ACS Publications. Most Trusted. Most Cited. Most Read
Highly Efficient Real-Time TRPV1 Screening Methodology for Effective Drug Candidates
My Activity
  • Open Access
Article

Highly Efficient Real-Time TRPV1 Screening Methodology for Effective Drug Candidates
Click to copy article linkArticle link copied!

  • Seong Gi Lim
    Seong Gi Lim
    Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of Korea
    More by Seong Gi Lim
  • Sung Eun Seo
    Sung Eun Seo
    Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of Korea
    Department of Civil and Environmental Engineering, Yonsei University, Seoul 03722, Republic of Korea
    More by Sung Eun Seo
  • Seongjae Jo
    Seongjae Jo
    Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of Korea
    More by Seongjae Jo
  • Kyung Ho Kim
    Kyung Ho Kim
    Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of Korea
    More by Kyung Ho Kim
  • Lina Kim
    Lina Kim
    Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of Korea
    More by Lina Kim
  • Oh Seok Kwon*
    Oh Seok Kwon
    Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of Korea
    Department of Biotechnology, University of Science & Technology (UST), Daejeon 34141, Republic of Korea
    College of Biotechnology and Bioengineering, Sungkyunkwan University, Suwon 16419, Republic of Korea
    *Email: [email protected]
    More by Oh Seok Kwon
Open PDFSupporting Information (1)

ACS Omega

Cite this: ACS Omega 2022, 7, 41, 36441–36447
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.2c04202
Published October 4, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

Transient receptor potential vanilloid 1 (TRPV1) agonists that bind to the vanilloid pocket are being actively studied in the pharmaceutical industry to develop novel treatments for chronic pain and cancer. To discover synthetic vanilloids without the side effect of capsaicin, a time-consuming process of drug candidate selection is essential to a myriad of chemical compounds. Herein, we propose a novel approach to field-effect transistors for the fast and facile screening of lead vanilloid compounds for the development of TRPV1-targeting medications. The graphene field-effect transistor was fabricated with human TRPV1 receptor protein as the bioprobe, and various analyses (SEM, Raman, and FT-IR) were utilized to verify successful manufacture. Simulations of TRPV1 with capsaicin, olvanil, and arvanil were conducted using AutoDock Vina/PyMOL to confirm the binding affinity. The interaction of the ligands with TRPV1 was detected via the fabricated platform, and the collected responses corresponded to the simulation analysis.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2022 The Authors. Published by American Chemical Society

1. Introduction

Click to copy section linkSection link copied!

Transient receptor potential vanilloid 1 (TRPV1) is a vanilloid receptor that is activated by pungent chemicals and noxious stimuli such as heat, toxin, and acidic pH. (1,2) Among agonists, the prototypical capsaicin interaction with TRPV1 results in a unique condition called desensitization, which is a lasting refractory period after vanilloid treatment. (3) Recent studies have focused on vanilloids for the development of novel analgesics for chronic diseases, (4) because no stimulus of any kind is recognized after the application of capsaicin. (5,6) As a result, over-the-counter capsaicin creams are currently available to relieve arthritis pain, and 8% capsaicin patches can be prescribed for neuropathy. (7−9) However, the side effects of capsaicin limit its potential for further use in medicine, such as anticancer drugs. (10,11)
Capsaicin analogues without the side effect of pungency have been of great interest to the pharmaceutical industry, and synthetic vanilloids have been sought as candidates with high potential. (12) As a result, TRPV1 agonists are actively synthesized by altering the structure of capsaicin, which is divided into the phenol head, amide neck, and hydrophobic tail. (13) The key to a TRPV1 agonist study is to find a nonpungent compound that exhibits physiological behaviors similar to those of capsaicin. (14) To date, olvanil and arvanil are the most promising drug candidates because they bind to the desired vanilloid pocket yet are not pungent. In addition, they have a higher affinity for the TRPV1 receptor than capsaicin, implying higher medical efficacy. (14−16) The pursuit of finding effective TRPV1 agonists for the treatment of chronic pain and cancer is an active ongoing study in the field of drug development. (4,17)
Technological advances in the development of nanotechnology are being pursued for the facile selection of drug candidates because the discovery of lead chemicals requires highly demanding processes. The drug discovery process begins with a preliminary biological screening of multitudinous chemical compounds, which are synthesized and tested to assess their effectiveness. Nanotechnological approaches are advantageous for in vitro processes to identify compounds with potential utility because only trace amounts are needed for screening. Determining drug candidates among synthetic vanilloids that target the TRPV1 receptor can benefit from the introduction of nanotechnology, as it provides a rapid selection of potent therapeutic agents.
We developed a novel screening platform for the facile selection of TRPV1 agonists as therapeutic agents by demonstrating the potent efficacy of the compounds tested. The field-effect transistor with a graphene surface modified by the TRPV1 receptor collects responses in real time when the compounds are simply introduced into the platform. We described the platform manufacturing process and its validation and performed a molecular docking simulation using AutoDock Vina/PyMOL to determine the binding affinities of the tested synthetic vanilloids. Arvanil, olvanil, and capsaicin were selected as potential drug candidates, and their binding affinity and efficacy as TRPV1 agonists were studied. The collected responses of the vanilloid compounds were compared, and the corresponding calibration curves were provided (Scheme 1).

Scheme 1

Scheme 1. Schematic Illustration of the Overall Operation Process of the Fabricated Platform for the In Vitro Selection of Drug Candidates Using a TRPV1-Immobilized Graphene Field-Effect Transistor

2. Results and Discussion

Click to copy section linkSection link copied!

2.1. Characterization of the Electrode Components

Graphene was synthesized using bottom-up synthesis (Figure S1). The possible defects of graphene were identified using transmission electron microscopy (TEM). Figure 1a shows the selected area electron diffraction (SAED) pattern of the fabricated graphene, clearly revealing the hexagonal dot pattern of the graphene crystal structure. This indicates that graphene was produced with a high degree of crystallinity without an amorphous graphene oxide region. (18−20) Two sets of well-defined planes of graphene were recognized in the SAED pattern, with the inner and outer circles corresponding to the (100) and (110) planes, respectively. (18)

Figure 1

Figure 1. Identification of graphene and the TRPV1 bioprobe (a) SAED pattern of graphene. (b) Output curve of graphene. (c) SDS–PAGE (right) and Western blot (left) of the fabricated TRPV1 protein with the molecular weight of 98 kDa.

Figure 1b shows the output curve of the graphene-FET platform at different biases of the applied gate voltage. Since the linear regime is well demonstrated in the ohmic region, it was concluded that the contact between the drain/source terminals and graphene is adequate, and the charge carrier mobility is high. A typical trend of the graphene-FET with p-type characteristics was observed when a negative bias voltage was serially applied. As the negative bias was increased from −0.5 to −2.0 V, the amplitude of the negative current increased due to the increase in the carrier density of the transistor.
Figure 1c shows validation of the synthesis and purification of the TRPV1 protein using two analysis techniques with Coomassie blue staining: sodium dodecyl-sulfate polyacrylamide gel electrophoresis (SDS–PAGE) and Western blotting. (21) Due to the His tag at the N-terminus, the molecular weight of the synthesized protein was expected to exceed 95 kDa. (22) A band at 98 kDa was observed by SDS–PAGE using a 10% Tris-glycine gel, and the protein of interest was transferred onto a hydrophilic polyvinylidene fluoride (PVDF) membrane for the immunodetection of TRPV1. When the membrane was blocked with bovine serum albumin and incubated with His-tagged antibody, the same band at 98 kDa was observed, which was concluded to be the TRPV1 band.

2.2. Fabrication of the Field-Effect Transistor Template

A schematic diagram of the manufacture of the FET platform is shown in Figure 2a. After the graphene was transferred onto a silicon wafer, a microelectromechanical system (MEMS) was run to generate gate, source, and drain terminals (Figure S2). Surface modification was initiated by the formation of π–π interactions between graphene and the linker chemical PDI-diacid. (23) To further attach the bioprobe, the amidation reaction was conducted between the carboxyl group at the diacid terminal and the amine group of the TRPV1 protein.

Figure 2

Figure 2. Fabrication and identification of the electrode (a) schematic illustration of the fabrication process of the electrode, (b) color SEM image, (c) IV curves after each step of the surface modification, (d) transfer curve, (e) contact angle analysis, (f) FT-IR spectra, and (g) Raman spectra.

The scanning electrode microscopy (SEM) image of the manufactured electrode is shown in Figure 2b and colorization was applied for clarity in discrimination. The gold source and drain terminals are color coded yellow, graphene is black, and the wafer is blue. The graphene in black is located between the electrode, and the width is determined to be 56 μm, according to the inserted scale bar of 100 μm. Passivation was applied to prevent an unnecessary increase in resistance other than that of graphene because it was intended to collect responses from changes in the graphene surface.
In Figure 2c, the IV graph of the FET platform was measured after each process of surface modification of the fabricated electrodes was completed. The slope of the linear region is equal to the conductance, which is proportional to the mobility of the charge carriers. (24) A slight decrease in slope was observed from 0.424 to 0.394 after PDI-diacid formed a π–π interaction with the graphene substrate. However, a significant decrease to 0.187 was detected when the bioprobe was attached, as the macromolecule impairs the conductivity of the substrate.
The transfer curves for comparison after attaching the bioprobe are shown in Figure 2d. A negative shift in the voltage at the neutral point of charge (VCNP) was detected compared to the voltage of the pristine graphene. This is because the surface of graphene was modified with a protein with an isoelectric point of 6.9, which is negatively charged in pH 7.4 PBS buffer. (25) After introducing TRPV1 onto the substrate, the negative charge of the bioprobe caused an increase in the number of electrons, leading to a negative shift.
The hydrophilicity of the substrate surface after mounting the bioprobe was compared with that of pristine graphene and is shown in Figure 2e. Hydrophilicity was measured by the value of the contact angle, as they are in inverse proportion. (26,27) The contact angle of the water droplet on the graphene was measured to be 52.39°. After the hydrophobic TRPV1 adhered to the substrate, the contact angle increased to 78.01°. The change in contact angle from low to high indicates the completion of surface modification as it depends on the presence of hydrophobic proteins.
Fourier-transfer infrared (FT-IR) spectra of the pristine graphene and the modified graphene are compared in Figure 2f to confirm the bioprobe attachment. In pristine graphene, two peaks appeared at 1090 and 1270 cm–1 due to the oxidative defects of ether formation. After bioprobe attachment, the peaks corresponding to amide bonds were observed at 1530 cm–1 (amide II) and 1650 cm–1 (amide I). (28) Other absorption peaks that appear on biomaterials were also observed, such as the vibration of the hydroxyl group (vibration) at 3380 cm–1, asymmetric stretching of the ester at 1170 cm–1, and scissoring vibration of −CH2 at 1460 cm–1. (28)
Figure 2g shows Raman spectra to identify layers of graphene and confirm the attachment of the bioprobe. The pristine graphene exhibited two peaks at 1591 and 2690 cm–1, which were assigned to the G peak and secondary D peak (2D), respectively. The intensity ratio (I2D/IG) was calculated to be 1.52, which indicates a graphene monolayer structure. On the other hand, the peaks that appeared after bioprobe attachment were assigned as follows: amide I at 1595 cm–1, bending vibration of −CH2 at 1447 cm–1, and amide III from 1248–1364 cm–1. (29−31) Therefore, the bioprobe attachment was evidenced by the appearance of unique peaks in the protein structure.

2.3. Simulation of the Binding Mechanism of TRPV1 and Vanilloid Molecules

The binding affinity of vanilloids for the TRPV1 receptor protein was compared by using AutoDock Vina/PyMOL. Simulations were performed using natural vanilloid capsaicin and two actively studied nonpungent synthetic vanilloids, arvanil and olvanil. Figure 3 shows the affinity and ligand binding sites for each ligand within the vanilloid binding pocket of the receptor. A myriad of secondary bonds may be formed between ligands and the protein; however, only those involving tyrosine 511 and threonine 550 were considered because they are the major interaction sites. (32,33)Figure 3a shows the overall structure of the arvanil molecule in the vanilloid binding pocket, and details of the specific binding are further shown in Figure 3d,g. The affinity was calculated to be −6.7 kcal/mol, and the reason for the strong affinity was thought to be due to the following two hydrogen bonds involved in the interaction between TRPV1 and arvanil. One was a 2.9 Å long bond between the threonine residue and the oxygen atom at the neck of the vanilloid molecule. The other was a strong hydrogen bond with a length of 2.2 Å between the tyrosine residue and the hydrogen of the secondary amine of the arvanil neck. Figure 3b shows the geometric representation of the olvanil molecule, and its interaction with the TRPV1 protein is further shown in Figure 3e,h. The strong electronegative atoms of the olvanil neck structure were involved in the formation of hydrogen bonds with both T550 and Y511. Their lengths were 2.6 and 2.7 Å, respectively, and a high affinity of −6.3 kcal/mol was found between olvanil and the receptor protein. Capsaicin in the vanilloid pocket is shown in Figure 3c, and its binding details are shown in Figure 3f,i. The hydrogen atom of the amine group of the capsaicin neck formed a strong hydrogen bond to T550 with a length of 2.2 Å. However, the bond between the oxygen atom of the capsaicin head and the tyrosine residue was found to be a dipole–dipole interaction, but not a hydrogen bond because the bond length was 4.0 Å. The binding affinity of capsaicin was −6.2 kcal/mol, which was lower than that of synthetic vanilloids. Consistent with previous reports, (34) the results indicate that interactions with tyrosine residue correlate with ligand affinity, whereas threonine is essential for vanilloid pockets to recognize ligands.

Figure 3

Figure 3. Simulation of ligand–protein docking between TRPV1 and arvanil (a,d,g), olvanil (b,e,h), and capsaicin (c,f,i).

2.4. Real-Time Monitoring of Efficacy for Drug Candidates

The novel approach using a field-effect transistor was applied to collect the in vitro response of the synthesized vanilloids for screening drug target candidates. The fabricated FET platform collects the responses evoked by the interaction between the receptor protein and its vanilloid agonist. The efficacy of drug candidates is deducted from the limit of detection concentration and the amount of activity after the addition of the ligand. Meanwhile, no meaningful signal was detected after continuous insertion of ligand molecules into the platform using the pristine graphene compared to the TRPV1-anchored FET system. Therefore, the nonspecific binding of the ligands to graphene was ignored.
Since the medium for vanilloid insertion was a PBS solution with a pH of 7.4, TRPV1 with a value of pI 6.9 showed a negative charge. After injecting the vanilloid molecule into the chamber, an increase in current was detected as the ligand bound to the negatively charged receptor. The normalized sensitivity of vanilloids and the corresponding calibration curves using the Langmuir isotherm model (Figure S3) are illustrated in Figure 4. The interaction between arvanil and TRPV1 is shown in Figure 4a, and the corresponding calibration curve is shown in Figure 4d. When examining the response from arvanil, the strongest signals were collected with the lowest detection limit of 10 aM. In Figure 4b, the response from olvanil bound to the receptor was collected, and the calibration curve is shown in Figure 4e. For olvanil, meaningful signals were collected from 100 aM. Lastly, Figure 4c,f shows the responses and interaction between capsaicin and the receptor protein. The natural vanilloid capsaicin required the highest dose to bind TRPV1 in sufficient amounts to produce a detectable signal. Therefore, the efficacy of vanilloid for TRPV1 activation was in the order of arvanil, olvanil, and capsaicin. The deducted order was consistent with both the simulation results in Figure 3 and the results of previous studies. (15,35,36) On the other hand, it was concluded that synthetic vanilloids function better than capsaicin as drug candidates in that they were more available for binding to TRPV1 at low concentrations. In addition, synthetic vanilloids have been shown to be more effective because they showed higher sensitivity at lower concentrations in the fabricated FET systems.

Figure 4

Figure 4. Real-time monitoring and the corresponding calibrated response curves of arvanil (a,d), olvanil (b,e), and capsaicin (c,f) collected from the manufactured FET platform.

3. Conclusions

Click to copy section linkSection link copied!

In this study, we used the graphene-FET platform to observe the response of vanilloids to TRPV1 in real time and obtained data on its efficacy and potential as a new drug. The efficacy of the agonists was consistent with the order concluded from the simulations and previous studies, confirming the validity of the fabricated system: arvanil, olvanil, and capsaicin.
The manufactured platform is a novel methodology that can be applied to perform in vitro trials for drug candidate selection more quickly and easily. The screening of a large library of synthetic vanilloids to examine their efficacy in interacting with TRPV1 can be completed more quickly and easily when the fabricated platform is utilized. This study collected responses that were limited to measuring efficacy; however, further applications are possible by simply changing the experimental settings. For example, changing from PBS buffer to serum containing interfering species can produce results in a physiological environment.

4. Experimental Methods

Click to copy section linkSection link copied!

4.1. Synthesis of PDI-Diacid

Imidazole (1.60 g, 23.1 mmol) was placed in a 100 mL round-bottom flask and heated to 95 °C with constant stirring. When the initial compound became viscous, perylene-3,4,9,10-tetracarboxylic dianhydride (300 mg, 0.77 mmol) and aminobutyric acid (159 mg, 1.54 mmol) were added. The reaction was stirred for 16 h under an inert nitrogen atmosphere. The mixture was cooled to room temperature, and filter paper was used to remove the residue of perylene-3,4,9,10-tetracarboxylic dianhydride. Then, 1 M HNO3 was added dropwise to the filtrate to neutralize the product. After precipitation, distilled water was applied to the product for washing, and the product was dried in vacuo to obtain a dark red solid (402 mg, 93%).

4.2. Synthesis of TRPV1 Protein

The TRPV1 gene with a histidine tag (6× His) at the N terminus was cloned into the pET21b(+) vector, which was transformed into Escherichia coli BL21 (DE3) cells. The BL21 cells were first incubated in a Luria-Bertani medium with 100 mg/mL ampicillin, and then, 0.1 mM isopropyl β-d-thiogalactopyranoside was used for protein expression. After the harvest of the cultures, resuspension was conducted in 10 mL of 5% Triton X-100 in pH 7.4 PBS buffer solution and 20 min of ultrasonication was performed for lysis. The solution was centrifuged at 15,000g for 15 min and filtered with a 0.45 μM filter, and 8 M urea was added to refold the protein. The supernatant was inserted into 5 mL of HisTrap HP for purification of TRPV1 via fast protein liquid chromatography (FPLC). Buffer solutions containing 50 mM Tris-HCl (pH 8.0) and 0.5 M NaCl were utilized for FPLC and the imidazole concentrations of the solutions were as follows: 0 mM for the binding buffer, 20 mM for washing buffer, and 250 mM for elution buffer.

4.3. Simulation of Ligand–Protein Docking

Docking of vanilloids with the capsaicin receptor TRPV1 (PDB ID: 3J5R) was performed using AutoDock Vina 1.1.2. The vanilloid molecules were converted to structural data using ChemDraw 20.0 and optimized before docking. The vanilloid binding pocket was defined as a 30 Å long cube centered between the alpha carbons of the selected binding sites, tyrosine 511 and threonine 550. Secondary bonds between the receptor and agonists were identified, and their bond lengths were measured using PyMOL.

4.4. Real-Time Response from the FET Platform

Responses were collected from the Keithley 2612A source meter connected to the probe station (MS-TECH, model 4000A). Constant voltages for the drain/source and gate were applied during real-time monitoring (Vds = 10 mV and Vg = 0.1 V). (37−39) To construct a liquid phase transistor, a chamber containing 40 μL of pH 7.4 PBS solution was set up on the transistor. Each vanilloid solution was serially diluted from freshly prepared 1 mM stock solution and applied to the chamber.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.2c04202.

  • Temperature program for graphene synthesis; MEMS process for the electrode fabrication; and equations for response normalization and calibration (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Oh Seok Kwon - Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of KoreaDepartment of Biotechnology, University of Science & Technology (UST), Daejeon 34141, Republic of KoreaCollege of Biotechnology and Bioengineering, Sungkyunkwan University, Suwon 16419, Republic of KoreaOrcidhttps://orcid.org/0000-0002-5936-244X Email: [email protected]
  • Authors
    • Seong Gi Lim - Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of Korea
    • Sung Eun Seo - Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of KoreaDepartment of Civil and Environmental Engineering, Yonsei University, Seoul 03722, Republic of KoreaOrcidhttps://orcid.org/0000-0001-6261-9309
    • Seongjae Jo - Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of KoreaOrcidhttps://orcid.org/0000-0003-2864-9162
    • Kyung Ho Kim - Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of KoreaOrcidhttps://orcid.org/0000-0003-3587-0877
    • Lina Kim - Infectious Disease Research Center, Korea Research Institute of Bioscience and Biotechnology (KRIBB), Daejeon 34141, Republic of KoreaOrcidhttps://orcid.org/0000-0001-6460-1700
  • Author Contributions

    S.G.L., S.E.S., and S.J. contributed equally to this work. All authors have given approval to the final version of the manuscript.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

This work was supported by the Korea Institute of Planning and Evaluation for Technology in Food, Agriculture and Forestry (IPET) and Korea Smart Farm R&D Foundation (KosFarm) through the Smart Farm Innovation Technology Development Program, funded by the Ministry of Agriculture, Food and Rural Affairs (MAFRA) and Ministry of Science and ICT (MSIT); Rural Development Administration (RDA) (421020-03); the National Research Council of Science & Technology(NST) grant by the Korea government (MSIT) (no. CAP22011-000); Korea Ministry of Environment (MOE) through Technology Development Project for Safety Management of Household Chemical Products Program (or Project), funded by Korea Ministry of Environment (MOE) (2022002980005, 1485018893, 2022002980009, 1485018881); and the Korea Research Institute of Bioscience and Biotechnology (KRIBB) Research Initiative Program (1711134045).

Abbreviations

Click to copy section linkSection link copied!

TRPV1

transient receptor potential vanilloid 1

FET

field-effect transistor

References

Click to copy section linkSection link copied!

This article references 39 other publications.

  1. 1
    Cao, E.; Liao, M.; Cheng, Y.; Julius, D. TRPV1 Structures in Distinct Conformations Reveal Activation Mechanisms. Nature 2013, 504, 113118,  DOI: 10.1038/nature12823
  2. 2
    Bevan, S.; Quallo, T.; Andersson, D. A. Mammalian Transient Receptor Potential (TRP) Cation Channels; Nilius, B., Flockerzi, V., Eds.; Springer New York LLC, 2014; Vol. 222.
  3. 3
    Szallasi, A.; Blumberg, P. M. Vanilloid (Capsaicin) Receptors and Mechanisms. Pharmacol. Rev. 1999, 51, 159212
  4. 4
    O’Neill, J.; Brock, C.; Olesen, A. E.; Andresen, T.; Nilsson, M.; Dickenson, A. H. Unravelling the Mystery of Capsaicin: A Tool to Understand and Treat Pain. Pharmacol. Rev. 2012, 64, 939971,  DOI: 10.1124/pr.112.006163
  5. 5
    Premkumar, L. S.; Sikand, P. TRPV1 A Target for Next Generation Analgesics. Curr. Neuropharmacol. 2008, 6, 151163,  DOI: 10.2174/157015908784533888
  6. 6
    Knotkova, H.; Pappagallo, M.; Szallasi, A. Capsaicin (TRPV1 Agonist) Therapy for Pain Relief. Clin. J. Pain 2008, 24, 142154,  DOI: 10.1097/ajp.0b013e318158ed9e
  7. 7
    Mason, L.; Moore, R. A.; Derry, S.; Edwards, J. E.; McQuay, H. J. Systematic Review of Topical Capsaicin for the Treatment of Chronic Pain. Br. Med. J. 2004, 328, 991994,  DOI: 10.1136/bmj.38042.506748.ee
  8. 8
    Blair, H. A. Capsaicin 8% Dermal Patch: A Review in Peripheral Neuropathic Pain. Drugs 2018, 78, 14891500,  DOI: 10.1007/s40265-018-0982-7
  9. 9
    Peppin, J. F.; Pappagallo, M. Capsaicinoids in the Treatment of Neuropathic Pain: A Review. Ther. Adv. Neurol. Disord. 2014, 7, 2232,  DOI: 10.1177/1756285613501576
  10. 10
    Drewes, A. M.; Schipper, K. P.; Dimcevski, G.; Petersen, P.; Gregersen, H.; Funch-Jensen, P.; Arendt-Nielsen, L. Gut Pain and Hyperalgesia Induced by Capsaicin: A Human Experimental Model. Pain 2003, 104, 333341,  DOI: 10.1016/s0304-3959(03)00039-3
  11. 11
    LaMotte, R. .; Lundberg, L. E. R.; Torebjörk, H. E. Pain, Hyperalgesia and Activity in Nociceptive C Units in Humans after Intradermal Injection of Capsaicin. J. Physiol. 1992, 448, 749764,  DOI: 10.1113/jphysiol.1992.sp019068
  12. 12
    Karai, L.; Brown, D. C.; Mannes, A. J.; Connelly, S. T.; Brown, J.; Gandal, M.; Wellisch, O. M.; Neubert, J. K.; Olah, Z.; Iadarola, M. J. Deletion of Vanilloid Receptor 1-Expressing Primary Afferent Neurons for Pain Control. J. Clin. Invest. 2004, 113, 13441352,  DOI: 10.1172/jci20449
  13. 13
    Yang, F.; Zheng, J. Understand Spiciness: Mechanism of TRPV1 Channel Activation by Capsaicin. Protein Cell 2017, 8, 169177,  DOI: 10.1007/s13238-016-0353-7
  14. 14
    Appendino, G.; De Petrocellis, L.; Trevisani, M.; Minassi, A.; Daddario, N.; Moriello, A. S.; Gazzieri, D.; Ligresti, A.; Campi, B.; Fontana, G.; Pinna, C.; Geppetti, P.; Di Marzo, V. Development of the First Ultra-Potent “Capsaicinoid” Agonist at Transient Receptor Potential Vanilloid Type 1 (TRPV1) Channels and Its Therapeutic Potential. J. Pharmacol. Exp. Ther. 2005, 312, 561570,  DOI: 10.1124/jpet.104.074864
  15. 15
    Hurley, J. D.; Akers, A. T.; Friedman, J. R.; Nolan, N. A.; Brown, K. C.; Dasgupta, P. Non-Pungent Long Chain Capsaicin-Analogs Arvanil and Olvanil Display Better Anti-Invasive Activity than Capsaicin in Human Small Cell Lung Cancers. Cell Adhes. Migrat. 2017, 11, 8097,  DOI: 10.1080/19336918.2016.1187368
  16. 16
    Di Marzo, V.; Griffin, G.; De Petrocellis, L.; Brandi, I.; Bisogno, T.; Williams, W.; Grier, M. C.; Kulasegram, S.; Mahadevan, A. N. U.; Razdan, R. K.; Martin, B. R. A Structure/Activity Relationship Study on Arvanil, an Endocannabinoid and Vanilloid Hybrid. J. Pharmacol. Exp. Ther. 2002, 300, 984991,  DOI: 10.1124/jpet.300.3.984
  17. 17
    Chen, J.; Luan, Y.; Yu, R.; Zhang, Z.; Zhang, J.; Wang, W. Transient Receptor Potential (TRP) Channels, Promising Potential Diagnostic and Therapeutic Tools for Cancer. Biosci. Trends 2014, 8, 110,  DOI: 10.5582/bst.8.1
  18. 18
    Wang, G.; Juan, Y.; Jinsoo, P.; Xinglong, G.; Bei, W.; Hao, L.; Jane, Y. Facile Synthesis and Characterization of Graphene Nanosheets. J. Phys. Chem. C 2008, 112, 81928195,  DOI: 10.1021/jp710931h
  19. 19
    Robertson, A. W.; Warner, J. H. Hexagonal Single Crystal Domains of Few-Layer Graphene on Copper Foils. Nano Lett. 2011, 11, 11821189,  DOI: 10.1021/nl104142k
  20. 20
    Wang, G.; Zhang, M.; Zhu, Y.; Ding, G.; Jiang, D.; Guo, Q.; Liu, S.; Xie, X.; Chu, P. K.; Di, Z.; Wang, X. Direct Growth of Graphene Film on Germanium Substrate. Sci. Rep. 2013, 3, 2465,  DOI: 10.1038/srep02465
  21. 21
    Rosano, G. L.; Ceccarelli, E. A. Recombinant Protein Expression in Escherichia Coli: Advances and Challenges. Front. Microbiol. 2014, 5, 172,  DOI: 10.3389/fmicb.2014.00172
  22. 22
    Pingle, S. C.; Matta, J. A.; Ahern, G. P. Capsaicin Receptor: TRPV1 a Promiscuous TRP Channel; Flockerzi, V., Nilius, B., Eds.; Springer: Berlin, Heidelberg, 2007; Vol. 179.
  23. 23
    Hirsch, A.; Englert, J. M.; Hauke, F. Wet Chemical Functionalization of Graphene. Acc. Chem. Res. 2013, 46, 8796,  DOI: 10.1021/ar300116q
  24. 24
    Moon, J. S.; Antcliffe, M.; Seo, H. C.; Curtis, D.; Lin, S.; Schmitz, A.; Milosavljevic, I.; Kiselev, A. A.; Ross, R. S.; Gaskill, D. K.; Campbell, P. M.; Fitch, R. C.; Lee, K. M.; Asbeck, P. Ultra-Low Resistance Ohmic Contacts in Graphene Field Effect Transistors. Appl. Phys. Lett. 2012, 100, 203512,  DOI: 10.1063/1.4719579
  25. 25
    Kim, D. J.; Sohn, I. Y.; Jung, J. H.; Yoon, O. J.; Lee, N. E.; Park, J. S. Reduced Graphene Oxide Field-Effect Transistor for Label-Free Femtomolar Protein Detection. Biosens. Bioelectron. 2013, 41, 621626,  DOI: 10.1016/j.bios.2012.09.040
  26. 26
    Huhtamäki, T.; Tian, X.; Korhonen, J. T.; Ras, R. H. A. Surface-Wetting Characterization Using Contact-Angle Measurements. Nat. Protoc. 2018, 13, 15211538,  DOI: 10.1038/s41596-018-0003-z
  27. 27
    Marmur, A. Solid-Surface Characterization by Wetting. Annu. Rev. Mater. Res. 2009, 39, 473489,  DOI: 10.1146/annurev.matsci.38.060407.132425
  28. 28
    Wang, Q.; He, H.; Li, B.; Lin, H.; Zhang, Y.; Zhang, J.; Wang, Z. UV–Vis and ATR–FTIR Spectroscopic Investigations of Postmortem Interval Based on the Changes in Rabbit Plasma. PLoS One 2017, 12, e0182161  DOI: 10.1371/journal.pone.0182161
  29. 29
    Rygula, A.; Majzner, K.; Marzec, K. M.; Kaczor, A.; Pilarczyk, M.; Baranska, M. Raman Spectroscopy of Proteins: A Review. J. Raman Spectrosc. 2013, 44, 10611076,  DOI: 10.1002/jrs.4335
  30. 30
    Tuma, R. Raman Spectroscopy of Proteins: From Peptides to Large Assemblies. J. Raman Spectrosc. 2005, 36, 307319,  DOI: 10.1002/jrs.1323
  31. 31
    Sjöberg, B.; Foley, S.; Cardey, B.; Enescu, M. An Experimental and Theoretical Study of the Amino Acid Side Chain Raman Bands in Proteins. Spectrochim. Acta, Part A 2014, 128, 300311,  DOI: 10.1016/j.saa.2014.02.080
  32. 32
    Jara-Oseguera, A.; Simon, S. A.; Rosenbaum, T. TRPV1: ON THE ROAD TO PAIN RELIEF. Curr. Mol. Pharmacol. 2008, 1, 255269,  DOI: 10.2174/1874467210801030255
  33. 33
    Elokely, K.; Velisetty, P.; Delemotte, L.; Palovcak, E.; Klein, M. L.; Rohacs, T.; Carnevale, V. Understanding TRPV1 Activation by Ligands: Insights from the Binding Modes of Capsaicin and Resiniferatoxin. Proc. Natl. Acad. Sci. U.S.A. 2016, 113, E137E145,  DOI: 10.1073/pnas.1517288113
  34. 34
    Domene, C.; Darré, L.; Oakes, V.; Gonzalez-Resines, S. A Potential Route of Capsaicin to Its Binding Site in the TRPV1 Ion Channel. J. Chem. Inf. Model. 2022, 62, 24812489,  DOI: 10.1021/acs.jcim.1c01441
  35. 35
    Correll, C. C.; Phelps, P. T.; Anthes, J. C.; Umland, S.; Greenfeder, S. Cloning and Pharmacological Characterization of Mouse TRPV1. Neurosci. Lett. 2004, 370, 5560,  DOI: 10.1016/j.neulet.2004.07.058
  36. 36
    Smart, D.; Jerman, J. C.; Gunthorpe, M. J.; Brough, S. J.; Ranson, J.; Cairns, W.; Hayes, P. D.; Randall, A. D.; Davis, J. B. Characterisation Using FLIPR of Human Vanilloid VR1 Receptor Pharmacology. Eur. J. Pharmacol. 2001, 417, 51,  DOI: 10.1016/s0014-2999(01)00901-3
  37. 37
    Park, S. J.; Seo, S. E.; Kim, K. H.; Lee, S. H.; Kim, J.; Ha, S.; Song, H. S.; Lee, S. H.; Kwon, O. S. Real-time monitoring of geosmin based on an aptamer-conjugated graphene field-effect transistor. Biosens. Bioelectron. 2021, 174, 112804,  DOI: 10.1016/j.bios.2020.112804
  38. 38
    Kim, K. H.; Park, C. S.; Park, S. J.; Kim, J.; Seo, S. E.; An, J. E.; Ha, S.; Bae, J.; Phyo, S.; Lee, J.; Kim, K.; Moon, D.; Park, T. H.; Song, H. S.; Kwon, O. S. In-situ food spoilage monitoring using a wireless chemical receptor-conjugated graphene electronic nose. Biosens. Bioelectron. 2022, 200, 113908,  DOI: 10.1016/j.bios.2021.113908
  39. 39
    Lee, S. H.; Kim, K. H.; Seo, S. E.; Kim, M. I.; Park, S. J.; Kwon, O. S. Cytochrome C-decorated graphene field-effect transistor for highly sensitive hydrogen peroxide detection. J. Ind. Eng. Chem. 2020, 83, 2934,  DOI: 10.1016/j.jiec.2019.11.009

Cited By

Click to copy section linkSection link copied!

This article is cited by 6 publications.

  1. Kyung Ho Kim, Jisung Kwak, Sung Eun Seo, Siyoung Ha, Gyeong-Ji Kim, Sangwoo Lee, Sang Jun Sim, Yeon Kyung Lee, Na Ly Tran, Seung Ja Oh, Woo-Keun Kim, Hyun Seok Song, Oh Seok Kwon. TRPA1 nanovesicle-conjugated receptonics for rapid biocide screening. Journal of Hazardous Materials 2024, 477 , 135282. https://doi.org/10.1016/j.jhazmat.2024.135282
  2. Wen Li, Shuai Sun, Wanchao Chen, Haile Ma, Tingzhao Li, Zhong Zhang, Di Wu, Mengqiu Yan, Yan Yang. Exploring the taste presentation and receptor perception mechanism of salty peptides from Stropharia rugosoannulata based on molecular dynamics and thermodynamics simulation. Food Science and Human Wellness 2024, 13 (4) , 2277-2288. https://doi.org/10.26599/FSHW.2022.9250190
  3. Qi Li, Linghui Kong, Yu-Cai He, Cuiluan Ma. Efficient biosynthesis of vanillylamine via biological transamination of lignin-derived vanillin in a β-cyclodextrin-water medium. Sustainable Chemistry and Pharmacy 2024, 38 , 101437. https://doi.org/10.1016/j.scp.2024.101437
  4. Mujahid Sher, Ishtiaq Hussain, Farhat Ali Khan, Muhammad Zahoor, Wiaam Mujahid Sher, Muhammad Saqib Khalil, Muhammad Sulaiman, Riaz Ullah, Sumaira Naz, Essam A. Ali. Synthesis and characterization of capsaicin nanoparticles: An attempt to enhance its bioavailability and pharmacological actions. Green Processing and Synthesis 2024, 13 (1) https://doi.org/10.1515/gps-2023-0206
  5. Janayne L. Silva, Elandia A. Santos, Jacqueline I. Alvarez-Leite. Are We Ready to Recommend Capsaicin for Disorders Other Than Neuropathic Pain?. Nutrients 2023, 15 (20) , 4469. https://doi.org/10.3390/nu15204469
  6. Sung Eun Seo, Seong Gi Lim, Kyung Ho Kim, Jinyeong Kim, Chan Jae Shin, Soomin Kim, Lina Kim, Seung Hwan Lee, Song Yee Jang, Hyun Woo Oh, Hyang‐Ae Lee, Woo‐Keun Kim, Yoo Min Park, Kyoung G. Lee, Sang Hun Lee, Siyoung Ha, Oh Seok Kwon. Reusable Electronic Tongue Based on Transient Receptor Potential Vanilloid 1 Nanodisc‐Conjugated Graphene Field‐Effect Transistor for a Spiciness‐Related Pain Evaluation. Advanced Materials 2023, 35 (19) https://doi.org/10.1002/adma.202206198

ACS Omega

Cite this: ACS Omega 2022, 7, 41, 36441–36447
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.2c04202
Published October 4, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

1436

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Scheme 1

    Scheme 1. Schematic Illustration of the Overall Operation Process of the Fabricated Platform for the In Vitro Selection of Drug Candidates Using a TRPV1-Immobilized Graphene Field-Effect Transistor

    Figure 1

    Figure 1. Identification of graphene and the TRPV1 bioprobe (a) SAED pattern of graphene. (b) Output curve of graphene. (c) SDS–PAGE (right) and Western blot (left) of the fabricated TRPV1 protein with the molecular weight of 98 kDa.

    Figure 2

    Figure 2. Fabrication and identification of the electrode (a) schematic illustration of the fabrication process of the electrode, (b) color SEM image, (c) IV curves after each step of the surface modification, (d) transfer curve, (e) contact angle analysis, (f) FT-IR spectra, and (g) Raman spectra.

    Figure 3

    Figure 3. Simulation of ligand–protein docking between TRPV1 and arvanil (a,d,g), olvanil (b,e,h), and capsaicin (c,f,i).

    Figure 4

    Figure 4. Real-time monitoring and the corresponding calibrated response curves of arvanil (a,d), olvanil (b,e), and capsaicin (c,f) collected from the manufactured FET platform.

  • References


    This article references 39 other publications.

    1. 1
      Cao, E.; Liao, M.; Cheng, Y.; Julius, D. TRPV1 Structures in Distinct Conformations Reveal Activation Mechanisms. Nature 2013, 504, 113118,  DOI: 10.1038/nature12823
    2. 2
      Bevan, S.; Quallo, T.; Andersson, D. A. Mammalian Transient Receptor Potential (TRP) Cation Channels; Nilius, B., Flockerzi, V., Eds.; Springer New York LLC, 2014; Vol. 222.
    3. 3
      Szallasi, A.; Blumberg, P. M. Vanilloid (Capsaicin) Receptors and Mechanisms. Pharmacol. Rev. 1999, 51, 159212
    4. 4
      O’Neill, J.; Brock, C.; Olesen, A. E.; Andresen, T.; Nilsson, M.; Dickenson, A. H. Unravelling the Mystery of Capsaicin: A Tool to Understand and Treat Pain. Pharmacol. Rev. 2012, 64, 939971,  DOI: 10.1124/pr.112.006163
    5. 5
      Premkumar, L. S.; Sikand, P. TRPV1 A Target for Next Generation Analgesics. Curr. Neuropharmacol. 2008, 6, 151163,  DOI: 10.2174/157015908784533888
    6. 6
      Knotkova, H.; Pappagallo, M.; Szallasi, A. Capsaicin (TRPV1 Agonist) Therapy for Pain Relief. Clin. J. Pain 2008, 24, 142154,  DOI: 10.1097/ajp.0b013e318158ed9e
    7. 7
      Mason, L.; Moore, R. A.; Derry, S.; Edwards, J. E.; McQuay, H. J. Systematic Review of Topical Capsaicin for the Treatment of Chronic Pain. Br. Med. J. 2004, 328, 991994,  DOI: 10.1136/bmj.38042.506748.ee
    8. 8
      Blair, H. A. Capsaicin 8% Dermal Patch: A Review in Peripheral Neuropathic Pain. Drugs 2018, 78, 14891500,  DOI: 10.1007/s40265-018-0982-7
    9. 9
      Peppin, J. F.; Pappagallo, M. Capsaicinoids in the Treatment of Neuropathic Pain: A Review. Ther. Adv. Neurol. Disord. 2014, 7, 2232,  DOI: 10.1177/1756285613501576
    10. 10
      Drewes, A. M.; Schipper, K. P.; Dimcevski, G.; Petersen, P.; Gregersen, H.; Funch-Jensen, P.; Arendt-Nielsen, L. Gut Pain and Hyperalgesia Induced by Capsaicin: A Human Experimental Model. Pain 2003, 104, 333341,  DOI: 10.1016/s0304-3959(03)00039-3
    11. 11
      LaMotte, R. .; Lundberg, L. E. R.; Torebjörk, H. E. Pain, Hyperalgesia and Activity in Nociceptive C Units in Humans after Intradermal Injection of Capsaicin. J. Physiol. 1992, 448, 749764,  DOI: 10.1113/jphysiol.1992.sp019068
    12. 12
      Karai, L.; Brown, D. C.; Mannes, A. J.; Connelly, S. T.; Brown, J.; Gandal, M.; Wellisch, O. M.; Neubert, J. K.; Olah, Z.; Iadarola, M. J. Deletion of Vanilloid Receptor 1-Expressing Primary Afferent Neurons for Pain Control. J. Clin. Invest. 2004, 113, 13441352,  DOI: 10.1172/jci20449
    13. 13
      Yang, F.; Zheng, J. Understand Spiciness: Mechanism of TRPV1 Channel Activation by Capsaicin. Protein Cell 2017, 8, 169177,  DOI: 10.1007/s13238-016-0353-7
    14. 14
      Appendino, G.; De Petrocellis, L.; Trevisani, M.; Minassi, A.; Daddario, N.; Moriello, A. S.; Gazzieri, D.; Ligresti, A.; Campi, B.; Fontana, G.; Pinna, C.; Geppetti, P.; Di Marzo, V. Development of the First Ultra-Potent “Capsaicinoid” Agonist at Transient Receptor Potential Vanilloid Type 1 (TRPV1) Channels and Its Therapeutic Potential. J. Pharmacol. Exp. Ther. 2005, 312, 561570,  DOI: 10.1124/jpet.104.074864
    15. 15
      Hurley, J. D.; Akers, A. T.; Friedman, J. R.; Nolan, N. A.; Brown, K. C.; Dasgupta, P. Non-Pungent Long Chain Capsaicin-Analogs Arvanil and Olvanil Display Better Anti-Invasive Activity than Capsaicin in Human Small Cell Lung Cancers. Cell Adhes. Migrat. 2017, 11, 8097,  DOI: 10.1080/19336918.2016.1187368
    16. 16
      Di Marzo, V.; Griffin, G.; De Petrocellis, L.; Brandi, I.; Bisogno, T.; Williams, W.; Grier, M. C.; Kulasegram, S.; Mahadevan, A. N. U.; Razdan, R. K.; Martin, B. R. A Structure/Activity Relationship Study on Arvanil, an Endocannabinoid and Vanilloid Hybrid. J. Pharmacol. Exp. Ther. 2002, 300, 984991,  DOI: 10.1124/jpet.300.3.984
    17. 17
      Chen, J.; Luan, Y.; Yu, R.; Zhang, Z.; Zhang, J.; Wang, W. Transient Receptor Potential (TRP) Channels, Promising Potential Diagnostic and Therapeutic Tools for Cancer. Biosci. Trends 2014, 8, 110,  DOI: 10.5582/bst.8.1
    18. 18
      Wang, G.; Juan, Y.; Jinsoo, P.; Xinglong, G.; Bei, W.; Hao, L.; Jane, Y. Facile Synthesis and Characterization of Graphene Nanosheets. J. Phys. Chem. C 2008, 112, 81928195,  DOI: 10.1021/jp710931h
    19. 19
      Robertson, A. W.; Warner, J. H. Hexagonal Single Crystal Domains of Few-Layer Graphene on Copper Foils. Nano Lett. 2011, 11, 11821189,  DOI: 10.1021/nl104142k
    20. 20
      Wang, G.; Zhang, M.; Zhu, Y.; Ding, G.; Jiang, D.; Guo, Q.; Liu, S.; Xie, X.; Chu, P. K.; Di, Z.; Wang, X. Direct Growth of Graphene Film on Germanium Substrate. Sci. Rep. 2013, 3, 2465,  DOI: 10.1038/srep02465
    21. 21
      Rosano, G. L.; Ceccarelli, E. A. Recombinant Protein Expression in Escherichia Coli: Advances and Challenges. Front. Microbiol. 2014, 5, 172,  DOI: 10.3389/fmicb.2014.00172
    22. 22
      Pingle, S. C.; Matta, J. A.; Ahern, G. P. Capsaicin Receptor: TRPV1 a Promiscuous TRP Channel; Flockerzi, V., Nilius, B., Eds.; Springer: Berlin, Heidelberg, 2007; Vol. 179.
    23. 23
      Hirsch, A.; Englert, J. M.; Hauke, F. Wet Chemical Functionalization of Graphene. Acc. Chem. Res. 2013, 46, 8796,  DOI: 10.1021/ar300116q
    24. 24
      Moon, J. S.; Antcliffe, M.; Seo, H. C.; Curtis, D.; Lin, S.; Schmitz, A.; Milosavljevic, I.; Kiselev, A. A.; Ross, R. S.; Gaskill, D. K.; Campbell, P. M.; Fitch, R. C.; Lee, K. M.; Asbeck, P. Ultra-Low Resistance Ohmic Contacts in Graphene Field Effect Transistors. Appl. Phys. Lett. 2012, 100, 203512,  DOI: 10.1063/1.4719579
    25. 25
      Kim, D. J.; Sohn, I. Y.; Jung, J. H.; Yoon, O. J.; Lee, N. E.; Park, J. S. Reduced Graphene Oxide Field-Effect Transistor for Label-Free Femtomolar Protein Detection. Biosens. Bioelectron. 2013, 41, 621626,  DOI: 10.1016/j.bios.2012.09.040
    26. 26
      Huhtamäki, T.; Tian, X.; Korhonen, J. T.; Ras, R. H. A. Surface-Wetting Characterization Using Contact-Angle Measurements. Nat. Protoc. 2018, 13, 15211538,  DOI: 10.1038/s41596-018-0003-z
    27. 27
      Marmur, A. Solid-Surface Characterization by Wetting. Annu. Rev. Mater. Res. 2009, 39, 473489,  DOI: 10.1146/annurev.matsci.38.060407.132425
    28. 28
      Wang, Q.; He, H.; Li, B.; Lin, H.; Zhang, Y.; Zhang, J.; Wang, Z. UV–Vis and ATR–FTIR Spectroscopic Investigations of Postmortem Interval Based on the Changes in Rabbit Plasma. PLoS One 2017, 12, e0182161  DOI: 10.1371/journal.pone.0182161
    29. 29
      Rygula, A.; Majzner, K.; Marzec, K. M.; Kaczor, A.; Pilarczyk, M.; Baranska, M. Raman Spectroscopy of Proteins: A Review. J. Raman Spectrosc. 2013, 44, 10611076,  DOI: 10.1002/jrs.4335
    30. 30
      Tuma, R. Raman Spectroscopy of Proteins: From Peptides to Large Assemblies. J. Raman Spectrosc. 2005, 36, 307319,  DOI: 10.1002/jrs.1323
    31. 31
      Sjöberg, B.; Foley, S.; Cardey, B.; Enescu, M. An Experimental and Theoretical Study of the Amino Acid Side Chain Raman Bands in Proteins. Spectrochim. Acta, Part A 2014, 128, 300311,  DOI: 10.1016/j.saa.2014.02.080
    32. 32
      Jara-Oseguera, A.; Simon, S. A.; Rosenbaum, T. TRPV1: ON THE ROAD TO PAIN RELIEF. Curr. Mol. Pharmacol. 2008, 1, 255269,  DOI: 10.2174/1874467210801030255
    33. 33
      Elokely, K.; Velisetty, P.; Delemotte, L.; Palovcak, E.; Klein, M. L.; Rohacs, T.; Carnevale, V. Understanding TRPV1 Activation by Ligands: Insights from the Binding Modes of Capsaicin and Resiniferatoxin. Proc. Natl. Acad. Sci. U.S.A. 2016, 113, E137E145,  DOI: 10.1073/pnas.1517288113
    34. 34
      Domene, C.; Darré, L.; Oakes, V.; Gonzalez-Resines, S. A Potential Route of Capsaicin to Its Binding Site in the TRPV1 Ion Channel. J. Chem. Inf. Model. 2022, 62, 24812489,  DOI: 10.1021/acs.jcim.1c01441
    35. 35
      Correll, C. C.; Phelps, P. T.; Anthes, J. C.; Umland, S.; Greenfeder, S. Cloning and Pharmacological Characterization of Mouse TRPV1. Neurosci. Lett. 2004, 370, 5560,  DOI: 10.1016/j.neulet.2004.07.058
    36. 36
      Smart, D.; Jerman, J. C.; Gunthorpe, M. J.; Brough, S. J.; Ranson, J.; Cairns, W.; Hayes, P. D.; Randall, A. D.; Davis, J. B. Characterisation Using FLIPR of Human Vanilloid VR1 Receptor Pharmacology. Eur. J. Pharmacol. 2001, 417, 51,  DOI: 10.1016/s0014-2999(01)00901-3
    37. 37
      Park, S. J.; Seo, S. E.; Kim, K. H.; Lee, S. H.; Kim, J.; Ha, S.; Song, H. S.; Lee, S. H.; Kwon, O. S. Real-time monitoring of geosmin based on an aptamer-conjugated graphene field-effect transistor. Biosens. Bioelectron. 2021, 174, 112804,  DOI: 10.1016/j.bios.2020.112804
    38. 38
      Kim, K. H.; Park, C. S.; Park, S. J.; Kim, J.; Seo, S. E.; An, J. E.; Ha, S.; Bae, J.; Phyo, S.; Lee, J.; Kim, K.; Moon, D.; Park, T. H.; Song, H. S.; Kwon, O. S. In-situ food spoilage monitoring using a wireless chemical receptor-conjugated graphene electronic nose. Biosens. Bioelectron. 2022, 200, 113908,  DOI: 10.1016/j.bios.2021.113908
    39. 39
      Lee, S. H.; Kim, K. H.; Seo, S. E.; Kim, M. I.; Park, S. J.; Kwon, O. S. Cytochrome C-decorated graphene field-effect transistor for highly sensitive hydrogen peroxide detection. J. Ind. Eng. Chem. 2020, 83, 2934,  DOI: 10.1016/j.jiec.2019.11.009
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.2c04202.

    • Temperature program for graphene synthesis; MEMS process for the electrode fabrication; and equations for response normalization and calibration (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.