res. 37, 646–650 (1977).
  6. Rodrigues Filho, E. M. et al. Elevated cell-free plasma DNA level as an independent predictor of mortality in patients with severe traumatic brain injury. J. Neurotrauma 31, 1639–1646 (2014).
  7. Tsai, N.-W. et al. The value of serial plasma nuclear and mitochondrial DNA levels in patients with acute ischemic stroke. Clin. Chim. Acta 412, 476–479 (2011).
  8. Breitbach, S., Sterzing, B., Magallanes, C., Tug, S. & Simon, P.Direct measurement of cell-free DNA from serially collected capillary plasma during incremental exercise. J. Appl. Physiol. 117, 119–130 (2014).
  9. De Vlaminck, I. et al. Circulating cell-free DNA enables noninvasive diagnosis of heart transplant rejection. Sci. Transl Med. 6, 241ra77 (2014).
  10. De Vlaminck, I. et al. Noninvasive monitoring of infection and rejection after lung transplantation. Proc. Natl Acad. Sci. USA 112, 13336–13341 (2015).
  11. Lo, Y. M. D. et al. Presence of fetal DNA in maternal plasma and serum. Lancet 350, 485–487 (1997).
  12. Hyett, J. A. et al. Reduction in diagnostic and therapeutic interventions by non-invasive determination of fetal sex in early pregnancy. Prenat. Diagn. 25, 1111–1116 (2005).
  13. Saito, H., Sekizawa, A., Morimoto, T., Suzuki, M. & Yanaihara, T.Prenatal DNA diagnosis of a single-gene disorder from maternal plasma. Lancet 356, 1170 (2000).
  14. Lo, Y. M. D. et al. Digital PCR for the molecular detection of fetal chromosomal aneuploidy. Proc. Natl Acad. Sci. USA 104, 13116–13121 (2007).
  15. Allyse, M. et al. Non-invasive prenatal testing: a review of international implementation and challenges. Int. J. Womens Health 7, 113–126 (2015).
  16. Hill, M. et al. Evaluation of non-invasive prenatal testing (NIPT) for aneuploidy in an NHS setting: a reliable accurate prenatal non-invasive diagnosis (RAPID) protocol. BMC Pregnancy Childbirth 14, 229 (2014).
  17. Stroun, M. et al. Neoplastic characteristics of the DNA found in the plasma of cancer patients. Oncology 46, 318–322 (1989).
  18. Sidransky, D. et al. Identification of p53 gene mutations in bladder cancers and urine samples. Science 252, 706–709 (1991).
  19. Sidransky, D. et al. Identification of ras oncogene mutations in the stool of patients with curable colorectal tumors. Science 256, 102–105 (1992).
  20. Caldas, C. et al. Detection of K-ras mutations in the stool of patients with pancreatic adenocarcinoma and pancreatic ductal hyperplasia. Cancer Res. 54, 3568–3573 (1994).
  21. Mao, L., Hruban, R. H., Boyle, J. O., Tockman, M. & Sidransky, D.Detection of oncogene mutations in sputum precedes diagnosis of lung cancer. Cancer Res. 54, 1634–1637 (1994).
  22. Takeda, S., Ichii, S. & Nakamura, Y. Detection of K-ras mutation in sputum by mutant-allele-specific amplification (MASA). Hum. Mutat. 2, 112–117 (1993).
  23. Sorenson, G. D. et al. Soluble normal and mutated DNA-sequences from single-copy genes in human blood. Cancer Epidemiol. Biomarkers Prev. 3, 67–71 (1994).
  24. Swisher, E. M. et al. Tumor-specific p53 sequences in blood and peritoneal fluid of women with epithelial ovarian cancer. Am. J. Obstet. Gynecol. 193, 662–667 (2005).
  25. Kimura, H. et al. Detection of epidermal growth factor receptor mutations in serum as a predictor of the response to gefitinib in patients with non-small-cell lung cancer. Clin. Cancer Res. 12, 3915–3921 (2006).
  26. Sozzi, G., Musso, K., Ratcliffe, C., Goldstraw, P. & Pierotti, M. A.Detection of microsatellite alterations in plasma DNA of non-small cell lung cancer patients: a prospect for early diagnosis. Clin. Cancer Res. 5, 2689–2692 (1999).
  27. Lecomte, T. et al. Detection of free-circulating tumor-associated DNA in plasma of colorectal cancer patients and its association with prognosis. Int. J. Cancer 100, 542–548 (2002).
  28. Vogelstein, B. & Kinzler, K. W. Digital PCR. Proc. Natl Acad. Sci. USA 96, 9236–9241 (1999).
  29. Dressman, D., Yan, H., Traverso, G., Kinzler, K. W. & Vogelstein, B.Transforming single DNA molecules into fluorescent magnetic particles for detection and enumeration of genetic variations. Proc. Natl Acad. Sci. USA 100, 8817–8822 (2003).
  30. Diehl, F. et al. Detection and quantification of mutations in the plasma of patients with colorectal tumors. Proc. Natl Acad. Sci. USA 102, 16368–16373 (2005).
  31. Diehl, F. et al. Circulating mutant DNA to assess tumor dynamics. Nat. Med. 14, 985–990 (2008). This study demonstrates that ctDNA dynamics in patients undergoing treatment for colorectal cancer reflect tumour responses and progression, and that ctDNA detection after surgery represented a marker of residual disease and a prognostic factor.
  32. Kuang, Y. et al. Noninvasive detection of EGFR T790M in gefitinib or erlotinib resistant non-small cell lung cancer. Clin. Cancer Res.15, 2630–2636 (2009).
  33. Qian, X. et al. Circulating cell-free DNA has a high degree of specificity to detect exon 19 deletions and the single-point substitution mutation L858R in non-small cell lung cancer. Oncotarget 7, 29154–29165 (2016).
  34. Forshew, T. et al. Noninvasive identification and monitoring of cancer mutations by targeted deep sequencing of plasma DNA. Sci. Transl Med. 4, 136ra68 (2012). This study identifies cancer mutations directly in plasma using a next-generation sequencing-based assay, and applies this method to quantify and monitor multiple mutations.
  35. Dawson, S.-J. et al. Analysis of circulating tumor DNA to monitor metastatic breast cancer. N. Engl. J. Med. 368, 1199–1209 (2013).
  36. Leary, R. J. et al. Detection of chromosomal alterations in the circulation of cancer patients with whole-genome sequencing. Sci. Transl Med. 4, 162ra154 (2012). In this study, WGS of cfDNA from plasma identified copy number alterations and rearrangements.
  37. Chan, K. C. A. et al. Cancer genome scanning in plasma: detection of tumor-associated copy number aberrations, single-nucleotide variants, and tumoral heterogeneity by massively parallel sequencing. Clin. Chem. 59, 211–224 (2013).
  38. Heitzer, E. et al. Tumor-associated copy number changes in the circulation of patients with prostate cancer identified through whole-genome sequencing. Genome Med. 5, 30 (2013). This study demonstrates that WGS of plasma cfDNA at a shallow sequencing depth can establish copy number profiles in patients with prostate cancer with a rapid turnaround time.
  39. Murtaza, M. et al. Non-invasive analysis of acquired resistance to cancer therapy by sequencing of plasma DNA. Nature 497, 108–112 (2013). This paper describes the use of exome sequencing of cfDNA from serial plasma samples to study clonal evolution and to track ctDNA dynamics in high-burden disease.
  40. Overman, M. J. et al. Use of research biopsies in clinical trials: are risks and benefits adequately discussed? J. Clin. Oncol. 31, 17–22 (2013).
  41. VanderLaan, P. A. et al. Success and failure rates of tumor genotyping techniques in routine pathological samples with non-small-cell lung cancer. Lung Cancer 84, 39–44 (2014).
  42. Ellis, P. M. et al. Dacomitinib compared with placebo in pretreated patients with advanced or metastatic non-small-cell lung cancer (NCIC CTG BR.26): a double-blind, randomised, phase 3 trial. Lancet Oncol. 15, 1379–1388 (2014).
  43. Popper, H. H. Commentary on tumor heterogeneity. Transl Lung Cancer Res. 5, 433–435 (2016).
  44. De Mattos-Arruda, L. et al. Capturing intra-tumor genetic heterogeneity by de novo mutation profiling of circulating cell-free tumor DNA: a proof-of-principle. Ann. Oncol. 25, 1729–1735 (2014).
  45. Lebofsky, R. et al. Circulating tumor DNA as a non-invasive substitute to metastasis biopsy for tumor genotyping and personalized medicine in a prospective trial across all tumor types. Mol. Oncol. 9, 783–790 (2015).
  46. Jamal-Hanjani, M. et al. Detection of ubiquitous and heterogeneous mutations in cell-free DNA from patients with early-stage non-small-cell lung cancer. Ann. Oncol. 27, 862–867 (2016).
  47. Murtaza, M. et al. Multifocal clonal evolution characterized using circulating tumour DNA in a case of metastatic breast cancer. Nat. Commun. 6, 8760 (2015).
  48. Duffy, M. J. Serum tumor markers in breast cancer: are they of clinical value? Clin. Chem. 52, 345–351 (2006).
  49. Fazel, R. et al. Exposure to low-dose ionizing radiation from medical imaging procedures. N. Engl. J. Med. 361, 849–857 (2009).
  50. Thakur, B. K. et al. Double-stranded DNA in exosomes: a novel biomarker in cancer detection. Cell Res. 24, 766–769 (2014).
  51. Kahlert, C. et al. Identification of double-stranded genomic DNA spanning all chromosomes with mutated KRAS and p53 DNA in the serum exosomes of patients with pancreatic cancer. J. Biol. Chem. 289, 3869–3875 (2014).
  52. Jahr, S. et al. DNA fragments in the blood plasma of cancer patients: quantitations and evidence for their origin from apoptotic and necrotic cells. Cancer Res. 61, 1659–1665 (2001).
  53. Stroun, M., Lyautey, J., Lederrey, C., Olson-Sand, A. & Anker, P.About the possible origin and mechanism of circulating DNA: apoptosis and active DNA release. Clin. Chim. Acta 313, 139–142 (2001).
  54. Botezatu, I. et al. Genetic analysis of DNA excreted in urine: a new approach for detecting specific genomic DNA sequences from cells dying in an organism. Clin. Chem. 46, 1078–1084 (2000).
  55. Chan, K. C. A., Leung, S. F., Yeung, S. W., Chan, A. T. C. & Lo, Y. M. D. Quantitative analysis of the transrenal excretion of circulating EBV DNA in nasopharyngeal carcinoma patients. Clin. Cancer Res.14, 4809 (2008).
  56. Birkenkamp-Demtröder, K. et al. Genomic alterations in liquid biopsies from patients with bladder cancer. Eur. Urol. 70, 75–82 (2016).
  57. Su, Y.-H. et al. Human urine contains small, 150 to 250 nucleotide- sized, soluble DNA derived from the circulation and may be useful in the detection of colorectal cancer. J. Mol. Diagn.6, 101–107 (2004).
  58. Reckamp, K. L. et al. A highly sensitive and quantitative test platform for detection of NSCLC EGFR mutations in urine and plasma. J. Thorac. Oncol. 11, 1690–1700 (2016).
  59. Wang, Y. et al. Detection of tumor-derived DNA in cerebrospinal fluid of patients with primary tumors of the brain and spinal cord. Proc. Natl Acad. Sci. USA 112, 9704–9709 (2015).
  60. Pan, W., Gu, W., Nagpal, S., Gephart, M. H. & Quake, S. R. Brain tumor mutations detected in cerebral spinal fluid. Clin. Chem. 61, 514–522 (2015).
  61. De Mattos-Arruda, L. et al. Cerebrospinal fluid-derived circulating tumour DNA better represents the genomic alterations of brain tumours than plasma. Nat. Commun. 6, 8839 (2015).
  62. Rhodes, C. H., Honsinger, C. & Sorenson, G. D. PCR-detection of tumor-derived p53 DNA in cerebrospinal fluid. Am. J. Clin. Pathol.103, 404–408 (1995).
  63. Sriram, K. B. et al. Pleural fluid cell-free DNA integrity index to identify cytologically negative malignant pleural effusions including mesotheliomas. BMC Cancer 12, 428 (2012).
  64. Mithani, S. K. et al. Mitochondrial resequencing arrays detect tumor-specific mutations in salivary rinses of patients with head and neck cancer. Clin. Cancer Res. 13, 7335–7340 (2007).
  65. Snyder, M. W., Kircher, M., Hill, A. J., Daza, R. M. & Shendure, J.Cell-free DNA comprises an in vivo nucleosome footprint that informs its tissues-of-origin. Cell 164, 57–68 (2016).
  66. Lehmann-Werman, R. et al. Identification of tissue-specific cell death using methylation patterns of circulating DNA. Proc. Natl Acad. Sci. USA 113, E1826–E1834 (2016).
  67. Sun, K. et al. Plasma DNA tissue mapping by genome-wide methylation sequencing for noninvasive prenatal, cancer, and transplantation assessments. Proc. Natl Acad. Sci. USA 112, E5503–E5512 (2015). In this study the relative contributions of different tissues to the total cfDNA pool were analysed using methylation deconvolution.
  68. Lui, Y. Y. N. et al. Predominant hematopoietic origin of cell-free DNA in plasma and serum after sex-mismatched bone marrow transplantation. Clin. Chem. 48, 421–427 (2002).
  69. Tug, S. et al. Exercise-induced increases in cell free DNA in human plasma originate predominantly from cells of the haematopoietic lineage. Exerc. Immunol. Rev. 21, 164–173 (2015).
  70. To, E. W. H. et al. Rapid clearance of plasma Epstein–Barr virus DNA after surgical treatment of nasopharyngeal carcinoma. Clin. Cancer Res. 9, 3254–3259 (2003).
  71. Lo, Y. M. et al. Rapid clearance of fetal DNA from maternal plasma. Am. J. Hum. Genet. 64, 218–224 (1999).
  72. Wang, Y., Chen, M., Xiao, N. & Liu, H. Evaluation and comparison of in vitro degradation kinetics of DNA in serum, urine and saliva: a qualitative study. Gene 590, 142–148 (2016).
  73. Tamkovich, S. N. et al. Circulating DNA and DNase activity in human blood. Ann. NY Acad. Sci. 1075, 191–196 (2006).
  74. Tsui, N. B. Y. et al. High resolution size analysis of fetal DNA in the urine of pregnant women by paired-end massively parallel sequencing. PLoS ONE 7, e48319 (2012).
  75. Chused, T. M., Steinberg, A. D. & Talal, N. The clearance and localization of nucleic acids by New Zealand and normal mice. Clin. Exp. Immunol. 12, 465–476 (1972).
  76. Giacona, M. B. et al. Cell-free DNA in human blood plasma: length measurements in patients with pancreatic cancer and healthy controls. Pancreas 17, 89–97 (1998).
  77. Lo, Y. M. D. et al. Maternal plasma DNA sequencing reveals the genome-wide genetic and mutational profile of the fetus. Sci. Transl Med. 2, 61ra91 (2010).
  78. Thierry, A. R. et al. Origin and quantification of circulating DNA in mice with human colorectal cancer xenografts. Nucleic Acids Res.38, 6159–6175 (2010).
  79. Nadano, D., Yasuda, T. & Kishi, K. Measurement of deoxyribonuclease I activity in human tissues and body fluids by a single radial enzyme-diffusion method. Clin. Chem. 39, 448–452 (1993).
  80. Mouliere, F. et al. Circulating cell-free DNA from colorectal cancer patients may reveal high KRAS or BRAF mutation load. Transl Oncol. 6, 319–328 (2013).
  81. Underhill, H. R. et al. Fragment length of circulating tumor DNA. PLoS Genet. 12, 426–437 (2016).
  82. Gorges, T. M. et al. Cancer therapy monitoring in xenografts by quantitative analysis of circulating tumor DNA. Biomarkers 17, 498–506 (2012).
  83. Zheng, Y. W. L. et al. Nonhematopoietically derived DNA is shorter than hematopoietically derived DNA in plasma: a transplantation model. Clin. Chem. 58, 549–558 (2012).
  84. Jiang, P. & Lo, Y. M. D. The long and short of circulating cell-free DNA and the ins and outs of molecular diagnostics. Trends Genet.32, 360–371 (2016).
  85. Cheng, S. H. et al. Noninvasive prenatal testing by nanopore sequencing of maternal plasma DNA: feasibility assessment. Clin. Chem. 61, 1305–1306 (2015).
  86. Breitbach, S. et al. Direct quantification of cell-free, circulating DNA from unpurified plasma. PLoS ONE 9, e87838 (2014).
  87. Beránek, M. et al. Carrier molecules and extraction of circulating tumor DNA for next generation sequencing in colorectal cancer. Acta Medica (Hradec Kralove) 59, 54–58 (2016).
  88. Gansauge, M.-T. & Meyer, M. Single-stranded DNA library preparation for the sequencing of ancient or damaged DNA. Nat. Protoc. 8, 737–748 (2013).
  89. Burnham, P. et al. Single-stranded DNA library preparation uncovers the origin and diversity of ultrashort cell-free DNA in plasma. Sci. Rep. 6, 27859 (2016).
  90. Kostyuk, S. V. et al. Fragments of cell-free DNA increase transcription in human mesenchymal stem cells, activate TLR-dependent signal pathway, and suppress apoptosis. Biochem. (Mosc.) Suppl. Ser. B Biomed. Chem. 6, 68–74 (2012).
  91. Nishimoto, S. et al. Obesity-induced DNA released from adipocytes stimulates chronic adipose tissue inflammation and insulin resistance. Sci. Adv. 2, e1501332 (2016).
  92. Dvor˘áková, M. et al. DNA released by leukemic cells contributes to the disruption of the bone marrow microenvironment. Oncogene 32, 5201–5209 (2013).
  93. Gartler, S. M. Cellular uptake of deoxyribonucleic acid by human tissue culture cells. Nature 184, 1505–1506 (1959).
  94. García-Olmo, D. C. et al. Cell-free nucleic acids circulating in the plasma of colorectal cancer patients induce the oncogenic transformation of susceptible cultured cells. Cancer Res. 70, 560–567 (2010).
  95. Ju, Y. S. et al. Frequent somatic transfer of mitochondrial DNA into the nuclear genome of human cancer cells. Genome Res. 25, 814–824 (2015).
  96. Wu, D. Y., Ugozzoli, L., Pal, B. K. & Wallace, R. B. Allele-specific enzymatic amplification of beta-globin genomic DNA for diagnosis of sickle cell anemia. Proc. Natl Acad. Sci. USA 86, 2757–2760 (1989).
  97. Newton, C. R. et al. Analysis of any point mutation in DNA. The amplification refractory mutation system (ARMS). Nucleic Acids Res. 17, 2503–2516 (1989).
  98. QIAGEN. Therascreen EGFR Plasma RGQ PCR Kit Handbook — Version 2. https://www.qiagen.com/gb/resources/resourcedetail?id=eb32e329-3422-4eda-b3d6-e44ed787002a&lang=en (2014).
  99. US Food & Drug Administration. Premarket approval P150044 — Cobas EGFR MUTATION TEST V2. FDAhttp://www.accessdata.fda.gov/scripts/cdrh/cfdocs/cfpma/pma.cfm?id=P150044 (2016).
  100. Schiavon, G. et al. Analysis of ESR1 mutation in circulating tumor DNA demonstrates evolution during therapy for metastatic breast cancer. Sci. Transl Med. 7, 313ra182 (2015).
  101. Taly, V. et al. Multiplex picodroplet digital PCR to detect KRASmutations in circulating DNA from the plasma of colorectal cancer patients. Clin. Chem. 59, 1722–1731 (2013).
  102. Sacher, A. G. et al. Prospective validation of rapid plasma genotyping for the detection of EGFR and KRAS mutations in advanced lung cancer. JAMA Oncol. 2, 1014–1022 (2016).
  103. Yung, T. K. F. et al. Single-molecule detection of epidermal growth factor receptor mutations in plasma by microfluidics digital PCR in non-small cell lung cancer patients. Clin. Cancer Res. 15, 2076–2084 (2009).
  104. Mosko, M. J. et al. Ultrasensitive detection of multiplexed somatic mutations using MALDI-TOF mass spectrometry. J. Mol. Diagn. 18, 23–31 (2016).
  105. Marziali, A., Pel, J., Bizzotto, D. & Whitehead, L. A. Novel electrophoresis mechanism based on synchronous alternating drag perturbation. Electrophoresis 26, 82–90 (2005).
  106. Thompson, J. D., Shibahara, G., Rajan, S., Pel, J. & Marziali, A.Winnowing DNA for rare sequences: highly specific sequence and methylation based enrichment. PLoS ONE 7, e31597 (2012).
  107. Song, C. et al. Elimination of unaltered DNA in mixed clinical samples via nuclease-assisted minor-allele enrichment. Nucleic Acids Res. 44, 1–11 (2016).
  108. Li, J. et al. Replacing PCR with COLD-PCR enriches variant DNA sequences and redefines the sensitivity of genetic testing. Nat. Med. 14, 579–584 (2008).
  109. Guha, M., Castellanos-Rizaldos, E., Liu, P., Mamon, H. & Makrigiorgos, G. M. Differential strand separation at critical temperature: a minimally disruptive enrichment method for low-abundance unknown DNA mutations. Nucleic Acids Res. 41, e50 (2013).
  110. Newman, A. M. et al. An ultrasensitive method for quantitating circulating tumor DNA with broad patient coverage. Nat. Med. 20, 548–554 (2014).
  111. Lanman, R. B. et al. Analytical and clinical validation of a digital sequencing panel for quantitative, highly accurate evaluation of cell-free circulating tumor DNA. PLoS ONE 10, e0140712 (2015).
  112. Frenel, J.-S. et al. Serial next generation sequencing of circulating cell free DNA evaluating tumour clone response to molecularly targeted drug administration. Clin. Cancer Res. 21, 4586–4596 (2015).
  113. Narayan, A. et al. Ultrasensitive measurement of hotspot mutations in tumor DNA in blood using error-suppressed multiplexed deep sequencing. Cancer Res. 72, 3492–3498 (2012).
  114. Newman, A. M. et al. Integrated digital error suppression for improved detection of circulating tumor DNA. Nat. Biotechnol. 34, 547–555 (2016). This paper describes the use of an error suppression method for capture-enrichment sequencing that uses a combination of in silico elimination of stereotyped background artefacts and molecular barcoding.
  115. Kinde, I., Wu, J., Papadopoulos, N., Kinzler, K. W. & Vogelstein, B.Detection and quantification of rare mutations with massively parallel sequencing. Proc. Natl Acad. Sci. USA 108, 9530–9535 (2011).
  116. Rosenfeld, N., Forshew, T., Marass, F. & Murtaza, M. A method for detecting a genetic variant. World Intellectual Property Organization patent WO2016009224A1 (2016).
  117. Gale, D. et al. Analytical performance and validation of an enhanced TAm-Seq circulating tumor DNA sequencing assay. Inivata http://inivata.com/wp-content/uploads/2016/06/AACRposter2016.pdf (2016).
  118. Schwaederle, M. C. et al. Detection rate of actionable mutations in diverse cancers using a biopsy-free (blood) circulating tumor DNA assay. Oncotarget 33, 11004 (2015).
  119. Belic, J. et al. Rapid identification of plasma DNA samples with increased ctDNA levels by a modified FAST-SeqS approach. Clin. Chem. 61, 838–849 (2015).
  120. Kirkizlar, E. et al. Detection of clonal and subclonal copy-number variants in cell-free DNA from patients with breast cancer using a massively multiplexed PCR methodology. Transl Oncol. 8, 407–416 (2015).
  121. Parkinson, C. A. et al. Exploratory analysis of TP53 mutations in circulating tumour DNA as biomarkers of treatment response for patients with relapsed high-grade serous ovarian carcinoma: a retrospective study. PLoS Med. 13, e1002198 (2016).
  122. Thierry, A. R. et al. Clinical validation of the detection of KRAS and BRAF mutations from circulating tumor DNA. Nat. Med. 20, 430–435 (2014). This blinded prospective multicentre study compared KRAS and BRAF mutation status in tumours and plasma using an allele-specific quantitative PCR approach.
  123. Kamat, A. A. et al. Circulating cell-free DNA: a novel biomarker for response to therapy in ovarian carcinoma. Cancer Biol. Ther. 5, 1369–1374 (2006).
  124. Bettegowda, C. et al. Detection of circulating tumor DNA in early- and late-stage human malignancies. Sci. Transl Med. 6, 224ra24 (2014). In this study, a comprehensive quantitative analysis finds varying levels of ctDNA across patients with distinct cancer types and provides indicative ranges of ctDNA levels across the stages of disease.
  125. Nygaard, A. D., Garm Spindler, K.-L., Pallisgaard, N., Andersen, R. F. & Jakobsen, A. The prognostic value of KRAS mutated plasma DNA in advanced non-small cell lung cancer. Lung Cancer 79, 312–317 (2013).
  126. Wang, S. et al. Potential clinical significance of a plasma-based KRAS mutation analysis in patients with advanced non-small cell lung cancer. Clin. Cancer Res. 16, 1324–1330 (2010).
  127. Gautschi, O. et al. Origin and prognostic value of circulating KRASmutations in lung cancer patients. Cancer Lett. 254, 265–273 (2007).
  128. Santiago-Walker, A. et al. Correlation of BRAF mutation status in circulating-free DNA and tumor and association with clinical outcome across four BRAFi and MEKi clinical trials. Clin. Cancer Res. 22, 567–574 (2016).
  129. Gray, E. S. et al. Circulating tumor DNA to monitor treatment response and detect acquired resistance in patients with metastatic melanoma. Oncotarget 6, 42008–42018 (2015).
  130. Lipson, E. J. et al. Circulating tumor DNA analysis as a real-time method for monitoring tumor burden in melanoma patients undergoing treatment with immune checkpoint blockade. J. Immunother. Cancer 2, 42 (2014).
  131. Oshiro, C. et al. PIK3CA mutations in serum DNA are predictive of recurrence in primary breast cancer patients. Breast Cancer Res. Treat. 150, 299–307 (2015).
  132. Scherer, F. et al. Distinct biological subtypes and patterns of genome evolution in lymphoma revealed by circulating tumor DNA. Sci. Transl Med. 8, 364ra155 (2016).
  133. Office for National Statistics. Cancer survival by stage at diagnosis for England (experimental statistics): adults diagnosed 2012, 2013 and 2014 and followed up to 2015. ONShttps://www.ons.gov.uk/peoplepopulationandcommunity/healthandsocialcare/conditionsanddiseases/bulletins/cancersurvivalbystageatdiagnosisforenglandexperimentalstatistics/adultsdiagnosed20122013and2014andfollowedupto2015 (2016).
  134. Gormally, E. et al. TP53 and KRAS2 mutations in plasma DNA of healthy subjects and subsequent cancer occurrence: a prospective study. Cancer Res. 66, 6871–6876 (2006).
  135. Amant, F. et al. Presymptomatic identification of cancers in pregnant women during noninvasive prenatal testing. JAMA Oncol. 1, 814 (2015).
  136. Bianchi, D. W. et al. Noninvasive prenatal testing and incidental detection of occult maternal malignancies. JAMA 314, 162 (2015).
  137. Beaver, J. A. et al. Detection of cancer DNA in plasma of patients with early-stage breast cancer. Clin. Cancer Res. 20, 2643–2650 (2014).
  138. Cohen, P. A. et al. Abnormal plasma DNA profiles in early ovarian cancer using a non-invasive prenatal testing platform: implications for cancer screening. BMC Med. 14, 126 (2016).
  139. Gorges, T. M. et al. Enumeration and molecular characterization of tumor cells in lung cancer patients using a novel in vivo device for capturing circulating tumor cells. Clin. Cancer Res. 22, 2197–2206 (2016).
  140. Kinde, I. et al. Evaluation of DNA from the Papanicolaou test to detect ovarian and endometrial cancers. Sci. Transl Med. 5, 167ra4 (2013).
  141. Nair, N. et al. Genomic analysis of uterine lavage fluid detects early endometrial cancers and reveals a prevalent landscape of driver mutations in women without histopathologic evidence of cancer: a prospective cross-sectional study. PLoS Med. 13, e1002206 (2016).
  142. Ross-Innes, C. S. et al. Risk stratification of Barrett's oesophagus using a non-endoscopic sampling method coupled with a biomarker panel: a cohort study. Lancet Gastroenterol. Hepatol. 2, 23–31 (2016).
  143. Chan, K. C. A. et al. Early detection of nasopharyngeal carcinoma by plasma Epstein–Barr virus DNA analysis in a surveillance program. Cancer 119, 1838–1844 (2013).
  144. Guerrero-Preston, R. et al. Molecular triage of premalignant lesions in liquid-based cervical cytology and circulating cell-free DNA from urine, using a panel of methylated human papilloma virus and host genes. Cancer Prev. Res. 9, 915–924 (2016).
  145. Martincorena, I. et al. High burden and pervasive positive selection of somatic mutations in normal human skin. Science348, 880–886 (2015).
  146. Genovese, G. et al. Clonal hematopoiesis and blood-cancer risk inferred from blood DNA sequence. N. Engl. J. Med. 371, 2477–2487 (2014).
  147. Fernandez-Cuesta, L. et al. Identification of circulating tumor DNA for the early detection of small-cell lung cancer. eBioMedicine 10, 6–12 (2016).
  148. Rosenfeld, N. et al. MicroRNAs accurately identify cancer tissue origin. Nat. Biotechnol. 26, 462–469 (2008).
  149. Olsen, J. A. et al. A minimally-invasive blood-derived biomarker of oligodendrocyte cell-loss in multiple sclerosis. eBioMedicine 10, 227–235 (2016).
  150. Gerlinger, M. et al. Intratumor heterogeneity and branched evolution revealed by multiregion sequencing. N. Engl. J. Med.366, 883–892 (2012).
  151. Yates, L. R. et al. Subclonal diversification of primary breast cancer revealed by multiregion sequencing. Nat. Med. 21, 751–759 (2015).
  152. Turajlic, S. & Swanton, C. Metastasis as an evolutionary process. Science 352, 169–175 (2016).
  153. Piotrowska, Z. et al. Heterogeneity underlies the emergence of EGFRT790 wild-type clones following treatment of T790M-positive cancers with a third-generation EGFR inhibitor. Cancer Discov. 5, 37–54 (2015).
  154. De Mattos-Arruda, L. et al. Establishing the origin of metastatic deposits in the setting of multiple primary malignancies: the role of massively parallel sequencing. Mol. Oncol. 8, 150–158 (2014).
  155. Chan, K. C. A. et al. Noninvasive detection of cancer-associated genome-wide hypomethylation and copy number aberrations by plasma DNA bisulfite sequencing. Proc. Natl Acad. Sci. USA 110, 18761–18768 (2013).
  156. Kuo, Y.-B., Chen, J. S., Fand, C. W., Li, Y. S. & Chan, E. C.Comparison of KRAS mutation analysis of primary tumors and matched circulating cell-free DNA in plasmas of patients with colorectal cancer. Clin. Chim. Acta 433, 284–289 (2014).
  157. de Bruin, E. C. et al. Spatial and temporal diversity in genomic instability processes defines lung cancer evolution. Science 346, 251–256 (2014).
  158. Zhang, J. et al. Intratumor heterogeneity in localized lung adenocarcinomas delineated by multiregion sequencing. Science346, 256–259 (2014).
  159. Oxnard, G. R. et al. Association between plasma genotyping and outcomes of treatment with osimertinib (AZD9291) in advanced non-small-cell lung cancer. J. Clin. Oncol. 34, 3375–3382 (2016). In this retrospective analysis, patients with NSCLC who were positive for EGFRT790M in plasma show outcomes with osimertinib that were equivalent to those of patients classified as positive by a tissue-based assay, supporting the use of plasma analysis to avoid tumour biopsies in patients who are positive for EGFRT790M in plasma.
  160. Siravegna, G. et al. Clonal evolution and resistance to EGFR blockade in the blood of colorectal cancer patients. Nat. Med. 21, 795–801 (2015).
  161. Tie, J. et al. Circulating tumor DNA as an early marker of therapeutic response in patients with metastatic colorectal cancer. Ann. Oncol. 26, 1715–1722 (2015).
  162. Chabon, J. J. et al. Circulating tumour DNA profiling reveals heterogeneity of EGFR inhibitor resistance mechanisms in lung cancer patients. Nat. Commun. 7, 11815 (2016). In this study, ctDNA analysis by screening of a large panel of genes identifies multiple heterogeneous resistance mechanisms in patients with NSCLC treated with the EGFR inhibitor rociletinib.
  163. Karachaliou, N. et al. Association of EGFR L858R mutation in circulating free DNA with survival in the EURTAC trial. JAMA Oncol. 1, 149–157 (2015).
  164. Reck, M. et al. Circulating free tumor-derived DNA (ctDNA) determination of EGFR mutation status in European and Japanese patients with advanced NSCLC: the ASSESS study. J. Thorac. Oncol. 11, 1682–1689 (2016).
  165. Han, B. et al. Determining the prevalence of EGFR mutations in Asian and Russian patients (pts) with advanced non-small-cell lung cancer (aNSCLC) of adenocarcinoma (ADC) and non-ADC histology: IGNITE study. Ann. Oncol. 26, i29–i44 (2015).
  166. Douillard, J.-Y. et al. Gefitinib treatment in EGFR mutated Caucasian NSCLC. J. Thorac. Oncol. 9, 1345–1353 (2014).
  167. Normanno, N., Denis, M. G., Thress, K. S., Ratcliffe, M. & Reck, M.Guide to detecting epidermal growth factor receptor (EGFR) mutations in ctDNA of patients with advanced non-small-cell lung cancer. Oncotargethttp://dx.doi.org/10.18632/oncotarget.13915 (2016).
  168. Douillard, J.-Y. et al. First-line gefitinib in Caucasian EGFRmutation-positive NSCLC patients: a phase-IV, open-label, single-arm study. Br. J. Cancer 110, 55–62 (2014).
  169. Mok, T. S. et al. Osimertinib or platinum–pemetrexed in EGFRT790M-positive lung cancer. N. Engl. J. Med.http://dx.doi.org/10.1056/NEJMoa1612674 (2016).
  170. Sorensen, B. S. et al. Monitoring of epidermal growth factor receptor tyrosine kinase inhibitor-sensitizing and resistance mutations in the plasma DNA of patients with advanced non-small cell lung cancer during treatment with erlotinib. Cancer120, 3896–3901 (2014).
  171. Oxnard, G. R. et al. Noninvasive detection of response and resistance in EGFR-mutant lung cancer using quantitative next-generation genotyping of cell-free plasma DNA. Clin. Cancer Res.20, 1698–705 (2014).
  172. Thress, K. S. et al. EGFR mutation detection in ctDNA from NSCLC patient plasma: a cross-platform comparison of leading technologies to support the clinical development of AZD9291. Lung Cancer 90, 509–515 (2015).
  173. Remon, J. et al. Osimertinib benefit in EGFR-mutant NSCLC patients with T790M-mutation detected by circulating tumour DNA. Annc. Oncol. doi.dx.org:/10.1093/annonc/mdx017 (2017).
  174. European Medicines Agency. Iressa: public assessment report — product information. EMAhttp://www.ema.europa.eu/docs/en_GB/document_library/EPAR_-_Product_Information/human/001016/WC500036358.pdf(2016).
  175. European Medicines Agency. Tagrisso: public assessment report — product information. EMAhttp://www.ema.europa.eu/docs/en_GB/document_library/EPAR_-_Product_Information/human/004124/WC500202022.pdf(2016).
  176. US National Library of Medicine. ClinicalTrials.govhttps://clinicaltrials.gov/ct2/show/NCT02693535 (2017).
  177. Schwaederle, M. et al. Use of liquid biopsies in clinical oncology: pilot experience in 168 patients. Clin. Cancer Res. 22, 5497–5505 (2016).
  178. US National Library of Medicine. ClinicalTrials.govhttps://clinicaltrials.gov/ct2/show/NCT02140463 (2015).
  179. Remon, J. et al. Liquid biopsies for molecular profiling of mutations in non-small cell lung cancer (NSCLC) patients lacking tissue samples. Inivata http://inivata.com/wp-content/uploads/2016/09/MAP2016posterFINAL.pdf (2016).
  180. Institut Gustave Roussy. The prospective MOSCATO 01 trial demonstrates that molecular “portraits” improve outcome of patients with metastatic cancer. PRNewswirehttp://www.prnewswire.com/news-releases/the-prospective-moscato-01-trial-demonstrates-that-molecular-portraits-improve-outcome-of-patients-with-metastatic-cancer-594507371.html (2016).
  181. Pishvaian, M. J. et al. A pilot study evaluating concordance between blood-based and patient-matched tumor molecular testing within pancreatic cancer patients participating in the Know Your Tumor (KYT) initiative. Oncotargethttp://dx.doi.org/10.18632/oncotarget.13225 (2016).
  182. Heitzer, E. et al. Establishment of tumor-specific copy number alterations from plasma DNA of patients with cancer. Int. J. Cancer 133, 346–356 (2013).
  183. Mohan, S. et al. Changes in colorectal carcinoma genomes under anti-EGFR therapy identified by whole-genome plasma DNA sequencing. PLoS Genet. 10, e1004271 (2014).
  184. Kinde, I., Papadopoulos, N., Kinzler, K. W. & Vogelstein, B. FAST-SeqS: a simple and efficient method for the detection of aneuploidy by massively parallel sequencing. PLoS ONE 7, e41162 (2012).
  185. Abdueva, D. et al. Detection, frequency and actionability of recurrent copy number gains detected by non-invasive liquid biopsy of 3,942 lung and breast cancer samples. J. Clin. Oncol. 34, abstr. 11541 (2016).
  186. Jiang, P. et al. Lengthening and shortening of plasma DNA in hepatocellular carcinoma patients. Proc. Natl Acad. Sci. USA 112, E1317–E1325 (2015). In this study, a genome-wide analysis of plasma DNA from patients with hepatocellular carcinoma finds that short molecules of cfDNA are enriched for tumour-associated copy number alterations.
  187. Romanel, A. et al. Plasma AR and abiraterone-resistant prostate cancer. Sci. Transl Med. 7, 312re10 (2015).
  188. Paweletz, C. P. et al. Bias-corrected targeted next-generation sequencing for rapid, multiplexed detection of actionable alterations in cell-free DNA from advanced lung cancer patients. Clin. Cancer Res. 22, 915–922 (2016).
  189. Schreuer, M. et al. Quantitative assessment of BRAF V600 mutant circulating cell-free tumor DNA as a tool for therapeutic monitoring in metastatic melanoma patients treated with BRAF/MEK inhibitors. J. Transl Med. 14, 95 (2016).
  190. Marchetti, A. et al. Early prediction of response to tyrosine kinase inhibitors by quantification of EGFR mutations in plasma of NSCLC patients. J. Thorac. Oncol. 10, 1437–1443 (2015).
  191. Xi, L. et al. Circulating tumor DNA as an early indicator of response to T-cell transfer immunotherapy in metastatic melanoma. Clin. Cancer Res. 22, 5480–5486 (2016).
  192. Xin, Y. et al. Method for detecting BCR and TCR immune repertoire in blood plasma cfDNA. Chinese patent CN105087789A (2015).
  193. Tie, J. et al. Circulating tumor DNA analysis detects minimal residual disease and predicts recurrence in patients with stage II colon cancer. Sci. Transl Med. 8, 346ra92 (2016). This prospective cohort study of 230 patients with resected stage II colon cancer demonstrates that post-operative ctDNA detection predicts poor recurrence-free survival.
  194. Garcia-Murillas, I. et al. Mutation tracking in circulating tumor DNA predicts relapse in early breast cancer. Sci. Transl Med. 7, 302ra133 (2015). This study demonstrates the use of ctDNA for the detection of MRD in early-stage breast cancer following treatment with curative intent.
  195. Reinert, T. et al. Analysis of circulating tumour DNA to monitor disease burden following colorectal cancer surgery. Gut 65, 625–634 (2015).
  196. Olsson, E. et al. Serial monitoring of circulating tumor DNA in patients with primary breast cancer for detection of occult metastatic disease. EMBO Mol. Med. 7, 1034–1047 (2015).
  197. Leary, R. J. et al. Development of personalized tumor biomarkers using massively parallel sequencing. Sci. Transl Med. 2, 20ra14 (2010).
  198. Mcbride, D. J. et al. Use of cancer-specific genomic rearrangements to quantify disease burden in plasma from patients with solid tumors. Genes Chromosomes Cancer 49, 1062–1069 (2011).
  199. Russo, M. et al. Acquired resistance to the TRK inhibitor entrectinib in colorectal cancer. Cancer Discov. 6, 36–44 (2016).
  200. Misale, S. et al. Emergence of KRAS mutations and acquired resistance to anti-EGFR therapy in colorectal cancer. Nature 486, 532–536 (2012). In this study, serial ctDNA analysis identifies KRAS-mutant alleles in the plasma of cetuximab-treated patients 10 months before disease relapse was identified by imaging.
  201. Diaz, L. A. J. et al. The molecular evolution of acquired resistance to targeted EGFR blockade in colorectal cancers. Nature 486, 537–540 (2012).
  202. Ilié, M. & Hofman, P. Pros: can tissue biopsy be replaced by liquid biopsy? Transl Lung Cancer Res. 5, 420–423 (2016).
  203. Thress, K. S. et al. Acquired EGFR C797S mutation mediates resistance to AZD9291 in non-small cell lung cancer harboring EGFR T790M. Nat. Med. 21, 560–562 (2015). In this study, serial ctDNA monitoring of patients with lung cancer who were treated with AZD9291, an EGFR inhibitor, revealed a diversity of resistance mechanisms to therapy.
  204. Russo, M. et al. Tumor heterogeneity and lesion-specific response to targeted therapy in colorectal cancer. Cancer Discov. 6, 147–153 (2015).
  205. Morelli, M. P. et al. Characterizing the patterns of clonal selection in circulating tumor DNA from patients with colorectal cancer refractory to anti-EGFR treatment. Ann. Oncol. 26, 731–736 (2015).
  206. Gremel, G. et al. Distinct sub-clonal tumour responses to therapy revealed by circulating cell-free DNA. Ann. Oncol. 27, 1959–1965 (2016).
  207. Ulz, P. et al. Whole-genome plasma sequencing reveals focal amplifications as a driving force in metastatic prostate cancer. Nat. Commun. 7, 12008 (2016).
  208. European Medicines Agency. Draft guideline on good pharmacogenomic practice. EMA (2016).
  209. US National Library of Medicine. ClinicalTrials.govhttps://clinicaltrials.gov/ct2/show/NCT02284633 (2015).
  210. US National Library of Medicine. ClinicalTrials.govhttps://clinicaltrials.gov/ct2/show/NCT02743910 (2016).
  211. International Standard Randomised Controlled Trials Number Registry. isrctn.com http://www.isrctn.com/ISRCTN16945804(2016).
  212. Naik, S. H., Schumacher, T. N. & Perie, L. Cellular barcoding: a technical appraisal. Exp. Hematol. 42, 598–608 (2014).
  213. El Messaoudi, S. et al. Circulating DNA as a strong multimarker prognostic tool for metastatic colorectal cancer patient management care. Clin. Cancer Res. 22, 3067–3077 (2016).
  214. Spindler, K. L. G., Pallisgaard, N., Andersen, R. F. & Jakobsen, A.Changes in mutational status during third-line treatment for metastatic colorectal cancer — results of consecutive measurement of cell free DNA, KRAS and BRAF in the plasma. Int. J. Cancer 135, 2215–2222 (2014).
  215. Schwarzenbach, H., Hoon, D. S. B. & Pantel, K. Cell-free nucleic acids as biomarkers in cancer patients. Nat. Rev. Cancer 11, 426–437 (2011).
  216. Sullivan, R. J. et al. Plasma-based monitoring of BRAF mutations during therapy for malignant melanoma using combined exosomal RNA and cell-free DNA analysis. J. Clin. Oncol. 33, 9017 (2015).
  217. Ulz, P. et al. Inferring expressed genes by whole-genome sequencing of plasma DNA. Nat. Genet. 48, 1273–1278 (2016).
  218. Shao, H. et al. Chip-based analysis of exosomal mRNA mediating drug resistance in glioblastoma. Nat. Commun. 6, 6999 (2015).
  219. Alix-Panabières, C. & Pantel, K. Clinical applications of circulating tumor cells and circulating tumor DNA as liquid biopsy. Cancer Discov. 6, 479–491 (2016).
  220. Best, M. G. et al. RNA-seq of tumor-educated platelets enables blood-based pan-cancer, multiclass, and molecular pathway cancer diagnostics. Cancer Cell 28, 666–676 (2015).
  221. Gormally, E., Caboux, E., Vineis, P. & Hainaut, P. Circulating free DNA in plasma or serum as biomarker of carcinogenesis: practical aspects and biological significance. Mutat. Res. 635, 105–117 (2007).
  222. Webb, S. The cancer bloodhounds. Nat. Biotechnol. 34, 1090–1094 (2016).
  223. Long-Mira, E., Washetine, K. & Hofman, P. Sense and nonsense in the process of accreditation of a pathology laboratory. Virchows Arch. 468, 43–49 (2016).
  224. Melchior, L. et al. Multi-center evaluation of the novel fully-automated PCR-based Idylla BRAF mutation test on formalin-fixed paraffin-embedded tissue of malignant melanoma. Exp. Mol. Pathol. 99, 485–491 (2015).
  225. Janku, F. et al. BRAF mutation testing in cell-free DNA from the plasma of patients with advanced cancers using a rapid, automated molecular diagnostics system. Mol. Cancer Ther. 15, 1397–1404 (2016).
  226. Norris, A. L., Workman, R. E., Fan, Y., Eshleman, J. R. & Timp, W.Nanopore sequencing detects structural variants in cancer. Cancer Biol. Ther. 17, 246–253 (2016).
  227. Quick, J. et al. Real-time, portable genome sequencing for Ebola surveillance. Nature 530, 228–232 (2016).
  228. Wei, S. & Williams, Z. Rapid short-read sequencing and aneuploidy detection using minION nanopore technology. Genetics 202, 37–44 (2016).
  229. Ip, C. L. C. et al. MinION Analysis and Reference Consortium: phase 1 data release and analysis. F1000Res. 4, 1075 (2015).
  230. Loose, M., Malla, S. & Stout, M. Real-time selective sequencing using nanopore technology. Nat. Methods 13, 751–754 (2016).
  231. Umetani, N., Hiramatsu, S. & Hoon, D. S. B. Higher amount of free circulating DNA in serum than in plasma is not mainly caused by contaminated extraneous DNA during separation. Ann. NY Acad. Sci. 1075, 299–307 (2006).
  232. Willems, M., Moshage, H., Nevens, F., Fevery, J. & Yap, S. H. Plasma collected from heparinized blood is not suitable for HCV-RNA detection by conventional RT-PCR assay. J. Virol. Methods 42, 127–130 (1993).
  233. Jung, M., Klotzek, S., Lewandowski, M., Fleischhacker, M. & Jung, K. Changes in concentration of DNA in serum and plasma during storage of blood samples. Clin. Chem. 49, 1028–1029 (2003).
  234. Parpart-Li, S. et al. The effect of preservative and temperature on the analysis of circulating tumor DNA. Clin. Cancer Res.http://dx.doi.org/10.1158/1078-0432.CCR-16-1691 (2016).
  235. El Messaoudi, S., Rolet, F., Mouliere, F. & Thierry, A. R. Circulating cell free DNA: preanalytical considerations. Clin. Chim. Acta 424, 222–230 (2013).
  236. Sherwood, J. L. et al. Optimised pre-analytical methods improve KRAS mutation detection in circulating tumour DNA (ctDNA) from patients with non-small cell lung cancer (NSCLC). PLoS ONE 11, 1–14 (2016).
  237. Toro, P. V. et al. Comparison of cell stabilizing blood collection tubes for circulating plasma tumor DNA. Clin. Biochem. 48, 993–998 (2015).
  238. Kang, Q. et al. Comparative analysis of circulating tumor DNA stability in K3EDTA, Streck and CellSave blood collection tubes. Clin. Biochem. 49, 1354–1360 (2016).
  239. Medina Diaz, I. et al. Performance of Streck cfDNA blood collection tubes for liquid biopsy testing. PLoS ONE 11, e0166354 (2016).
  240. Norton, S. E., Lechner, J. M., Williams, T. & Fernando, M. R. A stabilizing reagent prevents cell-free DNA contamination by cellular DNA in plasma during blood sample storage and shipping as determined by digital PCR. Clin. Biochem. 46, 1561–1565 (2013).
  241. Dietz, S. et al. Low input whole-exome sequencing to determine the representation of the tumor exome in circulating DNA of non-small cell lung cancer patients. PLoS ONE 11, e0161012 (2016).
  242. Schmitt, M. W. et al. Detection of ultra-rare mutations by next-generation sequencing. Proc. Natl Acad. Sci. USA 109, 14508–14513 (2012).
  243. Zonta, E., Garlan, F., Perez-Toralla, K. & Taly, V. Multiplex detection of rare mutations by picoliter droplet based digital PCR: sensitivity and specificity considerations. PLoS ONE 11, e0159094 (2016).
  244. Kidess, E. et al. Mutation profiling of tumor DNA from plasma and tumor tissue of colorectal cancer patients with a novel, high-sensitivity multiplexed mutation detection platform. Oncotarget6, 2549–2561 (2015).