Next Article in Journal
Hybrid Domain Consistency Constraints-Based Deep Neural Network for Facial Expression Recognition
Next Article in Special Issue
High-Throughput Effect-Directed Monitoring Platform for Specific Toxicity Quantification of Unknown Waters: Lead-Caused Cell Damage as a Model Using a DNA Hybrid Chain-Reaction-Induced AuNPs@aptamer Self-Assembly Assay
Previous Article in Journal
Ratiometric Sensing of Glyphosate in Water Using Dual Fluorescent Carbon Dots
Previous Article in Special Issue
Interactions of Nanomaterials with Gut Microbiota and Their Applications in Cancer Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Anti-Fouling Strategies of Electrochemical Sensors for Tumor Markers

Department of Chemistry, Capital Normal University, Beijing 100048, China
*
Authors to whom correspondence should be addressed.
Sensors 2023, 23(11), 5202; https://doi.org/10.3390/s23115202
Submission received: 28 April 2023 / Revised: 26 May 2023 / Accepted: 27 May 2023 / Published: 30 May 2023
(This article belongs to the Special Issue Micro/Nano Biosensors and Devices)

Abstract

:
The early detection and prognosis of cancers require sensitive and accurate detection methods; with developments in medicine, electrochemical biosensors have been developed that can meet these clinical needs. However, the composition of biological samples represented by serum is complex; when substances undergo non-specific adsorption to an electrode and cause fouling, the sensitivity and accuracy of the electrochemical sensor are affected. In order to reduce the effects of fouling on electrochemical sensors, a variety of anti-fouling materials and methods have been developed, and enormous progress has been made over the past few decades. Herein, the recent advances in anti-fouling materials and strategies for using electrochemical sensors for tumor markers are reviewed; we focus on new anti-fouling methods that separate the immunorecognition and signal readout platforms.

1. Introduction

Cancer has become one of the leading causes of death in many countries and regions around the world [1,2]; it usually manifests clinically in the late stages of invasion of surrounding tissues. The effects of cancer treatment are limited, and the side effects are currently significant [3]. The mortality of cancer can be effectively reduced if early diagnoses of cancer can be made, the disease development can be determined, and corresponding treatment measures can be taken in time [4]. Therefore, the early detection of cancer before symptoms appear or metastasis occurs, coupled with early interventions, can prevent and improve survival rates; this is an important research topic [5].
Tumor markers refer to a series of characteristic biochemical indicators that exist more abundantly in tumor cells or are specifically secreted by tumor cells; they can directly reflect the condition of cancer, including the metabolites of tumor cells, products of gene expression disorder in tumor cells, and products of the apoptosis and disintegration of tumor cells. They show a correlation with the occurrence and development of various types of cancers, such as prostate cancer, liver cancer, lung cancer, colorectal cancer, and so on [6,7]. Tumor markers in human blood exist in low concentrations, and their changes are related to the occurrence, development, prognosis, and cure of tumors [8,9,10]. In the early stage of cancer, the relevant markers exhibit minimal fluctuation compared to a healthy body; high sensitivity is required to detect them in real samples [11].
The development of highly sensitive, accurate, and economical tumor marker detection methods is highly important for the early diagnosis and prognostic treatment of cancer [12,13]. In the past few decades, several sensitive assays for detecting tumor markers have been developed based on the specific immune response between antibodies and tumor markers, including immunofluorescence [14], colorimetric immunoassays [15], electrochemiluminescent immunoassays [16,17,18,19], radioimmunoassays [20], photoelectrochemical immunoassays [21,22,23,24,25,26], and electrochemical immunoassays [27,28,29,30,31,32,33,34,35,36]. Among these methods, electrochemical immunosensors are widely used because of their advantages of low costs, portability, high sensitivity, and fast response speeds, which have made electrochemical biosensors an essential basic tool for clinical diagnosis in many in situ and point-of-care (POC) situations [37,38]. Electrochemical biosensors identify biological functions in processes by reading out different electronic responses, such as differences in the current, charge, potential, and impedance, which correspond to modification and stimulation processes on the electrode’s surface. Experimental results can provide important information for describing and analyzing various biological processes [39]. After decades of development, electrochemical immunosensors have attracted the attention of researchers, and significant breakthroughs have been made [40].
The composition of biological samples is very complex, as they contain a variety of biomolecules in addition to the target analyte to be measured, such as proteins, peptides, fats, and hormones. In the process of conducting immunoassays, the non-specific adsorption of coexisting species will cause false positives, false-negative, and lowered signal-to-noise ratios [41,42,43], weakening the accuracy and sensitivity of immunosensors. The fouling caused by non-specific adsorption on the electrode’s surface is one of the major problems that need to be solved in the practical applications of electrochemical immunosensors. The construction of anti-fouling electrochemical immunosensors represents a crucial step from theoretical research to practical sample detection. Thus, electrodes are commonly modified with anti-fouling material, together with antibodies.
As the immune recognition process usually occurs on the electrode’s surface, biologically complex samples have the opportunity to contact and contaminate the electrode’s surface. The anti-fouling process is mainly intended to prevent the occurrence of non-specific adsorption through the hydrophilic, electro-neutral, or steric hindrance of anti-fouling materials in the immune recognition process after the addition of biological fluid samples, such as urine, serum, or plasma. In most cases, the method used to reduce non-specific adsorption involves applying anti-fouling materials directly onto the electrode’s surface. These materials include PEG and its derivatives [44,45,46,47,48,49], zwitterionic polymers [41,50,51,52], peptides [53,54,55,56,57], and hydrogels [58,59,60,61], among others [62,63,64]. A new method has been developed in recent years, whereby the immunorecognition and signal readout platforms are separated, and the anti-fouling material is used to modify magnetic beads instead of the electrode’s surface, resulting in good conductivity of the electrode [65,66,67,68,69,70,71,72]. The immunological recognition process occurs on the magnetic beads’ surface, taking advantage of the large specific surface area of functionalized magnetic beads to increase the sites for antibody binding. Furthermore, sandwich assays are much less sensitive to fouling [42]. The non-target substances in complex samples can be completely washed away and have no chance to contact the electrode. The possibility of electrode surface contamination can be completely eliminated. Fouling resistance is the most significant feature of this method.
In this review, we introduce the most frequently used anti-fouling materials, describing their principles and giving examples. Moreover, an outlook on the future of anti-fouling electrochemical immunosensors for tumor markers is also presented.

2. Different Anti-Fouling Strategies Based on Anti-Fouling Materials

2.1. Polyethylene Glycol (PEG) and Its Derivatives

Polyethylene glycol (PEG) and its derivatives have been widely used as anti-fouling materials since the 1970s [73]. After PEG is modified on the electrode’s surface, abundant hydroxyl groups at the end of the PEG chain form a hydration layer through hydrogen bonding. The hydration layer plays an important role in surface resistance to protein adsorption [52]. The water molecules in this hydration layer are relatively firmly fixed between the electrode interface and the serum due to the action of hydrogen bonding. It is difficult for the protein to break through the hydration layer and stick to the electrode’s surface [42].
Due to the poor electrical conductivity of PEG, the impedance of the modified electrode surface will increase, reducing the sensitivity of detection. The use of immunoprobes enables signal amplification and improves sensor performance through collaboration with anti-fouling sensing interfaces. Ma and his co-author designed an immunoprobe in which metal–organic framework 818 (MOF-818) served as a support to enrich the electrochemical signal. The anti-fouling sensing interface was self-assembled by HS-PEG-NH2 on gold–sulfur bonds on gold-plated GCE, and PEG-NH2 acted as a blocker to block non-reactive sites modified by antibodies. The strong hydrophilicity of PEG was utilized to form a dense hydration layer, which effectively reduced the adsorption of non-specific proteins on the electrode surface and achieved the expected anti-fouling effect. The results showed that when human serum epididymal protein 4 (HE-4), human immunoglobulin (IgG), carcinoma antigen 125 (CA12-5), and neuron-specific enolase (NSE) were selected as interferences, the corresponding currents produced were negligible [46]. The fabrication process of the electrochemical immunosensor is shown in Figure 1.
Poly(3,4-ethylenedioxythiophene) (PEDOT) and polyaniline (PANI) are widely used in electrochemical biosensors due to their stability in the oxidized state, ease of processing and excellent electrical conductivity [74,75]. They cross-linked PEG together to form the anti-fouling layer, which can effectively improve the resistance caused by modification of the anti-fouling layer. PEG can copolymerize with conductive polymers on the electrode surface [47] or be modified after the deposition of the conductive polymer [48] (Figure 2). The modification time of copolymerization is shorter, and the steps are fewer. However, attention should be paid to the content of copolymerization raw materials to generate composite materials with both anti-fouling and conductive functions. Moreover, copolymerization will cause a part of PEG to be entrapped in the conducting polymers; only the surface exposed part can exhibit anti-fouling ability with a low utilization rate. The second method requires the conductive polymer to have an anchoring functional group of PEG; otherwise, the conductive polymer surface is necessary to be modified to immobilize the bio-recognition molecules. Gold nanoparticles were deposited on PANI to assist PEG fixation, for the PANI had a few amino as anchoring functional groups [76].
Due to the magnetic beads’ contact with the serum rather than the electrode surface, it is also necessary to modify the magnetic beads’ surface with anti-fouling materials to avoid the occurrence of non-specific adsorption. Zhang and co-authors reported an ultra-sensitive protein serum assay at the zeptomolar level, which combined the silver mirror reaction. The separation of immunorecognition and signal readout platforms prevented noise interference and biofouling at the electrode interface, and PEG was used to resist the adsorption of non-specific proteins (Figure 3). This biosensor was utilized to detect h-IgG and exhibited a large dynamic range from 1 μg mL−1 to 100 ag mL−1 and an ultralow limit of detection (LOD) 6.31 ag mL−1 (0.04 zeptomoles mL−1) [77].

2.2. Zwitterionic Materials

Complex biological samples contain charged materials, such as proteins. The protein carries a net charge due to the presence of ionized side chains on the protein surface. At suitable pH surroundings for the detection of tumor markers, charged proteins are present in the sample. The non-selectivity electrostatic adsorption possibly occurs if the electrode surface has the opposite charge to proteins. Using electro-neutral material to modify the electrode is an effective means to reduce electrostatic adsorption. Zwitterionic materials can form a hydrated layer through hydrogen bonds while being electrically neutral integrally as a whole, which is the key to its anti-fouling performance [52,78,79,80].
Poly(sulfobetaine methacrylate) (PSBMA), for instance, has possessed equivalent positive and negative charged groups, the whole molecule is electric neutrality [50], and water molecules are strongly hydrogen-bonded at its interfaces [51]. Recent research reported that SBMA and dopamine (DA) monomers could undergo polymerization on the electrode surface, forming PDA-PSBMA film to resist anti-nonspecific adsorption (Figure 4). Polydopamine (PDA) was used as a strong adhesive to stabilize PSBMA on the electrode. Aptamers or antibodies with thiol groups could attach to the PDA via the Michael addition reaction. The PDA-PSBMA interface was used to track the dynamics of ascorbic acid (AA) in the rat brain both in a normal state and in Parkinson’s disease [41]. PDA-PSBMA had shown high oxidation resistance, hydrolysis stability, and anti-fouling properties. Nevertheless, betaine polymers require complex pretreatment when used at the interface.
Yao et al. developed a regenerating electrochemical detection platform that can be regenerated in 3 min. The anti-fouling immunomagnetic beads were obtained by modifying SBMA and antibodies on magnetic beads. SBMA provided the anti-fouling function for immunomagnetic beads, and the dual interface mode reduced the non-specific adsorption as well. Squamous cell carcinoma (SCCA) was detected as the model analyte, and carbohydrate antigen 19-9 (CA19-9), alpha-fetoprotein (AFP), prostate-specific antigen (PSA), carcinoembryonic antigen (CEA), and neuron-specific enolase (NSE) were tested as interfering substances, indicating a high specificity of the biosensor. The concentration of SCCA from 1 pg mL−1 to 1 μg mL−1 possesses a linear relationship with the logarithmic value of the normalized SWV signal intensity, and the LOD of this platform was 31.20 fg mL−1 [81].

2.3. Peptides

The natural biocompatibility of functionalized anti-fouling peptides makes them become a popular anti-fouling material and widely used in an electrochemical assay for tumor markers. Due to the polar functional groups and high hydrogen bonding ability of zwitterionic charges, peptides are strongly hydrated, which is one of the reasons for their anti-fouling function. Hydrophilic and amphiphilic peptides without charge also have anti-fouling properties; for instance, the anti-fouling segment of peptides based on the EK motif was designed as electroneutrality [82,83]. The electrically neutral or hydrophilic functionalized anti-fouling peptides can be designed. All these advantages make peptides a prime candidate for biodegradable anti-fouling materials [54,55,84].

2.3.1. Straight Peptides

Complex biological samples contain various enzymes. Peptides are prone to enzymatic degradation, which will reduce their anti-fouling effect. The anti-fouling performance can be improved by enhancing the resistance to enzymatic decomposition of peptides. Luo’s group designed a novel anti-fouling peptide with three D-amino acids in the N and C terminals (pD-Peptide), which made the anti-fouling peptide stable in human serum and resistant to enzymatic hydrolysis. The aptamer and pD-peptide were successively immobilized on the PANI-modified electrode with abundant amino groups. Due to the anti-enzymatic hydrolysis ability of D-amino acids at the two terminals of the peptide, the anti-fouling property of the pD-peptide-modified electrode was proved in different charge protein solutions and different concentrations of human serum samples. This modified electrode also enhanced long-term anti-fouling performance compared with normal L-peptide. This anti-fouling electrochemical biosensor demonstrated excellent selectivity with a wide linear range (from 0.1 fg mL−1 to 1.0 ng mL−1) and a low detection limit (0.031 fg mL−1). The result was measured with the electrochemiluminescence (ECL) assaying method used in hospitals and is closely matched, with the RSD ranging from 0.64% to 9.5% [85].
Another design idea for anti-fouling peptides is to develop the peptides with recognition sequences. The peptides which contain an electrical signal are cleaved by antigens and also have good specificity. Therefore, the electrode surface, where the aptamer was originally modified, for instance, can be used to modify the multifunctional peptide with the recognition segment. The tumor marker matrix metalloproteinase-7 (MMP-7) is a proteolytic enzyme present in serum that can specifically cleave the peptide sequence of RPLALWRSC. Based on that, an anti-fouling and sensitive biosensor was designed (Figure 5). The anti-fouling sequence was EKEKEK, which was electrically neutral to avoid non-specific adsorption. The pep/SA-GO-Pb2+/GCE surface was proven to have a remarkable anti-fouling property in different biological environments such as antigens (NSE, CEA, IgG, and PSA), cholesterol, amino acids, dopamine (DA), ascorbic acid (AA), and uric acid (AA). The peptides linked the hydrophilic interface with sodium alginate–graphene oxide-Pb2+ (SA-GO-Pb2+) gel and the novel carboxyl-rich pyrrole-doped ZIF loaded with urease (Urease @ZIF-Py). The anti-fouling effects of peptides and the cascade initiated by urease made this biosensor have a satisfactory linear relationship with the logarithm of the MMP-7 concentration range of 1 pg mL−1–100 ng mL−1, and LOD was 24.34 fg mL−1 (S/N = 3) [53].
PSA is a protease tumor marker that can specifically cleave the amino acid sequence “HSSKLQ”. “KEK” amino acid sequence is electro-neutral, which acts as an anti-fouling segment to reduce non-specific adsorption, and ferrocene can link terminally to provide an electrochemical signal. Based on this idea, an electrochemical sensing interface based on the oriented self-assembly of histidine-labeled peptides induced by Ni2+ for protease detection was developed. This sensor had significant anti-fouling performance and can be removed by pickling to realize the reconstruction of the recognition layer, in which the reproducibility was about 80% [86].

2.3.2. Peptides with Branches

In addition to the straight chain structure, anti-fouling peptides can also be designed with branches to form Y-shaped structures. The branches can be either both recognition sequences (Figure 6) [87,88] or recognition sequences with anti-fouling sequences, respectively (Figure 7) [89]. The Y-type peptide with a single recognition branch has one recognition branch and one anti-fouling branch, and its anti-fouling performance was better than that of the straight-type peptide, which was proved by the anti-fouling performances assays of BSA (negatively charged), Lys (positively charged), Mb (neutral in charge) and different concentrations of human serum. This biosensor was able to resist biofouling in various protein solutions and serum samples and could be the base for developing other biosensing systems for different targets [89]. Another research indicated that there is a synergistic effect between the two octapeptides, FYWHCLDE and FYCHTIDE, which makes a higher selectivity for IgG [90]. Based on that, an anti-fouling electrochemical sensor with dual-recognizing branch peptides was developed [87]. The detection limit of the sensor was 0.031 pg mL−1, which was two orders of magnitude lower than the biosensor with single-recognizing branches peptide. This biosensor could detect IgG in real biological samples, such as human serum, and the contamination of the sensing interface by complex biological media was effectively avoided. The peptide sequence of HSSKLQ can be specifically cleaved by PSA, and the GO-NB nanocomposite was modified onto the electrode surface to enhance the conductivity and specific surface area of the electrode. NB was used as an internal electrochemical probe to generate a reference signal, and the ratiometric signal of Fc/NB was used as the final sensing signal, which can provide higher sensitivity in addition to the anti-fouling function of the sensor [88].

2.4. Hydrogels

Hydrogels are cross-linked networks composed of hydrophilic polymers with abundant hydrogen bonds. With a large specific surface area and unique three-dimensional network structure, hydrogels can accommodate abundant modified materials to modify the electrodes. Therefore, hydrogels can load a large number of molecules with different functions and give them good conductivity, strong electrochemical signals, and excellent catalytic ability. At the same time, the micro-environment present in hydrogels can increase the stability and biological activity of biomolecules. Additionally, the high permeability of hydrogels can accelerate the transport of small molecules and ions, as well as the rapid transfer of electrons. These unique properties make hydrogels possess great potential in building electrochemical immunosensing interfaces [90,91,92,93].
As shown in Figure 8, a novel redox polyaniline-polythionine hydrogel (PANI-PThi gel) was fabricated as an ultrasensitive and protein-resistant label-free amperometric immunosensing platform for carcinoma antigen-125 (CA12-5). In the anti-fouling performance test, the PANI-PThi gel-modified electrode was immersed in PBS solution with different concentrations of serum for 12 h, and the current fluctuation range was not more than 4 μA, which proved that the PANI-PThi gel could significantly reduce the non-specific adsorption of proteins in human serum. Moreover, SWV curves indicated that the proposed hydrogel has a large and repeatable signal and could remain at 94.6% after a month. The current response of PANI-PThi gels can be amplified in the presence of H2O2, which contributes to a wide linear range for the immunosensor. The result of such biosensors agreed well with ELISA in the detection of clinical serum samples [94].
An immunoplatform with high specificity and anti-fouling performance could be established in the hydrophilicity of the hydrogel coupled with the specificity of the molecular imprinting technique [95]. Luo et al. synthesized a protein-imprinted hydrogel that has temperature-responsive PNiPAAm chains. The template protein IgG can easily remove without the aid of any washing solvents at 37 °C. The structure of the IgG-MIH/GCE will revert to its initial form when the temperature is back to 20 °C. IgG-MIH/GCE surfaces were incubated in a series of complex protein-containing solutions and HeLa cell medium, indicating that the IgG-MIH/GCE surfaces had long-term stability, anti-fouling ability, and the ability to prevent cell adhesion (Figure 9) [61].

2.5. Bovine Serum Albumin (BSA)

Serum albumins (SAs) are one of the most widely abundant protein components in the body, both in animals and human beings. SAs are well known as a stable and widely applicable anti-fouling material to resist non-specific protein adsorption and often are used as passivation or blockers in electrochemical biosensors due to the conservation of both amino acid sequences and a three-dimensional structure [96]. Immunomagnetic beads can take advantage of the huge specific surface area when contacting complex biological samples. After the antibody was modified on the surface of the immunomagnetic beads, BSA acted as a blocking agent to block any possible remaining active sites and prevent the non-specific adsorption of proteins in complex biological samples. The resulting biosensor still has good sensitivity and specificity performance and an anti-fouling effect [66,67,68,69,70,71,97].
Yin et al. designed an enzyme-Fenton reaction that triggered the destruction of Fe3+ cross-linked alginate hydrogel to improve the electrochemical immunosensor, as shown in Figure 10. Silica oxide-glucose oxidase nanocomposites were used as immunoprobes to realize the high efficiency of glucose oxidation and H2O2 generation. With the continuous supply of H2O2, the Fenton reaction was triggered. Fe3+ transformed to Fe2+ and destroyed the structure of Fe3+-alginate hydrogel, leading to a significant decrease in the interface resistance. Under optimal conditions, neuron-specific enolase (NSE) was quantitatively detected with a wide linear range from 1 pg mL−1 to 100 ng mL−1, a low detection limit of 0.447 pg mL−1 [68].

2.6. Other Anti-Fouling Materials

Owing to its compact, globular structure with a multitude of hydroxyl groups, hyperbranched polyglycerol (HPG) has great hydrophilicity. Through the electrochemical polymerization of corresponding monomers, PEDOT was polymerized with HPG to improve the conductivity. The hydration layers were formed due to the presence of abundant hydroxyl groups on the surface of PEDOT-HPG, which inhibits the adsorption of proteins and resists non-specific cell attachment. The research showed that the biosensor had good anti-fouling ability and sensitivity with a detection limit is 0.035 pg mL−1 (Figure 11) [62].
The cell membrane itself has good biocompatibility. Due to the structure of the phospholipid bilayer, the hydrophilic phospholipid head was exposed to the sample. In the meanwhile, the head of the phospholipid contains the negatively charged phosphate group and the positively charged quaternary ammonium group, making the whole surface electroneutrality [98]. The platelet membrane/Au nanoparticle/delaminated V2C nanosheet (PM/AuNPs/d-V2C)-modified electrode with antibacterial fouling ability depends on the steric hindrance of glycoproteins, the abundance of zwitterionic headgroups in phospholipid bilayers and the good hydrophilicity of the interface (Figure 12) [63].
Engrailed 2 (EN 2)protein is a biomarker of prostate cancer (PCa). Its isoelectric point is equal to 9.1, so EN 2 is positively charged at physiological pH (7–7.4), which inspired the authors to use positively charged histamine molecules to prevent non-specific adsorption of EN 2 on the pre-functionalized electrodes [64]. In previous reports, only three of the eleven commercially available ELISA kits passed the accuracy threshold, reflecting the difficulty of developing ELISA urine tests with adequate analytical performance [98]. As positively charged blockers, histamine molecules can effectively prevent EN 2 from non-specific adsorption on the pre-functionalized electrode due to electrostatic repulsion. The designed immunosensor is a promising alternative to ELISA assay.

3. Conclusions and Prospects

In conclusion, we have summarized the principle of anti-fouling materials commonly used and introduced the method which separated the immunorecognition and signal readout platforms. These anti-fouling materials are widely used for anti-fouling electrochemical sensors with the advantages of high hydrophilicity or electroneutrality. However, they have certain constraints in practical application. PEG and its derivatives were prone to oxidate into aldehydes and ethers by oxygen, resulting in polymer long-chain breakage and shedding. The deposition after the synthesis of anti-fouling conductive polymer monomers, such as SBMA-EDOT, is complex and time-consuming. Complex biological samples contain various enzymes which could make peptides enzymatic degradation. These nonconductive anti-fouling materials on the electrode surface will generate impedance to the immune layer of the electrode sensing interface (Table 1 and Table 2).
In addition to the limitations of the natural character of these anti-fouling materials, the type of non-specific adsorption also needs to be fully considered. Since electrostatic adsorption is not selective and directional, the overall electrification of other modified materials and anti-fouling materials on the electrode base should be fully considered in the design of the electrode substrate. Strictly speaking, anti-fouling peptides, for example, are electrically neutral only at the anti-fouling segment, while other amino acid sequences may be charged. Therefore, the distance between the charged part of the anti-fouling material and the complex biological samples is one of the factors affecting the anti-fouling performance.
The future trend of electrochemical biosensors is rapid detection in vivo. This requires anti-fouling sensors to have strong long-term stability and long-term anti-fouling performance, which is another big challenge in the development of anti-fouling biosensors. The ideal anti-fouling material can enhance the signal-to-noise ratio with a strong hydrophilic and electro-neutral structure, and it can also reduce the non-specific binding. After modification, it can stably exist on the electrode or immunomagnetic bead and does not affect the binding of antigen-antibody, nor does it react with antibodies or detection reagents. Most anti-fouling sensors are stable for only four weeks and need to be stored at 4 °C, and some need to be protected from oxidation by oxygen. Consequently, the development of a long-term stable anti-fouling material with hydrophilicity, electroneutrality, and strong conductivity simultaneously is particularly important. For instance, zwitterionic inorganic/organic polymers can achieve desired hydrophilic, electro-neutral structures and high resistance to enzymatic decomposition through molecular design. To improve the conductivity, zwitterionic polymers can introduce the conjugated groups or copolymerize with conductive polymers. For the method which separates the immunorecognition and signal readout platform, the prepared magnetic beads and immunoprobes individually packaged with modified or bare electrodes can be better for commercial promotion and use, which has low requirements for testing personnel.

Author Contributions

G.S. investigation, writing—original draft; H.H., project administration, supervision; Z.M., funding acquisition, project administration, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

The authors thank the funding of the National Natural Science Foundation of China (22172104) and the Joint Project of Beijing Municipal Education Commission, and the Beijing Natural Science Foundation (KZ202110028042).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Campuzano, S.; Serafín, V.; Gamella, M.; Pedrero, M.; Yáñez-Sedeño, P.; Pingarrón, J.M. Opportunities, Challenges, and Prospects in Electrochemical Biosensing of Circulating Tumor DNA and Its Specific Features. Sensors 2019, 19, 3762. [Google Scholar] [CrossRef] [PubMed]
  2. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, L.D.; Liu, Y.; Li, J.H. Self-phosphorylating deoxyribozyme initiated cascade enzymatic amplification for guanosine-5′-triphosphate detection. Anal. Chem. 2014, 86, 7907–7912. [Google Scholar] [CrossRef] [PubMed]
  4. Hassan, E.M.; DeRosa, M.C. Recent advances in cancer early detection and diagnosis: Role of nucleic acid based aptasensors. TrAC Trends Anal. Chem. 2020, 124, 115806. [Google Scholar]
  5. Crosby, D.; Bhatia, S.; Brindle, K.M.; Coussens, L.M.; Dive, C.; Emberton, M.; Esener, S.; Fitzgerald, C.; Gambhir, S.S.; Kuhn, P.; et al. Early detection of cancer. Science 2022, 375, eaay9040. [Google Scholar] [CrossRef] [PubMed]
  6. Tothill, I.E. Biosensors for cancer markers diagnosis. Semin. Cell Dev. Biol. 2009, 20, 55–62. [Google Scholar] [CrossRef]
  7. Labib, M.; Sargent, E.H.; Kelley, S.O. Electrochemical methods for the analysis of clinically relevant biomolecules. Chem. Rev. 2016, 116, 9001–9090. [Google Scholar] [CrossRef]
  8. Kwong, G.A.; Ghosh, S.; Gamboa, L.; Patriotis, C.; Srivastava, S.; Bhatia, S.N. Synthetic biomarkers: A twenty-first century path to early cancer detection. Nat. Rev. Cancer 2021, 21, 655–668. [Google Scholar] [CrossRef]
  9. Sadighbayan, D.; Sadighbayan, K.; Khosroushahi, A.Y.; Hasanzadeh, M. Recent advances on the DNA-based electrochemical biosensing of cancer biomarkers: Analytical approach. TrAC Trends Anal. Chem. 2019, 119, 115609. [Google Scholar] [CrossRef]
  10. Zhang, Z.J.; Sen, P.; Adhikari, B.R.; Li, Y.F.; Soleymani, L. Development of nucleic-acid-based electrochemical biosensors for clinical applications. Angew. Chem. Int. Ed. 2022, 61, e202212496. [Google Scholar]
  11. Arya, S.K.; Estrela, P. Electrochemical ELISA protein biosensing in undiluted serum using a polypyrrole-based platform. Sensors 2020, 20, 2857. [Google Scholar] [CrossRef]
  12. Pashayan, N.; Pharoah, P. The challenge of early detection in cancer. Science 2020, 368, 589–590. [Google Scholar] [CrossRef] [PubMed]
  13. Rebbeck, T.R.; Burns-White, K.; Chan, A.T.; Emmons, K.; Freedman, M.; Hunter, D.J.; Kraft, P.; Laden, F.; Mucci, L.; Parmigiani, G.; et al. Precision prevention and early detection of cancer: Fundamental principles. Cancer Discov. 2018, 8, 803–811. [Google Scholar] [CrossRef]
  14. Huang, X.L.; Zhan, S.N.; Xu, H.Y.; Meng, X.W.; Xiong, Y.H.; Chen, X.Y. Ultrasensitive fluorescence immunoassay for detection of ochratoxin A using catalase-mediated fluorescence quenching of CdTe QDs. Nanoscale 2016, 8, 9390. [Google Scholar] [CrossRef]
  15. Lai, W.Q.; Wei, Q.H.; Zhuang, J.Y.; Lu, M.H.; Tang, D.P. Fenton reaction-based colorimetric immunoassay for sensitive detection of brevetoxin B. Biosens. Bioelectron. 2016, 80, 249–256. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, C.; Han, Q.; Liu, P.K.; Zhang, G.; Song, L.; Zou, X.C.; Fu, Y.Z. Novel enhanced lanthanide electrochemiluminescence luminophores: Ce3+-doped TbPO4 facile synthesis and detection for mucin1. Anal. Chem. 2021, 93, 12289–12295. [Google Scholar] [CrossRef] [PubMed]
  17. Zhu, S.; Ran, P.Y.; Wu, J.L.; Chen, M.; Fu, Y.Z. An electrochemiluminesence chiral sensor for propranolol enantiomers based on functionalized graphite-like carbon nitride nanosheets. Electroanalysis 2020, 32, 185–190. [Google Scholar] [CrossRef]
  18. Fang, Y.; Wang, H.M.; Gu, Y.X.; Yu, L.; Wang, A.J.; Yuan, P.X.; Feng, J.J. Highly Enhanced Electrochemiluminescence Luminophore Generated by Zeolitic Imidazole Framework-8-Linked Porphyrin and Its Application for Thrombin Detection. Anal. Chem. 2020, 92, 3206–3212. [Google Scholar] [CrossRef]
  19. Hu, L.H.; Shi, T.F.; Chen, J.Y.; Cui, Q.Q.; Yu, H.; Wu, D.; Ma, H.M.; Wei, Q.; Ju, H.X. Dual-quenching electrochemiluminescence resonance energy transfer system from CoPd nanoparticles enhanced porous g-C3N4 to FeMOFs-sCuO for neuron-specific enolase immunosensing. Biosens. Bioelectron. 2023, 226, 115132. [Google Scholar] [CrossRef]
  20. Tyan, Y.C.; Jong, S.B.; Yang, M.H.; Wu, T.L.; Chung, T.W.; Liao, P.C.; Lin, C.Y.; Huang, Y.F. Utilizing isotope dilution-matrix-assisted laser desorption ionization-time of flight mass spectrometry as a reference procedure for the radioimmunoassay of serum thyroxine. Clin. Chim. Acta 2013, 420, 99–103. [Google Scholar] [CrossRef]
  21. Bai, Y.; Leng, D.Q.; Feng, T.; Kuang, X.; Fan, D.W.; Ren, X.; Li, Y.Y.; Wei, Q.; Ju, H.X. A split-type photoelectrochemical immunosensor based on a high-performance In2O3/BiVO4 photoelectrode modulated by a ZIF-8 protective layer. Sens. Actuators B Chem. 2023, 382, 133479. [Google Scholar] [CrossRef]
  22. Zhang, Y.Y.; Wu, T.T.; Cui, Q.Q.; Qu, Z.F.; Zhang, Y.; Ma, H.M.; Wei, Q. ReS2@Au NPs as signal labels quenching steady photocurrent generated by NiCo2O4/CdIn2S4/In2S3 heterojunction for sensitive detection of CYFRA 21-1. Biosens. Bioelectron. 2023, 222, 114992. [Google Scholar] [CrossRef] [PubMed]
  23. Li, X.F.; Ren, X.; Yang, L.; Wang, W.; Fan, D.W.; Kuang, X.; Sun, X.; Wei, Q.; Ju, H.X. Ru(dcbpy)32+-functionalized γ-cyclodextrin metal-organic frameworks as efficient electrochemiluminescence tags for the detection of CYFRA21-1 in human serum. Sens. Actuators B Chem. 2023, 378, 133152. [Google Scholar] [CrossRef]
  24. Chen, M.; Mo, F.J.; Meng, H.; Wang, C.; Guo, J.; Fu, Y.Z. Efficient Curing Sacrificial Agent-Induced Dual-Heterojunction Photoelectrochemical System for Highly Sensitive Immunoassay. Anal. Chem. 2021, 93, 2464–2470. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, M.; Guo, J.; Mo, F.J.; Yu, W.Q.; Fu, Y.G. Highly Sensitive Photoelectrochemical Immunosensor Based on Organic Multielectron Donor Nanocomposite as Signal Probe. Anal. Chem. 2022, 94, 17039–17045. [Google Scholar] [CrossRef]
  26. Song, P.; Wang, M.L.; Duan, Y.X.; Wang, A.J.; Xue, Y.D.; Mei, L.P.; Feng, J.J. Bifunctional photoelectrochemical aptasensor based on heterostructured Ag3PO4/Ag/TiO2 nanorod array for determination of two tumor markers. Microchim. Acta 2023, 190, 85. [Google Scholar] [CrossRef]
  27. Tang, Z.X.; Ma, Z.F. Ratiometric ultrasensitive electrochemical immunosensor based on redox substrate and immunoprobe. Sci. Rep. 2016, 6, 35440. [Google Scholar] [CrossRef] [PubMed]
  28. Zhang, C.; Zhang, N.N.; Xu, Y.; Feng, J.J.; Yao, T.; Wang, F.; Ma, Z.F.; Han, H.L. Fenton reaction-mediated dual-attenuation of signal for ultrasensitive amperometric immunoassay. Biosens. Bioelectron. 2021, 178, 113009. [Google Scholar] [CrossRef]
  29. Zheng, Y.; Ma, Z.F. Dual-reaction triggered sensitivity amplification for ultrasensitive peptide-cleavage based electrochemical detection of matrix metalloproteinase-7. Biosens. Bioelectron. 2018, 108, 46–52. [Google Scholar] [CrossRef]
  30. Feng, J.J.; Liang, X.Y.; Ma, Z.F. New immunoprobe: Dual-labeling ZIF-8 embellished with multifunctional bovine serum albumin lamella for electrochemical immunoassay of tumor marker. Biosens. Bioelectron. 2021, 175, 112853. [Google Scholar] [CrossRef]
  31. Chen, X.J.; Wang, Y.Z.; Zhang, Y.Y.; Chen, Z.H.; Liu, Y.; Li, Z.L.; Li, J.H. Sensitive Electrochemical Aptamer Biosensor for Dynamic Cell Surface N-Glycan Evaluation Featuring Multivalent Recognition and Signal Amplification on a Dendrimer-Graphene Electrode Interface. Anal. Chem. 2014, 86, 4278–4286. [Google Scholar] [CrossRef]
  32. Wang, R.; Liu, W.D.; Wang, A.J.; Xue, Y.D.; Wu, L.; Feng, J.J. A new label-free electrochemical immunosensor based on dendritic core-shell AuPd@Au nanocrystals for highly sensitive detection of prostate specific antigen. Biosens. Bioelectron. 2018, 99, 458–463. [Google Scholar] [CrossRef]
  33. Zhang, J.X.; Lv, C.L.; Tang, C.; Wang, A.J.; Mei, L.P.; Song, P.; Feng, J.J. Sandwich-type ultrasensitive immunosensing of breast cancer biomarker based on core-shell Au@PdAg dog-bone-like nanostructures and Au@PtRh nanorods. Sens. Actuators B Chem. 2023, 382, 133497. [Google Scholar] [CrossRef]
  34. Chen, Y.; Wang, A.J.; Yuan, P.X.; Luo, X.L.; Xue, Y.D.; Feng, J.J. Three dimensional sea-urchin-like PdAuCu nanocrystals/ferrocene-grafted-polylysine as an efficient probe to amplify the electrochemical signals for ultrasensitive immunoassay of carcinoembryonic antigen. Biosens. Bioelectron. 2019, 132, 294–301. [Google Scholar] [CrossRef]
  35. Li, B.; Li, Y.X.; Li, C.C.; Yang, J.H.; Liu, D.L.; Wang, H.B.; 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] [PubMed]
  36. Pan, H.J.; Zhou, Q.Y.; Gong, Y.C.; Jia, L.P.; Zhang, W.; Shang, L.; Xue, Q.W.; Wei, Q.; Wang, H.S.; Ma, R.N. A universal distance-independent ratiometric electrochemical biosensing strategy based on competitive host-guest interactions for matrix metalloproteinase-2 detection. Sens. Actuators B Chem. 2023, 378, 133144. [Google Scholar] [CrossRef]
  37. Campuzano, S.; Yáñez-Sedeño, P.; Pingarrón, J.M. Electrochemical Affinity Biosensors Based on Selected Nanostructures for Food and Environmental Monitoring. Sensors 2020, 20, 5125. [Google Scholar] [CrossRef]
  38. Russo, M.J.; Han, M.; Desroches, P.E.; Manasa, C.S.; Dennaoui, J.; Quigley, A.F.; Kapsa, R.M.I.; Moulton, S.E.; Guijt, R.M.; Greene, G.W.; et al. Anti-fouling Strategies for Electrochemical Biosensing: Mechanisms and Performance toward Point of Care Based Diagnostic Applications. ACS Sens. 2021, 6, 1482–1507. [Google Scholar] [CrossRef]
  39. Felix, F.S.; Angnes, L. Electrochemical immunosensors-A powerful tool for analytical applications. Biosens. Bioelectron. 2018, 102, 470–478. [Google Scholar] [CrossRef]
  40. Feng, J.J.; Yao, T.; Ma, Z.F. Recent advances of peroxidase-active nanozymes in electrochemical immunoassay. Sens. Diagn. 2023. [Google Scholar] [CrossRef]
  41. Feng, T.T.; Ji, W.L.; Zhang, Y.; Wu, F.; Tang, Q.; Wei, H.; Mao, L.Q.; Zhang, M.N. Zwitterionic Polydopamine Engineered Interface for In Vivo Sensing with High Biocompatibility. Angew. Chem. Int. Ed. 2020, 59, 23445–23449. [Google Scholar] [CrossRef] [PubMed]
  42. Jiang, C.; Wang, G.X.; Hein, R.; Liu, N.Z.; Luo, X.L.; Davis, J.J. Anti-fouling Strategies for Selective In Vitro and In Vivo Sensing. Chem. Rev. 2020, 120, 3852–3889. [Google Scholar] [CrossRef] [PubMed]
  43. Bolduc, O.R.; Pelletier, J.N.; Masson, J.F. SPR Biosensing in Crude Serum Using Ultralow Fouling Binary Patterned Peptide SAM. Anal. Chem. 2010, 82, 3699–3706. [Google Scholar] [CrossRef] [PubMed]
  44. Fan, B.B.; Fan, Q.; Hu, L.L.; Cui, M.; Wang, X.Y.; Ma, H.M.; Wei, Q. Polydopamine-PEG−Folic Acid Conjugate Film Engineered TiO2 Nanotube Arrays for Photoelectrochemical Sensing of Folate Binding Protein. ACS Appl. Mater. Interfaces 2020, 12, 1877–1884. [Google Scholar] [CrossRef]
  45. Henrya, O.Y.F.; Sanchez, J.L.A.; O’Sullivan, C.K. Bipodal PEGylated alkanethiol for the enhanced electrochemical detection of genetic markers involved in breast cancer. Biosens. Bioelectron. 2010, 26, 1500–1506. [Google Scholar] [CrossRef]
  46. Ma, B.C.; Yao, T.; Meng, X.Z.; Xu, Y.; Ma, Z.F.; Han, H.L. Novel Immunoprobe Based on MOF-818 Synergizing with an Anti-fouling Sensing Interface to Improve Immunosensors. ACS Sustain. Chem. Eng. 2022, 10, 12041–12047. [Google Scholar] [CrossRef]
  47. Cui, M.; Song, Z.L.; Wu, Y.M.; Guo, B.; Fan, X.J.; Luo, X.L. A highly sensitive biosensor for tumor maker alpha fetoprotein based on poly(ethylene glycol) doped conducting polymer PEDOT. Biosens. Bioelectron. 2016, 79, 736–741. [Google Scholar] [CrossRef]
  48. Hui, N.; Sun, X.T.; Song, Z.L.; Niu, S.Y.; Luo, X.L. Gold nanoparticles and polyethylene glycols functionalized conducting polyaniline nanowires for ultrasensitive and low fouling immunosensing of alpha-fetoprotein. Biosens. Bioelectron. 2016, 86, 143–149. [Google Scholar] [CrossRef]
  49. Meng, X.Z.; Xu, Y.; Ma, B.C.; Ma, Z.F.; Han, H.L. Anti-fouling materials decorated immunoprobe and electrochemical sensing interface to improve immunoassay. Chem. Eng. J. 2022, 450, 137954. [Google Scholar] [CrossRef]
  50. Xu, Z.Y.; Han, R.; Liu, N.Z.; Gao, F.X.; Luo, X.L. Electrochemical biosensors for the detection of carcinoembryonic antigen with low fouling and high sensitivity based on copolymerized polydopamine and zwitterionic polymer. Sens. Actuators B Chem. 2020, 319, 128253. [Google Scholar] [CrossRef]
  51. Leng, C.; Hung, H.; Sieggreen, O.A.; Li, Y.T.; Jiang, S.Y.; Chen, Z. Probing the Surface Hydration of Nonfouling Zwitterionic and Poly(ethylene glycol) Materials with Isotopic Dilution Spectroscopy. J. Phys. Chem. C 2015, 119, 8775–8780. [Google Scholar] [CrossRef]
  52. Chen, S.F.; Zheng, J.; Li, L.Y.; Jiang, S.Y. Strong Resistance of Phosphorylcholine Self-Assembled Monolayers to Protein Adsorption: Insights into Nonfouling Properties of Zwitterionic Materials. J. Am. Chem. Soc. 2005, 127, 14473–14478. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, Z.; Xu, Y.; Zhang, Y.; Ma, B.C.; Ma, Z.F.; Han, H.L. Anti-fouling and sensitive biosensor based on multifunctional peptide and urease@ZIFs for metal matrix protease-7. Sens. Actuators B Chem. 2022, 364, 131844. [Google Scholar] [CrossRef]
  54. Cui, M.; Ma, Y.H.; Wang, L.; Wang, Y.; Wang, S.; Luo, X.L. Anti-fouling sensors based on peptides for biomarker detection. TrAC Trends Anal. Chem. 2020, 127, 115903. [Google Scholar] [CrossRef]
  55. Cui, M.; Wang, Y.; Jiao, M.X.; Jayachandran, S.; Wu, Y.M.; Fan, X.J.; Luo, X.L. Mixed Self-Assembled Aptamer and Newly Designed Zwitterionic Peptide as Anti-fouling Biosensing Interface for Electrochemical Detection of alpha-Fetoprotein. ACS Sens. 2017, 2, 490–494. [Google Scholar] [CrossRef]
  56. Cui, M.; Wang, Y.; Wang, H.P.; Wu, Y.M.; Luo, X.L. A label-free electrochemical DNA biosensor for breast cancer marker BRCA1 based on self-assembled anti-fouling peptide monolayer. Sens. Actuators B Chem. 2017, 244, 742–749. [Google Scholar] [CrossRef]
  57. Han, R.; Li, Y.; Chen, M.; Li, W.T.; Ding, C.F.; Luo, X.L. Anti-fouling Electrochemical Biosensor Based on the Designed Functional Peptide and the Electrodeposited Conducting Polymer for CTC Analysis in Human Blood. Anal. Chem. 2022, 94, 2204–2211. [Google Scholar] [CrossRef] [PubMed]
  58. Wang, S.; Che, Z.M.; Guo, C.E.; Liu, Y.Q.; Yang, S.H.; Zhou, M.Y.; Gong, Y.H.; Li, T.D.; Cui, M.; Luo, X.L. A durable anti-fouling protein molecularly imprinted gel interface for human serum albumin detection and antibacterial application. Chem. Eng. J. 2021, 421, 129752. [Google Scholar] [CrossRef]
  59. Zhao, L.H.; Yin, S.; Ma, Z.F. Ca2+-Triggered pH-Response Sodium Alginate Hydrogel Precipitation for Amplified Sandwich-Type Impedimetric Immunosensor of Tumor Marker. ACS Sens. 2019, 4, 450–455. [Google Scholar] [CrossRef]
  60. Li, W.X.; Shu, D.; Han, H.L.; Ma, Z.F. An amperometric immunoprobe based on multifunctional nanogel for sensitive detection of tumor marker. Sens. Actuators B Chem. 2018, 273, 1451–1455. [Google Scholar] [CrossRef]
  61. Cui, M.; Gong, Y.H.; Du, M.G.; Wang, K.; Li, T.D.; Zhu, X.L.; Wang, S.; Luo, X.L. An anti-fouling electrochemical biosensor based on a protein imprinted hydrogel for human immunoglobulin G recognition in complex biological media. Sens. Actuators B Chem. 2021, 337, 129820. [Google Scholar] [CrossRef]
  62. Ma, L.Z.; Jayachandran, S.; Li, Z.M.; Song, Z.; Wang, W.; Luo, X.L. Anti-fouling and conducting PEDOT derivative grafted with polyglycerol for highly sensitive electrochemical protein detection in complex biological media. J. Electro. Anal. Chem. 2019, 840, 272–278. [Google Scholar] [CrossRef]
  63. Lian, M.L.; Shi, Y.Q.; Chen, L.X.; Qin, Y.G.; Zhang, W.; Zhao, J.B.; Chen, D. Cell Membrane and V2C MXene-Based Electrochemical Immunosensor with Enhanced Anti-fouling Capability for Detection of CD44. ACS Sens. 2022, 7, 2701–2709. [Google Scholar] [CrossRef] [PubMed]
  64. Mokni, M.; Tlili, A.; Attia, G.; Khaoulani, S.; Zerrouki, C.; Omezzine, A.; Othmane, A.; Bouslama, A.; Fourati, N. Novel sensitive immunosensor for the selective detection of Engrailed 2 urinary prostate cancer biomarker. Biosens. Bioelectron. 2022, 217, 114678. [Google Scholar] [CrossRef]
  65. Feng, J.J.; Yao, T.; Chu, C.S.; Ma, Z.F.; Han, H.L. Proton-responsive annunciator based on i-motif DNA structure modified metal organic frameworks for ameliorative construction of electrochemical immunosensing interface. J. Colloid Interface Sci. 2022, 608, 2050–2057. [Google Scholar] [CrossRef] [PubMed]
  66. Xu, Y.; Zheng, Y.; Ma, Z.F. ZIF-8 silencing shell removed by complexation competition reaction for ultrasensitive electrochemical immunoassay. Sens. Actuators B Chem. 2020, 307, 127647. [Google Scholar] [CrossRef]
  67. Yao, T.; Feng, J.J.; Chu, C.S.; Ma, Z.F.; Han, H.L. Cascade controlled release system based on pH-responsive ZIF-8 capsule and enzyme-responsive hyaluronic acid hydrogel for tumor marker detection using electro-readout-mode. Sens. Actuators B Chem. 2021, 348, 130701. [Google Scholar] [CrossRef]
  68. Yin, S.; Ma, Z.F. “Smart” sensing interface for the improvement of electrochemical immunosensor based on enzyme-Fenton reaction triggered destruction of Fe3+ cross-linked alginate hydrogel. Sens. Actuators B Chem. 2019, 281, 857–863. [Google Scholar] [CrossRef]
  69. Zhang, Z.; Xu, Y.; Ma, B.; Ma, Z.F.; Han, H.L. A novel electrochemical sensor based on process-formed laccase-like catalyst to degrade polyhydroquinone for tumor marker. Talanta 2021, 235, 122736. [Google Scholar] [CrossRef]
  70. Martína, C.M.; Gamellaa, M.; Pedreroa, M.; Montero-Calleb, A.; Barderasb, R.; Campuzanoa, S.; Pingarrón, J.M. Magnetic beads-based electrochemical immunosensing of HIF-1α, a biomarker of tumoral hypoxia. Sens. Actuators B Chem. 2020, 307, 127623. [Google Scholar] [CrossRef]
  71. Freitasa, M.; Nouwsa, H.P.A.; Keating, E.; Delerue-Matos, C. High-performance electrochemical immunomagnetic assay for breast cancer Analysis. Sens. Actuators B Chem. 2020, 308, 127667. [Google Scholar] [CrossRef]
  72. Koo, K.M.; Soda, N.; Shiddiky, M.J.A. Magnetic nanomaterial-based electrochemical biosensors for the detection of diverse circulating cancer biomarkers. Curr. Opin. Electrochem. 2021, 25, 100645. [Google Scholar] [CrossRef]
  73. Ostuni, E.; Chapman, R.G.; Holmlin, R.E.; Takayama, S.; Whitesides, G.M. A Survey of Structure-Property Relationships of Surfaces that Resist the Adsorption of Protein. Langmuir 2001, 17, 5605–5620. [Google Scholar] [CrossRef]
  74. Ahmed, S.; Ansari, A.; Haidyrah, A.S.; Chaudhary, A.A.; Imran, M.; Khan, A. Hierarchical Molecularly Imprinted Inverse Opal-Based Platforms for Highly Selective and Sensitive Determination of Histamine. ACS Appl. Polym. Mater. 2022, 4, 2783–2793. [Google Scholar] [CrossRef]
  75. Imran, M.; Ahmed, S.; Abdullah, A.Z.; Hakami, J.; Chaudhary, A.A.; Rudayni, H.A.; Khan, S.-U.; Khan, A.; Basher, N.S. Nanostructured material-based optical and electrochemical detection of amoxicillin antibiotic. Luminescence 2022, 1, 1–23. [Google Scholar] [CrossRef]
  76. Han, R.; Wang, G.X.; Xu, Z.Y.; Zhang, L.Y.; Li, Q.; Han, Y.F.; Luo, X.L. Designed anti-fouling peptides planted in conducting polymers through controlled partial doping for electrochemical detection of biomarkers in human serum. Biosens. Bioelectron. 2020, 164, 112317. [Google Scholar] [CrossRef] [PubMed]
  77. Zhang, Z.; Yao, T.; Han, H.L.; Ma, Z.F. Universal and High-Speed Zeptomolar Protein Serum Assay with Unprecedented Sensitivity. Anal. Chem. 2022, 94, 16231–16236. [Google Scholar] [CrossRef]
  78. Wang, Q.; Zhou, L.; Li, J.Z.; Li, Z.; Wang, T. Enhanced Interfacial Affinity of the Supercapacitor Electrode with a Hydrogel Electrolyte by a Preadsorbed Polyzwitterion Layer. Langmuir 2022, 38, 8614–8622. [Google Scholar] [CrossRef] [PubMed]
  79. Shao, Q.; Jiang, S.Y. Molecular Understanding and Design of Zwitterionic Materials. Adv. Mater. 2015, 27, 15–26. [Google Scholar] [CrossRef] [PubMed]
  80. Miodek, A.; Regan, E.M.; Bhalla, N.; Hopkins, N.A.E.; Goodchild, S.A.; Estrela, P. Optimisation and Characterisation of Anti-fouling Ternary SAM Layers for Impedance-Based Aptasensors. Sensors 2015, 15, 25015–25032. [Google Scholar] [CrossRef] [PubMed]
  81. Yao, T.; Feng, J.J.; Xiong, Q.C.; Chu, C.S.; Xu, Y.; Ma, Z.F.; Han, H.L. Regenerating electrochemical detection platform by electro-oxidation me-diated host-guest dissociation between 6-mercapto-6-deoxy-β-cyclodextrin and N,N-dimethylaminomethylferrocene. Chem. Eng. J. 2022, 439, 135599. [Google Scholar] [CrossRef]
  82. Ederth, T.; Lerm, M.; Orihuela, B.; Rittschof, D. Resistance of Zwitterionic Peptide Monolayers to Biofouling. Langmuir 2019, 35, 1818–1827. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, N.Z.; Hui, N.; Davis, J.J.; Luo, X.L. Low Fouling Protein Detection in Complex Biological Media Supported by a Designed Mul-tifunctional Peptide. ACS Sens. 2018, 3, 1210–1216. [Google Scholar] [CrossRef]
  84. Chen, S.F.; Cao, Z.Q.; Jiang, S.Y. Ultra-low fouling peptide surfaces derived from natural amino acids. Biomaterials 2009, 30, 5892–5896. [Google Scholar] [CrossRef] [PubMed]
  85. Zhao, S.J.; Liu, N.Z.; Wang, W.Q.; Xu, Z.Y.; Wu, Y.M.; Luo, X.L. An electrochemical biosensor for alpha-fetoprotein detection in human serum based on peptides containing isomer D-Amino acids with enhanced stability and anti-fouling property. Biosens. Bioelectron. 2021, 190, 113466. [Google Scholar] [CrossRef]
  86. Wang, F.; Gong, Y.C.; Xu, Y.; Ma, Z.F.; Han, H.L. Electrochemical sensing interface based on the oriented self-assembly of histidine labeled peptides induced by Ni2+ for protease detection. Biosens. Bioelectron. 2023, 230, 115259. [Google Scholar] [CrossRef]
  87. Li, Y.X.; Song, Z.; Chen, M.; Xu, Z.Y.; Zhao, S.J.; Xu, Y.Q.; Luo, X.L. Designed multifunctional peptides with two recognizing branches specific for one target to achieve highly sensitive and low fouling electrochemical protein assay in human serum. Anal. Chim. Acta 2022, 1208, 339841. [Google Scholar] [CrossRef]
  88. Chen, M.; Han, R.; Li, Y.; Luo, X.L. Nonfouling and ratiometric electrochemical detection of prostate specific antigen in whole serum. Anal. Chim. Acta 2022, 1224, 340191. [Google Scholar] [CrossRef]
  89. Chen, M.; Song, Z.; Han, R.; Li, Y.; Luo, X. Low fouling electrochemical biosensors based on designed Y-shaped peptides with anti-fouling and recognizing branches for the detection of IgG in human serum. Biosens. Bioelectron. 2021, 178, 113016. [Google Scholar] [CrossRef]
  90. Zhao, W.W.; Shi, Q.H.; Sun, Y. Dual-ligand affinity systems with octapeptide ligands for affinity chromatography of hIgG and monoclonal antibody. J Chromatogr. A 2014, 1369, 64–72. [Google Scholar] [CrossRef]
  91. Yin, S.; Ma, Z.F. Electrochemical immunoassay for tumor markers based on hydrogels. Expert Rev. Mol. Diagn. 2018, 18, 457–465. [Google Scholar] [CrossRef] [PubMed]
  92. Wang, H.Q.; Han, H.L.; Ma, Z.F. Conductive hydrogel composed of 1,3,5-benzenetricarboxylic acid and Fe3+ used as enhanced electrochemical immunosensing substrate for tumor biomarker. Bioelectrochemistry 2017, 114, 48–53. [Google Scholar] [CrossRef]
  93. Wang, H.Q.; Ma, Z.F. Ultrasensitive amperometric detection of the tumor biomarker cytokeratin antigen using a hydrogel composite consisting of phytic acid, Pb(II) ions and gold nanoparticles. Microchim. Acta 2017, 184, 1045–1050. [Google Scholar] [CrossRef]
  94. Zhao, L.H.; Ma, Z.F. Facile synthesis of polyaniline-polythionine redox hydrogel: Conductive, anti-fouling and enzyme-linked material for ultrasensitive label-free amperometric immunosensor toward carcinoma antigen-125. Anal. Chim. Acta 2018, 997, 60–66. [Google Scholar] [CrossRef]
  95. Ghaani, M.; Büyüktaş, D.; Carullo, D.; Farris, S. Development of a New Electrochemical Sensor Based on Molecularly Imprinted Biopolymer for Determination of 4,4-Methylene Diphenyl Diamine. Sensors 2023, 23, 46. [Google Scholar] [CrossRef] [PubMed]
  96. Li, Y.; Han, R.; Chen, M.; Zhang, L.Y.; Wang, G.X.; Luo, X.L. Bovine Serum Albumin-Cross-Linked Polyaniline Nanowires for Ultralow Fouling and Highly Sensitive Electrochemical Protein Quantification in Human Serum Samples. Anal. Chem. 2021, 93, 4326–4333. [Google Scholar] [CrossRef]
  97. Zhang, D.S.; Li, W.X.; Ma, Z.F.; Han, H.L. Improved ELISA for tumor marker detection using electro-readout-mode based on label triggered degradation of methylene blue. Biosens. Bioelectron. 2019, 126, 800–805. [Google Scholar] [CrossRef] [PubMed]
  98. Chatziharalambous, D.; Lygirou, V.; Latosinska, A.; Stravodimos, K.; Vlahou, A.; Jankowski, V.; Zoidakis, J. Analytical Performance of ELISA Assays in Urine: One More Bottleneck towards Biomarker Validation and Clinical Implementation. PLoS ONE 2016, 11, e0149471. [Google Scholar] [CrossRef]
Figure 1. The fabrication process of the electrochemical immunosensor based on MOF-818 synergizing with an anti-fouling sensing interface. Reprinted with permission from Ref. [46], Copyright @ 2022, American Chemical Society.
Figure 1. The fabrication process of the electrochemical immunosensor based on MOF-818 synergizing with an anti-fouling sensing interface. Reprinted with permission from Ref. [46], Copyright @ 2022, American Chemical Society.
Sensors 23 05202 g001
Figure 2. The fabrication process of the AFP immunosensor (a) PANI nanowires deposited on the GCE, (b) AuNPs electrodeposition, (c) PEG modification, (d) AFP antibody immobilization, (e) AFP target capturing, and (f) DPV current signal recording with/without AFP. Reprinted with permission from Ref. [48], Copyright @ 2016, Elsevier.
Figure 2. The fabrication process of the AFP immunosensor (a) PANI nanowires deposited on the GCE, (b) AuNPs electrodeposition, (c) PEG modification, (d) AFP antibody immobilization, (e) AFP target capturing, and (f) DPV current signal recording with/without AFP. Reprinted with permission from Ref. [48], Copyright @ 2016, Elsevier.
Sensors 23 05202 g002
Figure 3. Schematic presentation of EC Sensor at zeptomolar level. Reprinted with permission from Ref. [77], Copyright @ 2022, American Chemical Society.
Figure 3. Schematic presentation of EC Sensor at zeptomolar level. Reprinted with permission from Ref. [77], Copyright @ 2022, American Chemical Society.
Sensors 23 05202 g003
Figure 4. The fabrication process of the AFP immunosensor. Reprinted with permission from Ref. [50], Copyright @ 2020, Elsevier.
Figure 4. The fabrication process of the AFP immunosensor. Reprinted with permission from Ref. [50], Copyright @ 2020, Elsevier.
Sensors 23 05202 g004
Figure 5. The detection process of the anti-fouling electrochemical sensor for metal matrix protease-7. Reprinted with permission from Ref. [53], Copyright @ 2022, Elsevier.
Figure 5. The detection process of the anti-fouling electrochemical sensor for metal matrix protease-7. Reprinted with permission from Ref. [53], Copyright @ 2022, Elsevier.
Sensors 23 05202 g005
Figure 6. The fabrication process of the anti-fouling electrochemical biosensor based on the designed multifunctional peptide with two recognizing branches of the target IgG. Reprinted with permission from Ref. [88], Copyright @ 2022, Elsevier.
Figure 6. The fabrication process of the anti-fouling electrochemical biosensor based on the designed multifunctional peptide with two recognizing branches of the target IgG. Reprinted with permission from Ref. [88], Copyright @ 2022, Elsevier.
Sensors 23 05202 g006
Figure 7. The schematic diagram for the fabrication of anti-fouling biosensors based on designed Y-shaped peptides. Reprinted with permission from Ref. [89], Copyright @ 2021, Elsevier.
Figure 7. The schematic diagram for the fabrication of anti-fouling biosensors based on designed Y-shaped peptides. Reprinted with permission from Ref. [89], Copyright @ 2021, Elsevier.
Sensors 23 05202 g007
Figure 8. The schematic diagram of the label-free amperometric immunosensor for CA12-5. Reprinted with permission from Ref. [94], Copyright @ 2018, Elsevier.
Figure 8. The schematic diagram of the label-free amperometric immunosensor for CA12-5. Reprinted with permission from Ref. [94], Copyright @ 2018, Elsevier.
Sensors 23 05202 g008
Figure 9. Schematic fabrication protocol of the electrochemical biosensor based on IgG-imprinted hydrogel. Reprinted with permission from Ref. [61], Copyright @ 2021, Elsevier.
Figure 9. Schematic fabrication protocol of the electrochemical biosensor based on IgG-imprinted hydrogel. Reprinted with permission from Ref. [61], Copyright @ 2021, Elsevier.
Sensors 23 05202 g009
Figure 10. Schematic illustrations of the “smart” electrochemical sensing interface for NSE. Reprinted with permission from Ref. [68], Copyright @ 2019, Elsevier.
Figure 10. Schematic illustrations of the “smart” electrochemical sensing interface for NSE. Reprinted with permission from Ref. [68], Copyright @ 2019, Elsevier.
Sensors 23 05202 g010
Figure 11. Schematic illustration of the fabrication process of the AFP biosensor with PEDOT-HPG. Reprinted with permission from Ref. [62], Copyright @ 2019, Elsevier.
Figure 11. Schematic illustration of the fabrication process of the AFP biosensor with PEDOT-HPG. Reprinted with permission from Ref. [62], Copyright @ 2019, Elsevier.
Sensors 23 05202 g011
Figure 12. The fabrication process of the electrochemical immunosensor with anti-fouling capability for detection of CD44. Reprinted with permission from Ref. [63], Copyright @ 2022, American Chemical Society.
Figure 12. The fabrication process of the electrochemical immunosensor with anti-fouling capability for detection of CD44. Reprinted with permission from Ref. [63], Copyright @ 2022, American Chemical Society.
Sensors 23 05202 g012
Table 1. Summary of anti-fouling materials with principles and limitations.
Table 1. Summary of anti-fouling materials with principles and limitations.
Anti-Fouling MaterialsStrong HydrophilicityElectrically NeutralLimitations
PEG and its derivativesProne to oxidate into aldehydes and ethers by oxygen, resulting in polymer long-chain breakage and shedding
Zwitterionic materialsElectric field-sensitive or pH-sensitive zwitterions are used under limited conditions
Peptides Suffer from oxidative damage or protease degradation
Hydrogels Low mechanical strength, high water content, poor freezing resistance
BSA Cannot ensure full blocking of the active sites and may detach from the surfaces gradually
√ refers to positive in this principle.
Table 2. Summary of the biomarkers, anti-fouling material type, detection range, and detection limit of the anti-fouling electrochemical sensor.
Table 2. Summary of the biomarkers, anti-fouling material type, detection range, and detection limit of the anti-fouling electrochemical sensor.
BiomarkerAnti-Fouling MaterialDetection RangeDetection LimitYearRef.
FBPPEG1 × 10−3 to 5 × 102 ng mL−12 × 10−4 ng mL−12019[44]
CA72-4PEG1 μU mL−1 to 10 U mL−126.48 nU mL−12022[46]
AFPPEG0.001 fg mL−1 to 10 fg mL−10.0003 fg mL−12016[47]
AFPPEG0.01 pg mL−1–1.0 ng mL−10.007 pg mL−12016[48]
CA19-9PEG0.00001 to 100 U mL−11.03 μU mL−12022[49]
IgGPEG1 ng mL−1 to 100 ag mL−16.31 ag mL−1
(0.04 zeptomoles mL−1)
2022[77]
CEAZwitterionic0.01–10 pg mL−13.3 fg mL−12020[50]
SCCAZwitterionic1 pg mL−1 to 1 μg mL−131.20 fg mL−12022[81]
AFPZwitterionic Peptide10.0 fg mL−1 to 100.0 pg mL−13.1 fg mL−12017[55]
CA 15-3Peptides0.01–1000 U mL−13.34 mU mL−12020[76]
AFPPeptides0.1 fg mL−1 to 1.0 ng mL−10.03 fg mL−12021[85]
PSAPeptides1 pg mL−1 to 100 ng mL−111.8 fg mL−12023[86]
IgGPeptides0.1 pg mL−1 to 0.1 mg mL−10.031 pg mL−12022[87]
PSAPeptides5.0 pg mL−1 to 100 ng mL−11.26 pg mL−12022[88]
IgGPeptides100 pg mL−1 to 10 μg mL−132 pg mL−12021[89]
MMP-7Peptides0.1 pg mL−1–100 ng mL−124.34 fg mL−12022[53]
BRCA 1Peptides1.0 fM to 10.0 pM0.3 fM2017[56]
MCF-7 cancer cellsPeptides50–106 cells mL−117 cells mL−12022[57]
HSAHydrogels10−6–5 × 10−2 ng mL−10.03 × 10−6 ng mL−12021[58]
PSAHydrogels1.0 fg mL−1 to 100 ng mL−10.09 fg mL−12019[59]
NSEHydrogels0.01 to 1000 ng mL−14.6 pg mL−12018[60]
IgGHydrogels0.5–200.0 ng mL−10.03 ng mL−12021[61]
NSEHydrogels1 pg mL−1 to 200 ng mL−10.26 pg mL−12017[92]
CYFRA21-1Hydrogels50 fg mL−1 to 100 ng mL−138 fg mL−12017[93]
CA125Hydrogels0.0001 U mL−1 to 1 kU mL−10.00125 U mL−12018[94]
SCCABSA1 mg mL−1 to 1 pg mL−11.504 fg mL−12022[65]
CYFRA21-1BSA10 fg mL−1 to 1 μg mL−13.175 fg mL−12020[66]
PSABSA1.0 × 10−3–1.0 × 102 ng mL−137.27 fg mL−12021[67]
NSEBSA1 pg mL−1 to 100 ng mL−10.447 pg mL−12019[68]
HE4BSA1 pg mL−1 to 100 ng mL−10.302 pg mL−12021[69]
HIF-1αBSA0.25–10.0 ng mL−176 pg mL−12020[70]
HER 2-ECDBSA5.0 and 50 ng mL−1
50 and 100 ng mL−1
2.8 ng mL−1
3 cells mL−1
2020[71]
CA 125BSA0.1 mU mL−1 to 500 U mL−10.048 mU mL−12019[97]
AFPPolyglycerol0.10 pg mL−1–1.0 ng mL−10.035 pg mL−12019[62]
CD44
CD44 Cancer Cell
Cell Membrane0.5 ng mL−1 to 500 ng mL−1
103 to 106 cells mL−1
1.4 pg mL−1
140 cells mL−1
2022[63]
EN 2Histamine10−5 ng mL−1 to 1 μg mL−110−5 ng mL−12022[64]
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

Song, G.; Han, H.; Ma, Z. Anti-Fouling Strategies of Electrochemical Sensors for Tumor Markers. Sensors 2023, 23, 5202. https://doi.org/10.3390/s23115202

AMA Style

Song G, Han H, Ma Z. Anti-Fouling Strategies of Electrochemical Sensors for Tumor Markers. Sensors. 2023; 23(11):5202. https://doi.org/10.3390/s23115202

Chicago/Turabian Style

Song, Ge, Hongliang Han, and Zhanfang Ma. 2023. "Anti-Fouling Strategies of Electrochemical Sensors for Tumor Markers" Sensors 23, no. 11: 5202. https://doi.org/10.3390/s23115202

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