You are on page 1of 7

Biosensors and Bioelectronics 102 (2018) 610–616

Contents lists available at ScienceDirect

Biosensors and Bioelectronics


journal homepage: www.elsevier.com/locate/bios

Electrochemical detection of carcinoembryonic antigen T


a,b b,⁎ b,1 a b,c,⁎⁎
Xuefang Gu , Zhe She , Tianxiao Ma , Shu Tian , Heinz-Bernhard Kraatz
a
School of Chemistry and Chemical Engineering, Nantong University, Nantong 226007, PR China
b
Department of Physical and Environmental Science, University of Toronto Scarborough, 1265 Military Trail, Toronto, Ontario, Canada M1C 1A4
c
Department of Chemistry, University of Toronto, 80. St. George Street, Toronto, Canada M5S 3H6

A R T I C L E I N F O A B S T R A C T

Keywords: In this work, a sandwich-type electrochemical immunosensor for carcinoembryonic antigen (CEA) detection has
Ferrocene been constructed and tested. Unlike many other sensors using external electrochemical species in the electrolyte
Gold nanoparticles to generate an electrochemical signal, a ferrocene derivative has been integrated into the design of the sensor to
Carcinoembryonic antigen provide an internal reporting system, allowing detection of CEA in buffers and biological samples. Gold nano-
Electrochemical immunosensor
particles, which have been used to increase the conductivity of sensing surfaces, also carry immobilized sec-
ondary anti-CEA and a ferrocene derivative. The shelf life testing of the sensor shows good performance after
storage for 4 weeks. The sensor has been calibrated against different concentration of the target protein using
square wave voltammetry. The calibration curve has been obtained in the range of 0.05–20 ng mL−1, and the
detection limit for CEA is ~ 0.01 ng mL−1. The capability of the immunosensor has been verified by performing
detection of CEA in human serum samples.

1. Introduction immunoassay, based on the specificity of an antibody to its corre-


sponding antigen (Pei et al., 2013). Formation of an immunocomplex
The development of an effective diagnostic method for cancer de- has been quantified by colorimetry (Kim et al., 2009), fluorescence (He
tection, progression, and monitoring of treatment efficacy remains an et al., 2013), chemiluminescence (Wang et al., 2016; Zhao et al., 2016),
important bioanalytical target. With the development of proteomics, electrochemical methods (Chen et al., 2012; Gao et al., 2011; Han et al.,
serum tumor markers such as carcinoembryonic antigen (CEA), alpha- 2016; Lai et al., 2016; Wang et al., 2013; Weng et al., 2013; Yang et al.,
fetoprotein, CA125, CA19-9, and cytokeratin 19 fragment, have been 2017) and by surface-enhanced Raman scattering (SERS) (Chon et al.,
widely used as screening tools for high-risk populations in annual 2009). However, traditional fluorescence-based methods, such as en-
medical checkups (Han et al., 2008). The average CEA concentration in zyme-linked immunosorbent assay (ELISA), often suffer from low sen-
a healthy human is below 5 μg L−1, while serum CEA levels above sitivity due to a large optical background. Radioimmunoassays have the
20 μg L−1 indicates the presence of cancer (Han et al., 2017). This has potential to expose the operator to a potential safety hazard and require
been demonstrated for colon cancer (Vogel et al., 2001), breast cancer special waste treatment. While the main problem of Raman-based
(Han and Magliocco, 2016), lung cancer (Han et al., 2008), gastric readout methods is related to reproducibility issues (Gu et al., 2014).
cancer (Tachibana et al., 1998), and ovarian carcinoma (Kudoh et al., Amongst conventional immunoassay techniques, electrochemical
1999). The coincidence rate increases up to 78% when a combination of methods have attracted considerable interest, since they offer high
CEA and CYFRA21-1 are used for the diagnosis of non-small cell lung sensitivity, short diagnostic time, and the possibility of miniaturization.
cancer (Wu et al., 2007). In addition, the preoperative and periodic Electrochemical responses are readily quantifiable (Li et al., 2014b; Xu
postoperative CEA determination provides useful information in pre- et al., 2017). Detection by amperometric methods are most promising
dicting stage, tumor progression and recurrence (Liu et al., 2012). because of their large linear range, high sensitivity, and low detection
Therefore, a rapid, accurate and sensitive CEA detection method is of limit (Cevik et al., 2016; He et al., 2008; Ou et al., 2007; Ren et al.,
great interests for the community and may provide an approach to 2014; Zhang et al., 2017).
rapid and facile cancer screening. We have exploited ferrocene (Fc) derivatives as part of electro-
At present, the benchmark for single-protein measurement is by chemical sensors before. Examples of this application are the use of Fc


Corresponding author.
⁎⁎
Corresponding author at: Department of Physical and Environmental Science, University of Toronto Scarborough, 1265 Military Trail, Toronto, Ontario, Canada M1C 1A4.
E-mail addresses: zhe.she@utoronto.ca (Z. She), bernie.kraatz@utoronto.ca (H.-B. Kraatz).
1
Current address: Department of Chemistry, University of British Columbia, Vancouver Campus, 2036 Main Mall, Vancouver, BC, Canada V6T 1Z1.

https://doi.org/10.1016/j.bios.2017.12.014
Received 25 September 2017; Received in revised form 28 November 2017; Accepted 7 December 2017
Available online 08 December 2017
0956-5663/ © 2017 Elsevier B.V. All rights reserved.
X. Gu et al. Biosensors and Bioelectronics 102 (2018) 610–616

Scheme 1. Schematic illustration of (a) preparation of Fc-labeled


nanogold probes (Ab2-Au-Fc), and (b) the procedures for the
construction of sandwich-type electrochemical biosensor.

as a peptide conjugate for the detection of bacteria (Li et al., 2014b) and gold electrodes through an immunoreaction. Fc-SH molecules forming a
exploiting Fc as a redox relay for monitoring phosphorylation of pep- self-assembled monolayer (SAM) on the surface of AuNPs via a covalent
tides and proteins, as well as enhancing the electrochemical imaging of attachment process have greatly improved the stability and repeat-
protein receptors (She et al., 2017; Wang et al., 2015). For a protein ability of the signals in comparison with drop-casting method. In ad-
immunoassay, the most direct approach would be to conjugate Fc di- dition, direct electron transfer could also be achieved through the
rectly to antibodies. For example, Tang's group reported a proof-of- monolayer of Fc, thus generating a faster electron transport rate than
concept study of the determination of human IgG protein by integrating through the much thicker drop-dry films, which is essential in fabri-
an immunoreaction and hybridization chain reaction (HCR), where Fc cating an electrochemical sensor with high sensitivity and rapid re-
molecules were conjugated to well-designed hairpin DNAs for the HCR sponse (Love et al., 2005). The electrochemistry was used to monitor
that is triggered by initiator strands on the surface of gold nanoparticles the step-by-step fabrication of the sensor, followed by CEA detection by
(Zhang et al., 2012). Feng and coworkers fabricated a sandwich-type square-wave voltammetry. Results presented here show that the ap-
CEA immunosensor using Fc as electrochemical probe. The en- proach discussed here has a wide dynamic range and can be used to
capsulated HRP was to catalyze H2O2, which further promoted the detect CEA in human serum.
electron transfer of Fc-COOH and amplified the electrochemical signal,
and a linear range of 0.001–10 ng mL−1 with a detection limit of 2. Experimental sections
0.0002 ng mL−1 was achieved (Feng et al., 2016). However, no matter
what labeling approach is adopted, the labeling process itself is labor- 2.1. Materials and reagents
ious and time-consuming, and the efficiency of binding is usually quite
low due to steric hindrance. Carcinoembryonic antigen (CEA), monoclonal anti-CEA (designated
To solve these issues, various nanoprobes have been prepared that as Ab1), and polyclonal anti-CEA (designated as Ab2) were purchased
carry a large number of electroactive molecules, such as methylene from R&D system. Lipoic acid N-hydroxysuccinimide ester (LPA) was
bule, RuHex, and HRP, on a range of different nanocarriers, including purchased from Synchem UG & Co (See SI for the detail of other re-
gold nanoparticles (AuNPs) (Bai et al., 2017; Giljohann et al., 2010; agents and buffers used in this work).
Larguinho and Baptista, 2012; Liu et al., 2016; Xu et al., 2017), gra-
phene (Halder et al., 2017; Weng et al., 2013), carbon nanotubes (Gao
2.2. Preparation of Fc-labeled biofunctionalized AuNPs (Ab2-AuNPs-Fc)
et al., 2011; Han et al., 2016), chitosan (He et al., 2008), polymer
dendrimers (Cevik et al., 2016). This approach has been widely em-
Colloidal gold sol was synthesized by aqueous reduction of chlor-
ployed for the enhancement of the electrochemical signal, but is also
oauric acid with trisodium citrate according to the previous literature
fraud with problems including the loss of the label during the electro-
(Grabar et al., 1995) (See SI for the detail of preparation). Scheme 1a
chemical experiment. Additionally, it is also difficult to control the
illustrates the formation of the Fc-labeled nanogold probes (Ab2-
concentration of deposited electroactive materials, thus reducing the
AuNPs-Fc). Polyclonal anti-CEA was first immobilized on the surface of
reproducibility of the result (Gu et al., 2015, 2016). In this regard, in-
AuNPs through the electrostatic interaction between negative charged
troducing thiol groups (–SH or –SCH3) into Fc molecules might be a
AuNPs and the protonated amino group of antibody. The electroactive
feasible approach, because Fc–SH can be immobilized on the surface of
Fc molecules were then chemisorbed to the gap of the antibody via self-
gold nanoparticles via formation of covalent bonds between gold and
assembly, which was done by adding a few microliters of 0.2 M K2CO3
sulfur atoms, to form a stable and well-organized films (Love et al.,
to the gold sol for the adjustment of acidity to pH 8.5. Then, 150 μL of
2005), where nanoparticles act as nanocarriers and Fc serves as the
50 μg mL−1 Ab2 was added to 1 mL of the AuNPs solution, and the
signal amplifier.
mixture was incubated at room temperature for 2 h. Next, 10 μL Fc-SH
In this study, we have developed a sensitive and reproducible
ethanol solution (1 mM) was added to the mixture to react with Ab2-
electrochemical immunosensor for the detection of CEA. A sandwich-
AuNPs for 1 h. Finally, PEG8000 was added to a final concentration of
like immunocomplex containing electroactive Fc was the core of this
0.5% for stabilizing the AuNPs. The excess antibody and Fc were re-
sensor, gold nanoparticles (AuNPs) that modified by Fc groups and
moved by repeated centrifugation at 9000 rpm for 20 min and rinsed
carrying immobilized anti-CEA, were then connected on the surface of
with storage buffer. The final deposition (Ab2-AuNPs-Fc) was

611
X. Gu et al. Biosensors and Bioelectronics 102 (2018) 610–616

suspended in 3 mL of storage buffer and stored at 277 K for further use. resistance of the surface to [Fe(CN)6]3-/4-. After the bare Au electrode
The as-prepared Fc modified gold nanoparticles could be stably stored was modified with LPA (curve b), the Ret value increased to 1684 Ω,
in solution for more than three months without aggregation. which could be attributed to the insulating effect of the SAM film
formed on the electrode surface, confirming successful immobilization
2.3. Construction of electrochemical biosensor of the linker on the surface of Au electrodes. When the electrode was
further modified with antibody (curve c), the Ret value dramatically
Prior to the construction, the electrodes were immersed into a pir- increased to 3012 Ω (curve c), suggesting that the electron transfer was
anha solution (Volume ratio H2SO4/H2O2 = 3:1) for 30 s and rinsed hindered by the immobilized Ab1 molecules. Compared to the un-
with a large amount of water. The bare gold electrodes were polished blocked electrode, Ret value showed a slight decrease after the blocking
using aqueous slurries of alumina (1 down to 0.05 µm) to obtain a stage (curve d). This phenomenon was possibly due to the replacement
mirror surface and then followed by electrochemical cleaning. of nonspecifically absorbed antibody with the small blocking agent
Scheme 1b shows the different stages during the fabrication of the during the process. As expected, the size of the arc on EIS increased
proposed CEA immunosensor. The cleaned electrodes were im- again (curve e) on the recognition of CEA antigen. When the Ab2-
mediately immersed in a 1 mL of 2 mM LPA-NHS ethanol solution at AuNPs-Fc nanoprobes were captured onto the surface of the modified
277 K overnight. The LPA-NHS monolayer here served as the linker Au electrode, the Ret value decreased significantly to 1094 Ω (curve f),
between the anti-CEA and Au electrodes. The modified electrodes were which was ascribed to the fact that AuNPs facilitated the electron-
then thoroughly rinsed with ethanol and blown dry by N2. A 5 μL transfer. The change helps confirming formation of the im-
droplet of Ab1 (200 μg mL−1) in PBS buffer (pH 7.4) was carefully munocomplex sandwiches on the surface of Au electrodes.
placed over the surface of gold electrode and the electrode was then In order to further confirm the preparation process of the im-
placed at 277 K for 48 h. The Ab1-modified electrodes were rinsed with munosensor, CVs were carried out using Fe(CN)63-/4- as the redox
an excess volume of washing buffer and Milli-Q water and blown dry probe. Fig. 2B shows CVs of each step (same as Fig. 2a) during the
followed by soaking them in an ethanolamine solution for 2 h at room immunosensor construction in 5 mM [Fe(CN)6]3-/4- solution at a scan
temperature to block the unreacted LPA. The electrodes were incubated rate of 100 mV/s. As could be expected, the sequential modification of
in a solution of CEA antigen (different concentrations in 10 mM PBS the electrode blocked the electron transfer between the external redox
buffer) for 1 h at 310 K, rinsed again. Finally, the gold electrodes with probes and the modified electrode. For the CV on a bare Au electrode
Ab1-CEA antigen immunocomplexes were immersed in the Ab2-AuNPs- (curve a), a couple of reversible redox peaks at 0.302 and 0.234 V were
Fc solution for 45 min at 310 K to construct sandwich structures for the observed. When the Au electrode was modified with LPA (curve b), an
final electrochemical measurement. obvious decrease in peak current and an increase in the peak separation
were observed as the potential of peaks shifted outward. This sup-
2.4. Electrochemical measurements pression of current indicates the formation of a LPA monolayer,
blocking electron transfer between the Au surface and the redox couple.
Cyclic voltammograms (CV), electrochemical impedance spectro- Subsequently, the electrode was modified with antibody (curve c); a
scopy (EIS), Chronoamperometry (CA), and square wave voltammetry further decrease in peak current and an increase in the peak separation
(SWV) were recorded on a CHI-660C electrochemical station (CH were observed. However, after we blocked the unreacted active sites of
Instruments, Austin, Texas) with an enclosed Faraday cage. (See SI for LPA with ethanolamine to minimize non-specific recognitions, an ob-
the detailed electrochemical parameters). servable increase in the peak current and a narrower peak separation
compared with those of curve c occurred (curve d), due to the re-
3. Results and discussion placement of nonspecifically absorbed macromolecule Ab1 with EA. As
for the recognition CEA antigen (curve e) step, the current lowered
3.1. Characterization of the CEA immunosensor again, this was consistent with the enhanced electron-transfer barriers
introduced by successive modifications. Finally, after the im-
The step-by-step modification was followed up by TEM and the munoreaction with Fc-labeled nanoprobes (curve f), the peak current
images illustrate the formation of antibody modified nanoprobes. showed a dramatic increase (compare to curve e), and such an increase
Fig. 1a shows the TEM image of the relatively uniform AuNPs. As ob- in current could be basically ascribed to the fact that AuNPs facilitate
served in Fig. 1b, the TEM image of Fc-SH modified AuNPs, the dark the electron transfer (Larguinho and Baptista, 2012). The modification
center (AuNPs) was wrapped by a lighter color shell and the thickness has been monitored using both CV and EIS to make sure formation of
of the lighter parts was approximately 2 nm, suggesting the self-as- immunocomplex. These CV results are consistent with the above-
sembled Fc monolayer on the surface of AuNPs (Petrizza et al., 2016). mentioned EIS results, and consequently, all electrochemical results
Fig. 1c shows the morphology of Ab2-AuNPs-Fc. The thickness of the confirm the successful fabrication of the immunocomplex.
lighter parts increased to 5–6 nm, which was in good agreement with
the size of anti-CEA (Hammarström, 1999). Fig. 1d represents the SPR 3.2. Electrochemical response of the CEA immunosensor
spectra of the non-aggregated AuNPs and the antibody modified na-
noprobes, which occurred at 520 nm and 530 nm, respectively. Gen- Generally, the kinetic reversibility of a redox couple can be assessed
erally, the locations of SPR strongly depend on the shape, size and the according to the separation of the anodic and cathodic peak potentials
environment of the metal nanoparticles (Kelf et al., 2006). In this case, in CV. A control experiment was carried out by putting the Fc derivative
the red shift of the SPR band upon the adsorption of the anti-CEA was onto electrodes and investigated the redox behavior of the Fc firstly. We
most likely resulted from a change in the dielectric property of the have immersed a bare Au electrode in 2 mM Fc-SH ethanol solution for
media surrounding the AuNPs. 2 h, and then carried out the CV detection in PBS. A pair of well-defined
EIS is an informative technique to monitor the interfacial properties redox peaks was clearly observed, corresponding to the oxidation of
of modified electrodes by estimating parameters such as solution re- ferrocene to ferrocenium ion. The cathodic and anodic peaks are lo-
sistance (Rs), electron-transfer resistance (Ret), and double layer capa- cated at 0.294 V and 0.360 V (scan rate of 0.1 V s−1), respectively. The
citance (Cdl). The EIS measurements were performed in this study to peak-to-peak separation ΔEp = Epa −Epc = 66 mV, and the ratio of
follow the surface electrochemical property of each step of the im- the anodic peak current to the cathodic Ipa/Ipc was 1.32 ± 0.03, sug-
munosensor construction; the Nyquist plots results are shown in Fig. 2A gesting a quasi-reversible one-electron electrochemical reaction for the
(see Table S1 for the values of the equivalent circuit). The EIS of the immobilized Fc molecules. Fig. 3a shows the CVs of the Fc modified
bare Au electrode had a small arc (curve a), implying very low electrode with various scan rates ranging from 25 mV s−1 to 5 V s−1. As

612
X. Gu et al. Biosensors and Bioelectronics 102 (2018) 610–616

Fig. 1. TEM images of (a) AuNPs, (b) AuNPs-FcSH, (c) Ab2-


AuNPs-Fc, and UV/Vis spectra of AuNPs and Fc-labeled nanogold
probe.

can be seen, the anodic and cathodic peak currents increased simulta-
neously with increasing scan rate. The inset of Fig. 3a illustrates the
linear dependence of the anodic and cathodic peak currents on the scan
rates, indicating a surface-controlled process (Ou et al., 2007). More-
over, only a slight increase in the potential separation of the cathodic
and anodic peaks ranging from 25 mV s−1 to 5 V s−1 was observed,
suggesting a fast electron transfer rate between the self-assembled
monolayer of Fc molecules and the Au electrode.
The Fc-Gold electrode investigation was followed by probing the Fc
as part of the immunosensor. Fig. 3b presents the CV performance of an
integrated CEA immunosensor with 5 ng mL−1 CEA antigen as the
target. The redox peaks that resulted from Fc to Fc+ are well revealed
by CV. The cathodic and anodic peaks are located at 0.215 V and
0.289 V, respectively. It is clearly evident that in the presence of CEA,
the electrode exhibit reversible electrochemical signal from Fc, similar
to those obtained at Fc modified Au electrodes. Compared to Fig. 3a,
the redox potentials of the CEA immunosensor in Fig. 3b show a ne-
gative shift for ~ 80 mV. This could be attributed to the different way of
linking and different chemical environment around the Fc molecules
(Halder et al., 2017). As reported previously, the immunocomplex
formed on the electrode surface often had a poor electrical conductivity
and the long-range electron transfer would slow down the rate of
electron transfer, thus restrict their analytical performance in some
degree (Lai et al., 2016). In this work, though the peak-to-peak se-
paration had a slight increase, the 74 mV separation and a 1.23 Ipa/Ipc
ratio still implied a reversible and fast electron transfer. More im-
portant, the CV response of the immunosensor has not changed after 20
times of successive scans between 0.0 and 0.6 V, suggesting that the as-
Fig. 2. (A) EIS spectra and (B) cyclic voltammograms of [Fe(CN)6]3-/4- at the surface of prepared CEA biosensor is stable and the signals are repeatable and
the modified gold electrode after each modification step: (a) bare gold electrode; (b) after reliable. Overall, the results demonstrate that the attached Fc in na-
coating with LPA; (c) after immobilization of Ab1; (d) after blocking with EA; (e) after nogold probes can reversibly communicate fast electronic transporta-
immobilization of 1 ng mL−1 CEA and (f) after immobilization of Ab2-AuNPs-Fc. The CVs tion with the supporting electrode, and in turn the feasibility of the
were measured in a 10 mM HEPES buffer (pH7.4) containing 1 M NaClO4 and 5 mM/
immunosensor. Moreover, the surface coverage of the electroactive
5 mM [Fe(CN)6]3-/4- at a scan rate of 0.1 V s−1.

613
X. Gu et al. Biosensors and Bioelectronics 102 (2018) 610–616

Fig. 4. (a) Typical SWVs of the electrochemical immunobiosensor in 0.1 M PBS (pH~ 7.0)
Fig. 3. (a) CV evolutions of Au/Fc obtained in 0.1 M PBS (pH 7.0) containing at scan rates
solution containing increasing concentrations of CEA (0.05, 0.1, 0.5, 1.0, 5.0, 10.0,
0.025–5 V s−1 (from inner to outer curves: 0.025, 0.05, 0.075, 0.1, 0.15, 0.2, 0.25, 0.3,
20.0 ng mL−1), (b) the calibration curve of the electrochemical biosensor for determi-
0.35, 0.4, 0.45, 0.5, 0.6, 0.7, 0.8, 0.9, 1.0, 1.5, 2.0, 2.5, 3.0, 3.5, 4.0, 4.5, 5.0 V s−1), the
inset of a is the plot of current (Ip) vs. scan rate, (b) CV of the CEA biosensor in 0.1 M PBS nation of CEA standards. Each point is an average of five parallel determination results;
(pH 7.0) PBS at a scan rate 0.1 V s−1, the concentration of CEA is 5 ng mL−1. each error bar is the sample-to-sample standard deviation of the currents.

substance was calculated according to the Laviron Eq. (1) (Laviron, exhibited in Fig. S2a. That is, as the concentration of the nanogold
1974), and the amount of the immobilized Fc molecules was calculated probe increasing, the current increases accordingly and reached a pla-
as 2.43 × 10−9 mol·cm−2 (see SI for Fig. S1 and the detailed calcula- teau at 0.6 nM. Therefore, 0.6 nM of the nanogold probe was referred as
tion process) the optimum concentration in subsequent experiments.
In addition to the investigation on the concentrations, the incuba-
n2⋅F 2⋅A⋅Γ ⋅ν n⋅F ⋅Q⋅ν tion time was explored by monitoring the current response of the im-
Ip = =
4RT 4RT (1) munosensor. The SWV current of the immunosensor, shown in Fig. S2d,
increases with increasing incubation time used in the sandwich im-
3.3. Optimization of experimental conditions munoreaction and a constant value is reached at 45 min, indicating that
45 min is enough for the antigen-antibody recognition reaction.
In order to achieve the optimum analytical performance, the effect
of antibody concentration on the SWV current of the immunosensor was 3.4. Analytical performance
investigated. The relationship between the concentration of Ab1 and
current is shown in Fig. S2a (see SI). The peak current significantly Current-time (I/T) curves at different CEA concentrations were
increases with an increasing antibody concentration until the peak obtained to further investigate the electrochemical characteristics of
current reaches a plateau at 200 μg mL−1. Hence, the Ab1 concentra- this sensor. The signal was also analyzed towards quantifying the
tion of 200 μg mL−1 was selected in this work to have maximum concentration of CEA (see SI for Fig. S3 and the detailed description).
electrochemical response. The concentration of the Ab2 was also tested As the figure shows, the charging current of the system increases
to optimize the conditions. The current response as shown in Fig. S2b slightly with increased CEA concentration and the steady state current
increases quickly at the range of low concentration and reaches the can be reached within 8 s, indicating that the sensor has a faster elec-
highest value as the final concentration of Ab2 is 7.5 μg mL−1. How- tron transfer rate and a shorter response time (Han et al., 2016; Li et al.,
ever, when the concentration continues to increase, the current of the 2014a; Ye et al., 2013). Furthermore, calibration was attempted by
immunosensor shows a slight decrease. This might be due to the com- analyzing the I/T curves. It is found that the linear range of this
petitive adsorption between Ab2 and the labeled Fc molecules. Al- quantitative method was narrow (from 0.5 to 10 ng mL−1) and the
though a concentration of 7.5 μg mL−1 of Ab2 may not have highest detection limit (~ 0.2 ng mL−1) could not meet the detection require-
adsorption on gold nanoparticles, but it is a good balanced point be- ment. Therefore, SWV was used for the quantitative determination of
tween Ab2 and the labeled Fc molecules. The trend of current changes CEA in this work because it exhibited higher current sensitivity and a
in Fig. S2c (the effect of nanogold probe concentration) is similar to that better resolution than CV. Fig. 4a shows the voltammetric responses to

614
X. Gu et al. Biosensors and Bioelectronics 102 (2018) 610–616

Table 1
Comparison of the proposed method with reference methods for CEA determination.

Immunoassay substrate Readout method Linear range (ng mL−1) LOD ( ng mL−1) Ref.

Plasma treatment of paper chemiluminscene 0.1–80 0.03 (Zhao et al., 2016)


ZnS-CdS nanodots on MoS2 Electrochemiluminescence 0.05–20 0.03 (Wang et al., 2016)
Carbon nanotube decorated gold nanoclusters Electrochemistry, CV 0.5–5.0 0.1 (Gao et al., 2011)
Nanosilver-coated magnetic beads and gold-graphene nanolabels Electrochemistry, DPV 0.02–60 0.02 (Chen et al., 2012)
Polydopamine modified mesoporous silica nanoparticles Electrochemistry, DPV 0.01–40 0.002 (Wang et al., 2013)
Streptavidin-functionalized nitrogen-doped grapheme Electrochemistry, DPV 0.02–12 0.01 (Yang et al., 2017)
Palladium hybrid vanadium pentoxide/multiwalled carbon nanotubes Electrochemistry, Amperometric 0.01–10 0.003 (Han et al., 2016)
Ferrocene functionalized Fe3O4@SiO2 Electrochemistry, DPV 0.001–10 0.0002 (Feng et al., 2016)
Polymer containing both aldehyde and ferrocene functional groups Electrochemistry, DPV 0.1–20 0.07 (Zhang et al., 2017)
Anti-CEA/AuNP@nafion/FC@CHIT/GCE Electrochemistry, SWV 0.03–100 0.01 (Shi and Ma, 2011)
This work Electrochemistry, SWV 0.05–20 0.01 –

CEA with different concentrations. The electrochemical oxidation peak verify the detection ability of the immunosensor. Samples were made
of Fc is located at ca. 0.28 V and an increase in CEA concentration re- by adding varying volumes (100.0, 300.0, 500.0, 1000.0 μL) of
sults in the oxidation current increasing at the same time. Moreover, 1.0 μg mL−1 CEA standard solution firstly to 1.0 mL human serum and
there is a linear relationship between the current and CEA concentra- then diluted in a 100 mL volumetric flask with pH 7.4 PBS. SWV and
tion ranging from 0.05 to 20 ng mL−1 (Fig. 4b), with the linear re- the calibration curve obtained above were applied to determine the
gression equation expressed as y = 0.4494 x + 2.043 (r = 0.9968). concentration of each solution; the recovery and RSD are shown in
The limit of detection was calculated to be 0.01 ng mL−1, using IUPAC Table S2. The results suggest that the proposed immunoassay method
definition LOD = 3 sb/q, where sb is the standard deviation of the blank may have a great potential in the real-to-life sample analysis of CEA
signal, and q is the slope of the calibration curve. Comparing to the with high accuracy and good reliability.
calibration curves presented in the existing literatures, the im-
munosensor proposed in this study provides a relatively low limit of 4. Conclusion
detection. The comparison of different fabrication methods, the range
of linearity, and the detection limit of this method with other reports for In conclusion, we have demonstrated a method how to incorporate
CEA detection is presented in Table 1. the Fc molecules and AuNPs into a sensor towards sensitive detection of
To investigate the specificity of the immunosensor, a high con- CEA. The Fc-labeled AuNPs provide internal electrochemical signals,
centration (500 ng mL−1) of possible interferents (glucose, cysteine, therefore no redox species is needed in the electrolyte. Recognition
bovine serum albumin, human IgG, and prostate specific antigen) was antibody (anti-CEA, Ab2) loaded on the AuNPs offers the biological
introduced in two experiments. As shown in Fig. S4A, when sensor is detection/interaction capability. The Fc–AuNP-anti-CEA probes loaded
individually tested in CEA and control interfering samples, dominating a high amount of Fc molecules through self-assembly and realized the
signal was only observed in CEA solution. The test was then carried out purpose of signal amplification, thus the proposed immunosensor ex-
when interfering samples were each mixed into 10 ng mL−1 CEA so- hibited improved sensitivity. The detection limit was observed to be as
lution, respectively. The detection signal is very reliable that re- low as 0.01 ng mL−1. The sensor was also tested in biological samples
produced at very similar level of 7 µA as shown in Fig. S4b. and good recovery was observed. Moreover, the good specificity, re-
The repeatability of the electrochemical immunosensor was eval- producibility and long-term stability of this immunosensor not only
uated by calculating the relative standard deviation (RSD) of intra- offer an opportunity for the detection of CEA in low concentration, but
assay at 0.1, 1.0 and 10.0 ng mL−1 CEA concentration for 8 times of also provide a platform to develop various biosensors for many other
successive measurements. The RSDs of this sensor were 8.7%, 5.3% and similar proteins determination.
3.7%, respectively. The reproducibility was estimated by calculating
the inter-assay RSD from various batches (n = 5), for each assay a new Acknowledgments
freshly prepared nanogold probe was used. The inter-assay RSDs were
9.8%, 6.5%, and 4.2%, respectively. The results were acceptable for a Financial support from the Nature Science Foundation of China (No.
trace analysis. The stability of the sensor was tested by comparing 21505079), the Natural Science Foundation of Jiangsu Province (Grants
stored sensors with fresh ones. The Fc-AuNPs-Ab2 probe The sensors No BK20150402), and the Applied Research Program of Nantong City
was suspended in 10 mM phosphate sodium buffer solution, pH 7.4, (GY12015017, MS12016039) are gratefully acknowledged. X.Gu and S.
0.1% PEG8000 for 1–4 weeks before using in the detection. n PBS at Tian are also supported by the Jiangsu Overseas Research and Training
277 K and used for CEA detection 3 weeks later, and the current re- Program. H.-B. Kraatz is grateful for the financial support from Natural
sponse maintained about 91.8% of the original value(Fig. S5, see SI. Sciences and Engineering Research Council of Canada and from the
The good reproducibility and stability of this electrode could be asso- University of Toronto Scarborough.
ciated with a combination of following factors: (i) the good bio-
compatibility of gold nanoparticles; (ii) self-assembly method allows Fc Appendix A. Supporting information
molecules to functionalize the surface of gold nanoparticles via Au-S
covalent bonds, thus achieving a more stable film than physically fixed Supplementary data associated with this article can be found in the
ones; (iii) closely packed Fc molecules occupy most of the AuNPs after online version at http://dx.doi.org/10.1016/j.bios.2017.12.014.
the modification of antibody, making it difficult for the adsorption of
unwanted interference component; (iv) the internal electrochemical References
responses with no external redox species and enzyme conjugates
needed. Bai, R.-Y., Zhang, K.-L., Li, D.-L., Zhang, X., Liu, T.-Z., Liu, Y., Hu, R., Yang, Y.-H., 2017.
Chin. J. Anal. Chem. 45 (1), 48–55.
Cevik, E., Bahar, O., Senel, M., Abasiyanik, M.F., 2016. Biosens. Bioelectron. 86,
3.5. Preliminary analysis of real samples 1074–1079.
Chen, H., Tang, D., Zhang, B., Liu, B., Cui, Y., Chen, G., 2012. Talanta 91, 95–102.
Human serum samples with spiked CEA were prepared and used to Chon, H., Lee, S., Son, S.W., Oh, C.H., Choo, J., 2009. Anal. Chem. 81 (8), 3029–3034.

615
X. Gu et al. Biosensors and Bioelectronics 102 (2018) 610–616

Feng, T., Qiao, X., Wang, H., Sun, Z., Hong, C., 2016. Biosens. Bioelectron. 79 Liu, S., Zhou, J., Li, H., Yin, C., Lai, G., 2016. Electroanalysis 28 (12), 2993–2999.
(Supplement C), 48–54. Liu, Z.-P., Li, L.-M., Liu, X.-L., Zhang, D.-X., 2012. Oncol. Res. Treat. 35 (3), 108–113.
Gao, X., Zhang, Y., Chen, H., Chen, Z., Lin, X., 2011. Anal. Biochem. 414 (1), 70–76. Love, J.C., Estroff, L.A., Kriebel, J.K., Nuzzo, R.G., Whitesides, G.M., 2005. Chem. Rev.
Giljohann, D.A., Seferos, D.S., Daniel, W.L., Massich, M.D., Patel, P.C., Mirkin, C.A., 2010. 105 (4), 1103–1170.
Angew. Chem. Int. Ed. 49 (19), 3280–3294. Ou, C., Yuan, R., Chai, Y., Tang, M., Chai, R., He, X., 2007. Anal. Chim. Acta 603 (2),
Grabar, K.C., Freeman, R.G., Hommer, M.B., Natan, M.J., 1995. Anal. Chem. 67 (4), 205–213.
735–743. Pei, X., Zhang, B., Tang, J., Liu, B., Lai, W., Tang, D., 2013. Anal. Chim. Acta 758 (0),
Gu, X., Jiang, G., Jiang, G., Chen, T., Zhan, W., Li, X., Wu, S., Tian, S., 2015. Talanta 137, 1–18.
189–196. Petrizza, L., Genovese, D., Valenti, G., Iurlo, M., Fiorani, A., Paolucci, F., Rapino, S.,
Gu, X., Li, X., Wu, S., Shi, J., Jiang, G., Jiang, G., Tian, S., 2016. RSC Adv. 6 (10), Marcaccio, M., 2016. Electroanalysis 28 (11), 2777–2784.
8070–8078. Ren, K., Wu, J., Yan, F., Ju, H., 2014. Sci. Rep. 4, 4360.
Gu, X., Yan, Y., Jiang, G., Adkins, J., Shi, J., Jiang, G., Tian, S., 2014. Anal. Bioanal. She, Z., Topping, K., Dong, B., Shamsi, M.H., Kraatz, H.-B., 2017. Chem. Commun. 53,
Chem. 406 (7), 1885–1894. 2946–2949.
Halder, A., Zhang, M., Chi, Q., 2017. Biosens. Bioelectron. 87, 764–771. Shi, W., Ma, Z., 2011. Biosens. Bioelectron. 26 (6), 3068–3071.
Hammarström, S., 1999. Semin. Cancer Biol. 9 (2), 67–81. Tachibana, M., Takemoto, Y., Nakashima, Y., Kinugasa, S., Kotoh, T., Dhar, D.K., Kohno,
Han, H.S., Magliocco, A.M., 2016. Clin. Breast Cancer 16 (3), 166–179. H., Nagasue, N., 1998. J. Am. Coll. Surg. 187.
Han, J., Jiang, L., Li, F., Wang, P., Liu, Q., Dong, Y., Li, Y., Wei, Q., 2016. Biosens. Vogel, I., Francksen, H., Soeth, E., Henne-Bruns, D., Kremer, B., Juhl, H., 2001. Am. J.
Bioelectron. 77, 1104–1111. Surg. 181 (2), 188–193.
Han, J., Li, Y., Feng, J., Li, M., Wang, P., Chen, Z., Dong, Y., 2017. J. Electroanal. Chem. Wang, N., She, Z., Lin, Y.C., Martić, S., Mann, D.J., Kraatz, H.B., 2015. Chem. Eur. J. 21
786, 112–119. (13), 4988–4999.
Han, K.Q., Huang, G., Gao, C.F., Wang, X.-l., Ma, B., Sun, L.Q., Wei, Z.J., 2008. Am. J. Wang, X., Miao, J., Xia, Q., Yang, K., Huang, X., Zhao, W., Shen, J., 2013. Electrochim.
Clin. Oncol. 31 (2), 133–139. Acta 112, 473–479.
He, X., Yuan, R., Chai, Y., Shi, Y., 2008. J. Biochem. Biophys. Methods 70 (6), 823–829. Wang, Y.-L., Cao, J.-T., Chen, Y.-H., Liu, Y.-M., 2016. Anal. Methods 8 (26), 5242–5247.
He, Y., Tian, J., Hu, K., Zhang, J., Chen, S., Jiang, Y., Zhao, Y., Zhao, S., 2013. Anal. Chim. Weng, X., Cao, Q., Liang, L., Chen, J., You, C., Ruan, Y., Lin, H., Wu, L., 2013. Talanta
Acta 802, 67–73. 117, 359–365.
Kelf, T.A., Sugawara, Y., Cole, R.M., Baumberg, J.J., Abdelsalam, M.E., Cintra, S., Wu, G.-P., Jing, B., Zhao, Y.-J., Wang, E.-H., 2007. Acta Cytol. 51 (4), 679–680.
Mahajan, S., Russell, A.E., Bartlett, P.N., 2006. Phys. Rev. B 74 (24), 245415. Xu, S., Zhang, R., Zhao, W., Zhu, Y., Wei, W., Liu, X., Luo, J., 2017. Biosens. Bioelectron.
Kim, D., Daniel, W.L., Mirkin, C.A., 2009. Anal. Chem. 81 (21), 9183–9187. 92, 570–576.
Kudoh, K., Kikuchi, Y., Kita, T., Tode, T., Takano, M., Hirata, J., Mano, Y., Yamamoto, K., Yang, Z., Lan, Q., Li, J., Wu, J., Tang, Y., Hu, X., 2017. Biosens. Bioelectron. 89 (Pt 1),
Nagata, I., 1999. Gynecol. Obstet. Investig. 47 (1), 52–57. 312–318.
Lai, G., Cheng, H., Xin, D., Zhang, H., Yu, A., 2016. Anal. Chim. Acta 902, 189–195. Ye, D., Liang, G., Li, H., Luo, J., Zhang, S., Chen, H., Kong, J., 2013. Talanta 116,
Larguinho, M., Baptista, P.V., 2012. J. Proteom. 75 (10), 2811–2823. 223–230.
Laviron, E., 1974. J. Electroanal. Chem. Interfacial Electrochem. 52 (3), 355–393. Zhang, B., Liu, B., Tang, D., Niessner, R., Chen, G., Knopp, D., 2012. Anal. Chem. 84 (12),
Li, H., Wang, Y., Ye, D., Luo, J., Su, B., Zhang, S., Kong, J., 2014a. Talanta 127, 255–261. 5392–5399.
Li, Y., Afrasiabi, R., Fathi, F., Wang, N., Xiang, C., Love, R., She, Z., Kraatz, H.B., 2014b. Zhang, X., Shen, Y., Zhang, Y., Shen, G., Xiang, H., Long, X., 2017. Talanta 164, 483–489.
Biosens. Bioelectron. 58, 193–199. Zhao, M., Li, H., Liu, W., Guo, Y., Chu, W., 2016. Biosens. Bioelectron. 79, 581–588.

616

You might also like