Next Article in Journal
Functionalization of a Fully Integrated Electrophotonic Silicon Circuit for Biotin Sensing
Next Article in Special Issue
Interactions of Amino Group Functionalized Tetraphenylvinyl and DNA: A Label-Free “On-Off-On” Fluorescent Aptamer Sensor toward Ampicillin
Previous Article in Journal
Mental Stress Detection Using a Wearable In-Ear Plethysmography
Previous Article in Special Issue
Development of Gold-Nanoparticle-Based Lateral Flow Immunoassays for Rapid Detection of TB ESAT-6 and CFP-10
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Trends in Biosensing and Diagnostic Methods for Novel Cancer Biomarkers

by
Jagadeeswara Rao Bommi
1,
Shekher Kummari
2,
Kavitha Lakavath
2,
Reshmi A. Sukumaran
2,
Lakshmi R. Panicker
2,
Jean Louis Marty
3,* and
Kotagiri Yugender Goud
2,*
1
School of Medicine, Case Western Reserve University, Cleveland, OH 44106, USA
2
Department of Chemistry, Indian Institute of Technology Palakkad, Palakkad 678 557, Kerala, India
3
Université de Perpignan Via Domitia, 52 Avenue Paul Alduy, 66860 Perpignan, France
*
Authors to whom correspondence should be addressed.
Biosensors 2023, 13(3), 398; https://doi.org/10.3390/bios13030398
Submission received: 20 February 2023 / Revised: 13 March 2023 / Accepted: 15 March 2023 / Published: 18 March 2023
(This article belongs to the Special Issue Biosensing and Diagnosis)

Abstract

:
Cancer is one of the major public health issues in the world. It has become the second leading cause of death, with approximately 75% of cancer deaths transpiring in low- or middle-income countries. It causes a heavy global economic cost estimated at more than a trillion dollars per year. The most common cancers are breast, colon, rectum, prostate, and lung cancers. Many of these cancers can be treated effectively and cured if detected at the primary stage. Nowadays, around 50% of cancers are detected at late stages, leading to serious health complications and death. Early diagnosis of cancer diseases substantially increases the efficient treatment and high chances of survival. Biosensors are one of the potential screening methodologies useful in the early screening of cancer biomarkers. This review summarizes the recent findings about novel cancer biomarkers and their advantages over traditional biomarkers, and novel biosensing and diagnostic methods for them; thus, this review may be helpful in the early recognition and monitoring of treatment response of various human cancers.

1. Introduction

1.1. Cancer

One of the major life-threatening health issues in this world is cancer. It has become the second leading cause of death, around 75% of cancer deaths transpiring in low- to middle-income countries. It causes a heavy global economic cost estimated at more than trillion dollars per year [1,2,3]. In the course of cancer, cells can grow uncontrollably as well as expand to other parts of the organs in body. Cancer can occur from the transformation of normal cells into tumor cells. Tumor cells are classified as benign and malignant. Tumors that stick to their primary location without occupying distant parts of the body are called benign tumors; these are likely to grow slowly and are not problematic. Fibroids in the uterus are an example of benign tumors. Some of the benign tumors can change into malignant tumors (e.g., colon polyps); these can be removed surgically. Tumor cells that can grow uncontrollably and spread from their primary location to distant sites are called malignant tumors; these are cancerous (i.e., invade from other parts). Malignant tumors can rapidly spread to distant parts through blood or lymph stream; this process is called as metastasis. Omnipresent cancers are the primary reason of death in the patients with cancer. Most frequently metastasis can be found in brain, lungs, liver and bone [4].
Cancers can be grouped into different categories. Mostly, the group that the cancer belongs to is determined based on the type of cells or tissues it is producing. The following are some of the following important groups in this category.
Carcinoma: Most common type of cancer. It starts in epithelium, which is the tissue that lines or covers internal organs and passageways in the human body as well as skin. It appears in the form of tumors, which can form on lungs, breasts, skin, pancreas, kidneys, prostate, colon, etc. There are various subtypes, including adeno-carcinoma, SCC (squamous cell carcinoma), ductal carcinoma, TCC (transitional cell carcinoma) and (BCC) basal cell carcinoma.
Sarcoma: This is a type of cancer that arises in connective tissue and/or supportive tissue such as muscle, cartilage, bone, blood vessels or fat.
Leukemia: This type is also known as blood cancer/cancer of WBC (white blood cells). It can appear in tissues which can produce blood cells (e.g., bone marrow).
Myeloma and Lymphoma: This type of cancer occurs in the cells present in the immune system (myeloma: Starts in plasma cells present inside bone marrow; lymphoma: Starts in lymphatic system such as spleen, lymph nodes).
Spinal cord and brain cancers: These cancers occur in spinal cord and brain, and are also known as central nervous system cancers.
Multiple factors can generate cancer. These cancer-causing agents are termed carcinogens (Figure 1). These could be genetic factors (mutated genes pass from parents to children and cause various cancers, e.g., BRCA1 and BRCA2) and external factors (physical carcinogens: Ionizing and UV radiation; tobacco smoke, alcohol, aflatoxin and arsenic are the examples of chemical carcinogens; certain viruses, bacteria, fungus and some type parasites are considered biological carcinogens; life style factors: lack of exercise, smoking, nutrient imbalance) [5] (https://www.who.int/news-room/fact-sheets/detail/cancer) (accessed on 11 November 2022).

1.2. Importance of Cancer Diagnostics

Early diagnosis of cancer has held great assurance and intuitive interest in the medical community for over 100 years [6,7]. However, delayed identification and imperfect prognosis are major causes for very low survival rate in various patients with cancer [8]. Many cancers can be treated effectively and cured if detected at the primary stage. Patients diagnosed with cancer at earlier stages are not only likely to survive, but, importantly, they also experience better care, minimal treatment morbidity, and enhanced quality of life when compared with patients with late diagnosis [9,10]. Improving earlier detection of cancers is a complex and multifaceted process. Few recent patient behaviors could be helpful for the earlier diagnosis of various cancers. One is attending cancer screening procedures, which aim to identify the cancer in asymptomatic condition (e.g., mammography to detect breast cancer), and the other is promptly introducing the patient with any potential cancer symptoms to primary care providers [11]. The necessity of symptomatic presentation to primary care is emphasized by improving public perception of the early indications of cancer. In recent years, cancer awareness is also increasing in developing and lesser-income countries. Raising public awareness, promoting visits to a healthcare provider, and diagnosing a patient at an earlier stage of the cancer can provide the opportunity to offer better treatment [12,13,14]. Nowadays, around 50% of cancers are detected at late stages, leading to serious health complications and death [12]. Early diagnosis of cancer significantly increases the efficient treatment and high chances of survival.

1.3. Traditional Screening Techniques

Cancer screening techniques have significantly promoted the decline of morbidity as well as mortality of cancer. Methods to enhance the choice of candidates for cancer screening, to acknowledge the biological foundation of cancer formation and development of novel technologies for the tumor screening would allow for advancement in the tumor screening process over time [15,16]. Screening in healthy as well as high-risk populations provides the chance to detect cancer at an early stage and with an expanded chance for treatment. Nowadays, screening techniques play a crucial role in detecting specific cancer types. However, each screening technique has some limitations, and upgraded screening techniques are very much essential to identify cancer early in healthy populations [17,18,19]. Identifying tumors at their primary stage often delivers the finest probability for a cure, which is why it is always crucial for the general population to talk with their health care providers about the types of screening that might be required. Research studies explain that early screening process can save many lives by identifying cancer in primary stages. Various medical communities and patient advocacy groups have drafted guidelines for cancer screening. In general, primary health care providers use the following approaches for tumor diagnosis [20,21,22].
Physical exam: During this exam, a provider can perform a scan to detect any abnormalities in body, such as skin color change, formation of lumps or abnormal growth of a tissue or an organ. This may provide the indication of the tumor.
Laboratory tests: Some lab procedures such as blood and/or urine tests can help to detect any cancer-related anomalies in body. For example, simple blood work called complete blood cell count might disclose any abnormal number or type of WBC (white blood cells) in leukemia patients.
Genetic tests: These are also lab-based tests. In these procedures, cells or tissues are examined to observe any modifications in their genes and/or chromosomes. Any of those differences might be an indication that a person is at risk of encountering a particular problem or condition.
Imaging tests: Imaging tests can examine bones or any interior organs through noninvasive method. Common cancer diagnosing imaging methods are a CT (computerized tomography) scan, X-ray, PET (positron emission tomography) scan, ultrasound, and MRI (magnetic resonance imaging) [23,24].
Biopsy: During this procedure, samples of cells are collected for examining in the laboratory. Collection of a sample can be performed by a variety of methods. The suitable type of biopsy is based on the cancer type and cancer location in body. In many cases, biopsy is the exclusive method for conclusive cancer diagnosis [25,26,27].

1.3.1. Advantages of Traditional Cancer Screening

  • It can help to detect the cancer before it spreads, when it is easier to treat.
  • It can provide an advantage of early detection, which might lead to lesser recovery time and no intense treatment.
  • It can provide a better chance of survival.
  • It offers flexibility to start early treatment before symptoms appear.
  • It can also reassure a person if the screening result is normal [28,29,30].

1.3.2. Limitations of Traditional Cancer Screening

  • Sometimes a false-positive test result suggests a cancer-positive status, even though no cancer is present.
  • Sometimes false-negative test results may not detect cancer, even though it is present.
  • Some screening tests might lead to more detection tests and procedures that can be painful.
  • Over-diagnosis causes needless anxiety.
  • Some screenings might cause potential issues (e.g., colon cancer screening may cause tear in colon lining).
  • These screening methods are high cost.
  • Test availability is limited to metro cities only [28,29,30].
There is a strong necessity to develop rapid and affordable screening methods for cancer diagnosis. Cancer biomarker screening through clinical and point-of-care diagnostic methods is a promising tool for an early diagnosis of cancer.
At present, most of the available literature is specific to biomarkers for certain cancer types such as biomarkers for breast cancer [31], biomarkers for colorectal cancer [32], biomarkers for prostate cancer [33], biomarkers for ovarian cancer [34], biomarkers for cervical cancer [35], etc.
In this review, we focused mainly on more recent cancer-detecting methods such as cancer biomarker detection in various cancer types and ways in which this biosensor-mediated biosensing technology can show advantage over traditional detection methods. We also explored various types of cancer biomarkers, availability of traditional cancer biomarkers, recent developments in finding novel cancer biomarkers and their respective detection methods.

2. Cancer Biomarkers

The conventional diagnostic technologies such as MRI (magnetic resonance imaging), CT (computerized tomography) scan, ultrasound and biopsy were not effective for cancer detection at primary stages; this is because of their dependency on tumorigenic properties or phenotypic characters of a tumor [36,37,38,39,40]. Cancer is a very complex disease, with many epigenetic as well as genetic modifications which might alter the cell signaling process, related to development and resulting in tumorigenic transformation and malignancy [41]. For almost all cancer patients, researchers and clinicians expect tests or methods that might diagnose cancer significantly earlier, provide better prognosis, and that can allow for increased survival rates. Cancer markers have been used over the past few decades in the oncology field. Biomarkers are molecules of biologic emergence found in blood, tissues, various body fluids such as urine, cerebrospinal fluid, or different body tissues that are elevated is the indicative of an abnormal disease or condition with cancer. Human body responsiveness to any therapy can be observed and regulated through biomarkers. These also exist on or in cancer cells. Cancer biomarkers are possibly one of the most potential tools to detect cancer early [42,43,44,45,46].

2.1. Clinical Significance of Cancer Biomarkers

Cancer biomarkers can be utilized for cancer patient evaluation in different clinical levels, as well as disease screening, prognosis, diagnosis, staging, risk assessment, stratification, therapy planning and monitoring (Figure 2). Still, to date, several cancer markers have indicated poor validity and efficacy, especially in the most widespread cancers such as lung and breast cancers. Cancer biomarkers are bio-molecules necessary for remodeling throughout cancer which maintain excessive clinical relevance. These can be enzymes, iso-enzymes, nucleic acids, proteins, metabolites. Biomarkers are classified into three types based on their clinical advantage: prognostic, predictive and diagnostic. Prognostic biomarkers provide details about course of recurrence of the disease; patient response to the treatment is estimated by predictive biomarkers; disease detection can be performed by using diagnostic biomarkers [43,47,48,49]. The difference in the level of any unique biomarker in a cell or tissue is often used as evidence of tumor expansion. Biomarkers also play potential role in differentiating benign and malignant tumors and one type of malignancy from another type; specific biomarkers are helpful in unique settings, other biomarkers can be involved in multiple settings [48]. Biomarkers might be helpful to estimate a person’s chance of developing tumors/cancer. For example, a person having a solid family network (via ancestors) with ovarian cancer might receive a genetic test to decide whether they are acting as a carrier for a specific germ line modification or mutation, such as BRCA1, which could cause potential chance of developing breast and/or ovarian cancer [50,51]. Biomarkers might be helpful to determine malignancy in fit populations. A frequently utilized component for screening is PSA (prostate-specific antigen). It was approved by Food and Drug Administration (FDA) in 1986. Enhanced screening in male population above 50 years old lead to growth in the identification of prostate cancer. These kinds of traditional biomarkers also have limitations. In the previous decade, U.S. Preventive Services Task Force survey noted that an adequate documentation for common diagnosis with PSA [52,53,54]. Biomarkers were used to monitor prognosis and possibility of cancer reappearance irrespective of therapy/treatment. The clinical and pathological properties of a tumor could be useful for the prediction of cancers. Recently, modern techniques were used to evaluate prognosis of independent tumors; for example, a large number of genetic marks that had been matured in breast cancer might be useful to evaluate the identification for an individual patient depending on tumor assessment [55,56]. During breast cancer (metastatic) condition, circulating tumor cells are indicative of overall survival [57,58,59]. Biomarkers could be utilized as stimulus changers or prognostic factors for unique type of therapy, as well as to select the effective type of treatment. KRAS is a predictive biomarker for colorectal cancer; mutations occurring to KRAS in somatic cells are related to low response to anti-EGFR-mediated treatment [60,61]. Likewise, HER2 over-expression in the breast cancer as well as gastric cancer anticipates for stimulus to anti-Her2 drugs such as trastuzumab [62,63,64,65,66,67], and over-expression of estrogen receptor anticipates for stimulus to anti-endocrine therapy or treatment such as tamoxifen in breast cancer [68,69]. Identification of novel cancer biomarkers might help in quick and efficient diagnosis as well as monitoring of cancer progression.

2.2. Identification of Novel Cancer Biomarkers

Possible cancer biomarkers could be recognized through various approaches. An excellent way to recognize novel biomarkers mostly depends on biological nature of tumor and nearby environment of a tumor or metabolism of the drugs or biological products. With most recent studies and new information related to cancers and appearance of latest technology, cancer biomarker detection is performed frequently these days with applying a discovery approach. In this approach, few major areas of research such as gene expression arrays, proteomic technologies (mass spectroscopy, LC-MS/MS, MALDI-MS), and high throughput sequencing can be used to rapidly recognize unique biomarkers or pool of biomarkers which can show difference in the middle of cohorts. Expansion of sophisticated software algorithms for large data analysis has emerged in rapid advancement in the identification of novel cancer biomarkers. Openly available software programs can sort these data and compare the sequences to annotated genome databases to permit quantitative comparative evaluation of proteomes from multiple sources such as tumor area and nearby healthy tissues. Thus, over-expressed or down-regulated proteins in a cancer cell can be identified as putative cancer biomarkers. Huge amount of data produced with these technological methods mean that special attention needs to be directed toward both developing a study plan and conducting a large data analysis. Moreover, it is crucial to reduce the possibility of detecting relationships that are eventually determined to provide false-positive results. The most crucial features of biomarker improvement and identification to consider in depth include mindful study pattern to minimize any kind of bias, extensive evaluation, validation, and accurate communication of the results [48,70,71,72,73,74,75,76,77,78,79].

2.3. Cancer Biomarkers Currently Used in Clinical Settings

A cancer biomarker is a molecule existing inside and/or generated by tumor cells or surrounding cells in tissues or organs in stimulus to tumor or certain noncancerous (benign) situations, which can provide information about cancer, mostly which stage it is, whether it can respond to treatment, and what type of therapy might be useful. Here, few recent cancer biomarkers used in clinical practice are explained. New cancer biomarkers become available continuously, and they may not be explained below [27,40,44,80,81,82].

2.3.1. AFP (Alpha-Fetoprotein)

AFP (Alpha-fetoprotein) is one of the leading biomarker. Early fetal life of a baby (mostly yolk sac and liver) produces AFP during pregnancy. AFP can be detected in huge amounts in serum of the patients with specific tumors. According to Yuri Semenovich Tatarinov, AFP was first accepted as an antigen unique for human HCC (hepatocellular carcinoma) [83,84]. The scientific literature has explained that AFPs are classified into subtypes based on their dissimilar affinities to LCA (lens culinaris agglutinin), such as AFP-L1 (LCA unreactive AFP), AFP-L2 (LCA mild active AFP) and AFP-L3 (LCA active AFP). In healthy individual serum, an average level of AFP is less than 20 ng 8 mL−1 [85,86,87]. AFP is widely established for HCC (hepatocellular carcinoma) diagnosis. Moreover, in congenital tyrosinemia, cirrhosis, hepatitis (alcohol-induced), hepatitis (viral-induced), ataxia-telangiectasia syndrome or in several malignancies such as testicular cancer, liver cancer, gastric cancer and nasopharyngeal cancer elevated AFP levels may also be present. Hence, sensing the AFP values is completely mandatory in clinical settings. High recognition rates of molecular assays have been obtained in quantitative observation of AFP due to their specificity and unalterable affinity of the probes to molecular targets [88,89].

2.3.2. PSA (Prostate Specific Antigen)

PSA (prostate-specific antigen) was one of the first recognized cancer biomarkers, utilized to detect and screen prostate cancer in clinical setting. It has been shown that increased levels of PSA can directly relate to prostate cancer. Human regular PSA level is 4 ng/mL. According to the study by Smith et. al, almost 30% of individuals with PSA values higher than normal (range of 4.1–9.9 ng/mL) were diagnosed with prostate cancer [90]. Along with this, raised PSA values may also indicate benign tumors (non-fatal), prostatitis/prostate inflammation or benign prostatic hyperplasia. Therefore, elevated values of PSA may not consistently suggest malignant tumors. There is a fact that caused reasonable controversy about using regular PSA screening to detect prostate cancer. Small-sized tumors identified by PSA screening may grow very slowly; death caused by small tumor might not be feasible in an individual lifetime. Moreover, it is very expensive to treat such slow-growing tumors. Other frequent issue with PSA screening is false-positives. This limitation of PSA testing can be overcome by modern biosensing technology mediated by biosensors [79,84,91,92,93]. PSA detection can be performed by various methods; these traditional methods are time-consuming as well as expensive. Yang et al. explained a graphene oxide/ssDNA-based biosensor integrated with dual antibody-modified PLLA NPs to amplify electrochemical signals for the effective and rapid electrochemical capture of PSA in serum samples from prostate cancer patients. The detection limit for PSA was 1 ng/mL, which achieved a wide linear range of 1–100 ng/mL for PSA. This is one of the examples that shows the usefulness of modern biosensing technology mediated by biosensors [94].

2.3.3. RCAS1 (Receptor-Binding Cancer Antigen)

RCAS1 (receptor-binding cancer antigen) overexpression data has been described in many gastric carcinomas; it is related with progression of gastric cancer. Further, RCAS1 is also suggested as a cancer biomarker for poor prediction in breast, esophageal, endometrial, bladder cancers and is associated with tumor weakening in pharyngeal carcinoma and laryngeal cancer. RCAS overexpression is observed in several types of cancers. Thus, it serves as a potential biomarker for cancer detection and prediction [95,96,97,98].

2.3.4. CA 15-3 (Cancer Antigen 15-3)

The most predominant cancer marker for breast cancer identification as well as monitoring is cancer antigen 15-3; additional biomarkers that are related to breast cancer are CA 27.29, BRCA1, BRCA2 and (carcinoembryonic antigen) CEA [44,99,100]. This particular marker is frequently used on a clinical level to monitor the therapy for breast cancer in its advanced stages. During breast cancer, CA 15-3 values increased by 10, 20, 30 and 40 % at various stages first, second, third and fourth stage [49]. Tampellini et. al showed the connection between breast cancer and CA 15-3 levels, also mentioning that before treatment, patients with levels of 30U/mL had notably higher survival rate compared to patients with higher values. Raised CA 15-3 values correspond with extensive metastasis [101,102]. Other research data explained that raised CA 15-3 levels post cancer treatment can be a sign of disease recurrence. Nowadays, to determine breast cancer treatment, protocol CA 15-3 values are considered along with risk factors (negative) such as PR/ER condition as well as iHer-2, cancer stage and tumor dimension. Hepatitis, endometriosis, pelvic inflammatory disease, lactation and pregnancy are conditions other than cancer where CA 15-3 levels are increased [103].

2.3.5. Cancer–Testis (CT) Antigens

Cancer–testis antigens are a specific type of cancer biomarkers. They are expressed in various cancers. CT antigen expression is limited to male germ cells of the testis but not shown in normal adult cells. These antigens are also expressed in trophoblast and ovary cells. Therefore, CT antigens have been considered as possible immunogenic targets for cancer vaccines (cancer immunotherapies). CT antigen autoantibodies have been studied as potential cancer biomarkers [104]. NY-ESO-1 (NewYorkesophagealsquamouscellcarcinoma1) is encoded by CTAG-1B; this is the class of antigens with high immunogenic nature which induces very robust cellular and humoral immune response in NY-ESO-1-positive cancers. Antibody titer to NY-ESO-1 has been shown to relate with disease progression. One of the best benefits of using CT autoantibodies as tumor biomarkers is the fact that they are easy to obtain and are also more stable proteins present in serum compared to tissues obtained via biopsy. Thus, they can be useful for cancer progression and recurrence [105,106,107,108,109]. Major limitation of CT antigens is the fact that many cancers express CT antigens, and they are rarely tumor-specific. Biosensor technology can obtain a profile for CT antigen, which can enhance the use of these antigens in cancer prognosis and diagnosis [110,111,112].

2.3.6. CA 125

The rise in CA 125 levels is primarily related to ovarian cancer. It is further correlated with several different cancers such as cervix, lungs, breast, liver, pancreas, uterus, stomach and colon cancers. Enhanced levels of CA 125 are also observed in various non-pathological conditions such as menstruation and pregnancy [8,87,113]. A total of 90% of women with advanced stage ovarian cancer and 40% of humans with intra-abdominal malignant tumors also exhibit high CA 125 levels. Still, approximately 50% of patients diagnosed with primary stage ovarian cancer show normal CA 125 levels [87,88,89]. Other germ cell origin biomarkers such as alpha-fetoprotein/AFP, human chorionic gonadotrophin/HCG as well as Lactate dehydrogenase/LDH are also connected to ovarian cancer [114,115]. Increase in CA 125 levels can be used to detect the development of benign tumors into malignant tumors. Enhanced CA 125 levels are also used to identify treatment failure as well as disease recurrence (e.g., high CA 125 levels after bilateral salpingo-oophorectomy or total abdominal hysterectomy, which may occur after first line chemotherapy) [116,117,118,119,120]. Altogether, CA 125 is an extremely essential biomarker for detection of cancer, and also for cancer progression monitoring and treatment.

2.3.7. CA 19-9

This antigen was first identified in pancreatic and colon cancer patient’s serum in 1981. It is a Lewis antigen of the MUC1 glycoprotein [8,121]. CA 19-9 normal level in serum is less than 37 U/mL. In the recent decade, on a clinical level, CA 19-9 biomarker has become extremely useful for the diagnosis of pancreatic cancer. Normal human serum CA 19-9 levels can play an outstanding role in clinical diagnosis of urothelial and gastric cancers [122,123]. Consequently, there is a necessity to improve highly sensitive methods which can detect CA 19-9 values in patients with cancer.

2.3.8. Nse (Neuron-Specific Enolase)

This neuron-specific enolase is a popular and unique marker for SCLC as well as NSCLC non-small cell lung cancer. It has a crucial role in glycolysis; in 1980s, NSE expression was noted in SCLC cells. From that time, it has been used as potential biomarker for lung cancer, able to detect increased values of NSCLC and acting as crucial predictor for patient survival (it is not dependent on remaining prognostic factors) [124,125,126]. NSE is also a unique marker for neuro-endocrine cells. Raised NSE values in body fluids might be an indication of tumor proliferation and staging determination in some of brain tumors [127,128]. In the recent decade, the value of NSE for prognosis in cancer patients is debatable. Therefore, it is mandatory to improve sensitive techniques to perceive Nse values in patients with cancer.

2.3.9. Tdt (Terminal Deoxynucleotidyl Transferase)

Tdt is an intracellular marker which has detected in the bone marrow as well as blood (mononucleate) cells in leukemia patients during diagnosis. Overall, TdT values are remarkably raised in several lymphocytic lymphomas. Tdt might be helpful to identify specific leukemia type and supportive sign for the solution of therapy [129,130,131,132].

2.3.10. CYFRA21-1

Cytokeratin-19 fragments/CYFRA21-1 have been extensively studied in patients with NSCLC and are widely utilized as predictive, prognostic markers. They have 56% sensitivity as well as 88% specificity when the value is >1.5 ng/mL. Researchers used maximum cut off value for CYFRA21-1 similar to ≥3 ng/mL; it had shown increased specificity at 97%. It has potential capability in lung, esophagal cancer prediction. Raised values are certain but barely sensitive. These values are strongly connected with cancer metastasis. Recent reports explained that CYFRA21-1 is used as independent prediction factor for various phases of lung cancer. This might function as definitive distinction between benign and malignant lung cancer along with clinical information [133,134,135,136,137].
Table 1 demonstrates the summary of various cancer biomarkers available in the literature for the detection of cancer diseases. Significant biomarkers are assembled in the table including their respective cancer diseases and their advantages.

3. Importance of Finding Novel Bio-Sensing Methods to Detect Cancer Biomarkers

Biomarkers might have different molecular origins, such as changes to nucleic acids including RNA, DNA (amplification, translocation, point mutation, loss of heterozygosis) and protein (antibodies, tumor suppressors, oncogenes and hormones). Some of these cancer biomarkers are convincing and extremely crucial for early diagnosis of tumors. Some of the biomarkers need to exhibit adequate specificity and sensitivity for translation into clinical use or for monitoring of disease progression. In this area, biosensing technology can potentially play a crucial role to improve early diagnosis of cancer. The traditional PCR (polymerase chain reaction) or ELISA (enzyme-linked immunosorbent assay) techniques for cancer biomarker identification have some technical limitations, including utilization of costly chemicals and expensive machines in every single assay, which could delay the detection. Moreover, these types of techniques are not capable of constant monitoring in patients throughout treatment. Along with this, multiple pathways are interlinked with cancer cells and these cells express more than one biomarker. Therefore, simultaneous identification of different biomarkers for accurate prognosis and diagnosis is indispensable [8,42,48,138,139,140,141,142,143]. Biosensors offer great potential sensing methodology platforms for the detection of various cancer biomarkers. Specifically electrochemical and optical biosensors based on the affinity, chemical, bio-affinity recognition elements attract great interest. In the following section, we comprehensively discuss the electrochemical and optical sensing strategies for cancer disease diagnosis.
Table 1. Biomarkers used in cancer detection.
Table 1. Biomarkers used in cancer detection.
Tumor/Cancer BiomarkerType of Cancer/Infected LocationApplicationReferences
AFP Liver (HCC)Identifying recurrence,
treatment monitoring,
disease diagnosis
[144,145,146,147]
PSA Prostate gland Screening,
identifying recurrence,
treatment monitoring,
disease diagnosis
[148,149,150]
CA 15-3BreastTreatment monitoring[103,151,152]
CT antigensProstate,
liver,
lung,bladder,
skin
Diagnosis, prognosis[105,106]
CA27.29 Breast Monitoring[69,153]
RCAS1StomachDetection,
prognosis
[96,97,98]
CA 19-9 Pancreas,
colon
Treatment monitoring[122,154]
CEA (Carcinoembryonic antigen)Colon,
liver
Screening,
Identifying recurrence,
Treatment monitoring,
Disease prognosis
[155,156]
Calcitonin Thyroid glandTreatment monitoring,
Disease prognosis
[157]
ER & PgR
(Estrogen, progesterone receptors)
Breast Stratification [158,159,160]
HER2Lung, breast Monitoring therapy [64,161,162,163]
CA 125 Ovary Prognosis,
identifying recurrence,
treatment monitoring,
disease diagnosis
[88,89]
HCG-β Ovary,
testis
Diagnosis, staging,
identifying recurrence,
treatment monitoring
[164,165]
TdtBlood/leukemiaDiagnosis[129]
NSeLungPrognosis[125,166]
Thyroglobulin ThyroidTreatment monitoring[167,168]
PCA3 Prostate glandPrognosis [169]
NY-eSO-1 Skin/melanomaProgression monitoring[170]
EGFRLungDiagnosis and monitoring therapy[171,172]
KRAS, ALKLungDiagnosis and monitoring therapy[173,174]
CD30Blood/LeukemiaDiagnosis and prognosis[175,176]
NMP 22 Bladder Screening,
treatment monitoring,
disease prognosis
[177,178]
CYFRA21-1 EsophagusPrognosis,
Treatment monitoring
[179,180,181]
BCL2Blood
and breast
Diagnosis,
treatment plan
[182,183,184]
BCR-ABL fusion geneBone marrow,
blood
Prognosis,
treatment determination,
monitoring
[185,186]
CD20BloodTreatment determination[187]
CD22BloodTreatment determination,
diagnosis
[188]
CD25Blood Treatment determination[189]
FGFR2 & FGFR3BladderTreatment determination,
therapy
[190,191]
Fibrin-fibrinogenBladderTreatment determination,
monitoring
[192,193]
SMRPLeukemia Progression monitoring[194,195]
ROS1LungTreatment determination [196]
OVA1OvaryPrognosis[197]
VMABrainDiagnosis[198]

4. Analytical Diagnostics Methodologies for Cancer Biomarkers Screening

Analytical biosensing methodologies for the detection of various analyte molecules including cancer biomarkers, pharmaceutical drugs, and agricultural toxins are recently rapidly growing. These techniques have several unique advantages such as point-of-care diagnosis, miniaturized portable instrumentation, cost-effectiveness and, moreover, user-friendliness to the end users. In this current section, we discuss some of the recently published works related to the optical- and electrochemical-based biosensors for cancer biomarker detection.
O. Awatef et al. reported a selective, sensitive, and inexpensive aptamer-based SERS biosensor for detection of prostate-specific antigen in human serum. Here, they have been using 1D (1 Dimensional) Silicon nanowires as a transduction material due to their advantage large surface area. These materials prepared from N-doped Si(100) wafers by metal-assisted chemical etching method. AgNPs were deposited on SiNWs through the electroless deposition technique to enhance the optical signal properties of the sensor. AgNPs/SiNWs were further functionalized with self-assembled layer of hexanethiol by incubating the SERS substrate of MCH to avoid the non-specific binding of the analyte. The developed sensor exhibited a good response in the dynamic range from 0.1 to 20 μg·L−1 with a limit of detection of 0.1 μg·L−1. Moreover, this sensing platform selectively and sensitively detected PSA in spiked PBS solutions [199].
M. Sachin et al. developed the fabrication of a tailored biofunctionalized interdigitated capacitor electrode (Ti/Pt imprinted) for label-free PSA detection. This sensing platform exhibits rapid detection within 3 s, stability up to many weeks, reusability and reproducibility; it is also requires low volume and easy to operate. Here, interdigitated capacitor (IDC) chip was initially functionalized with Ti/Pt metal by e-beam deposition process. Electrode surface was activated by placing it into piranha solution to formation of hydroxyl groups on the surface. Then, further formation of an amine group on the surface by APTES solution was drop-costed, and glutaraldehyde solution was applied to generate an antibody through cross-linking. Later anti-PSA was fabricated on the electrode surface, and further immobilization of PSA onto the bio surface has performed to verify the interactions through capacitance. APTES and glutaraldehyde increase the positive capacitive response of IDC-based PSA biosensor and are treated as reference value. By using LCR meter, the change in capacitance variables with respect to changing in concentration of target protein has been calculated [200].
Jose Ribeiro et al. developed new biosensing methodology by merging two different techniques, surface plasmon resonance spectroscopy and electrochemical technique for the detection of breast cancer biomarker carbohydrate Antigen 15-3 to monitor disease progression. Two steps are mainly involved in this process: (i) direct SPR monitoring interaction between surface immobilized antibody and CA 15-3 antigen, performed until adsorption reached equilibrium, and (ii) electrochemical measurements at the SPR gold surface and the resulting immunosensor selective detection for the breast cancer biomarker CA 15-3 protein [201].
D. Haihan, et al. performed investigations to explore the construction of paper-based photo electrochemical (PEC) biosensors with 1D self-doping SnO2 nanotubes for selectively detection of alpha fetoprotein (AFP). (Figure 3) With the template consumption technique, paper-based 1D-domed SnO2 nanotubes have been created from template ZnO nanorods. Additionally, a method of Sn self-doping was suggested to make it easier to separate photo-induced charge carriers and improve the harvesting of visible light. Additionally, self-doping of Sn can reduce the recombination rate of charge carriers and narrow the band gap of SnO2 nanotubes, which results in a significant increase in photocurrent intensity under visible light illumination [202].
W. Qiong et al. reported the new class of 2D Ti3C2-MXene nanosheet-based SPR biosensor for detection of carcinoembryonic antigen (CEA) cancer biomarker with high specificity and reproducibility. A novel class of ultrathin Ti3C2-MXene possesses hydrophilic biocompatible surface, which can be used as a biosensing material. Ti3C2 -MXene nanosheets were coated with AuNPs using a chemical reduction method and further modified with SPA to improve detection sensitivity or orient purpose. Later immobilization of anti-monoclonal CEA(Ab1) has performed to capture the analyte CEA. Here, Ti3C2-MXene-based SPR sensing platform exhibited significant performance for detection of CEA in real serum samples [203].
Table 2 demonstrates the summary of various optical screening methods based on different recognition matrices for the screening of cancer biomarkers. Significant articles with the transduction techniques, bio/chemical recognition matrices, dynamic working calibration range and limit of detection are assembled in the table.

Electrochemical Sensing Methodologies

Electrochemical biosensing methodologies for the detection of various analyte molecules including cancer biomarkers, pharmaceutical drugs [218,219,220,221], and agricultural toxins are recently rapidly growing. These techniques have several unique advantages such as point-of-care diagnosis, miniaturized portable instrumentation, cost-effectiveness and, moreover, user-friendliness to the end users. In this current section, we discuss some of the recently published works related to the electrochemical-based biosensors for cancer biomarker detection.
Huiqing Yang’s team recently developed electrochemical aptasensor for the specific detection of carcinoembryonic antigen (CEA) biomarker. In this investigation, the authors proposed the future of a new antifouling material MXC-Fe3O4-Ru on functional 2D nanomaterial-modified magnetic gold electrode (MGE). The ferrocene-modified carcinoembryonic antigen aptamer sequences were immobilized on the MGE/modified electrode surface with amido bond chemistry. Electrochemical signal of ferrocene decreases and [Ru(NH3)6]3 signal fixed on the electrode remains unchanged. The ratio of the electrochemical signals of ferrocene and [Ru(NH3)6]3 is proportional to the CEA concentration. Even in the complex samples, biosensors can reach high accuracy, selectivity and sensitivity for the detection of targets because of excellent antifouling performance and good conductivity [222].
João G. Pacheco et. al. developed an electrochemical sensor based on molecularly imprinted polymer to monitor breast cancer biomarker CA 15-3. In this work, the screen-printed gold electrodes were modified with the MIP recognition matrices. This MIP-based sensor was demonstrated in the electrolytic solution hexacyanoferrate (II/III) as redox probe for measuring the CA protein MIP-binding interactions. Interestingly, the peak current density increases with respect to the CA 15-3 concentration in adynamic range between 5 and 50 U mL−1 with the detection limit of 1.5 U mL−1. The prepared MIP sensor is low cost and works efficiently for fast (15 min) analysis [223].
Other efforts have focused on the fabrication of label-free electrochemical aptasensor with the integration of microfluidic paper device for the specific detection of prostate specific antigen (PSA) in clinical samples. Screen-printed gold electrodes were fabricated with wax-printed technology. (Figure 4) Au-SPE surface is modified with the reduced graphene oxide and gold–thionine nanoparticle composites. Then, DNA aptamer was immobilized on top of the modified SPE. Afterward, the fabricated aptasensor was tested for the specific detection of PSA for diagnosing prostate cancer disease [224].
In another report, an electrochemical aptasensor was proposed by Leila Farzin’s research team for the specific detection of CA-125 cancer biomarker. The proposed sensing platform consists of polycrystalline nanofibers coupled with amidoxime-doped silver nanoparticles. The authors reported that hybrid nanomaterial-modified sensor surface helped in better immobilization of aptamer sequences and obtaining sensitive detection limits for CA-125 biomarker detection in ovarian cancer-infected patients [225].
Another interesting system that was established was a sandwich-type electrochemical aptasensor to measure carcinoembryonic antigen and cancer antigen 15-3(CA 15-3). The proposed sensing platform consists of a three-dimensional graphene gel embedded with gold nanoparticles (AuNPs/3DGH). This biosensing transducing layer helps in better immobilization of the redox-labelled aptamer sequences. Affinity interaction between aptamers and respective cancer biomarkers (CEA and CA 15-3) was recorded with square wave voltammetry methods by measuring the change in redox probe electroactive signals. This proposed sandwich aptasesing assay exhibited good limit of detection value in nanomolar linear range. The obtained results are comparable with the ELISA method [226].
In another report, a DNA nano-tweezer-based electrochemical sensor was developed for sensing and specific detection carcinoembryonic antigen biomarker. (Figure 5) DNA nano-tweezer is a DNA nanomachine used to enhance the sensing performance of electrochemical biosensor. Here, three-dimensional DNA nanomachine possesses the more active sites that could help in enhancement of the competence of reaction. This is the first kind of 3D DNA nanoprobe used in electrochemical sensing platform to obtain stable and reproducible results. Sensor exhibited good electroanalytical performance towards the target biomarker (CEA) with a detection limit of 4.88 fg mL−1 [227].
Some other designs have been recently proposed by Navid Taheri et. al. for multiplexed determination of biomarkers alpha-fetoprotein (AFP) and carcinoembryonic antigen (CEA). Here, the sensing assay consisted of an electropolymerized polypyrrole conducting polymer, methyl orange layer and a DMIP layer on FTO surface. The target biomarker and MIP interactions were recorded with the electrochemical impedance spectroscopy by measuring impedance values. The sensor exhibited promising results in the dynamic range of 5–104 and 10–104 pg mL−1 and detection limits of 1.6 and 3.3 pg mL−1 for CEA and AFP, respectively [228].
M. Samira et al. developed a novel method for ovarian cancer antigen detection by ultrasensitive flexible aptasensor based on functionalized CNT-reduced graphene oxide nanocomposite. (Figure 6) Reduced graphene oxide film was prepared by modified Hummer’s method. The CA 125 ssDNA aptamer sequences were immobilized on MWCNTs surface by amide bond formation. Then, fabrication of rGO was performed on the polymethyl methacrylate (PMMA) surface by using the polishing method. Here, gold source and drain electrode deposited on graphene film, and the surface of graphene was modified with MWCNTs/CA 125 aptamer through π-π interaction. Overall, this technique specifically and selectively detects the CA 125 biomarker from serum [208].
Table 3 summarizes the recently reported literature related to the electrochemical biosensor for the detection of various cancer biomarkers based on various recognition matrices for the detection of cancer biomarkers. The reports are clearly classified and assembled in the table with respect to their recognition matrices, dynamic working ranges and limits of detection.

5. Conclusions and Future Perspectives

Overall, this review summarizes the importance of cancer biomarker detection, cancer-causing environments, traditional available screening biomarkers for cancer diagnosis, and novel cancer biomarkers and their advantages over traditional biomarkers. Recent developments in analytical diagnostic strategies including electrochemical and optical transduction methods for cancer biomarkers screening are also addressed. A summary of different cancer biomarkers available in the literature for the detection of cancer diseases are displayed in Table 1, including their respective cancer diseases and their advantages. Summaries of optical and electrochemical screening methods based on various recognition matrices for the detection of cancer biomarkers are assembled in their respective tables.
Cancer biomarker biosensing assay development consists of several critical challenges, including biofluid separation, real sample analysis, sensitivity, multiplex detection and integration of miniaturized instrumentation [243]. Real sample analysis is the particularly significant challenge; we are aiming to detect the target analyte in presence of several protein molecules. To achieve this, we need to integrate the device with the micro/nanofluidic devices and use different coating layers to protect the assay reading from biofouling studies. A recent publication has explained in detailed manner the way in which the fouling occurs, ways to overcome this problem by using different anti-biofouling coatings, as well as the effect of the fouling on electrocatalytic responses [244]. Finally, the device needs to be integrated with the miniaturized instrumentations, and a specific app/software need to be developed to record recognition element and biomarker interactions. The recent review exclusively discussed the integration of biosensing strategies with the electronics devices and wirelessly operated mobile phones for point-of-care diagnostic applications [245]. These sensor integration challenges could be overcome with the help of interdisciplinary approaches, specifically through collaborations between chemistry, electronics, computer, and nanofabrication experts. This review could be help to the early career researchers who are working in the domains of chemistry, biotechnology, nanotechnology, cell biology and biosensors. Overall, the current review provides new insights to the researchers to develop novel biosensing methodologies for the detection of various cancer biomarkers.

Funding

Funding source—Department of Science and Technology, Science and Engineering Research Board (DST-SERB/RJF/2021/000113, Ramanujan Award), New Delhi, India, and the Indian Institute of Technology Palakkad for financial support and research facilities.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

K. Yugender Goud would like to thank the Department of Science and Technology, Science and Engineering Research Board (DST-SERB/RJF/2021/000113, Ramanujan Award), New Delhi, India, and the Indian Institute of Technology Palakkad for financial support and research facilities.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AFPAlpha-fetoprotein
PSAProstate-specific antigen
RCAS1Receptor-binding cancer antigen
CA 15-3Cancer antigen 15-3
CT antigenCancer–testis antigen
CA 125Cancer antigen 125
CA 19-9Cancer antigen 19-9
NseNeuron-specific enolase
TdtTerminal deoxynucleotidyl transferase
CYFRA21-1Cytokeratin-19 fragments
CarcinomaEpithelial cell cancer
SarcomaConnective tissue/bone cancer
LymphomaLymphatic system cancer
MyelomaPlasma cell cancer
LeukemiaBlood cancer
BRCA1Breast cancer gene 1
BRCA2Breast cancer gene 2

References

  1. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer Statistics, 2022. CA A Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef]
  2. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer Statistics, 2020. CA A Cancer J. Clin. 2020, 70, 7–30. [Google Scholar] [CrossRef]
  3. Yabroff, K.R.; Wu, X.C.; Negoita, S.; Stevens, J.; Coyle, L.; Zhao, J.; Mumphrey, B.J.; Jemal, A.; Ward, K.C. Association of the COVID-19 Pandemic with Patterns of Statewide Cancer Services. J. Natl. Cancer Inst. 2022, 114, 907–909. [Google Scholar] [CrossRef]
  4. Patel, A. Benign vs Malignant Tumors. JAMA Oncol. 2020, 6, 1488. [Google Scholar] [CrossRef]
  5. WHO Cancer and Carcinogens. 2022. Available online: https://www.who.int/news-room/fact-sheets/detail/cancer) (accessed on 11 November 2022).
  6. Croswell, J.M.; Ransohoff, D.F.; Kramer, B.S. Principles of Cancer Screening: Lessons from History and Study Design Issues. Semin. Oncol. 2010, 37, 202–215. [Google Scholar] [CrossRef] [Green Version]
  7. Hamilton, W.; Walter, F.M.; Rubin, G.; Neal, R.D. Improving early diagnosis of symptomatic cancer. Nat. Rev. Clin. Oncol. 2016, 13, 740–749. [Google Scholar] [CrossRef]
  8. Patel, J.; Patel, P. Biosensors and Biomarkers: Promising Tools for Cancer Diagnosis. Int. J. Biosens. Bioelectron. 2017, 3, 00072. [Google Scholar] [CrossRef] [Green Version]
  9. Neal, R.D.; Tharmanathan, P.; France, B.; Din, N.U.; Cotton, S.; Fallon-Ferguson, J.; Hamilton, W.; Hendry, A.; Hendry, M.; Lewis, R.; et al. Is Increased Time to Diagnosis and Treatment in Symptomatic Cancer Associated with Poorer Outcomes? Systematic Review. Br. J. Cancer 2015, 112, S92–S107. [Google Scholar] [CrossRef] [Green Version]
  10. Summerfield, C.; Smith, L.; Todd, O.; Renzi, C.; Lyratzopoulos, G.; Neal, R.D.; Jones, D. The Effect of Older Age and Frailty on the Time to Diagnosis of Cancer: A Connected Bradford Electronic Health Records Study. Cancers 2022, 14, 5666. [Google Scholar] [CrossRef] [PubMed]
  11. Elliss-Brookes, L.; McPhail, S.; Ives, A.; Greenslade, M.; Shelton, J.; Hiom, S.; Richards, M. Routes to Diagnosis for Cancer—Determining the Patient Journey Using Multiple Routine Data Sets. Br. J. Cancer 2012, 107, 1220–1226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Whitaker, K. Earlier Diagnosis: The Importance of Cancer Symptoms. Lancet Oncol. 2020, 21, 6–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Kennedy, M.P.T.; Cheyne, L.; Darby, M.; Plant, P.; Milton, R.; Robson, J.M.; Gill, A.; Malhotra, P.; Ashford-Turner, V.; Rodger, K.; et al. Lung Cancer Stage-Shift Following a Symptom Awareness Campaign. Thorax 2018, 73, 1128–1136. [Google Scholar] [CrossRef] [PubMed]
  14. Power, E.; Wardle, J. Change in Public Awareness of Symptoms and Perceived Barriers to Seeing a Doctor Following Be Clear on Cancer Campaigns in England. Br. J. Cancer 2015, 112, S22–S26. [Google Scholar] [CrossRef] [PubMed]
  15. Loud, J.T.; Murphy, J. Cancer Screening and Early Detection in the 21 St Century. Semin. Oncol. Nurs. 2017, 33, 121–128. [Google Scholar] [CrossRef] [Green Version]
  16. Hewitt, K.; Son, J.; Glencer, A.; Borowsky, A.D.; Cooperberg, M.R.; Esserman, L.J. The Evolution of Our Understanding of the Biology of Cancer Is the Key to Avoiding Overdiagnosis and Overtreatment. Cancer Epidemiol. Biomark. Prev. 2020, 29, 2463–2474. [Google Scholar] [CrossRef]
  17. Loomans-Kropp, H.A.; Umar, A. Cancer Prevention and Screening: The next Step in the Era of Precision Medicine. Npj Precis. Oncol. 2019, 3, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Shieh, Y.; Eklund, M.; Madlensky, L.; Sawyer, S.D.; Thompson, C.K.; Fiscalini, A.S.; Ziv, E.; Van’t Veer, L.J.; Esserman, L.J.; Tice, J.A. Breast Cancer Screening in the Precision Medicine Era: Risk-Based Screening in a Population-Based Trial. J. Natl. Cancer Inst. 2017, 109, djw290. [Google Scholar] [CrossRef] [Green Version]
  19. Esserman, L.J. The WISDOM Study: Breaking the Deadlock in the Breast Cancer Screening Debate. Npj Breast Cancer 2017, 3, 34. [Google Scholar] [CrossRef]
  20. Henley, S.J.; King, J.B.; German, R.R.; Richardson, L.C.; Plescia, M.; Centers for Disease Control and Prevention (CDC). Surveillance of Screening-Detected Cancers. MMWR Surveill. Summ. 2010, 59, 1–25. [Google Scholar]
  21. Negoita, S.; Feuer, E.J.; Mariotto, A.; Cronin, K.A.; Petkov, V.I.; Hussey, S.K.; Benard, V.; Henley, S.J.; Anderson, R.N.; Fedewa, S.; et al. Annual Report to the Nation on the Status of Cancer, Part II: Recent Changes in Prostate Cancer Trends and Disease Characteristics. Cancer 2018, 124, 2801–2814. [Google Scholar] [CrossRef] [Green Version]
  22. Elmore, L.W.; Greer, S.F.; Daniels, E.C.; Saxe, C.C.; Melner, M.H.; Krawiec, G.M.; Cance, W.G.; Phelps, W.C. Blueprint for Cancer Research: Critical Gaps and Opportunities. CA A Cancer J. Clin. 2021, 71, 107–139. [Google Scholar] [CrossRef]
  23. Schiffman, J.D.; Fisher, P.G.; Gibbs, P. Early Detection of Cancer: Past, Present, and Future. Am. Soc. Clin. Oncol. Educ. Book 2015, 35, 57–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Saini, S.; Stark, D.D.; Brady, T.J.; Wittenberg, J.; Ferrucci, J.T. Dynamic Spin-Echo MRI of Liver Cancer Using Gadolinium-DTPA: Animal Investigation. AJR Am. J. Roentgenol. 1986, 147, 357–362. [Google Scholar] [CrossRef] [PubMed]
  25. Lyman, D.F.; Bell, A.; Black, A.; Dingerdissen, H.; Cauley, E.; Gogate, N.; Liu, D.; Joseph, A.; Kahsay, R.; Crichton, D.J.; et al. Modeling and Integration of N-Glycan Biomarkers in a Comprehensive Biomarker Data Model. Glycobiology 2022, 32, 855–870. [Google Scholar] [CrossRef]
  26. Alimirzaie, S.; Bagherzadeh, M.; Akbari, M.R. Liquid Biopsy in Breast Cancer: A Comprehensive Review. Clin. Genet. 2019, 95, 643–660. [Google Scholar] [CrossRef] [PubMed]
  27. Nikanjam, M.; Kato, S.; Kurzrock, R. Liquid Biopsy: Current Technology and Clinical Applications. J. Hematol. Oncol. 2022, 15, 1–14. [Google Scholar] [CrossRef]
  28. Chen, X.; Gole, J.; Gore, A.; He, Q.; Lu, M.; Min, J.; Yuan, Z.; Yang, X.; Jiang, Y.; Zhang, T.; et al. Non-Invasive Early Detection of Cancer Four Years before Conventional Diagnosis Using a Blood Test. Nat. Commun. 2020, 11, 3475. [Google Scholar] [CrossRef] [PubMed]
  29. Hall, I.J.; Tangka, F.K.L.; Sabatino, S.A.; Thompson, T.D.; Graubard, B.I.; Breen, N. Patterns and Trends in Cancer Screening in the United States. Prev. Chronic Dis. 2018, 15, 170465. [Google Scholar] [CrossRef] [Green Version]
  30. Pulumati, A.; Pulumati, A.; Dwarakanath, B.S.; Verma, A.; Papineni, R.V.L. Technological advancements in cancer diagnostics: Improvements and limitations. Cancer Rep. 2023, 6, e1764. [Google Scholar] [CrossRef]
  31. Gamble, P.; Jaroensri, R.; Wang, H.; Tan, F.; Moran, M.; Brown, T.; Flament-Auvigne, I.; Rakha, E.A.; Toss, M.; Dabbs, D.J.; et al. Determining Breast Cancer Biomarker Status and Associated Morphological Features Using Deep Learning. Commun. Med. 2021, 1, 14. [Google Scholar] [CrossRef]
  32. Zygulska, A.L.; Pierzchalski, P. Novel Diagnostic Biomarkers in Colorectal Cancer. Int. J. Mol. Sci. 2022, 23, 852. [Google Scholar] [CrossRef] [PubMed]
  33. Porzycki, P.; Ciszkowicz, E. Modern Biomarkers in Prostate Cancer Diagnosis. Cent. Eur. J. Urol. 2020, 73, 300–306. [Google Scholar] [CrossRef]
  34. Atallah, G.A.; Aziz, N.H.A.; Teik, C.K.; Shafiee, M.N.; Kampan, N.C. New Predictive Biomarkers for Ovarian Cancer. Diagnostics 2021, 11, 465. [Google Scholar] [CrossRef] [PubMed]
  35. Molina, M.A.; Diatricch, L.C.; Quintana, M.C.; Melchers, W.J.G.; Andralojc, K.M. Cervical Cancer Risk Profiling: Molecular Biomarkers Predicting the Outcome of HrHPV Infection. Expert Rev. Mol. Diagn. 2020, 20, 1099–1120. [Google Scholar] [CrossRef]
  36. Hussain, T.; Nguyen, Q.T. Molecular Imaging for Cancer Diagnosis and Surgery. Adv. Drug Deliv. Rev. 2014, 66, 90. [Google Scholar] [CrossRef] [Green Version]
  37. Coleman, R.; Finkelstein, D.M.; Barrios, C.; Martin, M.; Iwata, H.; Hegg, R.; Glaspy, J.; Periañez, A.M.; Tonkin, K.; Deleu, I.; et al. Adjuvant Denosumab in Early Breast Cancer (D-CARE): An International, Multicentre, Randomised, Controlled, Phase 3 Trial. Lancet Oncol. 2020, 21, 60–72. [Google Scholar] [CrossRef]
  38. Taouli, B.; Beer, A.J.; Chenevert, T.; Collins, D.; Lehman, C.; Matos, C.; Padhani, A.R.; Rosenkrantz, A.B.; Shukla-Dave, A.; Sigmund, E.; et al. Diffusion-Weighted Imaging Outside the Brain: Consensus Statement From an ISMRM-Sponsored Workshop. J. Magn. Reason. Imaging 2016, 44, 521–540. [Google Scholar] [CrossRef]
  39. Padhani, A.R.; Liu, G.; Mu-Koh, D.; Chenevert, T.L.; Thoeny, H.C.; Takahara, T.; Dzik-Jurasz, A.; Ross, B.D.; Van Cauteren, M.; Collins, D.; et al. Diffusion-Weighted Magnetic Resonance Imaging as a Cancer Biomarker: Consensus and Recommendations. Neoplasia 2009, 1, 102–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Altintas, Z.; Tothill, I. Biomarkers and Biosensors for the Early Diagnosis of Lung Cancer. Sens. Actuators B Chem. 2013, 188, 988–998. [Google Scholar] [CrossRef]
  41. del Sol, A.; Balling, R.; Hood, L.; Galas, D. Diseases as Network Perturbations. Curr. Opin. Biotechnol. 2010, 21, 566–571. [Google Scholar] [CrossRef]
  42. Sarhadi, V.K.; Armengol, G. Molecular Biomarkers in Cancer. Biomolecules 2022, 12, 1021. [Google Scholar] [CrossRef]
  43. Jayanthi, V.S.A.; Das, A.B.; Saxena, U. Recent Advances in Biosensor Development for the Detection of Cancer Biomarkers. Biosens. Bioelectron. 2017, 91, 15–23. [Google Scholar] [CrossRef] [PubMed]
  44. Tothill, I.E. Biosensors for Cancer Markers Diagnosis. Semin. Cell Dev. Biol. 2009, 20, 55–62. [Google Scholar] [CrossRef]
  45. Fong, Z.V.; Winter, J.M. Biomarkers in Pancreatic Cancer: Diagnostic, Prognostic, and Predictive. Cancer J. 2012, 18, 530–538. [Google Scholar] [CrossRef] [PubMed]
  46. Basil, C.F.; Zhao, Y.; Zavaglia, K.; Jin, P.; Panelli, M.C.; Voiculescu, S.; Mandruzzato, S.; Lee, H.M.; Seliger, B.; Freedman, R.S.; et al. Common Cancer Biomarkers. Cancer Res. 2006, 66, 2953–2961. [Google Scholar] [CrossRef] [Green Version]
  47. Chowdhury, S.G.; Ray, R.; Karmakar, P. Exosomal miRNAs—a diagnostic biomarker acting as a guiding light in the diagnosis of prostate cancer. Funct. Integr. Genom. 2023, 23, 23. [Google Scholar] [CrossRef]
  48. Henry, N.L.; Hayes, D.F. Cancer Biomarkers. Mol. Oncol. 2012, 6, 140–146. [Google Scholar] [CrossRef] [Green Version]
  49. Somerfield, M.R. Clinical Practice Guidelines for the Use of Tumor Markers in Breast and Colorectal Cancer. Adopted on May 17, 1996 by the American Society of Clinical Oncology. J. Clin. Oncol. 1996, 14, 2843–2877. [Google Scholar] [CrossRef]
  50. Meindl, A. Comprehensive Analysis of 989 Patients with Breast or Ovarian Cancer Provides BRCA1 and BRCA2 Mutation Profiles and Frequencies for the German Population. Int. J. Cancer 2002, 97, 472–480. [Google Scholar] [CrossRef]
  51. Easton, D.F.; Ford, D.; Bishop, D.T.; Haites, N.; Milner, B.; Allan, L.; Easton, D.F.; Ponder, B.A.J.; Peto, J.; Smith, S.; et al. Breast and Ovarian Cancer Incidence in BRCA1-Mutation Carriers. Breast Cancer Linkage Consortium. Am. J. Hum. Genet. 1995, 56, 265. [Google Scholar]
  52. Grossman, D.C.; Curry, S.J.; Owens, D.K.; Bibbins-Domingo, K.; Caughey, A.B.; Davidson, K.W.; Doubeni, C.A.; Ebell, M.; Epling, J.W.; Kemper, A.R.; et al. Screening for Prostate Cancer USPreventive Services task force recommendation Statement. JAMA-J. Am. Med. Assoc. 2018, 319, 1901–1913. [Google Scholar] [CrossRef]
  53. Lin, K.; Lipsitz, R.; Miller, T.; Janakiraman, S. Benefits and Harms of Prostate-Specific Antigen Screening for Prostate Cancer: An Evidence Update for the U.S. Preventive Services Task Force. Ann. Intern. Med. 2008, 149, 192–199. [Google Scholar] [CrossRef]
  54. Calonge, N.; Petitti, D.B.; Dewitt, T.G.; Dietrich, A.J.; Gregory, K.D.; Harris, R.; Isham, G.J.; Lefevre, M.L.; Leipzig, R.; Loveland-Cherry, C.; et al. Screening for Prostate Cancer: U.S. Preventive Services Task Force Recommendation Statement. Ann. Intern. Med. 2008, 149, 185–191. [Google Scholar] [CrossRef]
  55. Van’t Veer, L.J.; Paik, S.; Hayes, D.F. Gene Expression Profiling of Breast Cancer: A New Tumor Marker. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2005, 23, 1631–1635. [Google Scholar] [CrossRef] [PubMed]
  56. Paik, S.; Shak, S.; Tang, G.; Kim, C.; Baker, J.; Cronin, M.; Baehner, F.L.; Walker, M.G.; Watson, D.; Park, T.; et al. A Multigene Assay to Predict Recurrence of Tamoxifen-Treated, Node-Negative Breast Cancer. N. Engl. J. Med. 2004, 351, 2817–2826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Cohen, J. Detection and Localization of Surgically Resectable Cancers with a Multi-Analyte Blood Test. Science 2018, 359, 926–930. [Google Scholar] [CrossRef] [Green Version]
  58. Cristofanilli, M.; Budd, G.T.; Ellis, M.J.; Stopeck, A.; Matera, J.; Miller, M.C.; Reuben, J.M.; Doyle, G.V.; Allard, W.J.; Terstappen, L.W.M.M.; et al. Circulating Tumor Cells, Disease Progression, and Survival in Metastatic Breast Cancer. N. Engl. J. Med. 2004, 351, 781–791. [Google Scholar] [CrossRef] [Green Version]
  59. Zhao, Q.; Li, B.; Gao, Q.; Luo, Y.; Ming, L. Prognostic Value of Epithelial-Mesenchymal Transition Circulating Tumor Cells in Female Breast Cancer: A Meta-Analysis. Front. Oncol. 2022, 12, 1024783. [Google Scholar] [CrossRef] [PubMed]
  60. Abe, S.; Kawai, K.; Nozawa, H.; Sasaki, K.; Murono, K.; Emoto, S.; Ozawa, T.; Yokoyama, Y.; Nagai, Y.; Anzai, H.; et al. Clinical Impact of Primary Tumor Sidedness and Sex on Unresectable Post-Recurrence Survival in Resected Pathological Stage II-III Colorectal Cancers: A Nationwide Multicenter Retrospective Study. BMC Cancer 2022, 22, 1–10. [Google Scholar] [CrossRef]
  61. Allegra, C.J.; Jessup, J.M.; Somerfield, M.R.; Hamilton, S.R.; Hammond, E.H.; Hayes, D.F.; McAllister, P.K.; Morton, R.F.; Schilsky, R.L. American Society of Clinical Oncology Provisional Clinical Opinion: Testing for KRAS Gene Mutations in Patients with Metastatic Colorectal Carcinoma to Predict Response to Anti-Epidermal Growth Factor Receptor Monoclonal Antibody Therapy. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2009, 27, 2091–2096. [Google Scholar] [CrossRef] [Green Version]
  62. Thuss-Patience, P.C.; Shah, M.A.; Ohtsu, A.; Van Cutsem, E.; Ajani, J.A.; Castro, H.; Mansoor, W.; Chung, H.C.; Bodoky, G.; Shitara, K.; et al. Trastuzumab Emtansine versus Taxane Use for Previously Treated HER2-Positive Locally Advanced or Metastatic Gastric or Gastro-Oesophageal Junction Adenocarcinoma (GATSBY): An International Randomised, Open-Label, Adaptive, Phase 2/3 Study. Lancet Oncol. 2017, 18, 640–653. [Google Scholar] [CrossRef] [PubMed]
  63. Bang, Y.J.; Van Cutsem, E.; Feyereislova, A.; Chung, H.C.; Shen, L.; Sawaki, A.; Lordick, F.; Ohtsu, A.; Omuro, Y.; Satoh, T.; et al. Trastuzumab in Combination with Chemotherapy versus Chemotherapy Alone for Treatment of HER2-Positive Advanced Gastric or Gastro-Oesophageal Junction Cancer (ToGA): A Phase 3, Open-Label, Randomised Controlled Trial. Lancet 2010, 376, 687–697. [Google Scholar] [CrossRef] [PubMed]
  64. Sawaki, M.; Taira, N.; Uemura, Y.; Saito, T.; Baba, S.; Kobayashi, K.; Kawashima, H.; Tsuneizumi, M.; Sagawa, N.; Bando, H.; et al. Adjuvant Trastuzumab without Chemotherapy for Treating Early HER2-Positive Breast Cancer in Older Patients: A Propensity Score-Adjusted Analysis of a Prospective Cohort Study. Breast 2022, 66, 245–254. [Google Scholar] [CrossRef] [PubMed]
  65. Piccart-Gebhart, M.J.; Procter, M.; Leyland-Jones, B.; Goldhirsch, A.; Untch, M.; Smith, I.; Gianni, L.; Baselga, J.; Bell, R.; Jackisch, C.; et al. Trastuzumab after Adjuvant Chemotherapy in HER2-Positive Breast Cancer. N. Engl. J. Med. 2005, 353, 1659–1672. [Google Scholar] [CrossRef] [Green Version]
  66. Perez, E.A.; Ballman, K.V.; Tenner, K.S.; Thompson, E.A.; Badve, S.S.; Bailey, H.; Baehner, F.L. Association of Stromal Tumor-Infiltrating Lymphocytes with Recurrence-Free Survival in the N9831 Adjuvant Trial in Patients with Early-Stage HER2-Positive Breast Cancer. JAMA Oncol. 2016, 2, 56–64. [Google Scholar] [CrossRef]
  67. Romond, E.H.; Perez, E.A.; Bryant, J.; Suman, V.J.; Geyer, C.E.; Davidson, N.E.; Tan-Chiu, E.; Martino, S.; Paik, S.; Kaufman, P.A.; et al. Trastuzumab plus Adjuvant Chemotherapy for Operable HER2-Positive Breast Cancer. N. Engl. J. Med. 2005, 353, 1673–1684. [Google Scholar] [CrossRef] [Green Version]
  68. Liu, J.; Gan, M.; Lin, Z.; Deng, Q.; Deng, J.; Zeng, B.; Shi, Y.; Ming, J. Clinical Features and Prognosis Analysis of Hormone Receptor-Positive, HER2-Negative Breast Cancer with Differential Expression Levels of Estrogen and Progesterone Receptors: A 10-Year Retrospective Study. Breast J. 2022, 2022, 5469163. [Google Scholar] [CrossRef]
  69. Abe, O.; Abe, R.; Enomoto, K.; Kikuchi, K.; Koyama, H.; Masuda, H.; Nomura, Y.; Ohashi, Y.; Sakai, K.; Sugimachi, K.; et al. Relevance of Breast Cancer Hormone Receptors and Other Factors to the Efficacy of Adjuvant Tamoxifen: Patient-Level Meta-Analysis of Randomised Trials. Lancet 2011, 378, 771–784. [Google Scholar] [CrossRef] [Green Version]
  70. Teutsch, S.M.; Bradley, L.A.; Palomaki, G.E.; Haddow, J.E.; Piper, M.; Calonge, N.; Dotson, W.D.; Douglas, M.P.; Berg, A.O. The Evaluation of Genomic Applications in Practice and Prevention (EGAPP) Initiative: Methods of the EGAPP Working Group. Genet. Med. Off. J. Am. Coll. Med. Genet. 2009, 11, 3–14. [Google Scholar] [CrossRef] [Green Version]
  71. Hanash, S.M.; Pitteri, S.J.; Faca, V.M. Mining the Plasma Proteome for Cancer Biomarkers. Nature 2008, 452, 571–579. [Google Scholar] [CrossRef]
  72. Mighton, C.; Shickh, S.; Aguda, V.; Krishnapillai, S.; Adi-Wauran, E.; Bombard, Y. From the Patient to the Population: Use of Genomics for Population Screening. Front. Genet. 2022, 13, 893832. [Google Scholar] [CrossRef]
  73. Kohsaka, S.; Tada, Y.; Ando, M.; Nakaguro, M.; Shirai, Y.; Ueno, T.; Kojima, S.; Hirai, H.; Saigusa, N.; Kano, S.; et al. Identification of Novel Prognostic and Predictive Biomarkers in Salivary Duct Carcinoma via Comprehensive Molecular Profiling. Npj Precis. Oncol. 2022, 6, 82. [Google Scholar] [CrossRef]
  74. Wang, N.; Zhang, H.; Li, D.; Jiang, C.; Zhao, H.; Teng, Y. Identification of Novel Biomarkers in Breast Cancer via Integrated Bioinformatics Analysis and Experimental Validation. Bioengineered 2021, 12, 12431. [Google Scholar] [CrossRef]
  75. Wu, M.; Li, Q.; Wang, H. Identification of Novel Biomarkers Associated with the Prognosis and Potential Pathogenesis of Breast Cancer via Integrated Bioinformatics Analysis. Technol. Cancer Res. Treat. 2021, 20, 1533033821992081. [Google Scholar] [CrossRef]
  76. Li, C.-J.; Lin, L.-T.; Chu, P.-Y.; Chiang, A.-J.; Tsai, H.-W.; Chiu, Y.-H.; Huang, M.-S.; Wen, Z.-H.; Tsui, K.-H.; Li, C.-J.; et al. Identification of Novel Biomarkers and Candidate Drug in Ovarian Cancer. J. Pers. Med. 2021, 11, 316. [Google Scholar] [CrossRef]
  77. Islam, S.U.; Ahmed, M.B.; Ahsan, H.; Lee, Y.S. Role of Biomarkers in Personalized Medicine. In Cancer Biomarkers in Diagnosis and Therapeutics; Springer: Singapore, 2022; pp. 249–275. [Google Scholar] [CrossRef]
  78. Janjua, K.A.; Shahzad, R.; Shehzad, A. Development of Novel Cancer Biomarkers for Diagnosis and Prognosis. In Cancer Biomarkers in Diagnosis and Therapeutics; Springer: Singapore, 2022; pp. 277–343. [Google Scholar] [CrossRef]
  79. Bohunicky, B.; Mousa, S.A. Biosensors: The New Wave in Cancer Diagnosis. Nanotechnol. Sci. Appl. 2011, 4, 1–10. [Google Scholar] [CrossRef] [Green Version]
  80. Prabhakar, B.; Shende, P.; Augustine, S. Current Trends and Emerging Diagnostic Techniques for Lung Cancer. Biomed. Pharmacother. 2018, 106, 1586–1599. [Google Scholar] [CrossRef]
  81. Chen, M.; Wu, D.; Tu, S.; Yang, C.; Chen, D.J.; Xu, Y. A Novel Biosensor for the Ultrasensitive Detection of the LncRNA Biomarker MALAT1 in Non-Small Cell Lung Cancer. Sci. Rep. 2021, 11, 3666. [Google Scholar] [CrossRef] [PubMed]
  82. Reddy, K.K.; Bandal, H.; Satyanarayana, M.; Goud, K.Y.; Gobi, K.V.; Jayaramudu, T.; Amalraj, J.; Kim, H. Recent Trends in Electrochemical Sensors for Vital Biomedical Markers Using Hybrid Nanostructured Materials. Adv. Sci. 2020, 7, 1902980. [Google Scholar] [CrossRef]
  83. Ferraro, S.; Bussetti, M.; Bassani, N.; Rossi, R.S.; Incarbone, G.P.; Bianchi, F.; Maggioni, M.; Runza, L.; Ceriotti, F.; Panteghini, M. Definition of Outcome-Based Prostate-Specific Antigen (PSA) Thresholds for Advanced Prostate Cancer Risk Prediction. Cancers 2021, 13, 3381. [Google Scholar] [CrossRef]
  84. Thompson, I.M.; Ankerst, D.P. Prostate-Specific Antigen in the Early Detection of Prostate Cancer. CMAJ Can. Med. Assoc. J. 2007, 176, 1853–1858. [Google Scholar] [CrossRef] [Green Version]
  85. Wu, M.; Liu, H.; Liu, Z.; Liu, C.; Zhang, A.; Li, N. Analysis of Serum Alpha-Fetoprotein (AFP) and AFP-L3 Levels by Protein Microarray. J. Int. Med. Res. 2018, 46, 4297–4305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Yuan, Q.; Song, J.; Yang, W.; Wang, H.; Huo, Q.; Yang, J.; Yu, X.; Liu, Y.; Xu, C.; Bao, H. The Effect of CA125 on Metastasis of Ovarian Cancer: Old Marker New Function. Oncotarget 2017, 8, 50015–50022. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Meyer, T.; Rustin, G.J. Role of Tumour Markers in Monitoring Epithelial Ovarian Cancer. Br. J. Cancer 2000, 82, 1535–1538. [Google Scholar] [CrossRef] [Green Version]
  88. Chianca, M.; Panichella, G.; Fabiani, I.; Giannoni, A.; L’Abbate, S.; Aimo, A.; Del Franco, A.; Vergaro, G.; Grigoratos, C.; Castiglione, V.; et al. Bidirectional Relationship Between Cancer and Heart Failure: Insights on Circulating Biomarkers. Front. Cardiovasc. Med. 2022, 9, 936654. [Google Scholar] [CrossRef]
  89. Tuxen, M.K.; Sölétormos, G.; Dombernowsky, P. Tumor Markers in the Management of Patients with Ovarian Cancer. Cancer Treat. Rev. 1995, 21, 215–245. [Google Scholar] [CrossRef] [PubMed]
  90. Smith, D.S.; Humphrey, P.A.; Catalona, W.J. The Early Detection of Prostate Carcinoma with Prostate Specific Antigen: The Washington University Experience. Cancer 1997, 80, 1852–1856. [Google Scholar] [CrossRef]
  91. Bill-Axelson, A.; Holmberg, L.; Ruutu, M.; Häggman, M.; Andersson, S.-O.; Bratell, S.; Spångberg, A.; Busch, C.; Nordling, S.; Garmo, H.; et al. Radical Prostatectomy versus Watchful Waiting in Early Prostate Cancer. N. Engl. J. Med. 2005, 352, 1977–1984. [Google Scholar] [CrossRef] [Green Version]
  92. Jaffee, E.M.; Van Dang, C.; Agus, D.B.; Alexander, B.M.; Anderson, K.C.; Ashworth, A.; Barker, A.D.; Bastani, R.; Bhatia, S.; Bluestone, J.A.; et al. Future Cancer Research Priorities in the USA: A Lancet Oncology Commission. Lancet Oncol. 2017, 18, e653–e706. [Google Scholar] [CrossRef]
  93. National Cancer Institute. Prostate-Specific Antigen (PSA) Test—NCI; National Cancer Institute: Bethesda, MD, USA, 2022. [Google Scholar]
  94. Kim, J.H.; Suh, Y.J.; Park, D.; Yim, H.; Kim, H.; Kim, H.J.; Yoon, D.S.; Hwang, K.S. Technological Advances in Electrochemical Biosensors for the Detection of Disease Biomarkers. Biomed. Eng. Lett. 2021, 11, 309–334. [Google Scholar] [CrossRef]
  95. Dutsch-Wicherek, M.; Tomaszewska, R.; Lazar, A.; Wicherek, L.; Skladzien, J. The Association between RCAS1 Expression in Laryngeal and Pharyngeal Cancer and Its Healthy Stroma with Cancer Relapse. BMC Cancer 2009, 9, 35. [Google Scholar] [CrossRef] [Green Version]
  96. Kazmierczak, W.; Lazar, A.; Tomaszewska, R.; Popiela, T.J.; Koper, K.; Wicherek, L.; Dutsch-Wicherek, M. Analysis of the Intensity of Immune Cell Infiltration and Immunoreactivity of RCAS1 in Diffuse Large B-Cell Lymphoma of the Palatine Tonsil and Its Microenvironment. Cell Tissue Res. 2015, 361, 823–831. [Google Scholar] [CrossRef] [Green Version]
  97. Kubokawa, M.; Nakashima, M.; Yao, T.; Ito, K.I.; Harada, N.; Nawata, H.; Watanabe, T. Aberrant Intracellular Localization of RCAS1 Is Associated with Tumor Progression of Gastric Cancer. Int. J. Oncol. 2001, 19, 695–700. [Google Scholar] [CrossRef] [PubMed]
  98. Naito, Y.; Okabe, Y.; Nagayama, M.; Nishinakagawa, T.; Taira, T.; Kawahara, A.; Hattori, S.; MacHida, K.; Ishida, Y.; Kaji, R.; et al. Accuracy of Differential Diagnosis for Pancreatic Cancer Is Improved in the Combination of RCAS1 and CEA Measurements and Cytology in Pancreatic Juice. Med. Mol. Morphol. 2011, 44, 86–92. [Google Scholar] [CrossRef] [PubMed]
  99. Duffy, M.J. Serum Tumor Markers in Breast Cancer: Are They of Clinical Value? Clin. Chem. 2006, 52, 345–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Hasan, D. Diagnostic Impact of CEA and CA 15-3 on Chemotherapy Monitoring of Breast Cancer Patients. J. Circ. Biomark. 2022, 11, 57–63. [Google Scholar] [CrossRef]
  101. Tampellini, M.; Berruti, A.; Gerbino, A.; Buniva, T.; Torta, M.; Gorzegno, G.; Faggiuolo, R.; Cannone, R.; Farris, A.; Destefanis, M.; et al. Relationship between CA 15-3 Serum Levels and Disease Extent in Predicting Overall Survival of Breast Cancer Patients with Newly Diagnosed Metastatic Disease. Br. J. Cancer 1997, 75, 698–702. [Google Scholar] [CrossRef]
  102. Lian, M.; Zhang, C.; Zhang, D.; Chen, P.; Yang, H.; Yang, Y.; Chen, S.; Hong, G. The Association of Five Preoperative Serum Tumor Markers and Pathological Features in Patients with Breast Cancer. J. Clin. Lab. Anal. 2019, 33, e22875. [Google Scholar] [CrossRef]
  103. Bast, J.; Ravdin, P.; Hayes, D.F.; Bates, S.; Fritsche, H.J.; Jessup, J.M.; Kemeny, N.; Locker, G.Y.; Mennel, R.G.; Somerfield, M.R. 2000 Update of Recommendations for the Use of Tumor Markers in Breast and Colorectal Cancer: Clinical Practice Guidelines of the American Society of Clinical Oncology. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2001, 19, 1865–1878. [Google Scholar] [CrossRef]
  104. Lin, M.J.; Svensson-Arvelund, J.; Lubitz, G.S.; Marabelle, A.; Melero, I.; Brown, B.D.; Brody, J.D. Cancer Vaccines: The next Immunotherapy Frontier. Nat. Cancer 2022, 3, 911–926. [Google Scholar] [CrossRef]
  105. Chow, S.; Berek, J.S.; Dorigo, O. Development of Therapeutic Vaccines for Ovarian Cancer. Vaccines 2020, 8, 657. [Google Scholar] [CrossRef] [PubMed]
  106. Jäger, E.; Chen, Y.T.; Drijfhout, J.W.; Karbach, J.; Ringhoffer, M.; Jäger, D.; Arand, M.; Wada, H.; Noguchi, Y.; Stockert, E.; et al. Simultaneous Humoral and Cellular Immune Response against Cancer-Testis Antigen NY-ESO-1: Definition of Human Histocompatibility Leukocyte Antigen (HLA)-A2-Binding Peptide Epitopes. J. Exp. Med. 1998, 187, 265–270. [Google Scholar] [CrossRef] [PubMed]
  107. Tsuji, K.; Hamada, T.; Uenaka, A.; Wada, H.; Sato, E.; Isobe, M.; Asagoe, K.; Yamasaki, O.; Shiku, H.; Ritter, G.; et al. Induction of Immune Response against NY-ESO-1 by CHP-NY-ESO-1 Vaccination and Immune Regulation in a Melanoma Patient. Cancer Immunol. Immunother. 2008, 57, 1429–1437. [Google Scholar] [CrossRef]
  108. Jäger, E.; Stockert, E.; Zidianakis, Z.; Chen, Y.T.; Karbach, J.; Jäger, D.; Arand, M.; Ritter, G.; Old, L.J.; Knuth, A. Humoral Immune Responses of Cancer Patients against “Cancer-Testis” Antigen NY-ESO-1: Correlation with Clinical Events. Int. J. Cancer 1999, 84, 506–510. [Google Scholar] [CrossRef]
  109. Karbach, J.; Neumann, A.; Atmaca, A.; Wahle, C.; Brand, K.; Von Boehmer, L.; Knuth, A.; Bender, A.; Ritter, G.; Old, L.J.; et al. Efficient in Vivo Priming by Vaccination with Recombinant NY-ESO-1 Protein and CpG in Antigen Naive Prostate Cancer Patients. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2011, 17, 861–870. [Google Scholar] [CrossRef] [Green Version]
  110. Stockert, E.; Jäger, E.; Chen, Y.T.; Scanlan, M.J.; Gout, I.; Karbach, J.; Arand, M.; Knuth, A.; Old, L.J. A Survey of the Humoral Immune Response of Cancer Patients to a Panel of Human Tumor Antigens. J. Exp. Med. 1998, 187, 1349–1354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Imai, N.; Ikeda, H.; Shiku, H. [Targeting Cancer Antigen (MAGE-A4, NY-ESO -1) for Immunotherapy]. Nihon Rinsho. Jpn. J. Clin. Med. 2012, 70, 2125–2129. [Google Scholar]
  112. Raza, A.; Merhi, M.; Inchakalody, V.P.; Krishnankutty, R.; Relecom, A.; Uddin, S.; Dermime, S. Unleashing the Immune Response to NY-ESO-1 Cancer Testis Antigen as a Potential Target for Cancer Immunotherapy. J. Transl. Med. J. Transl. Med. 2020, 18, 1–11. [Google Scholar] [CrossRef] [Green Version]
  113. Babaier, A.; Mal, H.; Alselwi, W.; Ghatage, P. Low-Grade Serous Carcinoma of the Ovary: The Current Status. Diagnostics 2022, 12, 458. [Google Scholar] [CrossRef]
  114. Neesham, D. Ovarian Cancer Screening. Aust. Fam. Physician 2007, 36, 126–128. [Google Scholar]
  115. Menon, U.; Gentry-Maharaj, A.; Hallett, R.; Ryan, A.; Burnell, M.; Sharma, A.; Lewis, S.; Davies, S.; Philpott, S.; Lopes, A.; et al. Sensitivity and Specificity of Multimodal and Ultrasound Screening for Ovarian Cancer, and Stage Distribution of Detected Cancers: Results of the Prevalence Screen of the UK Collaborative Trial of Ovarian Cancer Screening (UKCTOCS). Lancet Oncol. 2009, 10, 327–340. [Google Scholar] [CrossRef] [Green Version]
  116. Wong, N.K.; Easton, R.L.; Panico, M.; Sutton-Smith, M.; Morrison, J.C.; Lattanzio, F.A.; Morris, H.R.; Clark, G.F.; Dell, A.; Patankar, M.S. Characterization of the Oligosaccharides Associated with the Human Ovarian Tumor Marker CA125. J. Biol. Chem. 2003, 278, 28619–28634. [Google Scholar] [CrossRef] [Green Version]
  117. White, B.; Patterson, M.; Karnwal, S.; Brooks, C.L. Crystal Structure of a Human MUC16 SEA Domain Reveals Insight into the Nature of the CA125 Tumor Marker. Proteins Struct. Funct. Bioinform. 2022, 90, 1210–1218. [Google Scholar] [CrossRef]
  118. Wang, Y.-S.; Ren, S.-F.; Jiang, W.; Lu, J.-Q.; Zhang, X.-Y.; Li, X.-P.; Cao, R.; Xu, C.-J. CA125-Tn ELISA Assay Improves Specificity of Pre-Operative Diagnosis of Ovarian Cancer among Patients with Elevated Serum CA125 Levels. Ann. Transl. Med. 2021, 9, 788. [Google Scholar] [CrossRef] [PubMed]
  119. Seelenmeyer, C.; Wegehingel, S.; Lechner, J.; Nickel, W. The Cancer Antigen CA125 Represents a Novel Counter Receptor for Galectin-1. J. Cell Sci. 2003, 116, 1305–1318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Liang, L.; Fang, J.; Han, X.; Zhai, X.; Song, Y.; Lu, Y.; Zhang, Q.; Ma, R. Prognostic Value of CEA, CA19-9, CA125, CA724, and CA242 in Serum and Ascites in Pseudomyxoma Peritonei. Front. Oncol. 2021, 11, 594763. [Google Scholar] [CrossRef] [PubMed]
  121. Li, B.; Li, Y.; Li, C.; Yang, J.; Liu, D.; Wang, H.; Xu, R.; Zhang, Y.; Wei, Q. An ultrasensitive split-type electrochemical immunosensor based on controlled-release strategy for detection of CA19-9. Biosens. Bioelectron. 2023, 227, 115180. [Google Scholar] [CrossRef]
  122. Poruk, K.E.; Firpo, M.A.; Adler, D.G.; Mulvihill, S.J. Screening for Pancreatic Cancer: Why, How, and Who? Ann. Surg. 2013, 257, 17–26. [Google Scholar] [CrossRef] [Green Version]
  123. Badheeb, M.; Abdelrahim, A.; Esmail, A.; Umoru, G.; Abboud, K.; Al-Najjar, E.; Rasheed, G.; Alkhulaifawi, M.; Abudayyeh, A.; Abdelrahim, M. Pancreatic Tumorigenesis: Precursors, Genetic Risk Factors and Screening. Curr. Oncol. 2022, 29, 8693–8719. [Google Scholar] [CrossRef] [PubMed]
  124. Hatzakis, K.D.; Froudarakis, M.E.; Bouros, D.; Tzanakis, N.; Karkavitsas, N.; Siafakas, N.M. Prognostic Value of Serum Tumor Markers in Patients with Lung Cancer. Respiration 2002, 69, 25–29. [Google Scholar] [CrossRef]
  125. Ferrigno, D.; Buccheri, G.; Giordano, C. Neuron-Specific Enolase Is an Effective Tumour Marker in Non-Small Cell Lung Cancer (NSCLC). Lung Cancer 2003, 41, 311–320. [Google Scholar] [CrossRef] [PubMed]
  126. Gazdar, A.F.; Carney, D.N.; Nau, M.M.; Minna, J.D. Characterization of Variant Subclasses of Cell Lines Derived from Small Cell Lung Cancer Having Distinctive Biochemical, Morphological, and Growth Properties. Cancer Res. 1985, 45, 2924–2930. [Google Scholar] [PubMed]
  127. Isgrò, M.A.; Bottoni, P.; Scatena, R. Neuron-Specifi c Enolase as a Biomarker: Biochemical and Clinical Aspects. Adv. Exp. Med. Biol. 2015, 867, 125–143. [Google Scholar] [CrossRef]
  128. Tian, Z.; Liang, C.; Zhang, Z.; Wen, H.; Feng, H.; Ma, Q.; Liu, D.; Qiang, G. Prognostic Value of Neuron-Specific Enolase for Small Cell Lung Cancer: A Systematic Review and Meta-Analysis. World J. Surg. Oncol. 2020, 18, 1–8. [Google Scholar] [CrossRef] [PubMed]
  129. Campagnari, F.; Bombardieri, E.; de Braud, F.; Baldini, L.; Maiolo, A.T. Terminal Deoxynucleotidyl Transferase, Tdt, as a Marker for Leukemia and Lymphoma Cells. Int. J. Biol. Markers 1987, 2, 31–42. [Google Scholar] [CrossRef] [PubMed]
  130. Peng, Y.; Shang, H.; Zheng, Z.; Li, H.; Chen, W.; Xu, J. Self-activation of symbiotic multi-DNA machines transducing exponentially amplified fluorescence for ultrasensitively signaling of terminal deoxynucleotidyl transferase activity. Sens. Actuators B. Chem 2023, 380, 133400. [Google Scholar] [CrossRef]
  131. Sur, M.; AlArdati, H.; Ross, C.; Alowami, S. TdT Expression in Merkel Cell Carcinoma: Potential Diagnostic Pitfall with Blastic Hematological Malignancies and Expanded Immunohistochemical Analysis. Mod. Pathol. 2007, 20, 1113–1120. [Google Scholar] [CrossRef] [Green Version]
  132. Mathewson, R.C.; Kjeldsberg, C.R.; Perkins, S.L. Detection of Terminal Deoxynucleotidyl Transferase (TdT) in Nonhematopoietic Small Round Cell Tumors of Children. Pediatr. Pathol. Lab. Med. J. Soc. Pediatr. Pathol. Affil. Int. Paediatr. Pathol. Assoc. 1997, 17, 835–844. [Google Scholar]
  133. Chapman, M.H.; Sandanayake, N.S.; Andreola, F.; Dhar, D.K.; Webster, G.J.; Dooley, J.S.; Pereira, S.P. Circulating CYFRA 21-1 Is a Specific Diagnostic and Prognostic Biomarker in Biliary Tract Cancer. J. Clin. Exp. Hepatol. 2011, 1, 6–12. [Google Scholar] [CrossRef]
  134. Zhang, L.; Liu, D.; Li, L.; Pu, D.; Zhou, P.; Jing, Y.; Yu, H.; Wang, Y.; Zhu, Y.; He, Y.; et al. The Important Role of Circulating CYFRA21-1 in Metastasis Diagnosis and Prognostic Value Compared with Carcinoembryonic Antigen and Neuron-Specific Enolase in Lung Cancer Patients. BMC Cancer 2017, 17, 96. [Google Scholar] [CrossRef] [Green Version]
  135. Dal Bello, M.G.; Filiberti, R.A.; Alama, A.; Orengo, A.M.; Mussap, M.; Coco, S.; Vanni, I.; Boccardo, S.; Rijavec, E.; Genova, C.; et al. The Role of CEA, CYFRA21-1 and NSE in Monitoring Tumor Response to Nivolumab in Advanced Non-Small Cell Lung Cancer (NSCLC) Patients. J. Transl. Med. 2019, 17, 74. [Google Scholar] [CrossRef] [Green Version]
  136. Rapellino, M.; Niklinski, J.; Pecchio, F.; Furman, M.; Baldi, S.; Chyczewski, L.; Ruffini, E.; Chyczewska, E. CYFRA 21-1 as a Tumour Marker for Bronchogenic Carcinoma. Eur. Respir. J. 1995, 8, 407–410. [Google Scholar] [CrossRef] [Green Version]
  137. Dall’Olio, F.G.; Abbati, F.; Facchinetti, F.; Massucci, M.; Melotti, B.; Squadrilli, A.; Buti, S.; Formica, F.; Tiseo, M.; Ardizzoni, A. CEA and CYFRA 21-1 as Prognostic Biomarker and as a Tool for Treatment Monitoring in Advanced NSCLC Treated with Immune Checkpoint Inhibitors. Ther. Adv. Med. Oncol. 2020, 12, 1758835920952994. [Google Scholar] [CrossRef]
  138. Hayakawa, T.; Naruse, S.; Kitagawa, M.; Ishiguro, H.; Kondo, T.; Kurimoto, K.; Fukushima, M.; Takayama, T.; Horiguchi, Y.; Kuno, N.; et al. A Prospective Multicenter Trial Evaluating Diagnostic Validity of Multivariate Analysis and Individual Serum Marker in Differential Diagnosis of Pancreatic Cancer from Benign Pancreatic Diseases. Int. J. Pancreatol. Off. J. Int. Assoc. Pancreatol. 1999, 25, 23–29. [Google Scholar] [CrossRef]
  139. Chen, G.; Yu, M.; Cao, J.; Zhao, H.; Dai, Y.; Cong, Y.; Qiao, G. Identification of Candidate Biomarkers Correlated with Poor Prognosis of Breast Cancer Based on Bioinformatics Analysis. Bioengineered 2021, 12, 5149–5161. [Google Scholar] [CrossRef]
  140. Kumar, S.; Mohan, A.; Guleria, R. Biomarkers in Cancer Screening, Research and Detection: Present and Future: A Review. Biomarkers 2008, 11, 385–405. [Google Scholar] [CrossRef]
  141. Chatterjee, S.K.; Zetter, B.R. Cancer Biomarkers: Knowing the Present and Predicting the Future. Future Oncol. 2005, 1, 37–50. [Google Scholar] [CrossRef]
  142. Hayes, D.F. Biomarker Validation and Testing. Mol. Oncol. 2015, 9, 960–966. [Google Scholar] [CrossRef]
  143. Hayes, D.F.; Bast, R.C.; Desch, C.E.; Fritsche, H.; Kemeny, N.E.; Jessup, J.M.; Locker, G.Y.; Macdonald, J.S.; Mennel, R.G.; Norton, L.; et al. Tumor Marker Utility Grading System: A Framework to Evaluate Clinical Utility of Tumor Markers. J. Natl. Cancer Inst. 1996, 88, 1456–1466. [Google Scholar] [CrossRef] [Green Version]
  144. Hamanishi, J.; Mandai, M.; Ikeda, T.; Minami, M.; Kawaguchi, A.; Murayama, T.; Kanai, M.; Mori, Y.; Matsumoto, S.; Chikuma, S.; et al. Safety and Antitumor Activity of Anti-PD-1 Antibody, Nivolumab, in Patients with Platinum-Resistant Ovarian Cancer. J. Clin. Oncol. 2015, 33, 4015–4022. [Google Scholar] [CrossRef]
  145. Johnson, P.J.; Williams, R.; Johnson, P.J. Serum Alpha-Fetoprotein Estimations and Doubling Time in Hepatocellular Carcinoma: Influence of Therapy and Possible Value in Early Detection. J. Natl. Cancer Inst. 1980, 64, 1329–1332. [Google Scholar] [CrossRef] [PubMed]
  146. Ye, K.; Fan, Q.; Yuan, M.; Wang, D.; Xiao, L.; Long, G.; Chen, R.; Fang, T.; Li, Z.; Zhou, L. Prognostic Value of Postoperative Circulating Tumor DNA in Patients with Early- and Intermediate-Stage Hepatocellular Carcinoma. Front. Oncol. 2022, 12, 834992. [Google Scholar] [CrossRef] [PubMed]
  147. Chen, L.; Ho, D.W.Y.; Lee, N.P.Y.; Sun, S.; Lam, B.; Wong, K.F.; Yi, X.; Lau, G.K.; Ng, E.W.Y.; Poon, T.C.W.; et al. Enhanced Detection of Early Hepatocellular Carcinoma by Serum SELDI-TOF Proteomic Signature Combined with Alpha-Fetoprotein Marker. Ann. Surg. Oncol. 2010, 17, 2518–2525. [Google Scholar] [CrossRef] [Green Version]
  148. Vickers, A.J.; Brewster, S.F. PSA Velocity and Doubling Time in Diagnosis and Prognosis of Prostate Cancer. Br. J. Med. Surg. Urol. 2012, 5, 162–168. [Google Scholar] [CrossRef] [Green Version]
  149. Bjartell, A.S. Next-Generation Prostate-Specific Antigen Test: Ready to Use? Eur. Urol. 2013, 64, 700–702. [Google Scholar] [CrossRef]
  150. Dunn, M.W. Prostate Cancer Screening. Semin. Oncol. Nurs. 2017, 33, 156–164. [Google Scholar] [CrossRef] [PubMed]
  151. Molina, R.; Barak, V.; Van Dalen, A.; Duffy, M.J.; Einarsson, R.; Gion, M.; Goike, H.; Lamerz, R.; Nap, M.; Sölétormos, G.; et al. Tumor Markers in Breast Cancer- European Group on Tumor Markers Recommendations. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2005, 26, 281–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Leja, M.; Linē, A. Early Detection of Gastric Cancer beyond Endoscopy—New Methods. Best Pract. Res. Clin. Gastroenterol. 2021, 50–51, 101731. [Google Scholar] [CrossRef]
  153. Jensen, J.L.; Maclean, G.D.; Suresh, M.R.; Almeida, A.; Jette, D.; Lloyd, S.; Bodnar, D.; Krantz, M.; Longenecker, B.M. Possible Utility of Serum Determinations of CA 125 and CA 27.29 in Breast Cancer Management. Int. J. Biol. Markers 1991, 6, 1–6. [Google Scholar] [CrossRef]
  154. Ludwig, J.A.; Weinstein, J.N. Biomarkers in Cancer Staging, Prognosis and Treatment Selection. Nat. Rev. Cancer 2005, 5, 845–856. [Google Scholar] [CrossRef]
  155. von Kleist, S. The Clinical Value of the Tumor Markers CA 19/9 and Carcinoembryonic Antigen (CEA) in Colorectal Carcinomas: A Critical Comparison. Int. J. Biol. Markers 1986, 1, 3–8. [Google Scholar] [CrossRef] [PubMed]
  156. Koizumi, F.; Odagiri, H.; Fujimoto, H.; Kawamura, T.; Ishimori, A. [Clinical evaluation of four tumor markers (CEA, TPA, CA50 and CA72-4) in colorectal cancer]. Rinsho Byori. Jpn. J. Clin. Pathol. 1992, 40, 523–528. [Google Scholar]
  157. Bae, Y.J.; Schaab, M.; Kratzsch, J. Calcitonin as Biomarker for the Medullary Thyroid Carcinoma. Recent Results Cancer Res. Fortschr. Der Krebsforschung. Prog. Dans Les Rech. Sur Le Cancer 2015, 204, 117–137. [Google Scholar] [CrossRef]
  158. Paik, S.; Tang, G.; Shak, S.; Kim, C.; Baker, J.; Kim, W.; Cronin, M.; Baehner, F.L.; Watson, D.; Bryant, J.; et al. Gene Expression and Benefit of Chemotherapy in Women with Node-Negative, Estrogen Receptor-Positive Breast Cancer. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2006, 24, 3726–3734. [Google Scholar] [CrossRef] [PubMed]
  159. Porras, L.; Gorse, F.; Thiombane, N.K.; Gaboury, L.; Mader, S. CAXII Is a Surrogate Marker for Luminal Breast Tumors Regulated by ER and GATA3. Cancers 2022, 14, 5453. [Google Scholar] [CrossRef]
  160. McGuire, W.L.; Horwitz, K.B.; Pearson, O.H.; Segaloff, A. Current Status of Estrogen and Progesterone Receptors in Breast Cancer. Cancer 1977, 39, 2934–2947. [Google Scholar] [CrossRef]
  161. Varadan, V.; Kamalakaran, S.; Gilmore, H.; Banerjee, N.; Janevski, A.; Miskimen, K.L.S.; Williams, N.; Basavanhalli, A.; Madabhushi, A.; Lezon-Geyda, K.; et al. Brief-Exposure to Preoperative Bevacizumab Reveals a TGF-β Signature Predictive of Response in HER2-Negative Breast Cancers. Int. J. Cancer 2016, 138, 747–757. [Google Scholar] [CrossRef]
  162. Alexandrov, L.B.; Nik-Zainal, S.; Wedge, D.C.; Aparicio, S.A.J.R.; Behjati, S.; Biankin, A.V.; Bignell, G.R.; Bolli, N.; Borg, A.; Børresen-Dale, A.L.; et al. Signatures of Mutational Processes in Human Cancer. Nature 2013, 500, 415–421. [Google Scholar] [CrossRef] [Green Version]
  163. Bhatt, A.N.; Mathur, R.; Farooque, A.; Verma, A.; Dwarakanath, B.S. Cancer Biomarkers—Current Perspectives. Indian J. Med. Res. 2010, 132, 129–149. [Google Scholar]
  164. Perrier D’Hauterive, S.; Berndt, S.; Tsampalas, M.; Charlet-Renard, C.; Dubois, M.; Bourgain, C.; Hazout, A.; Foidart, J.M.; Geenen, V. Dialogue between Blastocyst HCG and Endometrial LH/HCG Receptor: Which Role in Implantation? Gynecol. Obstet. Investig. 2007, 64, 156–160. [Google Scholar] [CrossRef]
  165. Kurtzman, J.T.; Wilson, H.; Rao, C.V. A Proposed Role for HCG in Clinical Obstetrics. Semin. Reprod. Med. 2001, 19, 63–68. [Google Scholar] [CrossRef] [PubMed]
  166. Adewole, I.F.; Newlands, E.S. Neuron-Specific Enolase (NSE) as a Tumour Marker and Comparative Evaluation with Carcinoembryonic Antigen (CEA) in Small-Cell Lung Cancer. Med. Oncol. Tumor Pharmacother. 1987, 4, 11–15. [Google Scholar] [CrossRef] [PubMed]
  167. Darmawan, B.; Sari, M.; Susilo, S.; Kartamihardja, A.H.S. Preradioactive Iodine Thyroglobulin Levels as Predictors of Metastasis in Well-Differentiated Thyroid Carcinoma Patients. World J. Nucl. Med. 2022, 21, 296–301. [Google Scholar] [CrossRef]
  168. Mazzaferri, E.L.; Robbins, R.J.; Spencer, C.A.; Braverman, L.E.; Pacini, F.; Wartofsky, L.; Haugen, B.R.; Sherman, S.I.; Cooper, D.S.; Braunstein, G.D.; et al. A Consensus Report of the Role of Serum Thyroglobulin as a Monitoring Method for Low-Risk Patients with Papillary Thyroid Carcinoma. J. Clin. Endocrinol. Metab. 2003, 88, 1433–1441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Van Gils, M.P.M.Q.; Cornel, E.B.; Hessels, D.; Peelen, W.P.; Witjes, J.A.; Mulders, P.F.A.; Rittenhouse, H.G.; Schalken, J.A. Molecular PCA3 Diagnostics on Prostatic Fluid. Prostate 2007, 67, 881–887. [Google Scholar] [CrossRef] [PubMed]
  170. Thomas, R.; Al-Khadairi, G.; Roelands, J.; Hendrickx, W.; Dermime, S.; Bedognetti, D.; Decock, J. NY-ESO-1 Based Immunotherapy of Cancer: Current Perspectives. Front. Immunol. 2018, 9, 947. [Google Scholar] [CrossRef]
  171. Ellison, G.; Zhu, G.; Moulis, A.; Dearden, S.; Speake, G.; McCormack, R. EGFR Mutation Testing in Lung Cancer: A Review of Available Methods and Their Use for Analysis of Tumour Tissue and Cytology Samples. J. Clin. Pathol. 2013, 66, 79–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Da Cunha Santos, G.; Shepherd, F.A.; Tsao, M.S. EGFR Mutations and Lung Cancer. Annu. Rev. Pathol. 2011, 6, 49–69. [Google Scholar] [CrossRef] [Green Version]
  173. Lee, B.; Lee, T.; Lee, S.H.; Choi, Y.L.; Han, J. Clinicopathologic Characteristics of EGFR, KRAS, and ALK Alterations in 6595 Lung Cancers. Oncotarget 2016, 7, 23874. [Google Scholar] [CrossRef]
  174. Warth, A.; Penzel, R.; Lindenmaier, H.; Brandt, R.; Stenzinger, A.; Herpel, E.; Goeppert, B.; Thomas, M.; Herth, F.J.F.; Dienemann, H.; et al. EGFR, KRAS, BRAF and ALK Gene Alterations in Lung Adenocarcinomas: Patient Outcome, Interplay with Morphology and Immunophenotype. Eur. Respir. J. 2014, 43, 872–883. [Google Scholar] [CrossRef] [Green Version]
  175. Pierce, J.M.R.; Mehta, A. Diagnostic, Prognostic and Therapeutic Role of CD30 in Lymphoma. Expert Rev. Hematol. 2017, 10, 29–37. [Google Scholar] [CrossRef] [PubMed]
  176. van der Weyden, C.A.; Pileri, S.A.; Feldman, A.L.; Whisstock, J.; Prince, H.M. Understanding CD30 Biology and Therapeutic Targeting: A Historical Perspective Providing Insight into Future Directions. Blood Cancer J. 2017, 7, e603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Kumar, A.; Kumar, R.; Gupta, N.P. Comparison of NMP22 BladderChek Test and Urine Cytology for the Detection of Recurrent Bladder Cancer. Jpn. J. Clin. Oncol. 2006, 36, 172–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Kibar, Y.; Goktas, S.; Kilic, S.; Yaman, H.; Onguru, O.; Peker, A.F. Prognostic Value of Cytology, Nuclear Matrix Protein 22 (NMP22) Test, and Urinary Bladder Cancer II (UBC II) Test in Early Recurrent Transitional Cell Carcinoma of the Bladder. Ann. Clin. Lab. Sci. 2006, 36, 31–38. [Google Scholar] [PubMed]
  179. Mei, X.; Zhu, X.; Zuo, L.; Wu, H.; Guo, M.; Liu, C. Predictive Significance of CYFRA21-1, Squamous Cell Carcinoma Antigen and Carcinoembryonic Antigen for Lymph Node Metastasis in Patients with Esophageal Squamous Cancer. Int. J. Biol. Markers 2019, 34, 200–204. [Google Scholar] [CrossRef] [Green Version]
  180. Nakamura, T.; Ide, H.; Eguchi, R.; Hayashi, K.; Takasaki, K.; Watanabe, S. CYFRA 21–1 as a Tumor Marker for Squamous Cell Carcinoma of the Esophagus. Dis. Esophagus 1998, 11, 35–39. [Google Scholar] [CrossRef]
  181. Brockmann, J.G.; St Nottberg, H.; Glodny, B.; Heinecke, A.; Senninger, N.J. CYFRA 21-1 Serum Analysis in Patients with Esophageal Cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2000, 6, 4249–4252. [Google Scholar]
  182. McCarthy, B.A.; Boyle, E.; Xue, P.W.; Guzowski, D.; Paul, S.; Catera, R.; Trott, J.; Yan, X.J.; Croce, C.M.; Damle, R.; et al. Surface Expression of Bcl-2 in Chronic Lymphocytic Leukemia and Other B-Cell and Lymphomas without a Breakpoint t(14;18). Mol. Med. 2008, 14, 618–627. [Google Scholar] [CrossRef] [Green Version]
  183. Dawson, S.J.; Makretsov, N.; Blows, F.M.; Driver, K.E.; Provenzano, E.; Le Quesne, J.; Baglietto, L.; Severi, G.; Giles, G.G.; McLean, C.A.; et al. BCL2 in Breast Cancer: A Favourable Prognostic Marker across Molecular Subtypes and Independent of Adjuvant Therapy Received. Br. J. Cancer 2010, 103, 668–675. [Google Scholar] [CrossRef]
  184. Kulsoom, B.; Shamsi, T.S.; Afsar, N.A.; Memon, Z.; Ahmed, N.; Hasnain, S.N. Bax, Bcl-2, and Bax/Bcl-2 as Prognostic Markers in Acute Myeloid Leukemia: Are We Ready for Bcl-2-Directed Therapy? Cancer Manag. Res. 2018, 10, 403–416. [Google Scholar] [CrossRef] [Green Version]
  185. Gaafar, T.M.; Raafat, I.I.; Aly, A.A.; Abd El-Ghaffar Mohamed, N.; Farid, R.J.; Saad, N.E.; El-Hawary, R.; Mostafaa, N.; Ahmed, M.M. Detection of BCR/ABL Translocation in Bone Marrow Derived Mesenchymal Stem Cells in Egyptian CML Patients. Open Access Maced. J. Med. Sci. 2015, 3, 231–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Wu, S.Q.; Voelkerding, K.V.; Sabatini, L.; Chen, X.R.; Huang, J.; Meisner, L.F. Extensive Amplification of Bcr/Abl Fusion Genes Clustered on Three Marker Chromosomes in Human Leukemic Cell Line K-562. Leukemia 1995, 9, 858–862. [Google Scholar] [PubMed]
  187. Boyd, S.D.; Natkunam, Y.; Allen, J.R.; Warnke, R.A. Selective Immunophenotyping for Diagnosis of B-Cell Neoplasms: Immunohistochemistry and Flow Cytometry Strategies and Results. Appl. Immunohistochem. Mol. Morphol. 2013, 21, 116–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Matsushita, K.; Margulies, I.; Onda, M.; Nagata, S.; Stetler-Stevenson, M.; Kreitman, R.J. Soluble CD22 as a Tumor Marker for Hairy Cell Leukemia. Blood 2008, 112, 2272–2277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Li, J.; Ran, Q.; Xu, B.; Luo, X.; Song, S.; Xu, D.; Zhang, X. Role of CD25 Expression on Prognosis of Acute Myeloid Leukemia: A Literature Review and Meta-Analysis. PLOS ONE 2020, 15, e0236124. [Google Scholar] [CrossRef] [PubMed]
  190. Casadei, C.; Dizman, N.; Schepisi, G.; Cursano, M.C.; Basso, U.; Santini, D.; Pal, S.K.; De Giorgi, U. Targeted Therapies for Advanced Bladder Cancer: New Strategies with FGFR Inhibitors. Ther. Adv. Med. Oncol. 2019, 11, 1758835919890285. [Google Scholar] [CrossRef] [Green Version]
  191. Knowles, M.A. Role of FGFR3 in Urothelial Cell Carcinoma: Biomarker and Potential Therapeutic Target. World J. Urol. 2007, 25, 581–593. [Google Scholar] [CrossRef] [Green Version]
  192. Topsakal, M.; Karadeniz, T.; Anaç, M.; Dönmezer, S.; Besisik, A. Assessment of Fibrin-Fibrinogen Degradation Products (Accu-Dx) Test in Bladder Cancer Patients. Eur. Urol. 2001, 39, 287–291. [Google Scholar] [CrossRef]
  193. Glas, A.S.; Roos, D.; Deutekom, M.; Zwinderman, A.H.; Bossuyt, P.M.M.; Kurth, K.H. Tumor Markers in the Diagnosis of Primary Bladder Cancer. A Systematic Review. J. Urol. 2003, 169, 1975–1982. [Google Scholar] [CrossRef]
  194. Robinson, B.W.S.; Creaney, J.; Lake, R.; Nowak, A.; Musk, A.W.; De Klerk, N.; Winzell, P.; Hellstrom, K.E.; Hellstrom, I. Soluble Mesothelin-Related Protein—A Blood Test for Mesothelioma. Lung Cancer 2005, 49, S109–S111. [Google Scholar] [CrossRef]
  195. Kaeding, A.J.; Barwe, S.P.; Gopalakrishnapillai, A.; Ries, R.E.; Alonzo, T.A.; Gerbing, R.B.; Correnti, C.; Loken, M.R.; Broderson, L.E.; Pardo, L.; et al. Mesothelin Is a Novel Cell Surface Disease Marker and Potential Therapeutic Target in Acute Myeloid Leukemia. Blood Adv. 2021, 5, 2350–2361. [Google Scholar] [CrossRef] [PubMed]
  196. Bubendorf, L.; Büttner, R.; Al-Dayel, F.; Dietel, M.; Elmberger, G.; Kerr, K.; López-Ríos, F.; Marchetti, A.; Öz, B.; Pauwels, P.; et al. Testing for ROS1 in Non-Small Cell Lung Cancer: A Review with Recommendations. Virchows Arch. Int. J. Pathol. 2016, 469, 489–503. [Google Scholar] [CrossRef] [PubMed]
  197. Nolen, B.M.; Lokshin, A.E. Biomarker Testing for Ovarian Cancer: Clinical Utility of Multiplex Assays. Mol. Diagn. Ther. 2013, 17, 139–146. [Google Scholar] [CrossRef] [Green Version]
  198. Jelski, W.; Mroczko, B. Molecular and Circulating Biomarkers of Brain Tumors. Int. J. Mol. Sci. 2021, 22, 7039. [Google Scholar] [CrossRef] [PubMed]
  199. Ouhibi, A.; Raouafi, A.; Lorrain, N.; Guendouz, M.; Raouafi, N.; Moadhen, A. Functionalized SERS Substrate Based on Silicon Nanowires for Rapid Detection of Prostate Specific Antigen. Sens. Actuators B Chem. 2021, 330, 129352. [Google Scholar] [CrossRef]
  200. Mishra, S.; Kim, E.-S.; Sharma, P.K.; Wang, Z.-J.; Yang, S.-H.; Kaushik, A.K.; Wang, C.; Li, Y.; Kim, N.-Y. Tailored Biofunctionalized Biosensor for the Label-Free Sensing of Prostate-Specific Antigen. ACS Appl. Bio Mater. 2020, 3, 7821–7830. [Google Scholar] [CrossRef]
  201. Ribeiro, J.A.; Sales, M.G.F.; Pereira, C.M. Electrochemistry-Assisted Surface Plasmon Resonance Biosensor for Detection of CA 15–3. Anal. Chem. 2021, 93, 7815–7824. [Google Scholar] [CrossRef]
  202. Yu, H.; Tan, X.; Sun, S.; Zhang, L.; Gao, C.; Ge, S. Engineering Paper-Based Visible Light-Responsive Sn-Self Doped Domed SnO2 Nanotubes for Ultrasensitive Photoelectrochemical Sensor. Biosens. Bioelectron. 2021, 185, 113250. [Google Scholar] [CrossRef]
  203. Wu, Q.; Li, N.; Wang, Y.; Liu, Y.; Xu, Y.; Wei, S.; Wu, J.; Jia, G.; Fang, X.; Chen, F.; et al. A 2D Transition Metal Carbide MXene-Based SPR Biosensor for Ultrasensitive Carcinoembryonic Antigen Detection. Biosens. Bioelectron. 2019, 144, 111697. [Google Scholar] [CrossRef]
  204. Mo, G.; He, X.; Zhou, C.; Ya, D.; Feng, J.; Yu, C.; Deng, B. A Novel ECL Sensor Based on a Boronate Affinity Molecular Imprinting Technique and Functionalized SiO2@CQDs/AuNPs/MPBA Nanocomposites for Sensitive Determination of Alpha-Fetoprotein. Biosens. Bioelectron. 2019, 126, 558–564. [Google Scholar] [CrossRef]
  205. Li, S.; Liu, X.; Liu, S.; Guo, M.; Liu, C.; Pei, M. Fluorescence Sensing Strategy Based on Aptamer Recognition and Mismatched Catalytic Hairpin Assembly for Highly Sensitive Detection of Alpha-Fetoprotein. Anal. Chim. Acta 2021, 1141, 21–27. [Google Scholar] [CrossRef] [PubMed]
  206. Mo, G.; He, X.; Qin, D.; Meng, S.; Wu, Y.; Deng, B. Spatially-Resolved Dual-Potential Sandwich Electrochemiluminescence Immunosensor for the Simultaneous Determination of Carbohydrate Antigen 19–9 and Carbohydrate Antigen 24-2. Biosens. Bioelectron. 2021, 178, 113024. [Google Scholar] [CrossRef] [PubMed]
  207. Bahari, D.; Babamiri, B.; Salimi, A. An Eco-Friendly MIP-Solid Surface Fluorescence Immunosensor for Detection of CA 19-9 Tumor Marker Using Ni Nanocluster as an Emitter Labels. J. Iran. Chem. Soc. 2020, 17, 2283–2291. [Google Scholar] [CrossRef]
  208. Mansouri Majd, S.; Salimi, A. Ultrasensitive Flexible FET-Type Aptasensor for CA 125 Cancer Marker Detection Based on Carboxylated Multiwalled Carbon Nanotubes Immobilized onto Reduced Graphene Oxide Film. Anal. Chim. Acta 2018, 1000, 273–282. [Google Scholar] [CrossRef]
  209. Szymańska, B.; Lukaszewski, Z.; Hermanowicz-Szamatowicz, K.; Gorodkiewicz, E. A Biosensor for Determination of the Circulating Biomarker CA125/MUC16 by Surface Plasmon Resonance Imaging. Talanta 2020, 206, 120187. [Google Scholar] [CrossRef]
  210. Bai, Y.; Zhang, H.; Zhao, L.; Wang, Y.; Chen, X.; Zhai, H.; Tian, M.; Zhao, R.; Wang, T.; Xu, H.; et al. A Novel Aptasensor Based on HCR and G-Quadruplex DNAzyme for Fluorescence Detection of Carcinoembryonic Antigen. Talanta 2021, 221, 121451. [Google Scholar] [CrossRef]
  211. Yılmaz, M.; Bakhshpour, M.; Göktürk, I.; Pişkin, A.K.; Denizli, A. Quartz Crystal Microbalance (QCM) Based Biosensor Functionalized by HER2/Neu Antibody for Breast Cancer Cell Detection. Chemosensors 2021, 9, 80. [Google Scholar] [CrossRef]
  212. Tanaka, S.; Wadati, H.; Sato, K.; Yasuda, H.; Niioka, H. Red-Fluorescent Pt Nanoclusters for Detecting and Imaging HER2 in Breast Cancer Cells. ACS Omega 2020, 5, 23718–23723. [Google Scholar] [CrossRef]
  213. Xie, Y.; Si, S.; Deng, C.; Wu, H. A Sensitive “Turn-on” Fluorescent Biosensor for the Detection of Biomarkers Based on Three-Dimensional DNA Walker. Talanta 2023, 253, 124071. [Google Scholar] [CrossRef]
  214. Hu, S.; Xu, H.; Zhou, B.; Xu, S.; Shen, B.; Dong, B.; Yin, Z.; Xu, S.; Sun, L.; Lv, J.; et al. Double Stopband Bilayer Photonic Crystal Based Upconversion Fluorescence PSA Sensor. Sens. Actuators B Chem. 2021, 326, 128816. [Google Scholar] [CrossRef]
  215. Wu, Y.; Chen, X.; Wang, X.; Yang, M.; Xu, F.; Hou, C.; Huo, D. A Fluorescent Biosensor Based on Prismatic Hollow Metal-Polydopamine Frameworks and 6-Carboxyfluorescein (FAM)-Labeled Protein Aptamer for CA15-3 Detection. Sens. Actuators B Chem. 2021, 329, 129249. [Google Scholar] [CrossRef]
  216. Othman, H.O.; Salehnia, F.; Fakhri, N.; Hassan, R.; Hosseini, M.; Faizullah, A.; Ganjali, M.R.; Kazem Aghamir, S.M. A Highly Sensitive Fluorescent Immunosensor for Sensitive Detection of Nuclear Matrix Protein 22 as Biomarker for Early Stage Diagnosis of Bladder Cancer. RSC Adv. 2020, 10, 28865–28871. [Google Scholar] [CrossRef] [PubMed]
  217. Othman, H.O.; Salehnia, F.; Hosseini, M.; Hassan, R.; Faizullah, A.; Ganjali, M.R. Fluorescence Immunoassay Based on Nitrogen Doped Carbon Dots for the Detection of Human Nuclear Matrix Protein NMP22 as Biomarker for Early Stage Diagnosis of Bladder Cancer. Microchem. J. 2020, 157, 104966. [Google Scholar] [CrossRef]
  218. Satyanarayana, M.; Yugender Goud, K.; Koteshwara Reddy, K.; Vengatajalabathy Gobi, K. Conducting Polymer-Layered Carbon Nanotube as Sensor Interface for Electrochemical Detection of Dacarbazine In-Vitro. Electrocatalysis 2017, 8, 214–223. [Google Scholar] [CrossRef]
  219. Goud, K.Y.; Satyanarayana, M.; Hayat, A.; Gobi, K.V.; Marty, J.L. Nanomaterial-Based Electrochemical Sensors in Pharmaceutical Applications. In Nanoparticles in Pharmacotherapy; Elsevier: Amsterdam, The Netherlands, 2019; pp. 195–216. ISBN 9780128165041. [Google Scholar]
  220. Kummari, S.; Sunil Kumar, V.; Vengatajalabathy Gobi, K. Facile Electrochemically Reduced Graphene Oxide-Multi-walled Carbon Nanotube Nanocomposite as Sensitive Probe for In-vitro Determination of Nitrofurantoin in Biological Fluids. Electroanalysis 2020, 32, 2452–2462. [Google Scholar] [CrossRef]
  221. Shekher, K.; Sampath, K.; Vandini, S.; Satyanarayana, M.; Vengatajalabathy Gobi, K. Gold Nanoparticle Assimilated Polymer Layer on Carbon Nanotube Matrices for Sensitive Detection of Serotonin in Presence of Dopamine In-Vitro. Inorg. Chim. Acta 2023, 549, 121399. [Google Scholar] [CrossRef]
  222. Yang, H.; Xu, Y.; Hou, Q.; Xu, Q.; Ding, C. Magnetic Antifouling Material Based Ratiometric Electrochemical Biosensor for the Accurate Detection of CEA in Clinical Serum. Biosens. Bioelectron. 2022, 208, 114216. [Google Scholar] [CrossRef]
  223. Pacheco, J.G.; Silva, M.S.V.; Freitas, M.; Nouws, H.P.A.; Delerue-Matos, C. Molecularly Imprinted Electrochemical Sensor for the Point-of-Care Detection of a Breast Cancer Biomarker (CA 15-3). Sens. Actuators B Chem. 2018, 256, 905–912. [Google Scholar] [CrossRef] [Green Version]
  224. Wei, B.; Mao, K.; Liu, N.; Zhang, M.; Yang, Z. Graphene Nanocomposites Modified Electrochemical Aptamer Sensor for Rapid and Highly Sensitive Detection of Prostate Specific Antigen. Biosens. Bioelectron. 2018, 121, 41–46. [Google Scholar] [CrossRef]
  225. Farzin, L.; Sadjadi, S.; Shamsipur, M.; Sheibani, S.; Mousazadeh, M. hasan Employing AgNPs Doped Amidoxime-Modified Polyacrylonitrile (PAN-Oxime) Nanofibers for Target Induced Strand Displacement-Based Electrochemical Aptasensing of CA125 in Ovarian Cancer Patients. Mater. Sci. Eng. C 2019, 97, 679–687. [Google Scholar] [CrossRef]
  226. Shekari, Z.; Zare, H.R.; Falahati, A. Dual Assaying of Breast Cancer Biomarkers by Using a Sandwich–Type Electrochemical Aptasensor Based on a Gold Nanoparticles–3D Graphene Hydrogel Nanocomposite and Redox Probes Labeled Aptamers. Sens. Actuators B Chem. 2021, 332, 129515. [Google Scholar] [CrossRef]
  227. Ji, Y.; Guo, J.; Ye, B.; Peng, G.; Zhang, C.; Zou, L. An Ultrasensitive Carcinoembryonic Antigen Electrochemical Aptasensor Based on 3D DNA Nanoprobe and Exo III. Biosens. Bioelectron. 2022, 196, 113741. [Google Scholar] [CrossRef] [PubMed]
  228. Taheri, N.; Khoshsafar, H.; Ghanei, M.; Ghazvini, A.; Bagheri, H. Dual-Template Rectangular Nanotube Molecularly Imprinted Polypyrrole for Label-Free Impedimetric Sensing of AFP and CEA as Lung Cancer Biomarkers. Talanta 2022, 239, 123146. [Google Scholar] [CrossRef]
  229. Upan, J.; Youngvises, N.; Tuantranont, A.; Karuwan, C.; Banet, P.; Aubert, P.H.; Jakmunee, J. A Simple Label-Free Electrochemical Sensor for Sensitive Detection of Alpha-Fetoprotein Based on Specific Aptamer Immobilized Platinum Nanoparticles/Carboxylated-Graphene Oxide. Sci. Rep. 2021, 11, 13969. [Google Scholar] [CrossRef]
  230. Castaño-Guerrero, Y.; Romaguera-Barcelay, Y.; Moreira, F.T.C.; Brito, W.R.; Fortunato, E.; Sales, M.G.F. Poly(Thionine)-Modified Screen-Printed Electrodes for CA 19-9 Detection and Its Properties in Raman Spectroscopy. Chemosensors 2022, 10, 92. [Google Scholar] [CrossRef]
  231. Yola, M.L.; Atar, N. Carbohydrate Antigen 19-9 Electrochemical Immunosensor Based on 1D-MoS2 Nanorods/LiNb3O8 and Polyoxometalate-Incorporated Gold Nanoparticles. Microchem. J. 2021, 170, 106643. [Google Scholar] [CrossRef]
  232. Rebelo, T.S.C.R.; Costa, R.; Brandão, A.T.S.C.; Silva, A.F.; Sales, M.G.F.; Pereira, C.M. Molecularly Imprinted Polymer SPE Sensor for Analysis of CA-125 on Serum. Anal. Chim. Acta 2019, 1082, 126–135. [Google Scholar] [CrossRef]
  233. Wang, M.; Hu, M.; Li, Z.; He, L.; Song, Y.; Jia, Q.; Zhang, Z.; Du, M. Construction of Tb-MOF-on-Fe-MOF Conjugate as a Novel Platform for Ultrasensitive Detection of Carbohydrate Antigen 125 and Living Cancer Cells. Biosens. Bioelectron. 2019, 142, 111536. [Google Scholar] [CrossRef]
  234. Freitas, M.; Nouws, H.P.A.; Delerue-Matos, C. Electrochemical Sensing Platforms for HER2-ECD Breast Cancer Biomarker Detection. Electroanalysis 2019, 31, 121–128. [Google Scholar] [CrossRef] [Green Version]
  235. Li, X.; Shen, C.; Yang, M.; Rasooly, A. Polycytosine DNA Electric-Current-Generated Immunosensor for Electrochemical Detection of Human Epidermal Growth Factor Receptor 2 (HER2). Anal. Chem. 2018, 90, 4764–4769. [Google Scholar] [CrossRef]
  236. Wang, W.; Han, R.; Chen, M.; Luo, X. Antifouling Peptide Hydrogel Based Electrochemical Biosensors for Highly Sensitive Detection of Cancer Biomarker HER2 in Human Serum. Anal. Chem. 2021, 93, 7355–7361. [Google Scholar] [CrossRef] [PubMed]
  237. Pacheco, J.G.; Rebelo, P.; Freitas, M.; Nouws, H.P.A.; Delerue-Matos, C. Breast Cancer Biomarker (HER2-ECD) Detection Using a Molecularly Imprinted Electrochemical Sensor. Sens. Actuators B Chem. 2018, 273, 1008–1014. [Google Scholar] [CrossRef]
  238. Díaz-Fernández, A.; Miranda-Castro, R.; De-los-Santos-Álvarez, N.; Lobo-Castañón, M.J.; Estrela, P. Impedimetric Aptamer-Based Glycan PSA Score for Discrimination of Prostate Cancer from Other Prostate Diseases. Biosens. Bioelectron. 2021, 175, 112872. [Google Scholar] [CrossRef] [PubMed]
  239. Hasanzadeh, M.; Tagi, S.; Solhi, E.; Mokhtarzadeh, A.; Shadjou, N.; Eftekhari, A.; Mahboob, S. An Innovative Immunosensor for Ultrasensitive Detection of Breast Cancer Specific Carbohydrate (CA 15-3) in Unprocessed Human Plasma and MCF-7 Breast Cancer Cell Lysates Using Gold Nanospear Electrochemically Assembled onto Thiolated Graphene Quantum Dots. Int. J. Biol. Macromol. 2018, 114, 1008–1017. [Google Scholar] [CrossRef] [PubMed]
  240. Rong, S.; Zou, L.; Zhu, Y.; Zhang, Z.; Liu, H.; Zhang, Y.; Zhang, H.; Gao, H.; Guan, H.; Dong, J.; et al. 2D/3D Material Amplification Strategy for Disposable Label-Free Electrochemical Immunosensor Based on RGO-TEPA@Cu-MOFs@SiO2@AgNPs Composites for NMP22 Detection. Microchem. J. 2021, 168, 106410. [Google Scholar] [CrossRef]
  241. Zhao, S.; Zhang, Y.; Ding, S.; Fan, J.; Luo, Z.; Liu, K.; Shi, Q.; Liu, W.; Zang, G. A Highly Sensitive Label-Free Electrochemical Immunosensor Based on AuNPs-PtNPs-MOFs for Nuclear Matrix Protein 22 Analysis in Urine Sample. J. Electroanal. Chem. 2019, 834, 33–42. [Google Scholar] [CrossRef]
  242. Li, S.; Yue, S.; Yu, C.; Chen, Y.; Yuan, D.; Yu, Q. A Label-Free Immunosensor for the Detection of Nuclear Matrix Protein-22 Based on a Chrysanthemum-like Co-MOFs/CuAu NWs Nanocomposite. Analyst 2019, 144, 649–655. [Google Scholar] [CrossRef]
  243. Goud, K.Y.; Reddy, K.K.; Khorshed, A.; Kumar, V.S.; Mishra, R.K.; Oraby, M.; Ibrahim, A.H.; Kim, H.; Gobi, K.V. Electrochemical Diagnostics of Infectious Viral Diseases: Trends and Challenges. Biosens. Bioelectron. 2021, 180, 113112. [Google Scholar] [CrossRef]
  244. Ruiz-Valdepeñas Montiel, V.; Sempionatto, J.R.; Esteban-Fernández de Ávila, B.; Whitworth, A.; Campuzano, S.; Pingarrón, J.M.; Wang, J. Delayed Sensor Activation Based on Transient Coatings: Biofouling Protection in Complex Biofluids. J. Am. Chem. Soc. 2018, 140, 14050–14053. [Google Scholar] [CrossRef]
  245. Ghafar-Zadeh, E. Wireless Integrated Biosensors for Point-of-Care Diagnostic Applications. Sensors 2015, 15, 3236–3261. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Cancer causing factors/carcinogens.
Figure 1. Cancer causing factors/carcinogens.
Biosensors 13 00398 g001
Figure 2. Cancer biomarker clinical applications.
Figure 2. Cancer biomarker clinical applications.
Biosensors 13 00398 g002
Figure 3. Schematic illustration of the construction of paper-based Sn-doped SnO2-x sample and assay procedures for the specific detection of cancer biomarker alpha fetoprotein. Reproduced with the permission. [202] Copyright 2021, Elsevier.
Figure 3. Schematic illustration of the construction of paper-based Sn-doped SnO2-x sample and assay procedures for the specific detection of cancer biomarker alpha fetoprotein. Reproduced with the permission. [202] Copyright 2021, Elsevier.
Biosensors 13 00398 g003
Figure 4. Schematic representation of microfluidic paper-based electrochemical aptasensor for the specific detection of prostate-specific antigen (PSA) in clinical samples. Reproduced with the permission. [224]. Copyright 2018, Elsevier.
Figure 4. Schematic representation of microfluidic paper-based electrochemical aptasensor for the specific detection of prostate-specific antigen (PSA) in clinical samples. Reproduced with the permission. [224]. Copyright 2018, Elsevier.
Biosensors 13 00398 g004
Figure 5. Schematic representation of label-free 3D DNA nanoprobe DNA tweezers-based electrochemical sensor for the detection of carcinoembryonic antigen (CEA) biomarker. Reproduced with the permission [227]. Copyright 2018, Elsevier.
Figure 5. Schematic representation of label-free 3D DNA nanoprobe DNA tweezers-based electrochemical sensor for the detection of carcinoembryonic antigen (CEA) biomarker. Reproduced with the permission [227]. Copyright 2018, Elsevier.
Biosensors 13 00398 g005
Figure 6. Schematic representation of rGO-FET-based electrochemical aptasensor for the specific detection of CA 125 ssDNA biomarker. Reproduced with the permission [208]. Copyright 2018, Elsevier.
Figure 6. Schematic representation of rGO-FET-based electrochemical aptasensor for the specific detection of CA 125 ssDNA biomarker. Reproduced with the permission [208]. Copyright 2018, Elsevier.
Biosensors 13 00398 g006
Table 2. Optical sensing methodologies for cancer biomarker detection.
Table 2. Optical sensing methodologies for cancer biomarker detection.
No.BiomarkerRecognitionMethodLinear RangeLODRef.
1AFPSiO2@CQDs/AuNPs/MPBAECL0.001–1000 ng/mL0.0004 ng/mL[204]
2AFP1D SnO2 NTsPEC10 pg mL−1–200 ng mL−13.84 pg mL−1[202]
3AFPaptamer-MCHAFluorescence0.1 ng mL−1–10 mgmL−10.033 ng. mL −1[205]
4CA 19-9luminol-AgNPs@ZIF-67ECL0.0001–10 U/mL31 μU/mL[206]
5CA 19-9Ni NCs-Ab2Fluorescence0.001–48 ng mL−10.00013 ng mL−1[207]
6CA 125rGO-based FET-type aptasensorFluorescence1.0 × 10−9-1.0 U/mL5.0 × 10−10 U/mL[208]
7CA 125CA 125/MUC16SPRI2.2–150 U/ml-[209]
8CEATi3C2-MXene/AuNPs/SPASPR2 × 10−16–2 × 10−8 M0.07 fM[203]
9CEAHCR and G-quadruplex
DNAzyme
Fluorescence0.25–1.5 nM0.2 nM[210]
10HER2nanoparticle coated QCMQCM10–500 cells/mL10 cells/mL[211]
11HER2PtAmiFluorescence--[212]
12HER23D DNA walkerFluorescence0.5–5 ng mL−10.01 ng mL−1[213]
13PSANaYF4:Yb3+, Er3+ UCNPs and NaYF4:Yb3+, Er3+@NaYF4:Yb3+ UCNPsFluorescence0.1 ng/mL–10 ng/mL0.01 ng/mL[214]
14PSAanti-PSA/MCH/ AgNPs/SiNWsSERS0.1–20 μg·L−10.1 μg·L−1[199]
15PSAIDCFET0.1−10.0 μL/mL-[200]
16CA-15-3PHMPFFluorescence2.56 × 10−5–1.28 U mL−12.56 × 10−5 U mL−1[215]
17CA-15-3SPR gold substratesSPR-0.0998 U mL−1[201]
18NMP 22orange emitting quantum dot CdTe/CdSFluorescence2–22 pg mL−10.05 pg mL−1[216]
19NMP 22NCDsFluorescence1.3–16.3 ng/mL0.047 ng/mL[217]
AFP—alpha-fetoprotein; ECL—electrochemiluminescence; PEC—photoelectrochemical; SPRI—Surface Plasmon Resonance Imaging; QCM—quartz crystal microbalance; PtAmi—red-emitting exchanged Pt nanoclusters; UCNPs—upconversion nanoparticles; IDC—interdigitated capacitor; PHMPF—prismatic hollow Metal-polydopamine frameworks; NCDs-—nitrogen-doped carbon dots.
Table 3. Electrochemical sensing methodologies for cancer biomarker detection.
Table 3. Electrochemical sensing methodologies for cancer biomarker detection.
No.BiomarkerRecognitionMethodLinear RangeLODRef.
1AFPPtNPs/GO-COOHSWV3.0–30 ng mL−11.22 ng mL−1[229]
2AFPFTO/PPy-MO DMIPEIS10–104 pg mL−13.3 pg mL−1[228]
3CA 19-9Au-SPE/THDPV0.010–10 U/mL-[230]
4CA 19-91DMoS2 nanorods/LiNb3O8 and AuNPs@POMDPV0.1–10.0 µU mL− 10.030 µU mL− 1[231]
5CA 125MIP@AuSPESWV0.01 and 500 U/mL0.1 U/mL[232]
6CA 125ITO/Ag NPs–PAN-oxime NFs/aptamer/c-DNA–MBDPV0.01 to 350 UmL−10.0042 UmL−1[225]
7CA 125Tb-MOF-on- Fe-MOFEIS1 × 102−1 × 105 cell·mL−119 cell·mL−1[233]
8CEAMXC-Fe3O4-RuDPV1 pg/mL–1 μg/mL0.62 pg/mL[222]
9CEAIEC-BADPV-4.88 fg mL−1[227]
10HER2SPCE-MWCNT/AuNPLSV7.5 and 50 ng/mL0.16 ng/mL[234]
11HER2polycytosine DNA (dC20)SWV0.001−1 ng/mL0.5 pg/mL[235]
12HER2GCE/PEDOT/Gel/Ab/HER2 DPV0.1 ng mL−1–1.0 μg mL−145 pg mL−1[236]
13HER2MIP-AuSPEDPV10–70 ng/mL1.6 ng/L[237]
14PSAAuNPs/rGO/THI-aptamerDPV0.05 to 200 ng mL−110 pg mL−1[224]
15PSAaptamer PSAG-1EIS0.64–62.5 ng/mL-[238]
16CA-15-3AuNPs/3DGHDPV5.0 × 10–2–100.0 U mL–1-[226]
17CA-15-3MIP/Au-SPEDPV5–50 U mL−11.5 U mL−1[223]
18CA-15-3CysA/Au NSs/GQDsSWV0.16–125 U/mL0.11 U/ml[239]
19NMP 22Cu-MOFs@SiO2@AgNPsDPV0.1 pg∙mL−1–1000 ng∙mL−133.33 fg∙mL−1[240]
20NMP 22AuNPs-PtNPs-MOFsDPV0.005 ng·mL−1
–20 ng·mL−1
1.7 pg·mL−1[241]
21NMP 22Co-MOFs/CuAu NWsAmperometric0.1 pg mL−1–1 ng mL−133 fg mL−1[242]
DMIP—dual-template molecularly imprinted polymer; TH—thionine; IEC-BA—ingenious electrochemical aptamer biosensor; PSA—prostate-specific antigen; AuNPs/3DGH—gold nanoparticle three-dimensional graphene hydrogel.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rao Bommi, J.; Kummari, S.; Lakavath, K.; Sukumaran, R.A.; Panicker, L.R.; Marty, J.L.; Yugender Goud, K. Recent Trends in Biosensing and Diagnostic Methods for Novel Cancer Biomarkers. Biosensors 2023, 13, 398. https://doi.org/10.3390/bios13030398

AMA Style

Rao Bommi J, Kummari S, Lakavath K, Sukumaran RA, Panicker LR, Marty JL, Yugender Goud K. Recent Trends in Biosensing and Diagnostic Methods for Novel Cancer Biomarkers. Biosensors. 2023; 13(3):398. https://doi.org/10.3390/bios13030398

Chicago/Turabian Style

Rao Bommi, Jagadeeswara, Shekher Kummari, Kavitha Lakavath, Reshmi A. Sukumaran, Lakshmi R. Panicker, Jean Louis Marty, and Kotagiri Yugender Goud. 2023. "Recent Trends in Biosensing and Diagnostic Methods for Novel Cancer Biomarkers" Biosensors 13, no. 3: 398. https://doi.org/10.3390/bios13030398

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

Article Metrics

Back to TopTop