You are on page 1of 9

Electrochimica Acta 192 (2016) 158–166

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Reversible Switched Detection of Dihydroxybenzenes Using a


Temperature-sensitive Electrochemical Sensing Film
Yuanqing Zhoua,1, Chao Chena,1, Jia Zhaob , Junjie Feia,* , Yonglan Dinga , Yuanli Caia,c
a
Key Laboratory of Environmentally Friendly Chemistry and Applications of Ministry of Education, College of Chemistry, Xiangtan University, Xiangtan
411105, PR China
b
College of Resource and Environment, Hunan Agricultural University, Changsha 410128, PR China
c
Department of Polymer Science and Engineering, Soochow University, Suzhou 215123, PR China

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

Article history: A composite sensing film (PGS), consisting of poly(N-isopropylacrylamide)101-b-poly(2-acrylamidoethyl


Received 6 November 2015 benzoate)37 (PNIPAM101-b-PAAEB37), graphene oxides (GO) and short multi-walled carbon nanotubes
Received in revised form 21 January 2016 (SMWCNs), was fabricated and modified onto a working electrode. The sensing film served as a reversible
Accepted 25 January 2016
switch for electrochemical detection, with the switching behaviour responding to thermal stimuli. Cyclic
Available online 28 January 2016
voltammetry of hydroquinone (HQ) and catechol (CC) at the PGS film-modified electrode displayed large
peak currents when the temperature was above the lower critical solution temperature (LCST) of
Keywords:
PNIPAM101-b-PAAEB37. These large currents disappeared at low temperature. Interestingly, the composite
graphene oxide
multi-walled carbon nanotube
film showed reversed electrochemical “on/off” behaviour as compared to previously reported switchable
electrochemical sensor electrodes, which were modified only with temperature-responsive polymers. This behaviour can be
temperature-responsive polymer attributed to the temperature-dependent phase transition of PNIPAM101-b-PAAEB37 and cooperative
electrochemical “on/off” detection effects of the other two functional components (SMWCNs and GO). The repeatable “on/off” switching of
the voltammetric responses of HQ/CC on the PGS-modified electrode were achieved via regulating the
solution temperature. This research provides a new type of temperature-controlled switchable electrode
with potential applications in the design of novel sensors, fuel cells and electronics.
ã 2016 Elsevier Ltd. All rights reserved.

1. Introduction Stimuli-responsive polymers, which exhibit changes in confor-


mation or structure due to external stimuli [12–17], provide
Chemically modified electrodes with reversible “on/off” switch- another approach to switchable electrode construction. Electrode
ing or tunable interfacial properties toward external physicochem- surfaces coated with responsive polymers can significantly change
ical signals are intriguing systems for fundamental research, and their interfacial properties (e.g., charged state, permeability, and
offer great promise in such fields as switchable biosensors for pH [18–20]) in response to external signals such as light,
biomedical applications [1,2], switchable fuel cells [3], electro- temperature or chemical/biochemical input, and show tunable
chemical information processing systems [4], and other electro- or switchable electrochemical behaviours [21–27]. Tam [19]
chemical devices [5]. One approach to switchable electrode reported that indium tin oxide (ITO) electrodes modified with
construction involves applying metallic or magnetic species in poly(4-vinyl pyridine) (P4VP) brushes showed switchable interfa-
electrochemical devices. The electrode interfacial properties (e.g., cial activity toward redox species. This was due to the high
conductivity) can be significantly tuned by variable potentials or an permeability for redox species of the protonated, swollen P4VP
external magnetic field [6–11]. For instance, Cu2+ ions in a poly brush state (electrochemical “on” state) and the impermeability of
(acrylic acid)/CdS matrix immobilized on an Au electrode surface the deprotonated, hydrophobic, shrunken P4VP brush state
were reduced to metallic clusters and then oxidized back to the (electrochemical “off” state). Temperature-sensitive polymers
cationic state by altering the potentials, which caused dramatic can undergo swelling–shrinking phase transitions in response to
changes in the conductivity of the polymer layer [7]. thermal stimuli [28,29], and have attracted interest as tempera-
ture-controlled electrochemical switch devices [30,31]. Poly(N-
isopropylacrylamide) (PNIPAM) has been extensively studied as a
temperature-sensitive polymer in the sensor field. Its LCST is
* Corresponding author. Tel.: +86 731 58298876; fax: +86 731 58292251.
E-mail address: fei_junjie@xtu.edu.cn (J. Fei). approximately 32  C [32,33], just within the manageable temper-
1
These author contributed equally to this work. ature range for conventional electrochemical experiments.

http://dx.doi.org/10.1016/j.electacta.2016.01.183
0013-4686/ ã 2016 Elsevier Ltd. All rights reserved.
Y. Zhou et al. / Electrochimica Acta 192 (2016) 158–166 159

PNIPAM films fabricated on electrode surfaces show a reversible, at the PGS-modified electrode were switched off when PNIPAM101-
temperature-controlled “on-off” property toward redox probes. b-PAAEB37 was swollen, but switched on while PNIPAM101-b-
Specifically, electrochemical responses of redox probes appeared PAAEB37 was shrunken (shown in Scheme 1). This switchable
when the PNIPAM brush was in the swollen state (below 32  C) but behaviour was reversed from that of previously reported switchable
disappeared when the PNIPAM brush was in the shrunken state electrodes (modified only by stimuli-responsive polymers), which
(above 32  C) [24]. were switched off in their hydrophobic shrunken state and
Carbon nanomaterials such as fullerenes, carbon nanotubes, switched on in their swollen state [19,24]. The opposite “on/off”
and graphene exhibit unique optical, electronic, and catalytic behaviour of the composite PGS electrode can be attributed to the
properties. In the sensor field, carbon nanomaterials used in cooperative, synergistic effect of PNIPAM101-b-PAAEB37, GO and
electrode modification can accelerate signal transduction through SMWCNs in response to temperature stimuli. The PGS electrode also
catalytic activity, conductivity, and biocompatibility [34–37]. showed reversible and repeatable electrochemical “on-off” proper-
Graphene oxide (GO), a carbon nanomaterial, possesses a two- ties through controlling the solution temperature.
dimensional, graphene-like structure with oxygen-containing
functional groups such as epoxide,  OH, and COOH. Due to 2. Experimental section
this structure, GO exhibits good hydrophobicity, moderate
conductivity, high chemical stability and excellent electrochemical 2.1. Reagents and Materials
properties, and can act as a support for the deposition of inorganic
nanocatalysts or polymers. GO is widely used in the sensing field to SMWCNs (OD <8 nm, Length: 0.5–2 mm, Purity >95 wt%) were
facilitate or mediate the charge/electron transfer between the obtained from the Chinese Academy of Sciences, Chengdu Organic
electroactive species and the electrode surface [38,39]. The Chemistry Company. Hydroquinone (HQ) and catechol (CC) were
integration of stimuli-responsive polymers and carbon nano- purchased from Shunyi Chemical Company of China and used
materials generates hybrid systems that combine the properties of without further purification. GO was prepared according to
the nanomaterials with stimuli-responsive functions, to obtain previous procedures [50]. PNIPAM101-b-PAAEB37 was prepared
switchable or tunable electronic, photonic and catalytic properties according to previous procedures [51]. Synthesis details of
[40–45]. To date, hybrid materials have attracted substantial PNIPAM101-b-PAAEB37 and its 1H NMR and GPC data are presented
attention and have extensive applications in the sensor field in the Supplementary Material, Fig. S1 and Fig. S2. All other
[46,47]. A major goal of such studies is linked to achieving chemical reagents were of analytical grade. Water was purified
synergistic efforts to bring advanced properties to the hybrid twice by distillation.
materials, which can improve upon their individual properties
[48]. 2.2. Fabrication of the Modified Electrodes
In this paper, a hybrid material consisting of the temperature-
sensitive copolymer PNIPAM101-b-PAAEB37, GO and SMWCNs was Glassy carbon electrodes (GCE) were carefully polished with
utilized to develop a temperature-controlled switch electrode. 0.5-mm alumina on a polish cloth, followed by sequential cleaning
PNIPAM101-b-PAAEB37 contains benzene rings, which can be for 5 min with ethanol and double-distilled water in an ultrasonic
noncovalently attached to GO surfaces through p–p stacking bath, then dried in air before use.
interactions, which can improve the stability of the hybrid film [49]. The GO solution (45 mL, 2 mg mL1) was added to a PNIPAM101-b-
The electrochemical responses of hydroquinone (HQ)/catechol (CC) PAAEB37 aqueous solution (75 mL,10 mg mL1), and the mixture was

Scheme 1. Reversible, temperature-controlled “on/off” electrochemical response of HQ/CC at the PGS interface.
160 Y. Zhou et al. / Electrochimica Acta 192 (2016) 158–166

sonicated for 10 min in an ice-water bath to yield a suspension polymer composite film on the GCE. The PG electrode was obtained
solution of PNIPAM101-b-PAAEB37/GO (PG). The resulting dispersion by dropping 6 mL of the PG solution on the surface of a GCE.
solution was stable at 5  C in the refrigerator for months with no Similarly, the GS electrode was obtained by dropping the mixed
precipitation. SMWCNs (0.6 mg) were added to PNIPAM101-b- solution of GO and SMWCNs on a GCE.
PAAEB37 solution (1 mL, 10 mg mL1), and this mixture required
shaking immediately preceding further use. The PNIPAM101-b- 2.3. Electrochemical measurements
PAAEB37/GO/SMWCNs (PGS) solution was formed by mixing the
PNIPAM101-b-PAAEB37/SMWCNs (PS) solution with the PG solution Electrochemical experiments were performed at a CHI
at a 1:1 (v/v) ratio. Then, 6 mL of the PGS solution was placed onto the 630 electrochemical workstation using a conventional three-
surface of a clean GCE and dried at room temperature to form a electrode cell. One of the modified electrodes was used as the

Fig. 1. SEM images of (a) GO, (b) SMWCNs, (c) PG composite film, (d) PS composite film, and the (e) PGS composite film. Inset of (b): 40000 image of SMWCNs.
Y. Zhou et al. / Electrochimica Acta 192 (2016) 158–166 161

working electrode, a platinum wire was employed as the counter


electrode and a saturated calomel electrode served as the reference
electrode. During the electrochemical measurements, all electro-
lyte solutions were purged with high-purity nitrogen for at least
20 min and were then maintained under nitrogen atmosphere. A
Zahner Zennium electrochemical workstation (ZAHNER-ELEKTRIK
GmbH & Co. KG) was employed for electrochemical impedance
spectra (EIS) measurements. A phosphate buffer solution was used
as the supporting electrolyte. The surface morphology of different
composite films was observed by using a JSM-6610Lv scanning
electron microscopy system (SEM, JEOL, Japan) at an accelerating
voltage of 30 V. Prior to SEM analysis, samples were coated with a
gold layer.

3. Results and Discussion

3.1. SEM Characterization

The surface morphologies of GO, SMWCNs, PG, PS, and PGS


films are depicted in Fig. 1. GO shows a lamellar shape with a folded
structure and resembles a piece of folded paper (Fig. 1a). The
SMWCNs display an expected uniform nanofiber structure.
Because the SMWCNs were not functionalized, some agglomera-
tion appeared, and the dispersion of the SMWCNs in the
PNIPAM101-b-PAAEB37 matrix became inhomogeneous, as the
arrows show in Fig. 1d. PG is a homogeneous composite film
(Fig. 1c), indicating that GO sheets are homogeneously imbedded
into the polymer matrix, and confirming the good compatibility
and strong interfacial adhesion between GO and the PNIPAM101-b-
PAAEB37 matrix. Unlike the surface morphologies of PG and PS, PGS
appears to be piled up in thick cakes (Fig. 1e), clearly indicating the
presence of SMWCNs aggregates in this film, which increases the Fig. 2. A. Electrochemical impedance spectra obtained at different modified
film thickness. electrodes in supporting electrolyte (20  C). a(!) PGS/GCE; b(*) PNIPAM101-b-
PAAEB37/GCE; c(^) PG/GCE; d(~) GO/GCE. B. Electrochemical impedance spectra
obtained at the PGS/GCE upon stepwise heating of the supporting electrolyte from
3.2. EIS study of Different Electrodes and EIS Temperature Effect of the 26 to 40  C. Frequency range: 0.1 Hz to 100 KHz; Supporting electrolyte: 0.1 mol L1
KCl containing 5 mmol L1 K3[Fe(CN)6]/K4[Fe(CN)6] (1:1).
PGS Electrode

EIS is often used as a tool to investigate electron transfer


blocking effects on electrode surfaces in various sensors or
interfacial electron transfer rate. The result also shows that
biosensors, and the switchable/tunable operation of electrochem-
SMWCNs are successfully embedded in the composite film.
ical gates based on polymer films [52]. The semicircle diameter in
The temperature-dependent EIS spectra of PGS electrodes were
the complex EIS plane plot of Z“ against Z‘ reflects the electron
also studied in the presence of redox probes [Fe(CN)6]3/4
transfer resistance (Rct). Based on the Rct of different modified
(Fig. 2B). Upon stepwise heating of the solution from 26  C to 30  C,
electrodes, we can obtain their interfacial impedance change,
the semicircle diameters of the EIS decreased slightly, but a sharp
which is related to interfacial properties such as permeability and
decrease in the semicircle diameters was observed upon heating to
conductivity. Fig. 2A shows the EIS results of different modified
32  C. The semicircle diameters remain smaller, with little change,
electrodes toward soluble redox probes [Fe(CN)6]3/4. The Rct at
upon continued heating, indicating a low interfacial electron
PNIPAM101-b-PAAEB37/GCE (Fig. 2A curve b) is smaller than that
transfer resistance of PGS/GCE above 32  C and a large electron
obtained at PG/GCE (Fig. 2A curve c), reflecting fast electron
transfer resistance below 32  C. The EIS spectra also show that
transfer kinetics at the first electrode. Because PNIPAM101-b-
modified electrodes without PNIPAM101-b-PAAEB37 presented no
PAAEB37 is swollen at the 20  C testing temperature, diffusional
temperature-responsive behaviour from 26  C to 36  C. The
redox probes [Fe(CN)6]3/4 can permeate from the swollen
changes of the Rct values in the PGS film above and below 32  C
polymer film to the electrode surface [19,24], resulting in enhanced
may therefore be attributed to the temperature-responsive phase
electron transfer kinetics at PNIPAM101-b-PAAEB37/GCE. The PG
transition behaviour of PNIPAM101-b-PAAEB37 at 32  C (Fig. S3).
composite film consisting of GO and PNIPAM101-b-PAAEB37
Upon stepwise heating above 32  C, PNIPAM101-b-PAAEB37 under-
displays the largest Rct effect (Fig. 2A curve c) among the four
goes hydrophobic shrinkage, allowing the SMWCNs embedded in
modified electrodes, although the GO film exhibits the fastest
the film to act as conductive channels, resulting in a decreased Rct.
interfacial electron transfer (Fig. 2A curve d). The large Rct effect of
Below 32  C, PNIPAM101-b-PAAEB37 is in a swollen state, leading to
PG/GCE may be attributed to the poor permeability of GO toward
a widened distance among GO sheets that cuts off conductive
redox probes because of its two-dimensional planar structure and
channels, causing a higher Rct. This EIS result shows that a PGS-
the poor conductivity of PNIPAM101-b-PAAEB37. When SMWCNs
modified electrode can act as a temperature-controlled interfacial
are embedded in a PG film, the Rct decreases dramatically (Fig. 2A,
resistance switch, and it would be beneficial to study the
curve a). SMWCNs possess high electrical conductivity, so the
switchable electrocatalytic performance of the PGS composite
presence of SMWCNs in the composite film enhances the
toward redox species such as HQ/CC.
162 Y. Zhou et al. / Electrochimica Acta 192 (2016) 158–166

2.0

A 2.0 1.5

Current/1e-6A
c
1.0
0.5 b
a
1.5 0.0
Current / 1e-4A

-0.5

-1.0

1.0 -1.5
e
0.8 0.6 0.4 0.2 0.0 -0.2 -0.4 -0.6 -0.8
E/ V

0.5 d
a b c
0.0

-0.5

-1.0

0.8 0.6 0.4 0.2 0.0 -0.2 -0.4 -0.6 -0.8


E/ V
Fig. 4. Cyclic voltammograms of HQ at the PGS electrode in 0.1 mol L1 PBS solution

B 2.0 over a pH range of 5.0  8.5. Scan rate: 75 mV s1. Testing temperature: 37  C. Inset is
2.0 1.5 the plot of oxidative peak potential (^), reductive peak potential (&), and midpoint
potential (~) against pH in the pH range of 5.0  8.5.
Current/1e-6A

1.0 g
0.5 f
1.5
Current / 1e-4A

a
0.0

-0.5 When increasing scan rates from 10 to 220 mV s1, both the
1.0 -1.0
h reduction and oxidation peak currents increase linearly with the
0.8 0.6 0.4 0.2 0.0 -0.2 -0.4 -0.6 -0.8
E/ V increasing square root of the scan rates (v1/2) (inset of Fig. S5),
0.5 d demonstrating the typical diffusion-controlled process of HQ at the
PGS electrode.
0.0 f g a The influence of electrolyte pH on HQ peak potentials at the PGS
electrode was also studied. Fig. 4 shows that an increase of solution
-0.5 pH from 5 to 8.5 causes a negative shift in both the reduction and
0 0
oxidation peak potentials. Plots of Epa,Epc and E0 (E0 = (Epa + Epc)/2)
0

-1.0 as a function of solution pH are shown in the inset of Fig. 4. The E0


points from pH 5 to pH 8.5 result in a best-fit straight line with a
slope of 0.076 and an intercept of 0.56 (R = 0.9916). The
0.8 0.6 0.4 0.2 0.0 -0.2 -0.4 -0.6 -0.8 experimental result of 76 mV pH1 is consistent with a 2H+-2e
E/ V transfer in the electrochemical reaction of HQ. However, the DEp
values of HQ at the PGS electrode vary with the pH of the
Fig. 3. A. Cyclic voltammograms of (a) bare GCE, (b) bare GCE + 2.7  105 mol L1
HQ, (c) PG/GCE + 2.7  105 mol L1 HQ, (d) PGS/GCE, and (e) PGS/GCE + 2.7  105
electrolyte solution. From pH 6.5 to 7.5, the DEp values are low with
mol L1 HQ. B. Cyclic voltammograms of (a) bare GCE, (f) bare GCE + 2.7  105 a pair of symmetric redox peaks, indicating a dramatic increase in
mol L1 CC, (g) PG/GCE + 2.7  105 mol L1 CC, (d) PGS/GCE, and (h) PGS/GCE + 2.7 the electron transfer rate and good reversibility of the HQ
 105 mol L1 CC. Solution temperature: 37  C. Scan rate: 75 mV s1. electrochemical reaction. However, DEp values increase with
increasing acidity or alkalinity of the electrolyte solution along
3.3. Electrochemical Behaviours of HQ and CC at the PGS Electrode with increased asymmetry in the CV peaks, indicating a slower
electron transfer rate and a more irreversible electrochemical
The cyclic voltammetry curves (CV) of HQ and CC at the bare, reaction of HQ at the PGS electrode. The influence of pH on CC peak
PG/GCE and PGS/GCE electrodes are presented in Fig. 3. A pair of potentials is similar to the results for HQ under the same test
redox peaks for both HQ and CC is observed at the three electrodes. conditions. The optimal pH range for detecting HQ/CC at the PGS
However, for the PG and bare electrodes, the CV responses are electrode thus appears to be from 6.5 to 7.5.
severely suppressed with remarkably decreased peak currents and The asymmetric voltammograms in acidic and alkaline PBS
increased DEp (DEp = Epa  Epc) (Fig. 3, curves b, c, f and g), solution can be described by the apparent charge transfer
indicating sluggish electron transfer kinetics. The CV response coefficient b. The kinetics of a redox couple described by the
current at the PGS electrode has a narrow DEp (Fig. 3, curves e and following equations is related to b [53]:
h) and is much larger than those at the bare GCE and PG electrodes,
kox = k0exp[bnf(E-E0 )]
0

demonstrating that the presence of SMWCNs in the composite film


enhances the electrochemical response and facilitates electron
transfer between HQ/CC and the PGS electrode, consistent with the
kred = k0exp[-anf(E-E0 )]
0

EIS results. The oxidation/reduction peaks at the PGS electrode


appear at +79/–35 mV for HQ (Fig. 3A, curve e) and +189/ + 75 mV where kox and kred are the rate constants for the oxidation and
for CC (Fig. 3B, curve h). The HQ and CC peaks are well-separated, reduction reactions at potential E, respectively. k is the standard
indicating simultaneous detection of CC and HQ isomers in their rate constant, b is the apparent charge transfer coefficient, a = 1-b,
mixed solution using the PGS electrode (detailed results are n is the number of electrons transferred in the redox reaction, and
0
presented in Supplementary Material, Fig. S4). The CV responses of E0 is the formal potential of the redox couple. Because of the
HQ at different scan rates (10  220 mV s1) at the PGS electrode exponential form of b in these equations, minor changes in the
are shown in Fig. S5. A pair of redox peaks is obtained in each cycle. transfer coefficient should result in a dramatic change in the
Y. Zhou et al. / Electrochimica Acta 192 (2016) 158–166 163

relative oxidation and reduction rates, and these changes may control the switching electrochemical reaction of HQ at the PGS
affect the voltammetric peak shape. Changes in the transfer electrode.
coefficient are often associated with different mechanistic steps in When PNIPAM101-b-PAAEB37 was removed from the film, the
the electron transfer process. Fig. 4 displays the voltammetric results show that there is no obvious difference in CV peak currents
symmetry differences of acidic, neutral and alkaline PBS solutions, from 24 to 40  C for the GS electrode (shown in Fig. S6), indicating
suggesting different b values and different electrochemical that the temperature-controlled electrochemical “on-off” switch-
mechanistic steps for HQ at the PGS electrode. This result is also ing behaviour of HQ at the PGS electrode may be attributed to the
consistent with the ‘scheme of squares’ model proposed by Jacq temperature-responsive phase transition behaviour of PNIPAM101-
[54]. This model was used to interpret the properties of quinone/ b-PAAEB37 at 32  C. The different phase states of PNIPAM101-b-
hydroquinone redox systems, and indicated that the electrochem- PAAEB37 in the PGS film above and below 32  C cause the film to
ical reaction pathway depends heavily on the pKa values of the display differential electrochemical catalytic activity toward HQ.
species within the model, as well as the pH of the local Tam [19] and Song [24] reported that the “on-off” switching
environment. electrochemical property at electrode surfaces modified by
stimuli-responsive polymers was related to the variable perme-
3.4. Temperature Effects on HQ/CC Electrochemical Behaviours at the ability of the modified surfaces for diffusional redox probes. The
PGS Electrode penetrative difference was caused by “swelling-shrinking” phase
transitions of stimuli-responsive polymers toward external stim-
The electrochemical responses of HQ at the PGS electrode at uli. A polymer brush in the swollen state was permeable to the
varied temperatures are shown in Fig. 5. No obvious peak currents redox probe, so the electrochemical response appeared, whereas in
are observed when the temperature of the HQ solution is below the shrunken state, the dense polymer brush blocked transport of
30  C, whereas a pair of peaks located at about +79/–35 mV is the probes to the electrode surface, and the electrochemical
obtained above 32  C. This pair of peaks belongs to the HQ response was inhibited. In our study, the temperature-based “on-
electrochemical response. This result demonstrates that the HQ off” switching behaviour of PGS surfaces is reversed from the above
electrochemical reaction is switched on above 32  C and switched result, and we are thus calling it an inverse electrochemical
off below 32  C (Fig. 5A), indicating that temperature can be used to temperature-switch. In the shrunken state of PNIPAM101-b-
PAAEB37 above 32  C, HQ displays a pair of well-defined redox
peaks (shown in Fig. 5A), indicating that the PGS electrode is active
toward HQ; in the swollen state below 32  C, the CV responses of
HQ are significantly suppressed, and the PGS electrode is inactive
(shown in Scheme 1). This opposite “on-off” switching behaviour
may be attributed to two key factors of the electrochemical activity
of the modified electrode toward redox species. One is interfacial
permeability, and the other is interfacial electron transfer rate. PGS
films possess poor permeability toward HQ because of the two-
dimensional planar structure of GO, which hinders HQ from
permeating to the GCE surface regardless of the swollen or
shrunken state of the PNIPAM101-b-PAAEB37 film. Therefore, the
electrochemical “on-off” switching activity of the PGS film is likely
related to its interfacial electron transfer rate. At T < 32  C, long
chains of PNIPAM101-b-PAAEB37 in the swollen state result in a
widened distance between GO sheets, introducing a large
interfacial electron transfer resistance, as measured by EIS

Fig. 5. A. Temperature-dependent cyclic voltammograms of HQ at the PGS


electrode in 0.1 mol L1 PBS solution (pH 7.0) from 24 to 40  C. B. The oxidation
and reduction peak current of HQ at the PGS electrode at different temperatures. Fig. 6. Differential pulse voltammetry (DPV) of HQ and CC at the PGS-modified
Scan rate: 75 mV s1; Concentration of HQ: 2.7  105 mol L1. electrode in 0.1 mol L1 PBS (pH 7.0) at temperatures from 24 to 40  C.
164 Y. Zhou et al. / Electrochimica Acta 192 (2016) 158–166

(Fig. 2B). Thus, the electrochemical responses of HQ are sup-


pressed. At T > 32  C, when PNIPAM101-b-PAAEB37 is in the
shrunken state, the distance between GO sheets shortens such
that SMWCNs embedded in the composite film can act as
conductive channels. This increases the electron transfer rate,
and the electrochemical responses of HQ are switched on. The
inverse electrochemical “on-off” switching behaviour of the PGS
film is thus attributed to the synergistic responses of PNIPAM101-b-
PAAEB37, GO and SMWCNs to temperature. When the HQ solution
temperature is raised from 24 to 30  C and from 32 to 40  C, the
peak currents show a small increasing trend (Fig. 5B), which may
be related to the increased diffusion behaviours of HQ or minor
adjustments of the PNIPAM101-b-PAAEB37 conformation upon
temperature changes in these ranges.
Fig. 6 shows differential pulse voltammograms of HQ and CC at
the PGS electrode at temperatures from 24 to 40  C. Highly
resolved and sensitive current responses of both HQ and CC are
obtained by differential pulse voltammetry (DPV) at the PGS
electrodes above 32  C, implying potential for simultaneous
detection of these compounds above 32  C by DPV. The current
responses of HQ and CC decrease sharply below 32  C and Fig. 8. Reproducibly switching “on–off” DPV responses of HQ/CC at the PGS
electrode between 26  C and 36  C for eight cycles. Concentrations of HQ and CC:
2.7  105 mol L1.

disappear completely at 24  C, similar to CV responses of HQ at


varied temperature. Therefore, the temperature of the HQ and CC
solution can also be used to control the electrochemical “on-off”
switching reaction of HQ and CC at the PGS electrode. The reasons
for the electrochemical “on-off” reaction of HQ and CC are the same
as discussed above.

3.5. Electrochemical “On-Off” Behaviours at the PGS Electrode

Because the temperature-related phase transition process of


the PNIPAM chain is completely reversible [32], we studied the
reversibility of the temperature-related HQ electrochemical “on-
off” behaviours at the PGS electrode. Two typical temperatures,
26  C and 36  C, were selected to explore this behaviour (Fig. 7A). At
36  C, HQ displays a reversible CV peak pair with relatively large
peak current, indicating that the PGS film is in an “on” state; at
26  C, the CV response is significantly reduced and can hardly be
observed, and the film is in an “off” state. Reproducibly switching
“on-off” CV responses of HQ were achieved through repeated
heating and cooling of the PBS solution between 36  C and 26  C for
eight cycles, with no substantial degradation of the peak current
observed (Fig. 7B). When HQ and CC coexist in 0.1 mol L1 PBS (pH
7.0), the reversible and repeatable switching detection of HQ and
CC at the PGS electrode can be achieved via cyclic changes of the
solution temperature between 36  C and 26  C (Fig. 8). The
electrochemical reactions of HQ and CC at the PGS electrode are
very sensitive to the environmental temperature and present
reversible “on-off” behaviours. The reversible transition of the
electrochemical “on-off” switching activities of the PGS film
demonstrates the possibility of modulating the electro-catalytic
activity by temperature. This finding can potentially be applied in
switchable sensors and switchable fuel cells.

4. Conclusions

Electrode surfaces modified by a temperature-responsive


polymer composite sensing film consisting of PNIPAM101-b-
PAAEB37, GO and SMWCNs displayed a reversible and switchable
interface property toward temperature change. CV and DPV study
Fig. 7. A. Reproducibly switching “on–off” CV response of HQ at the PGS electrode of HQ and CC at the PGS electrode indicated that reversible peak
between 26  C and 36  C, Concentration of HQ: 2.7  105 mol L1. Scan rate:
pairs were observed above 32  C, where the PGS composite film
75 mV s1. B. Dependence of CV Ipa on solution temperature switched between 26  C
and 36  C.
presented an electrochemical “switching on” state toward HQ and
Y. Zhou et al. / Electrochimica Acta 192 (2016) 158–166 165

CC. Because no obvious peaks appeared below 32  C, the film [17] X. Jiang, R. Wang, Y. Ren, J. Yin, Responsive Polymer Nanoparticles Formed by
displayed a “switching off” state. The switchable property can be Poly ether amine Containing Coumarin Units and a Poly ethylene oxide Short
Chain, Langmuir 25 (2009) 9629–9632.
attributed to the synergistic effects of all functional components of [18] E. Katz, B. Willner, I. Willner, Light-Controlled Electron Transfer Reactions at
the PGS film. Photo Isomerizable Monolayer Electrodes by Means of Electrostatic
Interactions: Active Interfaces for the Amperometric Transduction of Recorded
Optical Signals, Biosensors & Bioelectronics 12 (1997) 703–719.
Acknowledgments [19] T.K. Tam, M. Pita, O. Trotsenko, M. Motornov, I. Tokarev, J. Halámek, S. Minko, E.
Katz, Reversible Closing of an Electrode Interface Functionalized with a
Polymer Brush by an Electrochemical Signal, Langmuir 26 (2010) 4506–4513.
This research was financially supported by the NSF of China [20] Z. Yin, J. Zhang, L.-P. Jiang, J.-J. Zhu, Thermosensitive Behavior of Poly(N-
(Grants No. 21475114 and 21275123), the Program for Changjiang isopropyl acrylamide) and Release of Incorporated Hemoglobin, J. Phys. Chem.
C 113 (2009) 16104–16109.
Scholars and Innovative Research Team in University (1337304),
[21] T.K. Tam, J. Zhou, M. Pita, M. Ornatska, S. Minko, E. Katz, Biochemically
Project of Hunan provincial natural science Foundation of China Controlled Bioelectrocatalytic Interface, J. Am. Chem. Soc. 130 (2008)
(14JJ1019), and the Research fund for the doctoral program of 10888–10889.
higher education of china, Ministry of education of China [22] O. Yehezkeli, M. Moshe, R.T. Vered, Y. Feng, Y. Li, H. Tian, I. Willner, Switchable
Photochemical/Electrochemical Wiring of Glucose Oxidase with Electrodes,
(20134301110005). Analyst 135 (2010) 474–476.
[23] M.L. Dagan, E. Katz, I. Willner, Amperometric Transduction of Optical Signals
Recorded by Organized Monolayers of Photoisomerizable Biomaterials on Au
Appendix A. Supplementary data Electrodes, J. Am. Chem. Soc 116 (1994) 7913–7914.
[24] S. Song, N. Hu, On-Off Switchable Bioelectrocatalysis Synergistically
Controlled by Temperature and Sodium Sulfate Concentration Based on Poly
Supplementary data associated with this article can be (N-isopropylacrylamide) Films, J. Phys. Chem. B 114 (2010) 5940–5945.
found, in the online version, at http://dx.doi.org/10.1016/j. [25] R. Gabai, N. Sallacan, V. Chegel, T. Bourenko, E. Katz, I. Willner, Characterization
electacta.2016.01.183. of the Swelling of Acrylamidophenylboronic Acid-Acrylamide Hydrogels upon
Interaction with Glucose by Faradaic Impedance Spectroscopy,
Chronopotentiometry, Quartz-Crystal Microbalance (QCM), and Surface
References Plasmon Resonance (SPR) Experiments, J. Phys. Chem. B 105 (2001) 8196–
8202.
[26] T.K. Tam, M. Pita, M. Motornov, I. Tokarev, S. Minko, E. Katz, Modified
[1] I. Willner, M.L. Dagan, S.M. Tibbon, E. Katz, Bioelectrocatalyzed Amperometric
Electrodes with Switchable Selectivity for Cationic and Anionic Redox Species,
Transduction of Recorded Optical Signals Using Monolayer-Modified Au-
Electroanalysis 22 (2010) 35–40.
Electrodes, J. Am. Chem. Soc. 117 (1995) 6581–6592.
[27] J. Zhou, X. Lu, J. Hu, J. Li, Reversible Immobilization and Direct Electron Transfer
[2] Y. Dou, J. Han, T. Wang, M. Wei, D.G. Evans, X. Duan, Temperature-Controlled
of Cytochrome c on a pH-Sensitive Polymer Interface, Chem. Eur. J. 13 (2007)
Electrochemical Switch Based on Layered Double Hydroxide/Poly(N-
2847–2853.
isopropylacrylamide) Ultrathin Films Fabricated via Layer-by-Layer Assembly,
[28] E.S. Gil, S.M. Hudson, Stimuli-Responsive Polymers and their Bioconjugates,
Langmuir 28 (2012) 9535–9542.
Progress in Polymer Science 29 (2004) 1173–1222.
[3] J. Wang, M. Musameh, R. Laocharoensuk, O.G. García, J. Oni, D. Gervasio, Pt/Ru-
[29] S. Ahn, R.M. Kasi, S.-C. Kim, N. Sharmaa, Y. Zhou, Stimuli-Responsive Polymer
Functionalized Magnetic Spheres for a Magnetic-Field Stimulated Methanol
Gels, Soft Matter 4 (2008) 1151–1157.
and Oxygen Redox Processes: Towards On-Demand Activation of Fuel Cells,
[30] J. Zhou, G. Wang, J. Hu, X. Lu, J. Li, Temperature, ionic strength and pH induced
Electrochem. Commun. 8 (2006) 1106–1110.
electrochemical switching of smart polymer interfaces, Chem. Commun. 46
[4] A.N. Shipway E. Katz, I. Willner, Molecular Memory and Processing Devices in
(2006) 4820–4822.
Solution and on Surfaces, In: J.-P. Sauvage, V. Amendola, R. Ballardini, V.
[31] J. Zhou, J. Liu, G. Wang, X. Lu, Z. Wen, J. Li, Poly(N-isopropylacryl amide)
Balzani, A. Credi, L. Fabbrizzi, etc., Molecular Machines and Motors, Structure
Interfaces with Dissimilar Thermo-Responsive Behavior for Controlling Ion
and bonding 99 (2001) pp. 237–281.
Permeation and Immobilization, Adv. Funct. Mater. 17 (2007) 3377–3382.
[5] Y. Leroux, J.C. Lacroix, C. Fave, G. Trippe, N. Félidj, J. Aubard, A. Hohenau, J.R.
[32] S. Fujishige, K. Kubota, I. Ando, Phase Transition of Aqueous Solutions of Poly
Krenn, Tunable Electrochemical Switch of the Optical Properties of Metallic
(N-iso propylacrylamide) and Poly(N-isopropylmethacrylamide), J. Phys.
Nanoparticles, ACS Nano 2 (2008) 728–732.
Chem. 93 (1989) 3311–3313.
[6] E. Katz, R. Baron, I. Willner, Magnetoswitchable Electrochemistry Gated by
[33] Y. Katsumoto, T. Tanaka, H. Sato, Y. Ozaki, Conformational Change of Poly(N-
Alkyl-Chain—Functionalized Magnetic Nanoparticles: Control of Diffusional
isopropyl acrylamide) during the Coil-Globule Transition Investigated by
and Surface-Confined Electrochemical Processes, J. Am. Chem. Soc 127 (2005)
Attenuated Total Reflection/Infrared Spectroscopy and Density Functional
4060–4070.
Theory Calculation, J. Phys. Chem. A 106 (2002) 3429–3435.
[7] L.S.H. Ichia, Z. Cheglakov, I. Willner, Electroswitchable Photoelectrochemistry
[34] Y. Liu, Y. Liu, H. Feng, Y. Wu, L. Joshi, X. Zeng, J. Li, Layer-by-layer assembly of
by Cu2+- Polyacrylic Acid/CdS-Nanoparticle Assemblies, J. Phys. Chem. B 108
chemical reduced graphene and carbon nanotubes for sensitive
(2004) 11–15.
electrochemical immunoassay, Biosensors & Bioelectronics 35 (2012) 63–68.
[8] E. Katz, I. Willner, A Biofuel Cell with Electrochemically Switchable and
[35] C. Zhu, G. Yang, H. Li, D. Du, Y. Lin, Electrochemical Sensors and Biosensors
Tunable Power Output, J. Am. Chem. Soc. 125 (2003) 6803–6813.
Based on Nanomaterials and Nanostructures, Anal. Chem. 87 (2015) 230–249.
[9] X. Yang, F. Xiao, H. Lin, F. Wu, D. Chen, Z. Wu, A Novel H2O2 Biosensor Based on
[36] S. Kochmann, T. Hirsch, O.S. Wolfbeis, Graphenes in Chemical Sensors and
Fe3O4-Au Magnetic Nanoparticles Coated Horseradish Peroxidase and
Biosensors, Trends in Analytical Chemistry 39 (2012) 87–113.
Graphene Sheets-Nafion Film Modified Screen-Printed Carbon Electrode,
[37] G.A. Rivas, M.D. Rubianes, M.C. Rodríguez, N.F. Ferreyra, G.L. Luque, M.L.
Electrochim. Acta 109 (2013) 750–755.
Pedano, S.A. Miscoria, C. Parrado, Carbon nanotubes for electrochemical
[10] Y. Hu, Z. Zhang, H. Zhang, L. Luo, S. Yao, Selective and Sensitive Molecularly
biosensing, Talanta 74 (2007) 291–307.
Imprinted Sol-Gel Film-Based Electrochemical Sensor Combining
Mecaptoacetic Acid Modified PbS Nanoparticles with Fe3O4@Au-Multi-Walled [38] Y. Wang, J. Lu, L. Tang, H. Chang, J. Li, Graphene Oxide Amplified
Carbon Nanotubes-Chitosan, J. Solid State Electrochem. 16 (2012) 857–867. Electrogenerated Chemiluminescence of Quantum Dots and Its Selective
[11] M. Arvand, M. Hassannezhad, Magnetic Core-Shell Fe3O4@SiO2/MWCNT Sensing for Glutathione from Thiol-Containing Compounds, Anal. Chem. 81
Nanocomposite Modified Carbon Paste Electrode for Amplified (2009) 9710–9715.
Electrochemical Sensing of Uric Acid, Mater. Sci. Eng. C 36 (2014) 160–167. [39] D. Chen, H. Feng, J. Li, Graphene Oxide: Preparation, Functionalization, and
[12] Y. Ding, X. Ye, G. Zhang, Microcalorimetric Investigation on Aggregation and Electrochemical Applications, Chem. Rev. 112 (2012) 6027–6053.
Dissolution of Poly(N-isopropylacrylamide) Chains in Water, Macromolecules [40] J. Li, X. Hong, Y. Liu, D. Li, Y.-W. Wang, J.-H. Li, Y.-B. Bai, T.-J. Li, Highly
38 (2005) 904–908. Photoluminescent CdTe/Poly(N-isopropylacrylamide) Temperature-Sensitive
[13] A.C.W. Lau, C. Wu, Thermally Sensitive and Biocompatible Poly(N-vinyl Gels, Adv. Mater. 17 (2005) 163–166.
caprolactam): Synthesis and Characterization of High Molar Mass Linear [41] Y. Zhang, Y. Wen, Y. Liu, D. Li, J. Li, Functionalization of single-walled carbon
Chains, Macromolecules 32 (1999) 581–584. nanotubes with Prussian blue, Electrochem. Commun. 6 (2004) 1180–1184.
[14] S. Sanjuan, Y. Tran, Stimuli-Responsive Interfaces Using Random [42] J.-O. You, D.T. Auguste, Conductive, Physiologically Responsive Hydrogels,
Polyampholyte Brushes, Macromolecules 41 (2008) 8721–8728. Langmuir 26 (2010) 4607–4612.
[15] S. Samanta, J. Locklin, Formation of Photochromic Spiropyran Polymer Brushes [43] A. Riedinger, M.P. Leal, S.R. Deka, C. George, I.R. Franchini, A. Falqui, R.
via Surface-Initiated, Ring-Opening Metathesis Polymerization: Reversible Cingolani, T. Pellegrino, Nanohybrids Based on pH-Responsive Hydrogels and
Photocontrol of Wetting Behavior and Solvent Dependent Morphology Inorganic Nanoparticles for Drug Delivery and Sensor Applications, Nano Lett.
Changes, Langmuir 24 (2008) 9558–9565. 11 (2011) 3136–3141.
[16] A. Garcia, M. Marquez, T. Cai, R. Rosario, Z. Hu, D. Gust, M. Hayes, S.A. Vail, C.-D. [44] E. Katza, I. Willner, A Quinone-Functionalized Electrode in Conjunction with
Park, Photo-, Thermally, and pH-Responsive Microgels, Langmuir 23 (2007) Hydrophobic Magnetic Nanoparticles Acts as a Write-Read-Erase Information
224–229. Storage System, Chem. Commun. 45 (2005) 5641–5643.
166 Y. Zhou et al. / Electrochimica Acta 192 (2016) 158–166

[45] X. Zhao, Y. Liu, J. Lu, J. Zhou, J. Li, Temperature-Responsive Polymer/Carbon [50] W.S. Hummers Jr., R.E. Offeman, Preparation of Graphitic Oxide, J. Am. Chem.
Nanotube Hybrids: Smart Conductive Nanocomposite Films for Modulating Soc. 80 (1958) 1339.
the Bioelectrocatalysis of NADH, Chem. Eur. J. 18 (2012) 3687–3694. [51] L. Lu, H. Zhang, N. Yang, Y. Cai, Toward Rapid and Well-Controlled Ambient
[46] A. Kaushik, R. Kumar, S.K. Arya, M. Nair, B.D. Malhotra, S. Bhansali, Organic- Temperature RAFT Polymerization under UV–vis Radiation: Effect of Radiation
Inorganic Hybrid Nanocomposite-Based Gas Sensors for Environmental Wave Range, Macromolecules 39 (2006) 3770–3776.
Monitoring, Chem. Rev. 115 (2015) 4571–4606. [52] I. Tokarev, M. Orlov, E. Katz, S. Minko, An Electrochemical Gate Based on a
[47] L. Li, S. Xu, Z. Du, Y. Gao, J. Li, T. Wang, Electrografted Poly(N-mercaptoethyl Stimuli-Responsive Membrane Associated with an Electrode Surface, J. Phys.
acrylamide) and Au Nanoparticles-Based Organic/Inorganic Film: A Platform Chem. B 111 (2007) 12141–12145.
for the High-Performance Electrochemical Biosensors, Chem. Asian J. 5 (2010) [53] T.G. Strein, A.G. Ewing, Laser Activation of Microdisk Electrodes and Digital
919–924. Examined by Fast-Scan Rate Voltammetry and Digita Simulation, Anal. Chem
[48] A. Walcarius, Electrochemical Applications of Silica-Based Organic-Inorganic 66 (1994) 3864–3872.
Hybrid Materials, Chem. Mater. 13 (2001) 3351–3372. [54] Jacq Jean, Square scheme. Establishment and Discussion of the General
[49] A.-M.J. Haque, H. Park, D. Sung, S. Jon, S.-Y. Choi, K. Kim, An Electrochemically Equation for the Current Density-Potential Curve in the Stationary and
Reduced Graphene Oxide-Based Electrochemical Immunosensing Platform for Convective Diffusion Regime. J. Electro, Anal. Chem 29 (1971) 149–180.
Ultrasensitive Antigen Detection, Anal. Chem 84 (2012) 1871–1878.

You might also like