Next Article in Journal
Exploring the Potential Applications of Engineered Borophene in Nanobiosensing and Theranostics
Next Article in Special Issue
Computation of Vascular Parameters: Implementing Methodology and Performance Analysis
Previous Article in Journal
Biosensors for Monitoring of Biologically Relevant Molecules
Previous Article in Special Issue
Wearable Insulin Biosensors for Diabetes Management: Advances and Challenges
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Organic Electrochemical Transistor Immuno-Sensors for Spike Protein Early Detection

by
Mario Barra
1,*,
Giovanna Tomaiuolo
2,3,
Valeria Rachela Villella
2,3,
Speranza Esposito
2,3,
Aris Liboà
4,5,
Pasquale D’Angelo
4,
Simone Luigi Marasso
4,6,
Matteo Cocuzza
4,6,
Valentina Bertana
6,
Elena Camilli
6 and
Valentina Preziosi
2,3,*
1
CNR-SPIN, c/o Department of Physics ‘‘Ettore Pancini’’, P.le Tecchio, 80, 80125 Napoli, Italy
2
Department of Chemical, Materials and Production Engineering—University Federico II, P.le Tecchio 80, 80125 Napoli, Italy
3
CEINGE, Advanced Biotechnologies, 80145 Napoli, Italy
4
IMEM-CNR, Parco Area delle Scienze 37/A, 43124 Parma, Italy
5
Graduate School in Science and Technologies of Materials and Department of Physics, University of Parma, Parco Area delle Scienze, 7/A, 43121 Parma, Italy
6
ChiLab, Department of Applied Science and Technology, Politecnico di Torino, 10129 Torino, Italy
*
Authors to whom correspondence should be addressed.
Biosensors 2023, 13(7), 739; https://doi.org/10.3390/bios13070739
Submission received: 15 June 2023 / Revised: 3 July 2023 / Accepted: 5 July 2023 / Published: 17 July 2023
(This article belongs to the Special Issue Advances in Biosensors for Health-Care and Diagnostics)

Abstract

:
The global COVID-19 pandemic has had severe consequences from the social and economic perspectives, compelling the scientific community to focus on the development of effective diagnostics that can combine a fast response and accurate sensitivity/specificity performance. Presently available commercial antigen-detecting rapid diagnostic tests (Ag-RDTs) are very fast, but still face significant criticisms, mainly related to their inability to amplify the protein signal. This translates to a limited sensitive outcome and, hence, a reduced ability to hamper the spread of SARS-CoV-2 infection. To answer the urgent need for novel platforms for the early, specific and highly sensitive detection of the virus, this paper deals with the use of organic electrochemical transistors (OECTs) as very efficient ion–electron converters and amplifiers for the detection of spike proteins and their femtomolar concentration. The electrical response of the investigated OECTs was carefully analyzed, and the changes in the parameters associated with the transconductance (i.e., the slope of the transfer curves) in the gate voltage range between 0 and 0.3 V were found to be more clearly correlated with the spike protein concentration. Moreover, the functionalization of OECT-based biosensors with anti-spike and anti-nucleocapside proteins, the major proteins involved in the disease, demonstrated the specificity of these devices, whose potentialities should also be considered in light of the recent upsurge of the so-called “long COVID” syndrome.

1. Introduction

The uncontrolled and devastating spread of coronavirus (COVID-19 disease) that struck the world in the last two years has led to the need to find new, rapid and effective solutions to manage and reduce its diffusion. The optimization of oro-nasopharyngeal tests for the early detection of proteins involved in SARS-CoV-2 infection has been a focus of the global scientific community [1,2]. One of the main target proteins used to identify the presence of SARS-CoV-2 is represented by the spike protein, which, located at the virus surface, plays a fundamental role in penetrating host cells and initiating infection. At this stage, however, it has been recognized that the commercially available infection detection methods still exhibit limited precision, leading to false-positive results and the out-of-control diffusion of the virus [3,4,5,6]. With the progressive evolution of the COVID-19 outbreak, moreover, it has been also observed that, even after the resolution of the main symptomatology, several patients (about 10% of the initially infected population) continue to suffer from relevant symptoms, which can affect multiple organs with different degrees of impact [7]. This syndrome has been generally termed “long COVID” and is the subject of an intense global investigation at present. In this context, it was recently demonstrated that, since SARS-CoV-2 viral reservoirs may endure in the body for several months, the spike protein is among the main biomarkers that could be monitored to analyze the evolution of “long COVID” [8]. These findings further boost interest in the development of reliable biosensing platforms with the ability to combine high specificity, an excellent sensitivity and a very fast response for the detection of spike proteins.
One of the most promising approaches to address these urgent issues is given by the emerging field of organic bioelectronics, featuring outstanding performances in terms of ionic-to-electronic signal transduction and amplification. Organic electrochemical transistors (OECTs) [9,10,11] have been widely investigated as biosensors in many applications, from the detection of biomarkers [12,13,14,15,16] to neural activity [17,18,19], as well as in more general biomedical research [20,21,22,23]. Such devices exhibit a quite simple architecture, made of three electrodes (source, drain and gate) and a conductive polymer (PEDOT:PSS in our case) active layer. The polymer channel connects the source and drain contacts and is immersed in an electrolyte where the gate electrode is also positioned. The application of a voltage VGS to the gate leads to the injection of ions from the electrolyte solution into the polymeric channel, thus reversibly modifying the electronic current (IDS) flowing between the drain and source. A PEDOT:PSS channel changes its electronic conductivity in response to the penetration of cations; hence, the related OECTs work in depletion mode following the application of positive VG voltages produces a consistent decrease in the IDS current. Significantly, such devices can operate at voltages lower than 1 V, making it possible for them to interact with biological fluids in the absence of relevant hydrolysis or molecule denaturation effects. However, specific care should be taken to preserve the control of the OECT response within a complex environment, including ions, nutrients, proteins, etc. In this regard, the authors have recently investigated the OECT response in human blood, outlining the effect of the corpuscular component (i.e., red blood cells) with respect to that of the liquid matrix (i.e., plasma) [24]. For biosensing applications, the OECT’s functionalization towards a specific species is one of the key points guaranteeing a proper selective response. In the last two years, various studies have focused on the early detection of antibodies or antigens involved in COVID-19, mostly regarding the spike protein [25,26]. Some recent papers analyzed the role of another of the four structural proteins involved in the infection mechanism, the nucleocapside (NP). Such a protein plays a key role in viral assembly, replication, and the host immune response regulation, making it a subject of great interest for the scientific community [27,28].
In this work, we investigate the response of PEDOT:PSS-based OECTs, driven by antibody-functionalized gate electrodes for the detection of the SARS-CoV-2 spike protein at very low concentrations, comprising femto and nanomolar values. Our study focused on the identification of the specific features of the OECT response, as this is more suitable when establishing a direct correlation with the specific concentration of the SARS-CoV-2 spike protein. Moreover, we also demonstrate that the adopted detection approach preserves its specificity behavior when analyzed as compared to the SARS-CoV-2 NP (nucleocapside) antibody.

2. Materials and Methods

2.1. Materials and Sample Preparation

Acetone, isopropanol, 3-mercaptopropionic acid (3-MPA), 11-mercaptoundecanoic acid (11-MUA), 1-ethyl-3-(3 dimethylaminopropyl) carbodiimide (EDC), N-hydroxysulfosuccinimide sodium salt (sulfo-NHS), ethanolamine, Tween 20 and phosphate-buffered saline (PBS) solution (C = 10 mM) were all purchased from Sigma–Aldrich (Missouri, MO, USA). Bovine serum albumin (BSA) was purchased by Microgem (Pozzuoli, Italy).
Primary antibodies anti-SARS-CoV-2-spike-RBD region antibody, produced in rabbit (anti-spike protein), SARS-CoV-2 NP (nucleocapside) antibody, produced in mouse and rabbit (anti-NP protein), and S protein-antigen (spike-RBD protein), i.e., SARS-CoV-2 spike Receptor Binding Domain, were purchased by Merck (Rome, Italy). Secondary antibody donkey anti-rabbit-Alexa Fluor 488 was purchased by Life-technologies (California, CA, USA).

2.2. OECT Fabrication

OECTs were fabricated using a well-established protocol already described in detail elsewhere [24,29,30,31], which is briefly reviewed here. A solution of PEDOT:PSS layer (Clevios PH 1000, doped with 5 volume % ethylene glycol, 0.1 volume % dodecyl benzene sulfonic acid, and 1 wt% of GOPS (3-glycidyloxypropyl)trimethoxysilane) was spin-coated on the source and drain electrodes and subsequently patterned via photolithography (microchemicals AZ9260 resist) and etching with O2 plasma. A polydimethylsiloxane (PDMS) chamber, with an internal volume of 150 μL, was aligned with the PEDOT:PSS channel and irreversibly bonded on the surface (Figure 1a,b).
Electrodes were fabricated by e-beam evaporation (ULVAC EBX-14D) through the deposition of Ti/Au films (thickness 10 nm/100 nm) on Si wafer (100) finished with 1 μm of thermal oxide. The final device aspect ratio is W/L = 6 mm/200 μm = 30 (where W is the width and L is the length of the channel). The planar gate electrodes were made by a Ti/Au—10 nm/100 nm film. The latter was deposited by the evaporation systems already used for the source and drain electrodes, on a Si wafer (100) finished with 1 μm of thermal oxide (SiO2). Finally, they were diced in 5 × 5 or 5 × 10 mm2 and functionalized using the protocol described in the following.

2.2.1. Gate Electrode Biofunctionalization

Gate surface was firstly cleaned in an ultrasonic bath with acetone and isopropanol for 10 min each, then the cleaned electrode was immersed in a 10 mM 3-MPA and 11-MUA (10:1 molar ratio) ethanol solution for 24 h at room temperature. This step allowed for the creation of the chemical SAM layer, exposing carboxylic groups. Then, such groups were activated by immersing the electrode in a water solution of EDC (200 mM) and NHS (50 mM) for 2 h at 25 °C. After each step, the surface was rinsed with distilled water to remove any residues and dried with air flux.
To bond the antibodies on the surface, the gate electrode was immersed in anti-spike protein solution (C = 10 nM in PBS) for 2 h at 25 °C, and then washed with 0.1% Tween PBS solution. After binding with antibodies, the gate electrode was treated with ethanolamine (1 M in PBS) and with BSA (1 μM in PBS) to saturate the remaining nonspecific sites (Figure 1c). To investigate OECT selectivity, gate electrodes were also functionalized with anti-NP protein solution at the same concentration (C = 10 nM in PBS).
The quality of the biofunctionalization protocol was tested (Section 3.1 with Figure 1 andFigure S1) by immersing gate electrodes in a solution with secondary antibodies conjugated to Alexa Fluor® 488 dye (donkey anti-rabbit—Alexa Fluor 488 at dilution of 1:200, purchased by Thermofisher, Waltham, MA, USA) for 1 h at room temperature. Then, the samples were observed with confocal laser scanning microscopy (CLSM—Zeiss LSM5 Pascal) technique to assess the binding between the primary and secondary antibodies. Moreover, gate electrodes functionalized with anti-spike antibodies were analyzed by Fourier Transform Infrared Spectroscopy (FTIR) techniques. The latter was performed using Agilent Cary 630 FTIR (Agilent Technologies, Inc., Santa Clara, CA, USA), working with an ATR module in the range between 4000 and 650 nm. Before any FTIR spectrum acquisition, the samples were carefully dried on a very thin N2 flow for 5′ to reduce the moisture. During the analysis, more than 10 spectra were collected to find the best working conditions, i.e., environmental parameters, number of scans, resolution, overall instrument working energy and to perform several comparisons with different backgrounds.

2.2.2. OECT Characterization

A Keithley two-channel multimeter (Keithley 2602B) operating by Labview code was used to characterize the electrical response of the OECT devices. Output curves were obtained by applying a voltage between gate and source, VGS, ranging between −0.6 and 0.6 V with a step of 0.1 V, and measuring drain-source (IDS) and gate-source (IGS) current, as a function of drain-source voltage, VDS (in the range between 0 and −0.6 V). Then, curves were transferred, reporting IDS by varying VGS between −0.6 and 0.6 V with a step of 0.025 V at VDS = −0.3 V, were recorded for all the investigated functionalized gate electrodes (i.e., with anti-spike or anti-NP proteins). In the transfer curves, each point was acquired every 2 s (Δt = 2 s). For comparison, curves at different times (Δt = 0.5 and 5 s) are reported in the Supplementary Information (Figure S3).
This protocol eas applied by using the gate electrodes functionalized with antibodies to drive the OECT, while PBS 10 mM was employed as electrolyte. Then, the functionalized gates were incubated with a spike-RBD protein from femto to nanomolar concentrations for about 20 min at room temperature and, after careful rinsing with bi-distilled water, the same electrodes were used to repeat the OECT measurement again in PBS 10 mM.
Before starting any biosensing experiment, the OECT channels were immersed in bi-distilled water for at least 2 h. Moreover, a set of preliminary transfer curves were recorded to stabilize the OECT response in PBS 10 mM. A data analysis of transfer curves could describe the OECT electrical response in terms of the current modulation values expressed as (IDS − I0)/I0, where I0 is the baseline current, and transconductance gm = δIDS/δVGS, which represents the transduction efficiency that is related to the channel current slope as a function of the voltage VGS. For all the experiments with bio-functionalized gate electrodes, particular care was paid to keep the immersed gating area fixed in place (Av = 8 mm2).
The mean and standard deviation values for all transfer curves were achieved by acquiring five consecutive transfer curves and discarding the first one (generally affected by a major level of variability) to extract the various statistical parameters.

3. Results

3.1. Validation of the Gate Electrode Biofunctionalization Process

A validation step of the biofunctionalization process was initially carried out through immunofluorescence to demonstrate the presence of antibodies on the gate surface. In particular, gate electrodes previously incubated with primary antibodies were immersed in solutions with secondary antibodies conjugated to Alexa Fluor® 488 dye. Since such molecules have a specific affinity to the primary antibodies, they can unequivocally demonstrate their presence on the gate surface (Figure 1d,e). Before the testing step, the electrodes were rinsed with PBS to remove any impurities and dried with cleaned air. Hence, confocal laser scanning microscopy (CLSM) images recorded before (Figure 1d) and after (Figure 1e) the functionalization step confirmed the effectiveness of the adopted protocol. A further analysis of the gate functionalization was conducted by the Fourier Transform Infrared Spectroscopy (FTIR) technique. In particular, FTIR spectra were obtained from surfaces before and after the antibody functionalization process. As shown in Figure S1, Sample A was taken from a bare cleaned Au layer; meanwhile, Sample B was recorded for a gold surface once the functionalization process with an anti-spike protein was completed. The spectrum obtained on Sample A is fully consistent with a bare gold surface, as no relevant peak was observed in the considered range. Conversely, Sample B exhibited many different peaks all across the spectrum, as the antibody itself is a protein rich in functional groups that are visible in the IR range. In this context, FTIR spectroscopy can provide important information regarding proteins’ secondary structure [32]. According to the literature, increasing the size of the molecules induces a minor sensitivity, but important information about the presence of the antibody and the success of the process can still be achieved [33]. In Figure S1, two main peaks in the secondary structure are present. The former, namely, Amide A, is located at 3252 nm and indicates the N-H stretching in resonance; meanwhile, the latter, namely, Amide I, is found at 1654 nm and shows the C=O stretching vibration. These peaks provide strong evidence of the correct bonding of the antibody on the underneath layer, representing significant markers of the presence of proteins [33,34].

3.2. OECT Initial Characterization

Before the biosensing experiments, the OECTs investigated in this work were carefully characterized to set the optimal operating conditions. Figure 2 reports a general view of a typical response achieved with an un-functionalized (bare) gold gate electrode. As shown from the output curves in Figure 2a (with VGS ranging between −0.6 V and 0.6 V, while VDS was varied between 0 and −0.6 V), the devices behave correctly as depletion-mode transistors with the IDS current (always in the range of mA) decreasing (in absolute value) following the application of positive VGS. In the output curves, only the linear and the triode regions can be observed, and this feature is to be ascribed to the gating condition that was adopted. Indeed, because of the size of the gate-immersed area (~8 mm2, see Materials and Methods section) and considering the volumetric capacitance (CV ~ 40·F⋅cm−3), which can be associated with the entire electrolyte/PEDOT:PSS distributed interface, the response of the investigated OECT is mainly dictated by the electric double-layer (EDL) capacitance (CG) at the gate–electrolyte interface (i.e., CG ≪ CV) [35,36]. In the framework of the model of Bernards and Malliaras [37], this condition provides a very large value (higher than 1 V) of the so-called pinch-off (Vp) voltage, which determines the achievement of the current saturation phenomenon when VGS = 0 V and VDS < −|Vp|.
Figure 2b shows a set of transfer curves recorded at different VDS, while, in Figure 2c, the same curves are represented as normalized with respect to the IDS value measured at VGS = −0.6 V. Finally, Figure 2d reports the corresponding transconductance (gm) values (gm = δIDS/δVGS).
These plots clearly show that by increasing the absolute value of the applied VDS, both the overall IDS modulation and gm tend to rise. In particular, the transconductance receives values in the range of mS, confirming the excellent capabilities of OECT while operating as amplifying elements [17]. It is important to outline that, at a larger |VDS|, the gm curves exhibit the characteristic presence of a broad peak, for which the position and maximum value can be modified as a function of VDS. For all the transfer curves, moreover, a consistent rise in gm when VGS approaches 0.6 V can be observed, with this behavior becoming more and more pronounced at a larger |VDS|. As shown in Figure S2 of the Supplementary File, this trend is accompanied by larger values of the gate current (IGATE). Provided this analysis, for the biosensing experiments discussed in the following section, all the transfer curves were recorded by keeping a fixed VDS = −0.3 V. This choice was motivated by the search for a good trade-off between sufficiently large values of gm and IDS modulation with the presence of an extended VGS region (from negative values and until 0.3/0.4 V), where gm assumes a mild and increasing dependence on VGS. As discussed in the next section, this last feature can be used to simplify the analysis of the OECT performances in terms of the detection abilities of spike-RBD protein. Significantly, under the application of VDS = −0.3 V, IGATE values remain constantly lower than 200 nA (Figure S2b), being four orders of magnitude smaller than the corresponding IDS current throughout the analyzed VGS range. Since the specific OECT response is determined by the synergistic combination of an electronic and an ionic circuit featuring typically different time dynamics as a function of the gating conditions and the active channel size [35,37], the effect of different sweep times in the transfer curve recording was also analyzed. Basically, various values of the time delay (Δt = 0.5, 2 and 5 s) between the VGS application (step ΔVGS = 0.02 V) and the measurement of the corresponding IDS value were taken, with the goal of assessing the related impact on the transfer curves and the corresponding gm values. According to the results summarized in Figure S3, if the measurements are performed too fast (namely, with Δt = 0.5 s), the devices are unable to reach a steady condition. Hence, the eventual modulation and gm values (Figure 3b) are consistently reduced. In consideration of these findings, for all the measurements discussed hereafter, ∆t = 2 s was selected as being able to provide a good trade-off between the speed and accuracy of the experiments. In the absence of Faradaic reactions at the gold gating surface, the gate current (IGATE) plot in Figure S3c confirms the fundamental capacitive nature (i.e., displacement current) of IGATE which, indeed, is strongly dependent on the overall sweep time (i.e., the faster the measurement, the larger the IGATE ∝ δVGS/δt).

3.3. OECT Detection of Spike-RBD Proteins

For all the experiments devoted to the detection of SARS-CoV-2 spike RBD protein, before any incubation step, the OECTs were preliminarily characterized in PBS. The so-achieved transfer curves (as average of four curves consecutively measured with VDS = −0.3 V; see the Materials and Methods section) were then compared with analogous data acquired after the incubation (for 20 min) in solutions containing variable concentrations of the spike-RBD protein. According to this protocol, and similar to other experimental studies [25], the gate electrode represents the sensing and disposable components of the overall device, while the PEDOT:PSS channel can be used as an ion-to-electron amplifying transducer in combination with different gating surfaces.
Following the aforementioned procedure, Figure 3a–c shows the results of three experiments at 57 fM, 57 pM and 57 nM spike-RBD protein concentrations (i.e., 1.4 pg/mL, 1.4 ng/mL, 1.4 µg/mL, respectively). Firstly, it should be mentioned that a shift in the current baseline (i.e., the almost constant IDS current measured at very negative VGS values) was observed in most cases, but without any clear correlation with the incubation step and the corresponding spike-RBD protein concentration. This effect is likely related to underlying ion diffusion processes, which can slightly modify the overall conductivity of the PEDOT:PSS channel over time during the different measurements.
More interestingly, as suggested by the acquired transfer curves and their normalized counterparts, shown in Figure 3 (top and bottom panels, respectively), the incubation process tends to also modify the general shape of the transfer curves. The most evident feature is the reduction in the IDS slope, which is mainly observed in the intermediate VGS range between slightly negative values and 0.3/0.4 V. For very negative VGS values, the IDS slope remains rather low, and is quite unchanged following the incubation procedure. Conversely, as previously discussed (see Section 3.2), when VGS exceeds 0.3 V, the IDS growth with VGS becomes very steep and, in this case, turns out to be less reproducibly affected by the antibody–antigen binding occurrence. A further and clearer evidence of the IDS slope decrease in the intermediate VGS region, related to the incubation procedure, is given in Figure S4a–c (see the Supplementary File) directly showing the corresponding transconductance (gm) reduction.
For the sake of completeness, Figure 4 summarizes the results of an experiment where a gate functionalized with anti-SARS-CoV-2 NP (anti-NP) antibody was incubated in the spike-RBD protein solution with nanomolar (57 nM) concentration. At odds with the observations commented for Figure 3, here, the incubation step produced less characteristic changes and the related transconductance values were found to be slightly increased (Figure S4d) in the intermediate VGS region following the incubation process. As a whole, these observations confirm that the OECT behaves coherently following the non-specific interaction of the anti-NP protein and the spike-RBD protein.
Figure 5 offers a summary of all the previously performed experiments, demonstrating the ability to identify the different concentrations of the spike-RBD protein by driving the OECTS using gate electrodes functionalized with the proper antibody. Figure 5a reports the IDS modulation values given by the absolute value of (IF − I0/I0), where IF and I0 are the IDS currents in the transfer curves recorded at VGS = 0.6 V and VGS = −0.5 V, respectively. Here, the mean values from different experiments are indicated for any spike-RBD protein concentration, with the standard deviations being the error bars. The bottom panel of Figure 5a represents the percent variation in any IDS modulation value with respect to that measured prior to the corresponding incubation (BLANK). As shown, this analysis reveals the progressive reduction in the IDS modulation at an increasing spike-RBD protein concentration, with the percent decrease approaching 30% for the concentrations in the nanomolar range. It is worth highlighting that the IDS modulation value is largely used in the literature as a main parameter in biosensing experiments focused on the use of OECT [25,38] and that, similarly to what is observed here, percent variations ranging between a few units and tens per cent are typically reported for analyses dealing with antigen concentrations that are variable over several orders of magnitude [21,29]. In general, the IDS modulation reduction can be explained by considering that the incubation process provides a further decrease in the CG (i.e., the capacitance between gate and electrolyte) values, consequently attenuating the gate electrode’s overall ability to modulate the IDS current.
Following the aforementioned discussion about the incubation-induced changes in the slope of the transfer curves, Figure 5b reports the transconductance values estimated at VGS = +0.2 V (top panel), VGS = 0 V (middle panel) and VGS = −0.2 V (bottom panel) as a function of the spike-RBD protein concentration. Here, the decreasing trend is only clear for the gm values at VGS = +0.2 V with a variation of up to 30%, as achieved for the nanomolar concentration, in comparison to the initial value. This plot highlights the major sensitivity of the OECT response versus the antigen concentration in the intermediate VGS region, in comparison with what is observed for lower values of VGS (i.e., 0 and −0.2 V).
Moreover, in the graphs of Figure 5a,b (top panel), the IDS modulation and transconductance at VGS = +0.2 V values obtained for the experiments carried out with anti-NP protein functionalized gate electrode before and after incubation in a solution with 57 nM of spike-RBD protein are also included. In this case, because of the lack of a antigen–antibody complex, the IDS modulation and gm values before and after the incubation are completely indistinguishable.
With the aim of setting up an alternative and potentially more direct protocol to demonstrate the detection of the spike-RBD protein from the analysis of the OECT response, we elaborated the data through a different procedure focused on the modelling of the transfer curves region between −0.5 and 0.3 V. Hence, all the transfer curves in this VGS range were modelled by a simple second-order polynomial (IDS = α + β · V + γ · V 2 ) . Figure S5 in the Supplementary File presents examples of this fitting procedure applied to the transfer curves initially introduced in Figure 3 and Figure 4a. The excellent quality of the fitting curves demonstrates the properness of the approach, which can model both the initial linear and the following quadratic dependence of IDS on VGS.
Hence, the values of the β and γ fitting parameters averaged for all the experiments performed at different spike-RBD protein concentrations are plotted in Figure 6a,b, respectively. In particular, while resembling the behavior of gm at VGS = 0 and −0.2 V, even if the β fitting parameter (Figure 6a) is significantly reduced in the lowest concentration (i.e., fM) range in comparison with the BLANK test (i.e., prior to any incubation step [36]), it does not exhibit a regular dependence on the spike-RBD protein concentration, particularly for the values in the pico- and nano-molar ranges. Conversely, the γ fitting parameter quite satisfactorily reproduces the behavior achieved for the gm value estimated at the single VGS = 0.2 V, displaying a monotonously decaying behavior as a function of the spike-RBD protein concentration for the anti-spike protein functionalized gate electrode. In this case, no significant variations were observed before and after incubation in a solution with 57 nM of spike-RBD protein solution when a gate functionalized with anti-NP protein was employed. Hence, this approach confirms that the VGS region in which IDS shows a square dependence on VGS is the most sensitive to the changes produced by the gate incubation process. It should be outlined that the progressive occurrence of the gm reduction in the OECT response, as a consequence of a protein detection at the gate/electrolyte interface, has been reported in other works, confirming that transconductance behavior is strictly related to the specific VGS region [25]. This feature points to the need for alternative theoretical frameworks that can provide more insightful descriptions of the OECT electrical behavior [39,40].
In summary, in this work, we assessed the selective detection of spike-RBD proteins down to femto-molar range by using OECTs based on PEDOT:PSS channels. In addition to the overall IDS modulation values, our analysis revealed that the electrical parameters associated to the transconductance (i.e., the slope of the IDS vs VGS transfer curves) and the VGS region (between 0 and 0.3 V) of the OECT response are more clearly affected by the gate incubation process. The overall time for the testing procedure is about 30 min, with 20 min being required for the incubation step and the remaining time needed to perform the electrical analysis consisting of the recording of a number of consecutive transfer curves to be averaged. Given the discussed findings, at this stage, the sensitivity performances is estimated to be approximately in the range of few tens of nanomolars of the antigen concentration (corresponding to few hundreds of ng/mL of the spike-RBD protein). Further improvements can be achieved by optimizing the functionalization protocol or using alternative probes for antigen recognition [25]. Actually, as reported by a large number of works discussed in the literature, the sensitivity performances of devices conceived for the spike protein detection are strongly dependent (i.e., over several orders of magnitudes) on the specific sensing schemes [1,2]. It should be also mentioned that, in the analyzed configuration, the organic active channel immersed in PBS can be used for multiple experiments, in combination with different properly functionalized gate electrodes upon their incubation in the solution of interest. This approach simplifies the use of these devices which, owing to their low applied voltages, are suitable for further miniaturization and can be integrated in a compact and portable measurement set-up. Going beyond the precise diagnosis of coronavirus, all these features are very promising to the development of cheap, fast, and user-friendly tools for point-of-care analysis where OECT could be embedded in microfluidic–OECT-integrated platforms.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/bios13070739/s1, Figure S1: FITR spectra recorded for a gold surface before (SAMPLE A) and after (SAMPLE B), showing the complete functionalization process with anti-spike antibody; Figure S2: IGATE curves corresponding to the output (a) and transfer curves (b) reported, respectively, in Figure 2a,b of the main text; Figure S3: (a) Transfer curves achieved (VDS = −0.3V) using different values of the time delay between the VGS application and the IDS recording, where the corresponding transconductance (gm) and IGATE values are reported in (b) and (c), respectively; Figure S4: (a–c) Transconductance (gm) curves as a function of VGS extracted from the transfer curves shown, respectively, in Figure 3a–c of the main text. (d) gm curves evaluated from the transfer curves shown in Figure 4a; Figure S5: OECT transfer curves (scatter symbols) with VDS = −0.3 V and related fitting curves with the equation IDS = ( α + β · V + γ · V 2 ) . (solid lines) achieved using a functionalized gate with anti-spike protein before incubation and after incubation in solutions with (a) 57 femtomolar (fM); (b) 57 picomolar (pM) and (c) 57 nanomolar (nM) spike-RBD protein; (d) Corresponding transfer curves and fitting lines obtanied for a functionalized gate with nucleocapside (NP) antibody before incubation and after incubation with 57 nanomolar spike-RBD protein solution.

Author Contributions

Conceptualization, M.B., V.R.V., S.E. and V.P.; methodology, M.B., V.P., V.R.V., S.E., S.L.M., V.B. and E.C.; validation, M.B, G.T., P.D., S.L.M. and V.P., formal analysis, M.B. and P.D.; investigation, M.B., G.T., A.L., P.D., S.L.M., V.B., E.C. and V.P.; data curation, V.R.V., S.E., A.L. and P.D.; writing—original draft preparation, M.B., V.R.V., S.E., P.D. and V.P.; writing—review and editing, M.B., G.T., M.C.; supervision, M.B., M.C. and V.P.; project administration, V.P.; funding acquisition, V.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by FISR COVID 2020 (FISR2020IP_01635). The authors acknowledge also the financial support from Università degli Studi di Napoli Federico II (Programma per il finanziamento della ricerca di Ateneo, D.R. n. 2055 del 15 maggio 2022).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article or supplementary materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fabiani, L.; Caratelli, V.; Fiore, L.; Scognamiglio, V.; Antonacci, A.; Fillo, S.; De Santis, R.; Monte, A.; Bortone, M.; Moscone, D. State of the art on the SARS-CoV-2 toolkit for antigen detection: One year later. Biosensors 2021, 11, 310. [Google Scholar] [CrossRef] [PubMed]
  2. Antiochia, R. Electrochemical biosensors for SARS-CoV-2 detection: Voltametric or impedimetric transduction? Bioelectrochemistry 2022, 147, 108190. [Google Scholar] [CrossRef] [PubMed]
  3. Healy, B.; Khan, A.; Metezai, H.; Blyth, I.; Asad, H. The impact of false positive COVID-19 results in an area of low prevalence. Clin. Med. 2021, 21, e54. [Google Scholar] [CrossRef] [PubMed]
  4. Jia, X.; Xiao, L.; Liu, Y. False negative RT-PCR and false positive antibody tests–concern and solutions in the diagnosis of COVID-19. J. Infect. 2021, 82, 414–451. [Google Scholar] [CrossRef] [PubMed]
  5. Mouliou, D.S.; Gourgoulianis, K.I. False-positive and false-negative COVID-19 cases: Respiratory prevention and management strategies, vaccination, and further perspectives. Expert Rev. Respir. Med. 2021, 15, 993–1002. [Google Scholar] [CrossRef]
  6. Surkova, E.; Nikolayevskyy, V.; Drobniewski, F. False-positive COVID-19 results: Hidden problems and costs. Lancet Respir. Med. 2020, 8, 1167–1168. [Google Scholar] [CrossRef]
  7. Davis, H.E.; McCorkell, L.; Vogel, J.M.; Topol, E.J. Long COVID: Major findings, mechanisms and recommendations. Nat. Rev. Microbiol. 2023, 21, 133–146. [Google Scholar] [CrossRef]
  8. Swank, Z.; Senussi, Y.; Manickas-Hill, Z.; Yu, X.G.; Li, J.Z.; Alter, G.; Walt, D.R. Persistent circulating severe acute respiratory syndrome coronavirus 2 spike is associated with post-acute coronavirus disease 2019 sequelae. Clin. Infect. Dis. 2023, 76, e487–e490. [Google Scholar] [CrossRef]
  9. Rivnay, J.; Inal, S.; Salleo, A.; Owens, R.M.; Berggren, M.; Malliaras, G.G. Organic electrochemical transistors. Nat. Rev. Mater. 2018, 3, 17086. [Google Scholar] [CrossRef]
  10. Ohayon, D.; Inal, S. Organic Bioelectronics: Organic Bioelectronics: From Functional Materials to Next-Generation Devices and Power Sources. Adv. Mater. 2020, 32, 2070267. [Google Scholar]
  11. Tarabella, G.; Pezzella, A.; Romeo, A.; D’Angelo, P.; Coppedè, N.; Calicchio, M.; d’Ischia, M.; Mosca, R.; Iannotta, S. Irreversible evolution of eumelanin redox states detected by an organic electrochemical transistor: En route to bioelectronics and biosensing. J. Mater. Chem. B 2013, 1, 3843–3849. [Google Scholar] [CrossRef] [PubMed]
  12. Fu, Y.; Wang, N.; Yang, A.; Law, H.K.w.; Li, L.; Yan, F. Highly sensitive detection of protein biomarkers with organic electrochemical transistors. Adv. Mater. 2017, 29, 1703787. [Google Scholar] [CrossRef] [PubMed]
  13. Fu, Y.; Wang, N.; Yang, A.; Xu, Z.; Zhang, W.; Liu, H.; Law, H.K.-W.; Yan, F. Ultrasensitive detection of ribonucleic acid biomarkers using portable sensing platforms based on organic electrochemical transistors. Anal. Chem. 2021, 93, 14359–14364. [Google Scholar] [CrossRef]
  14. Burtscher, B.; Manco Urbina, P.A.; Diacci, C.; Borghi, S.; Pinti, M.; Cossarizza, A.; Salvarani, C.; Berggren, M.; Biscarini, F.; Simon, D.T. Sensing Inflammation Biomarkers with Electrolyte-Gated Organic Electronic Transistors. Adv. Healthc. Mater. 2021, 10, 2100955. [Google Scholar] [CrossRef] [PubMed]
  15. Brennan, D.; Galvin, P. Flexible substrate sensors for multiplex biomarker monitoring. MRS Commun. 2018, 8, 627–641. [Google Scholar] [CrossRef] [Green Version]
  16. Hu, J.; Wei, W.; Ke, S.; Zeng, X.; Lin, P. A novel and sensitive sarcosine biosensor based on organic electrochemical transistor. Electrochim. Acta 2019, 307, 100–106. [Google Scholar] [CrossRef]
  17. Khodagholy, D.; Rivnay, J.; Sessolo, M.; Gurfinkel, M.; Leleux, P.; Jimison, L.H.; Stavrinidou, E.; Herve, T.; Sanaur, S.; Owens, R.M. High transconductance organic electrochemical transistors. Nat. Commun. 2013, 4, 2133. [Google Scholar] [CrossRef] [Green Version]
  18. Harikesh, P.C.; Yang, C.-Y.; Tu, D.; Gerasimov, J.Y.; Dar, A.M.; Armada-Moreira, A.; Massetti, M.; Kroon, R.; Bliman, D.; Olsson, R. Organic electrochemical neurons and synapses with ion mediated spiking. Nat. Commun. 2022, 13, 901. [Google Scholar] [CrossRef]
  19. Go, G.T.; Lee, Y.; Seo, D.G.; Lee, T.W. Organic Neuroelectronics: From Neural Interfaces to Neuroprosthetics. Adv. Mater. 2022, 34, 2201864. [Google Scholar] [CrossRef]
  20. Rivnay, J.; Owens, R.M.; Malliaras, G.G. The rise of organic bioelectronics. Chem. Mater. 2013, 26, 679–685. [Google Scholar] [CrossRef]
  21. Owens, R.M.; Malliaras, G.G. Organic electronics at the interface with biology. MRS Bull. 2010, 35, 449–456. [Google Scholar] [CrossRef] [Green Version]
  22. Simon, D.T.; Gabrielsson, E.O.; Tybrandt, K.; Berggren, M. Organic bioelectronics: Bridging the signaling gap between biology and technology. Chem. Rev. 2016, 116, 13009–13041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Liao, C.; Zhang, M.; Niu, L.; Zheng, Z.; Yan, F. Highly selective and sensitive glucose sensors based on organic electrochemical transistors with graphene-modified gate electrodes. J. Mater. Chem. B 2013, 1, 3820–3829. [Google Scholar] [CrossRef] [PubMed]
  24. Preziosi, V.; Barra, M.; Tomaiuolo, G.; D’Angelo, P.; Marasso, S.L.; Verna, A.; Cocuzza, M.; Cassinese, A.; Guido, S. Organic electrochemical transistors as novel biosensing platforms to study the electrical response of whole blood and plasma. J. Mater. Chem. B 2022, 10, 87–95. [Google Scholar] [CrossRef]
  25. Guo, K.; Wustoni, S.; Koklu, A.; Díaz-Galicia, E.; Moser, M.; Hama, A.; Alqahtani, A.A.; Ahmad, A.N.; Alhamlan, F.S.; Shuaib, M. Rapid single-molecule detection of COVID-19 and MERS antigens via nanobody-functionalized organic electrochemical transistors. Nat. Biomed. Eng. 2021, 5, 666–677. [Google Scholar] [CrossRef]
  26. Liu, H.; Yang, A.; Song, J.; Wang, N.; Lam, P.; Li, Y.; Law, H.K.-w.; Yan, F. Ultrafast, sensitive, and portable detection of COVID-19 IgG using flexible organic electrochemical transistors. Sci. Adv. 2021, 7, eabg8387. [Google Scholar] [CrossRef]
  27. Bai, Z.; Cao, Y.; Liu, W.; Li, J. The SARS-CoV-2 nucleocapsid protein and its role in viral structure, biological functions, and a potential target for drug or vaccine mitigation. Viruses 2021, 13, 1115. [Google Scholar] [CrossRef]
  28. Burbelo, P.D.; Riedo, F.X.; Morishima, C.; Rawlings, S.; Smith, D.; Das, S.; Strich, J.R.; Chertow, D.S.; Davey, R.T., Jr.; Cohen, J.I. Detection of nucleocapsid antibody to SARS-CoV-2 is more sensitive than antibody to spike protein in COVID-19 patients. MedRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  29. Preziosi, V.; Barra, M.; Perazzo, A.; Tarabella, G.; Agostino, R.; Marasso, S.L.; D’Angelo, P.; Iannotta, S.; Cassinese, A.; Guido, S. Monitoring emulsion microstructure by Organic Electrochemical Transistors. J. Mater. Chem. C 2017, 5, 10. [Google Scholar] [CrossRef]
  30. Preziosi, V.; Tarabella, G.; D’Angelo, P.; Romeo, A.; Barra, M.; Guido, S.; Cassinese, A.; Iannotta, S. Real-time monitoring of self-assembling wormlike micelle formation by organic transistors. RSC Adv. 2015, 5, 8. [Google Scholar] [CrossRef]
  31. Battistoni, S.; Verna, A.; Marasso, L.S.; Cocuzza, M.; Erokhin, V. On the Interpretation of Hysteresis Loop for Electronic and Ionic Currents in Organic Memristive Devices. Phys. Status Solid 2020, 217, 1900985. [Google Scholar] [CrossRef]
  32. De Meutter, J.; Goormaghtigh, E. Evaluation of protein secondary structure from FTIR spectra improved after partial deuteration. Eur. Biophys. J. 2021, 50, 613–628. [Google Scholar] [CrossRef] [PubMed]
  33. Fabian, H.; Mäntele, W. Infrared spectroscopy of proteins. In Handbook of Vibrational Spectroscopy; Wiley: New York, NY, USA, 2006. [Google Scholar]
  34. Miller, L.M.; Bourassa, M.W.; Smith, R.J. FTIR spectroscopic imaging of protein aggregation in living cells. Biochim. Biophys. Acta (BBA)-Biomembr. 2013, 1828, 2339–2346. [Google Scholar] [CrossRef] [Green Version]
  35. Friedlein, J.T.; McLeod, R.R.; Rivnay, J. Device physics of organic electrochemical transistors. Org. Electron. 2018, 63, 398–414. [Google Scholar] [CrossRef]
  36. Preziosi, V.; Barra, M.; Villella, V.R.; Esposito, S.; D’Angelo, P.; Marasso, S.L.; Cocuzza, M.; Cassinese, A.; Guido, S. Immuno-Sensing at Ultra-Low Concentration of TG2 Protein by Organic Electrochemical Transistors. Biosensors 2023, 13, 448. [Google Scholar] [CrossRef] [PubMed]
  37. Bernards, D.A.; Malliaras, G.G. Steady-State and Transient Behavior of Organic Electrochemical Transistors. Adv. Funct. Mater. 2007, 17, 3538–3544. [Google Scholar] [CrossRef]
  38. Macchia, E.; Romele, P.; Manoli, K.; Ghittorelli, M.; Magliulo, M.; Kovács-Vajna, Z.M.; Torricelli, F.; Torsi, L. Ultra-sensitive protein detection with organic electrochemical transistors printed on plastic substrates. Flex. Print. Electron. 2018, 3, 034002. [Google Scholar] [CrossRef]
  39. Nissa, J.; Janson, P.; Simon, D.T.; Berggren, M. Expanding the understanding of organic electrochemical transistor function. Appl. Phys. Lett. 2021, 118, 053301. [Google Scholar] [CrossRef]
  40. Cucchi, M.; Weissbach, A.; Bongartz, L.M.; Kantelberg, R.; Tseng, H.; Kleemann, H.; Leo, K. Thermodynamics of organic electrochemical transistors. Nat. Commun. 2022, 13, 4514. [Google Scholar] [CrossRef]
Figure 1. (a,b) Pictures of the layout of the employed OECT devices and of the external gate electrodes; (c) cartoon of the gate electrode functionalization; CLSM gate images before (d) and after (e) incubation with secondary antibodies conjugated to Alexa Fluor® 488 dye.
Figure 1. (a,b) Pictures of the layout of the employed OECT devices and of the external gate electrodes; (c) cartoon of the gate electrode functionalization; CLSM gate images before (d) and after (e) incubation with secondary antibodies conjugated to Alexa Fluor® 488 dye.
Biosensors 13 00739 g001
Figure 2. (a) Output curves reporting IDS as a function of VDS and applying VGS between −0.6 and 0.6 V; (b) OECT transfer curves and (c) normalized transfer curves, with respect to the IDS current at VGS = −0.6 V, measured as a function of VGS with VDS ranging from −0.6 to −0.1 V; (d) transconductance gm as a function of VGS with VDS ranging between 0.1 and −0.6 V.
Figure 2. (a) Output curves reporting IDS as a function of VDS and applying VGS between −0.6 and 0.6 V; (b) OECT transfer curves and (c) normalized transfer curves, with respect to the IDS current at VGS = −0.6 V, measured as a function of VGS with VDS ranging from −0.6 to −0.1 V; (d) transconductance gm as a function of VGS with VDS ranging between 0.1 and −0.6 V.
Biosensors 13 00739 g002
Figure 3. Transfer curves (top panels) and normalized transfer curves with respect to the IDS current recorded at VGS = −0.5 V (bottom panels) measured as a function of VGS by using a functionalized gate with anti-spike protein before incubation and after incubation in solutions with (a) 57 femtomolar (fM) spike-RBD protein; (b) 57 picomolar (pM) spike-RBD protein; (c) 57 nanomolar (nM) spike-RBD protein; (d) cartoon reporting the spike antibody–antigen binding event.
Figure 3. Transfer curves (top panels) and normalized transfer curves with respect to the IDS current recorded at VGS = −0.5 V (bottom panels) measured as a function of VGS by using a functionalized gate with anti-spike protein before incubation and after incubation in solutions with (a) 57 femtomolar (fM) spike-RBD protein; (b) 57 picomolar (pM) spike-RBD protein; (c) 57 nanomolar (nM) spike-RBD protein; (d) cartoon reporting the spike antibody–antigen binding event.
Biosensors 13 00739 g003
Figure 4. (a) Transfer curves (top panel) and normalized transfer curves with respect to the IDS current recorded at VGS = −0.5 V (bottom panel) as a function of VGS by using a functionalized gate with nucleocapside (NP) antibody before and after incubation in 57 nanomolar spike-RBD protein solution; (b) cartoon of the anti-NP protein and spike-RBD protein unbound configuration.
Figure 4. (a) Transfer curves (top panel) and normalized transfer curves with respect to the IDS current recorded at VGS = −0.5 V (bottom panel) as a function of VGS by using a functionalized gate with nucleocapside (NP) antibody before and after incubation in 57 nanomolar spike-RBD protein solution; (b) cartoon of the anti-NP protein and spike-RBD protein unbound configuration.
Biosensors 13 00739 g004
Figure 5. (a) OECT IDS modulation (top panel) and its percent variation (bottom panel) as a function of spike- RBD protein concentration for a gate functionalized with spike and nucleocapside antibody. (b) Transconductance (gm) values estimated at different VGS values (VGS = +0.2 V top, VGS = 0 V middle, VGS = −0.2 V bottom) as a function of spike-RBD protein concentration for gate functionalized with spike antibody. The gm values estimated at VGS = 0.2 V, achieved for gate functionalized with nucleocapside antibody, are reported in the top panel of Figure 5b.
Figure 5. (a) OECT IDS modulation (top panel) and its percent variation (bottom panel) as a function of spike- RBD protein concentration for a gate functionalized with spike and nucleocapside antibody. (b) Transconductance (gm) values estimated at different VGS values (VGS = +0.2 V top, VGS = 0 V middle, VGS = −0.2 V bottom) as a function of spike-RBD protein concentration for gate functionalized with spike antibody. The gm values estimated at VGS = 0.2 V, achieved for gate functionalized with nucleocapside antibody, are reported in the top panel of Figure 5b.
Biosensors 13 00739 g005
Figure 6. β and γ fitting parameters (top panels) and the corresponding percent variations (bottom panels) are reported in (a,b), respectively, as a function of the spike-RBD protein concentration for gate electrodes functionalized with spike and nucleocapside antibodies.
Figure 6. β and γ fitting parameters (top panels) and the corresponding percent variations (bottom panels) are reported in (a,b), respectively, as a function of the spike-RBD protein concentration for gate electrodes functionalized with spike and nucleocapside antibodies.
Biosensors 13 00739 g006
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

Barra, M.; Tomaiuolo, G.; Villella, V.R.; Esposito, S.; Liboà, A.; D’Angelo, P.; Marasso, S.L.; Cocuzza, M.; Bertana, V.; Camilli, E.; et al. Organic Electrochemical Transistor Immuno-Sensors for Spike Protein Early Detection. Biosensors 2023, 13, 739. https://doi.org/10.3390/bios13070739

AMA Style

Barra M, Tomaiuolo G, Villella VR, Esposito S, Liboà A, D’Angelo P, Marasso SL, Cocuzza M, Bertana V, Camilli E, et al. Organic Electrochemical Transistor Immuno-Sensors for Spike Protein Early Detection. Biosensors. 2023; 13(7):739. https://doi.org/10.3390/bios13070739

Chicago/Turabian Style

Barra, Mario, Giovanna Tomaiuolo, Valeria Rachela Villella, Speranza Esposito, Aris Liboà, Pasquale D’Angelo, Simone Luigi Marasso, Matteo Cocuzza, Valentina Bertana, Elena Camilli, and et al. 2023. "Organic Electrochemical Transistor Immuno-Sensors for Spike Protein Early Detection" Biosensors 13, no. 7: 739. https://doi.org/10.3390/bios13070739

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