Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Sensors and Actuators B 256 (2018) 976–991

Contents lists available at ScienceDirect

Sensors and Actuators B: Chemical


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

Review

Graphene-based devices for measuring pH


P. Salvo a,b,∗ , B. Melai b , N. Calisi b , C. Paoletti b , F. Bellagambi b , A. Kirchhain b ,
M.G. Trivella a , R. Fuoco b , F. Di Francesco b
a
Institute of Clinical Physiology, National Research Council, Via Moruzzi 1, 56124, Pisa, Italy
b
Department of Chemistry and Industrial Chemistry, University of Pisa, Via Moruzzi 13, 56124, Pisa, Italy

a r t i c l e i n f o a b s t r a c t

Article history: pH measuring and monitoring is fundamental to understand or control many chemical processes in bio-
Received 17 March 2017 logical, industrial or environmental fields. Potentiometric measurements by a glass electrode is the most
Received in revised form 4 October 2017 common method to measure pH, although single-use paper strips are also widely used. Other methods
Accepted 6 October 2017
include the use of hydrogen, quinhydron, and antimony electrodes, the imaging using pH-sensitive indica-
Available online 12 October 2017
tors such as dyes or proteins, and the use of ion-selective field effect transistor (ISFET). Due to the chemical
reactivity of both sides of its 2D structure, nanometer thickness, high electron mobility, high reactivity
Keywords:
to oxygen groups such as OH− , and ultrafast optical response, graphene has the potential to be used for
Graphene
pH
the fabrication of nanoscale, wide-range, high-sensitivity and flexible pH sensors. This review describes
ISFET how graphene, graphene oxide and reduced graphene oxide can be used to fabricate pH-sensitive devices
Quantum dots (e.g. solution-gated FETs, solid-gate FETs, electrochemical sensors, and pH-sensitive quantum dots). The
Electrochemical sensors various configurations are reported along with the advantages and current limitations.
© 2017 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 977
2. Graphene-based pH sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 977
2.1. Solution-gated ISFETs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 977
2.2. Solution-gated ISFETs on a flexible substrate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 981
2.3. Solid gate and three dimensional FETs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 981
2.4. Graphene quantum dots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 982
3. Graphene oxide-based pH sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 983
3.1. Optical GO-based pH sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 984
3.2. Electrochemical GO-based pH sensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 986
4. Reduced graphene oxide-based pH sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 986
4.1. rGO FETs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 987
5. Conclusions and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 987
Competing financial interests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 989
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 989
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 989
Biographies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 991

∗ Corresponding author at: Institute of Clinical Physiology, National Research Council, Via Moruzzi 1, 56124, Pisa, Italy.
E-mail address: pietro.salvo@gmail.com (P. Salvo).

https://doi.org/10.1016/j.snb.2017.10.037
0925-4005/© 2017 Elsevier B.V. All rights reserved.
P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991 977

1. Introduction quencies. Graphene has also a large theoretical specific surface


area (2630 m2 g−1 ), and high Young’s modulus (∼1.0 TPa), ther-
pH governs the life of many living systems and chemical mal conductivity (∼5000 W m−1 K−1 ) and optical transmittance
reactions in nature. Biology and microbiology (e.g. monitoring (∼97.7%). Graphene can be a promising candidate for pH sens-
of bacteria, DNA, enzymes, aminoacids and cells), environmen- ing for its high reactivity to oxygen bearing groups such as OH-
tal protection (e.g. characterization of waste and sea water), food and a better electrochemical response than glassy carbon elec-
and beverage quality control (e.g. monitoring of fermentation trodes, graphite and CNTs. Furthermore, optical sensing can also
processes, characterization of fruit, drinking water and vegeta- benefit from the use of graphene, which is a highly efficient fluores-
bles), pharmaceutical, cosmetics (quality control of creams, gels cence quencher [34,37,38]. The first section of this review focuses
and shampoos), medicine (e.g. monitoring and analysis of culture on the chemico-physical properties and mechanisms that underlie
media, cancer, blood and cytoplasm) and industry (e.g. chemical graphene’s sensitivity to pH as well as pH sensors exploiting such
baths for paints and plating) are some of the fields where pH sensors properties.
have been extensively used to understand the nature of chemical Graphene oxide (GO) resulting from the oxidation of graphite
processes, as well as to monitor quality and control safety [1–6]. is the precursor of reduced graphene oxide (rGO), i.e. graphene
pH is defined as the negative decimal logarithm of the H+ activ- obtained by the exfoliation and reduction of GO [39–41]. GO sensi-
ity, which approaches the H+ concentration in diluted solutions. tivity to pH depends on the surface hydroxyl (OH), carboxyl (COOH)
The basics of pH theory and measurement are reported in several and epoxy groups (COC) that protonate/de-protonate when pH
publications [7–9]. changes: GO and its applications for pH sensing are discussed in
Although paper test strips and glass electrodes remain the most the second section of this review. Although reduced graphene oxide
widespread pH sensors, many studies have focused on developing shares many of graphene’s physical and chemical properties, there
less fragile, miniaturized, and biocompatible sensors with higher are differences such as the presence of defects and residual oxy-
sensitivity. In fact, the subjective reading of pH with paper strips is gen atoms not removed by reduction [37] that justify discussion
not precise, whereas the glass electrode is fragile, not suitable for in a separate section. In the literature, graphene, GO and rGO are
miniaturization and not particularly accurate for strongly basic or often mixed with other compounds to obtain pH sensitive mate-
acidic media [10]. rials, mainly to improve the mechanical structure or the electrical
The pH value can be measured through various approaches, conductivity of the resulting composites.
e.g. potentiometry, optical measurements, and ion-selective field This review only reports devices where graphene, GO or rGO are
effect transistors (ISFETs). ISFETs are currently the most effective used as the main pH sensitive element. We provide an overview of
alternative to glass electrodes, although scientific literature reports research efforts in the last decade and trends for the development
problems such as amplified noise due to the high input impedance and enhancement of graphene-based pH sensors.
and poor reproducibility. However, stable and low drift pH-ISFETs
®
suitable for long use are on the market, such as the DuraFET sensor
®
by Honeywell. DuraFET ’s sensing element is only a few millime- 2. Graphene-based pH sensors
tres wide and is suitable for industrial applications [11]. At the
frontier of research, the nanometre-size quantum dots could be The main application of graphene as pH sensitive material is
complementary to ISFETs as pH sensors. A multi-purpose and high in pH-ISFET because this configuration offers several advantages
performance micro/nano sensor is the Holy Grail for pH measure- compared with a glass electrode. A pH-ISFET is smaller and, in the
ment and the search for the most suitable sensitive materials is same conditions of use, lasts longer than a glass electrode. The pack-
currently ongoing. age of a pH-ISFET is more robust than a glass electrode, thus more
Despite the high number of materials and devices reported durable and less prone to break. The glass electrode needs to be
in the literature, only a few have been marketed. The most chal- stored wet and the bulb kept with a sufficient amount of electrolyte
lenging hurdles are the degradation of materials over time, low solution. Conversely, a pH-ISFET can be stored dry and does not
sensitivity, drift and difficulties in mass production. Illustrative pH- include any bulb-like part that needs to be hydrated with an elec-
sensitive materials reported in the literature include polyaniline trolyte solution. However, especially for biological measurements
(PANI), polypyrrole (PPy), polyparaphenylenediamine, poly(vinyl such as the pH of cells, the graphene quantum dots promise to be a
pyridine), poly(methacrylic acid) (PMAA), poly(acrylic acid) (PAA), reliable option because of their small size and non-contact (optical)
palladium oxide, aluminium oxide, riboflavin, anthraquinone- measurement.
sulfonate, fluorescent dyes such as rhodamine, silicates and
silanes, 6-carboxyfluorescein, as well as seminaphthorhodafluor
(SNARF)/seminaphthofluorescein (SNAFL), hydrogels such as 2.1. Solution-gated ISFETs
hydroxyethyl methacrylate/dimethylaminoethyl methacrylate
(HEMA-DMAEMA) and poly(vinyl alcohol)/poly(acrylic acid) A typical configuration of a graphene-based ISFET is the
(PVA-PAA) [12–31]. solution-gated FET (SGFET), where the graphene is used as the
After the isolation of graphene from highly oriented pyrolytic conduction channel. The SGFET resembles a three-electrode elec-
graphite in 2004 at the University of Manchester [32], many trochemical cell, where the electrolyte solution is in direct contact
researchers started to investigate how to exploit the extraordi- with graphene (working electrode), the gate voltage (VG ) is applied
nary electrical, chemical, mechanical and optical properties of this through a reference electrode (Ag/AgCl, calomel or Pt), and a Pt
material for sensing and biosensing [33–36]. Graphene is a single counter electrode completes the device. Although the counter
atomic plane of sp2 -bonded carbon atoms. The electrical proper- electrode is often absent, this top-gated FET with Ti/Au or Cr/Au
ties and the high electron mobility (200,000 cm2 V−1 s−1 ) make contacts for source and drain is a common configuration for pH
graphene an ideal material for high-performance transistors. As measurement (Fig. 1a). The most common methods for fabricat-
the gate length is reduced, the short channel effects (e.g. punch- ing the graphene channel are the exfoliation of graphite and the
through, velocity saturation, impact ionization, hot electron effect synthesis by chemical vapour deposition (CVD) [42–48]. After CVD
and scattering) degrade the transistor performances. However, on a substrate such as copper, the graphene film is coated with
these effects could be reduced by the monoatomic thickness and a protective material, e.g. poly(methyl methacrylate) (PMMA) or
high carrier mobility of graphene, and lead to high cut-off fre- poly(dimethylsiloxane) (PDMS), then the substrate is etched and
978 P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991

Fig. 1. a) Three-electrode configuration of a pH-measuring FET. Although the most common configuration is with VG applied through a Ag/AgCl reference electrode and no
counter electrode, this ISFET provides the most accurate measurement; b) example of the VDirac shift when pH increases. The area to the left of VDirac is associated with an
excess of holes in the graphene channel, and with an excess of electrons to the right. The determination of the VDirac allows the measurement of pH; c) Distribution of ions
and donors (holes or electrons) at the graphene-electrolyte solution at basic and acid pH.

the film is transferred onto the ISFET substrate, e.g. SiO2 /Si or 6H- sodium dodecyl sulfate (SDS) and, after sonication, the dispersion
SiC. was poured onto a paper substrate through a metal mask under
In 2008, Ang et al. proposed an SGFET to measure pH includ- vacuum. After rinsing with deionized water and drying at 80 ◦ C for
ing an epitaxial graphene layer grown on 6H-SiC [49]. Ang et al. 20 min, six U-shaped graphene sensors were fabricated. A paraffin
hypothesized conductivity (␴) of the graphene channel to be modu- pen was used to define a closed measurement area around each
lated by the adsorption of hydroxyl (OH− ) and hydroxonium (H3 O+ ) sensor. The highest fabrication repeatability and electrical stability
ions onto the surface. They used a VG ∈ [−1, 1] V and a constant were obtained by depositing 0.5 mg of graphene dispersion through
drain–source voltage (VDS ) of −1 V, and observed that OH− bound vacuum filtration, which resulted in a sensing layer thickness of
at the inner Helmholtz plane of the graphene/solution interface about 100 ␮m. The response was measured from pH 1 to about
attracted their counterions in graphene in the case negative VG , 10.5 by dispensing 10 ␮L drops of different buffer solutions in each
thus inducing a p-type doping. On the other hand, n-type dop- of the six measurement areas. At VDS ∈ [−1, 1] V, the sensitivity
ing was induced by the H3 O+ adsorption in the case of positive was about 30.8 /pH and the correlation between resistance and
VG or acid pH. As pH increased from 2 to 12, the Dirac voltage pH resulted in a regression line with a coefficient of determination
(VDirac , i.e. the voltage at which the number of holes and electrons R2 ∼= 0.93.
is balanced and the channel conductance is minimum) showed a Following the same structure as Fig. 1a, Ohno et al. described a
positive shift (Fig. 1b, c), thus allowing the pH to be measured. Sur- FET with a single layer of graphene and a leakage current lower than
prisingly, for negative VG and 3–4 graphene layers, a sensitivity of 10 nA [52]. The conductance was measured for two FET configura-
about 99 mV/pH was found, which is greater than the Nernst limit tions: top-gate and back-gate, where the back-gate is an n-doped
of 59.16 mV/pH. Si substrate. The top-gate capacitance was three times greater than
Lei et al. fabricated a chemiresistor, i.e. a gate-free FET, with in the back-gate configuration, therefore confirming the top-gate
a graphene sheet exfoliated from graphite and deposited onto a as the preferred option for sensing. The VDirac positively and lin-
SiO2 /Si substrate [50]. For a constant current of 10 ␮A and an initial early shifted at increasing pH from 4 to 8.2. For VG ∈ [−50, 50]
resistance of 86 k, the sensitivity of the device from pH 4 to 10 mV, the sensitivity was about 30 mV/pH. The authors pointed out
was approximately 2.13 k/pH. The resolution was calculated as that the VDirac may shift at different pH values because of charged
the ratio between the maximum standard deviation of resistance impurities scattering, which can be caused, for example, by the
for multiple measurements and the sensitivity of the device. The residual resist used in the photolithographic fabrication process,
resolution at basic pH and acid pH were 0.33 and 0.97, respectively. or graphene defects.
This behaviour may be explained assuming that the OH− are more Ristein et al. used a monolayer of graphene on a 6H–SiC single
ordered on the inner Helmholtz plane than the H3 O+ . This sensor crystal [53]. The solution-gated FET (SGFET) had a leakage current
was reusable despite a hysteresis of about 2 k. The authors sug- lower than 1 nA and linear I–V characteristics up to |VDS | = 0.1 V. The
gested that reusability was probably related to the easy removal drain-source current (IDS ) versus VG was measured at VDS = −50 mV
of electrolytes from graphene due to its extremely low thickness, from pH 3 to 12. The VDirac shift per unit pH, i.e. the sensitivity,
although the paper did not report any precise value or further was 19 ± 1 mV/pH. The authors compared the conductivity of the
explanation. Lee et al. proposed another example of a gate-free graphene channel (3.5 × 10−5 S) with the value reported by Ang
FET, this time on a paper substrate [51]. Graphene was dispersed in et al. (3.5 × 10−4 S) [49] since their transistors were similar. The
P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991 979

Table 1 face. They suggested pH measurements should only be performed


Concentrations and ionic strengths of the solutions used to determine the effects of
in diluted solutions, as ionic charge screening is more significant
different electrolytes. The buffers were obtained in 0.05 M phosphate solutions, and
0.05 M potassium was present in chloride and sodium solutions (adapted from Ref. at high concentrations. In fact, their tests with five and ten times
[55]). diluted sodium solutions yielded the typical positive VDirac shift for
increasing pH, whereas the sensor increased its sensitivity from −7
Electrolyte pH Concentration (M) Ionic strength (M)
to −27 mV/pH in a 0.1 M potassium solution.
6 0.0056 0.1224 Heller et al. expanded this research by using different KCl
Na+ 7 0.0279 0.2116
concentrations buffered at pH 7 and demonstrating a linear depen-
8 0.0463 0.2852
dence between the VDirac shift and the log(KCl) concentration with
6 0.0556 0.1224
respect to the VDirac normalized to its value at 1 M KCl in the
K+ 7 0.0779 0.2116
8 0.0963 0.2852
range [1 mM–1 M] [56]. Similar results were obtained with LiCl
at pH 3. At high ionic strength, the double layer capacitance at
6 0.0444 0.2112
the interface graphene-buffer increases and the surface potential
Cl− 7 0.0221 0.2558
8 0.0037 0.2926 decreases. At pH 7, the VDirac shift changed by −42.7 mV/decade
with the buffer concentration. At pH 3, there was a sign inversion
(+18.9 mV/decade), probably because of the supporting SiO2 layer.
difference of one order of magnitude in conductivity was probably In fact, SiO2 has ionized silanol and silylamine groups, such as SiO− ,
because Ang et al. tested their SGFET at VG > 0.3 V, which Ristein SiOH2 + and SiNH3 + , which can negatively or positively charge the
et al. found to be the maximum voltage that would prevent defects surface. The effect of different ions on the measurement of pH was
in the graphene layer. These defects probably caused the formation confirmed by testing Li+ , K+ and tetraethylammonium (TEA+ ), but
of holes, which then led to a super-Nernstian sensitivity. an explanation of this effect has still to be found. However, Heller
Cheng et al. etched a SiO2 substrate to suspend a graphene layer et al. hypothesized that the ionic radius or a specific adsorption
shorter than 1 ␮m between the Cr/Au contacts of drain and the onto graphene could play a major role.
source [54]. The resulting FET led to a VDirac of 0.1 ± 0.2 V and to a Fu et al. pointed out that the range of graphene sensitivities to
larger conductivity at negative VG (majority of holes) compared to pH reported in the literature was very large (12–99 mV/pH) [57].
positive VG (majority of electrons). Compared with a standard con- They thus set up an experiment with the three-electrode configura-
figuration, there was a net improvement in the FET characteristics tion shown in Fig. 1a. Graphene was exfoliated or grown on copper
with the suspended graphene: the SNR increased by 14 dB at fre- by chemical vapour deposition and then transferred to the SiO2 /Si
quencies below 1 kHz, and the sensitivity increased 1.5 and 2 times substrate. For VDS ∈ [10, 50] mV and VG ∈ [0.1, 0.8] V, the leak-
for holes and electrons, respectively, leading to a symmetric ␴-VG age current (1 nA) was negligible compared with the measured IDS .
curve. These improvements were ascribed to a reduced scattering, Although the VDirac shifted positively at increasing pH (as in previ-
but a full explanation was not supplied. This FET had a sensitivity ous works), the sensitivity was only 6 ± 1 mV/pH from pH 5 to 10.
of about 17 mV/pH when tested with buffer solutions from pH 6 to The hydrophobicity of the graphene surface was then increased by
9 and VG ∈ [−50, 50] mV. adding fluorobenzene for 30 s and drying the FET. In this case, the
Studies on the effects of different electrolyte solutions have sensitivity was almost zero from pH 4 to 10. On the other hand,
further contributed to the comprehension of the mechanisms reg- when an Al2 O3 layer thinner than 2 nm was deposited (atomic
ulating the graphene response to pH. Lee et al. investigated the layer deposition, 100 ◦ C) on graphene, the sensitivity increased to
VDirac shift when an SGFET was exposed to sodium hydroxide 17 mV/pH from pH 3 to 8. According to Fu et al., OH functional
(NaOH), hydrochloric acid (HCl) and potassium hydroxide (KOH) groups from Al2 O3 turned graphene partially hydrophilic. Their
[55]. A graphene film was grown onto a copper film in a furnace conclusion was that a defect-free graphene is not able to sense
at 1000 ◦ C in presence of a mixture of CH4 and H2 . PMMA dis- pH, and that graphene response to pH depends on the presence of
persed in chlorobenzene was used to coat and transfer the graphene imperfections, such as the hydroxyl and carbonyl groups. The sen-
layer onto a Si/SiO2 wafer after copper etching. The wafer had pre- sitivities of 99 and 20 mV/pH reported in [49] and [53], respectively,
patterned source, drain and Cr/AU electrodes. PMMA was removed probably depended on a high leakage gate current.
by rinsing the device with a solution of acetone and deionized Kwon et al. investigated the role of defects in graphene mesh
water, whereas graphene was patterned with standard photolitho- FETs (GM-FETs) [58]. The graphene mesh structure was synthesized
graphic techniques. Solutions at different pH values in the range using silica spheres as a growth mask to avoid photolithographic
[6–8] were obtained with different concentrations of Na+ , Cl− patterning, thereby reducing the disorder and contamination
and K+ in 0.05 M phosphate solutions (Table 1). While the SGETs of carbon atoms. A hexagonal close-packed monolayer of silica
reported in other works had positive VDirac shifts at increasing pH, spheres covered a copper foil as well as the graphene mesh aroused
Lee et al.’s transistor showed a negative VDirac shift which depended from the dissociation of carbon atoms at the copper–silica interface
on the ionic composition and strength of the buffer solutions. At (Fig. 2a). For FET comparison, a single layer graphene was grown
VDS = 50 mV, VG ∈ [0, 1] and for pH 6, 7 and 8, the pH sensitivities onto a copper by CVD (Fig. 2b). The resulting mesh was a hexagonal
were −78, −38 and −7 mV/pH in buffer solutions containing dif- array of circular holes with an average diameter and interspacing
ferent concentrations of Na+ , Cl− and K+ respectively, whereas the of 300 and 600 nm, respectively. At different pH values from 6.55
sensitivity was +69 mV/pH in reference buffers (0.05 M). Therefore, to 8.25 and with VDS = 50 mV, the single layer graphene FETS and
a negative shift for increasing pH was observed in the lab-made GM-FETs had a sensitivity of about 16 and 90 mV/pH, respectively.
buffers containing the electrolytes and a positive shift in commer- The authors suggested that the GM-FETs super-Nernstian sensitiv-
cial reference buffers. This result could not be explained in terms ity could depend on a high number of unsaturated carbon atoms
of ionic strength as the response for sodium and potassium buffers at edge defects, which could adsorb or probably bind the H+ ions.
were very different, although the ionic strengths were identical. Although the formation of covalent bonds improved sensitivity, the
The type of ion played an important role, as the VDirac shift was irreversibility of this reaction could limit the use of GM-FETs. In fact,
more sensitive to the concentration of Na+ than to the concentra- sensitivity decreased to about 7 mV/pH after five cycles of 3 min
tions of Cl− or K+ . Lee et al. hypothesized that the shift in VDirac was tests of the GM-FETs in solutions at pH 8.25, 7.85, 7.40, 6.95, and
a consequence of the ionic charge screening, which depends on ion 6.55. Therefore, it is essential to control graphene defects in order
size, degree of hydration, and affinity towards the graphene sur- to obtain reproducible and reliable results.
980 P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991

Fig. 2. Schematics (top) and SEM (bottom) images of a) graphene mesh and b) single-layer graphene (b). By dissociation of carbon atoms at the copper–silica interface, a
graphene mesh was grown on a copper foil covered with a hexagonally close-packed monolayer of silica spheres (a), whereas a single-layer graphene was grown by CVD on
a copper foil (b) (Adapted with permission from [58]. Copyright 2016 American Chemical Society); c) Microscopic image of graphene FET after EBL edge patterning (left);
atomic force microscopre (AFM) image to show graphene nanoribbons (right) (Adapted with permission from [59]. Copyright 2013 American Chemical Society); d) Field
effect transistor with a solid gate under a HfO2 layer; e) Three dimensional field effect transistor where a graphene foam coated with a 20 nm HfO2 layer is the pH-sensitive
material (Adapted with permission from [69]).

As the edges of graphene are chemically reactive, Tan et al. pat- tance at the interface between graphene and an electrolyte solution
terned the graphene layer by electron-beam lithography (EBL) and is given by the sum of the reciprocals of the Helmholtz layer capac-
oxygen plasma into four columns of 99.1 ± 1.5 nm wide and 5.5 ␮m itance (CH ) and the space-charge capacitance of graphene (CG ).
long nanoribbons (GNRs) to improve the number of hydroxyl The typical assumption is that CH » CG so that the Mott-Schottky
groups (Fig. 2c) [59]. With a constant VDS = 10 mV, sensitivity equation becomes:
improved after patterning from about 6.5 to 23.6 mV/pH in the
1 2
 kT

pH range [6–8]. One advantage of the edges defects is that unlike = V − VFB − (1)
CG2 εε0 eN e
defects in the basal plane they do not affect the current flow in the
graphene channel. Tan et al. also found that the improvement did where ε is the dielectric constant of graphene, ε0 is the vacuum per-
not depend on impurities such as residual PMMA used in the pho- mittivity, e is the electron charge, N is the donor density (holes or
tolithographic process, since similar results were obtained when electrons), V is the applied potential, VFB is the flat-band potential
hydrogen silsesquioxane (HSQ) was used as the photoresist. (i.e. the potential at which the conduction and the valence bands
Zhang et al. combined the ideas exposed in [54] and [59] in order of graphene are flat), k is the Boltzmann constant, and T the tem-
to improve the graphene sensitivity to pH and fabricate a suspended perature. When pH changed from 2 to 10, VFB = VFB (pH = 10) −
GNR ISFET for measuring pH and cancer markers in solution [60]. VFB (pH = 2) for pristine graphene was much lower (0.1 V) than for
In this study, the width of GNRs was 50 nm and the GNR ISFET was the anodized graphene (0.41 V). This result was reflected in a pH
powered with VDS = 1 V, whereas VG ∈ [−1.8, 1.8] V. The IDS was sensitivity of 12.5 mV/pH for pristine graphene and 51.3 mV/pH for
normalized with respect to the current I0 at pH 9 and presented as anodized graphene. The better performance of anodized graphene
the ratio (IDS − I0 )/I0 . From pH 5 up to 9, the suspended GNR FET was thus depended on the presence of the hydroxyl and carboxylic
1.5 times more sensitive (about 25 mV/pH) than an unsuspended groups.
GNR FET and a normal SGFET. However, as reported in [53], those Zaifuddin et al. presented an FET that was p-doped in the range
voltages could lead to the formation of defects capable of altering VG ∈ [−0.2 to 0.2] V and with a leakage current lower than 1 nA
the measurement. [63]. A monolayer graphene was obtained by the decomposition of
The electrochemical characteristics of crystalline epitaxial CH4 in a CVD reactor onto a Cu/SiO2 /Si substrate. After etching the
graphene grown on SiC were studied in [61] using a three-electrode copper with FeCl3 , graphene was transferred onto SiO2 /Si and an
electrochemical cell consisting of a graphene coated working elec- array of FETs were fabricated using standard microlithography. At
trode, an Ag/AgCl reference electrode with saturated KCl and a Pt VG = 0 V and VDS = 0.1 V, the sensitivity to pH versus time was tested
wire as counter electrode. Graphene was anodized at 2 V in a pH 7 for four FETs with different buffer solutions and, normalizing to the
phosphate buffer solution to introduce oxygenated surface groups current value at pH 4, there was an IDS shift of about 3% and 10%
such as C O and C O H. X-ray photoelectron spectroscopy (XPS) at pH 5.1 and 8.2, respectively. One of the four FETs had smaller
established that long anodization times (e.g. 500 s) led to the func- current changes than the other FETS and the authors suggested
tionalization of the edges with hydroxyl and carboxylic groups. The that this result could depend on the chemical residues left during
hydroxyl and carboxylic groups increased the capacitive charging the transfer process, but the type of resist was not reported.
current, thus fastening the electron transfer kinetics. The sensitiv- Lin et al. used CVD on copper and nickel foils to grow graphene,
ity to pH was determined with the Mott-Schottky equation, which which was then suspended on an array of Cr/Pb sources and drains
is commonly used to study the electrochemical interface between onto a SiO2 /Si substrate [64]. The channel width was 20 ␮m. The
a semiconductor and an electrolyte solution [62]. The total capaci- graphene grown onto nickel (GrNi ) was a continuous film with a
P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991 981

low number of defects and consisted of less than 10 layers, whereas were almost equal, the authors concluded that the substrate does
the graphene grown on copper (GrCu ) was a quasi-continuous film not influence the pH response. The PEN SGFET showed a good
consisting of a single-crystal monolayer. Since IDS had an exponen- reversibility over repeated pH measurement cycles (decreasing and
tial decay over time, the response was measured as (IDS -IDSin )/IDSin , increasing ramps), but hysteresis was present. Mailly-Giacchetti
where IDSin was the initial value. The GrNi was tested at pH 7, 10, 11 et al. also studied the effect of the surface transfer doping due to
and 13 and showed a linearly increasing response for increasing pH the exchange of electrons between the adsorbed ions and graphene.
(4.75% at pH 7, 45.1% at pH 13, and 0.1 pH resolution). At VG = 0 V The electron transfer reaction rate of graphene was decreased by
and VDS = 0.1 V, the IDSin was not constant at different pH values fabricating a SGFET on octadecyltrichlorosilane (OTS), as described
(e.g. about 250 ␮A at pH 7 and 1.5 mA at pH 13), thus indicating an in [67]. The OTS-SGET sensitivity was 18 mV/pH, comparable with
irreversible adsorption of ions on graphene. The GrCu was used at those of SiO2 /Si and PEN SGFETs, thus suggesting a negligible sur-
pH 7.35, 7.59 and 7.86 with sensitivities of 34.5%, 47.6%, and 57.4%, face transfer doping.
respectively. The resolution was 0.01 pH. However, their paper does
not report the number of measurements nor the error bars on plots.
2.3. Solid gate and three dimensional FETs
The adsorption of ions onto a graphene surface and desorption
have been studied via the effect on graphene conductivity. Kiani
SGFETs are characterized via a direct application of VG to the
et al. proposed an analytical model to explain and predict the effect
electrolyte solution, which is the gate dielectric. However, this
of electron exchange for a monolayer GNR [65]. The model used
implies that any variability of the solution can affect the sensor
a typical ISFET configuration with gold contacts, SiO2 /Si substrate
response. The fabrication of a solid gate onto the SiO2 substrate with
and an Ag/AgCl reference electrode. The conductance (G) of a GNR
a dielectric layer sandwiched between this gate and the graphene
can be written as:
 channel (Fig. 2d) is another FET configuration that improves the
3q2 3a3 t 3 kB T   integration and microfabrication, as no cumbersome gate electrode
G= F−1/2 () + F−1/2 (−) (2) (compared with the transistor dimensions) is needed.
hl
Zhu et al. proposed a solid gate FET with HfO2 as dielectric for pH
where q is the electron charge, a is the graphene lattice constant ∂IDS
(0.246 nm), t = 2.7 eV is the tight-binding energy for the nearest sensing [68]. For a FET, the transconductance gm = expresses
∂VG
neighbour C C atoms, kB is the Boltzmann constant, T is the tem- the sensitivity and is directly proportional to the capacitance of the
perature, h is the Planck constant, l is the graphene channel length, dielectric (Cox ). Since HfO2 is a high-dielectric constant material
 = (EF − Eg /kB T) is the normalized Fermi energy (EF is the Fermi level (k ≈ 25), a high capacitance (i.e. Cox (HfO2 ) = k tεoxo ≈ 1.1 ␮F/cm2 ) is
and Eg is the band gap energy), and F−1/2 is Fermi-Dirac integral of obtained for a thickness (tox ) of just 20 nm. The choice of a high-k
order −1/2. The conductance at different pH (GpH ) was modelled as gate dielectric allows the thickness of the dielectric and the leakage
P
GpH = pH G. current to be reduced, thus improving sensitivity. With a VG from
The parameter P is the pH sensing factor and, at T = 25 ◦ C, can 0.6 to 1.6 V, this solid gate FET had a sensitivity of about 57.5 mV/pH
be calculated as P = ˛ ln (pH) + ˇ, where ␣ and ˇ are 2.7318 and in the pH interval 5.3–9.3 and a VDirac positively shifting at increas-
4.5044, respectively. This model showed a good agreement with ing pH. The FET was tested at VG = 0.75 V for real-time sensing, but
experimental data, also with data reported in [57]. Although the no information on repeatability and hysteresis was reported.
model was not sufficiently precise to fit the VDirac , the pH sens- A further possible improvement to the FET configuration is the
ing factor can be used to predict the conductance of the graphene adoption of a 3D graphene channel to enhance the transistor per-
channel when exposed to an electrolyte solution. formances and reduce the power consumption [69–71]. A graphene
foam, i.e. a 3D network of single and double layer graphene, was
2.2. Solution-gated ISFETs on a flexible substrate grown on copper and then transferred onto a glass substrate.
Although the foam shrank after etching the copper, the 3D structure
The SiO2 /Si is widely used as substrate for the fabrication was preserved. The Ti/Au source and drain had a contact resistance
of graphene-based ISFETs. However, an attempt has been made of about 600 /䊐, a high value probably due to the non-uniform
to produce these ISFETs on flexible substrates to develop a shrinkage of the foam. The top side of the graphene foam was coated
technology capable of providing lightweight, bendable and/or with a layer of 20 nm of HfO2 and a PDMS well was created around
stretchable, potentially environment-friendly and cost effective the channel. The pH sensor was tested with Dulbecco phosphate
®
devices. Mailly-Giacchetti et al. compared the performances of an buffered saline (DPBS) and blood serum of a male Sprague Dawley
SGFET fabricated on SiO2 /Si with a similar device on a 125 ␮m thick rat. The pH of DPBS (150 mM) was changed by adding NaOH or HCl
poly(ethylene 2,6-naphthalenedicarboxylate) (PEN) substrate [66]. to obtain solution in the pH range [3–9]. The pH of the rat serum was
Monolayer graphene films were grown on copper foils by CVD and 7.7, but different pH values in the range [6.4–7.7] were obtained by
then transferred onto SiO2 /Si and PEN. Two SiO2 /Si SGFETs were dilution with HCl (0.03 M) through a nafion membrane. The nafion
fabricated with and without moderate PMMA leftover in order to membrane is only permeable to protons, thus the serum pH was
study the effects of residues due to the microlithographic pro- adjusted with the H+ of HCl. The length and width of the transis-
cess. The leakage current for both SiO2 /Si and PEN SGFETs was tor were 0.6 and 5 mm, respectively. With a VG ∈ [−1.5, 1.5] V,
about 1 nA. The tests on pH in the range [4–8] were performed the leakage current was about 3 orders of magnitude lower than
at VDS = 50 mV and VG ∈ [−0.1, 0.5] V. The SiO2 /Si SGFETs with IDS . Fig. 2e shows the 3D graphene FET. The VDirac positively shifted
(SGFET-SiO2 -PMMA) and without (SGFET-SiO2 ) residues showed from acid to basic pH and the sensitivity was 71 ± 7 mV/pH. The
a positive VDirac shift at increasing pH with similar sensitivi- authors attributed this super-Nernstian sensitivity to the combi-
ties, i.e. 21 and 22 mV/pH respectively. The SGFET-SiO2 was more nation of 3D graphene foam (larger sensing surface than for a 2D
p-doped than the SGFET-SiO2 -PMMA, probably because of the surface) and HfO2 (high density of binding sites for hydrogen ions),
annealing step at 500 ◦ C in H2 /Ar to remove PMMA. The PEN with a major effect of the large sensing surface. The real-time mon-
SGFET had a lower carrier mobility (300 cm2 V−1 s−1 ) than the itoring was verified at VS = −0.5 V and VDS = 0.5 V for pH decreasing
SiO2 /Si SGFET (1250 cm2 V−1 s−1 ) and this difference was ascribed from 9 to 3 and then back to 9. Although the high ionic strength of
to the higher scattering in PEN than in SiO2 because of the higher the background electrolytes (289 mM), this pH responsive was also
root mean square (RMS) surface roughness (∼5–10 nm for PEN responsive when tested with the rat serum, with a positive VDirac
and ∼0.4–0.5 nm for SiO2 ). Nevertheless, since the sensitivities shift toward positive voltages when pH increased from 6.4 to 7.7.
982 P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991

2.4. Graphene quantum dots Another synthetic approach for GQDs entailed the electrolysis
of a graphite rod in a 5 mL dimethylsulfoxide (DMSO) solution with
Graphene quantum dots (GQDs) are low toxicity and high sur- 0.01 M tetrabutylammonium perchlorate (TBAP) [86]. These GQDs
face area particles with diameters below 100 nm, typically between consisted of approximately 1–3 graphene layers with an average
3 and 20 nm. GQDs show quantum confinement effects and a non- diameter of about 10.6 nm. Sensitivity to pH was demonstrated in
zero bandgap that can be tuned during preparation to modify the a wide range, i.e. [1–14], with a particular behaviour at pH > 11.
photoluminescence (PL). The absorption spectra have a character- From 1 to 11, the PL intensity peak shifted from 522 to 575 nm
istic peak at about 230 nm and a typical PL quantum yield lower due to the deprotonation of the functional groups, e.g. COOH and
than approximately 30%. Their luminescence properties can be OH, which is in common with most previous pH sensors. However,
improved by new fabrication methods (e.g. water-phase molecu- from pH 12 to 14, the PL had new absorption peaks at 520–560 nm
lar fusion and hydrothermal methods), doping or functionalization and a strong red emission centred at about 625 nm. The emission
with organic, polymeric, inorganic, or biological molecules [72–77]. at 625 nm was stable even when the excitation wavelengths were
Although GQDs have only recently begun to show their potential changed from 360 to 560 nm. The FTIR and XPS spectra showed the
for sensing applications, they have already been used in imag- presence of a lactone structure that aroused by the tertiary alco-
ing of living cells, as probes for single-stranded deoxyribonucleic hols reacting with nearby carboxylic acids at the edges of GQDs.
acid (ssDNA), detection of human immunoglobulin (IgG), adenosine In strong basic pH conditions, Yuan et al. verified that the lactone
triphosphate (ATP), thiols, and cleavage of DNA. Furthermore, the turned to quinone and caused the modification of the PL absorp-
chemical synthesis of GQDs introduces oxygen-containing groups, tion spectra (Fig. 3c). In fact, after GQDs were reduced by NaBH4 ,
e.g. carboxylic acid groups, and defects at the edges and in the which is known to eliminate lactone in graphene, there was a blue
basal plane, which make GQDs a promising material for pH sensing fluorescence instead of a red emission at basic pH and the solution
[78–82]. colour turned to yellow instead of red at pH 13. From 1 to 13, the
Huang et al. reported a wide PL emission range, from ultraviolet GQDs were pH-responsive from 10 up to 80 ◦ C and had a reversible
C to blue, for a GQD diameter between 1 and 2 nm, whereas a PL red fluorescence response over six consecutive heating-cooling cycles
shift and a decreasing band gap were observed for larger diameters in this temperature range. The in vivo tests with human cervical
[83]. Furthermore, a blue shift was observed when pH moved from cancer cells (HeLa) showed visible colour changes with a culture
basic to acidic (Fig. 3a), and this was ascribed to the moving up pH increasing from 5 to 9.
of the exciton energy level. The authors suggested that the optical Park et al. grafted blue-emitting poly(7-(4-
excitation could induce two kinds of excitons, i.e. within the whole (acryloyloxy)butoxy)coumarin)-b-poly(N-isopropylacrylamide)
GQD or in the edge microstructure. At decreased pH values, the (P7AC-b-PNIPAAm) to green-emitting 10 nm GQDs (block copoly-
passivation of the edge microstructure was responsible for the blue mer integrated GQDs, bcp-GQDs) [87]. At room temperature, the
shift, whereas the wide emission spectra for small diameter size PL spectra of bcp-GQDs showed two emission peaks following an
depended on the strong quantum confinement due to the excitons excitement at 365 nm, a first peak at 410 nm corresponding to
confined within the GQD. the P7AC-b-PNIPAAm, and the other at 505 nm associated with
Wu et al. synthesized nitrogen-doped GQDs (N-GQDs) using the GQDs. The blue emission peak at 410 nm did not change with
citric acid (CA) and dicyandiamide (DCD) as the carbon and nitro- pH, whereas the green emission peak changed because of the
gen sources, respectively (Fig. 3b) [84]. The N-GQDs were 1–2 protonation and deprotonation of the oxygen-based functional
graphene layers thick with an average diameter of 2.3 nm. The groups. However, the intensity ratio I510 /I410 plotted versus pH
Fourier transform infrared spectroscopy (FTIR) and the XPS showed from 1 to 11 did not allow several pH values to be resolved, thus
the presence of COOH, OH and NH2 groups. In buffer solu- this promising idea should be further expanded with additional
tions from pH 2 to 9 and at an excitation wavelength ␭ex = 365 nm, work in order to be of practical application.
the fluorescence intensity of N-GQDs was linearly proportional to Song et al. fabricated GQDs via a modified graphite intercalated
pH and reversible during multiple cycles between pH 2 and 9. The compounds (GICs) method [88] and investigated the pH response
zeta potential versus pH showed that these N-GQDs are positively [89]. Two GQDs diameters were investigated, i.e. GQD-A (2 ± 1 nm)
charged for pH < 2.7 and negatively charged for higher values (sen- and GQD-B (18 ± 2 nm), with an oxygen content of ∼5 at% and
sitivity ∼ −4 mV/pH), which is a similar behaviour to that of amino ∼8 at%, respectively. The FTIR spectra showed that the oxygen con-
acids. These N-GQDs were also used to measure the pH of river tent was mostly due to the carboxyl groups and was lower than
water, rain and tap water and gave comparable results with those that of typical GQDs obtained from GO and rGO. In particular, GQD-
of a commercial pH-meter. The biocompatibility was assessed with A had a lower content than GQD-B. The carboxylic acid peak was
human epidermoid cancer cell lines (Hep-2) at an N-GQDs concen- absent in GQD-A and much lower in GQD-B. Raman spectra indi-
tration of 1 mg mL−1 for 24 h. The N-GQDs did not significantly alter cated that these GQDs had a few defects and GQD-A had a stronger
the viability of cells and were permeable to cells membranes with sp2 hybridization than GQD-B. GQDs were tested with excitation
an increasing fluorescence intensity from pH 5.7 to 8. wavelengths from 250 to 390 nm and at pH 2, 7 and 12. For GQD-
In 2015, Shi et al. used 3,4-dihydroxy-l-phenylalanine(L-DOPA) A, the maximum PL intensity was at ␭ex = 315 nm at all pH values.
as the nitrogen source and the acronym N-OGQDs to stress that The PL intensity decreased with increasing pH and the spectra were
these GQDs were rich in oxygen atoms [85]. Their material had an blue-shifted, sharper and with higher intensities at pH 2 than at pH
average diameter of 12.5 nm and contained 66.58 wt% C, 4.01 wt% N, 7 and 12, probably because of the protonation effect at basic pH
5.77 wt% H, and 23.64 wt% O. The N-OGQDs were stable in high- on the edges and defect sites of GQD-A. Conversely, GQD-B had the
ionic strength buffers (concentration of NaCl up to 1 M) and the PL lowest PL intensity at pH 2 and the maximum at pH 7 and 12 for
intensity linearly increased in the pH range [3–8] (␭ex = 346 nm). ␭ex = 360 nm, with an increasing trend from basic to acidic pH.
Given the fluorescence intensity in arbitrary units (a.u.), the sen- The nanosheets synthesized by Joseph et al. were very close
sitivity was ≈80 a.u./pH and can be ascribed to the reactivity of to nanodots, with average length and width of 135 and 137 nm,
the amide and carboxylic groups. Although the N-OGQDs seemed respectively, and a thickness of about 1–3 graphene layers [90].
potentially suitable for measuring pH in the physiological range, These nanosheets were obtained by the aqueous exfoliation of
further tests are needed to evaluate parameters such as repeatabil- graphite through the interaction with an antibacterial protein, the
ity and accuracy. lysozyme (LYS). Since the isoelectric point (IEP) of LYS depends
on the solution pH, the correlation between the LYS-nanosheets
P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991 983

Fig. 3. a) PH-dependent exciton absorption/emission of GQDs from basic to acidic conditions (Adapted with permission from [83]. Copyright 2016 American Chemical
Society); b) synthesis of the GQDs by hydrothermal treatment using citric acid and dicyandiamide as the carbon and nitrogen sources (Adapted from [84] with permission
from The Royal Society of Chemistry); c) (top) Schematic representation of the local structural transformation of GQDs, synthesized by electrolysis in a DMSO and TBAP
solution, from lactone (left) to quinone (right) modelled by theoretical calculation; (bottom) The structural model of the GQDs under neutral (pH = 7) and strong basic
(pH ≥ 12) conditions (Adapted from [86] with permission from The Royal Society of Chemistry).

and the zeta potential was studied at different pHs from 1 to 13. accepted model was developed by Lerf and Klinowski [93,94] and
The LYS-nanosheets IEP was about 8 with an initial pH of approx- describes GO as graphitic platelets characterized by epoxy and
imately 4. The zeta potential allowed the pH to be discriminated hydroxyl groups on the basal plane, and carboxyl groups on the
only between pH 1 and 5. At pH values below the IEP, the amines edges. The large number of highly polar groups makes GO highly
in the LYS-nanosheets led to a global positive charge that repelled hydrophilic and a good candidate for measuring the pH of solu-
the nanosheets from aggregations, whereas above the IEP the pre- tions. The pH dependence of GO in aqueous solutions has been
dominant effect was ascribed to the carboxylate ions that led to investigated and the change in GO hydrophilicity has generally
a global negative charge and a fine dispersion. On the other hand, been ascribed to the protonation-deprotonation of the carboxyl
around the IEP, the LYS-nanosheets aggregated because of a charge groups [95–97]. At high pH, the hydrophilicity is enhanced and
depletion. Interestingly, when tested for cytotoxicity with mouse GO should behave like a surfactant, but since the surface tension
embryonic fibroblasts (NIH-3T3) and three different cancer cells is weak, it dissolves in water like a regular salt. At low pH, the
lines, namely human colorectal cancer cells (HCT-116), HeLa and degree of deprotonation is low and GO forms suspended aggre-
squamous carcinoma cells (SCC-7), the LYS-nanosheets were less gates with a sandwich-like structure GO-water-GO surrounded by
toxic for the NIH-3T3 than for the cancer cells. Although the cyto- water molecules.
toxicity mechanism needs further investigation, these nanosheets Wang et al. prepared a GO colloid using a modified Hummers
showed promising properties for pH sensing. method [98] and studied the role of pH, adjusted with HCl solu-
It is worth noting that the response of GQDs to pH variations tion and aqueous ammonia, in the formation of a hydrogel [99].
showed a pH dependency, but the trueness and precision of the The GO colloid aggregated into a cylindrical hydrogel at pH ≤5 or
measured values should be further demonstrated. ≥10, whereas a black dispersion was obtained otherwise. This result
depended on i) the degree of dissociation of carboxyl groups, whose
3. Graphene oxide-based pH sensors charge state controlled the attraction-repulsion of GO sheets, ii) the
concentration of GO, and iii) the use of HCl and ammonia, which
Unlike graphene, GO is an electrical insulator due to its dis- reduced the thickness of the electrical double layer of the GO sheets.
rupted sp2 bonding network, though there are several methods With another modified version of the Hummers method, Qin
to restore the electrical conducitivity [91]. Although the precise et al. synthesized and dispersed about 20–50 mg of GO in 1 mL of
chemical structure is still under investigation [92], the most widely purified water to obtain a self-assembled GO hydrogel without any
984 P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991

pre-processing [100]. The average size of the GO nanosheets was film was found to be highly biocompatible in vitro when in contact
about 500 nm, and the thickness was between 0.911 and 0.607 nm, with cultures of RAW264.7, MDA and MC3T3-E1 cell lines.
respectively. The hydrogel was obtained with a minimum GO con- Chen et al. monitored the vis-near infrared (vis-NIR) fluores-
centration of 30 mg/mL. The gel formation was pH-dependent and cence of GO to determine the changes in the extracellular pH during
only occurred for acidic pH. The GO hydrogel was stable when the growth and metabolism of normal and cancer cells [105]. In buffer
GO concentration exceeded 5 wt%, probably because the ␲–␲ stack- solutions, the fluorescence intensity at 650 nm of the single-layer
ing and the hydrophobic interactions inhibited the pH-response. GO nanosheets had a reversible sigmoidal response when the pH
The ionization of the carboxyl groups, which increases with pH changed from 2 to 12 (Fig. 4c). The pH changes were also moni-
and reduces the number of the hydrogen bonds, was also exploited tored in cultures of chronic granulocytic leukaemia (CGL) cancer
in hydrogels consisting of GO and poly(N-isopropylacrylamide) cell (32D-BA) and normal mouse cells (32Dc-13). This test showed
[101], or GO and starch-based superabsorbent nanocomposites that the GO optical nanosheets were in linear correlation with a
(SANCs), i.e. magnetic iron oxide nanoparticles and starch-g- commercial pH meter (R2 = 0.99) between pH 6.6 and 7.7. The vis-
poly(acrylic acid-o-acrylamide) [102]. NIR fluorescence of GO was identical at different ionic strengths for
NaCl, KCl and KNO3 . The vis-NIR fluorescence intensity decreased
as the ionic strength increased from 0 to 10 mmol·L−1 and was
3.1. Optical GO-based pH sensors constant for further increases. The reversible response of the GO
vis-NIR fluorescence upon pH and ionic strength variations was
Because of its insulator-like behaviour, the use of GO as pH explained by the passivation of the radiative defects embedded
sensor is generally associated with optical analysis. Lv et al. ultra- in the graphene network. Probably, when pH and ionic strength
sonicated 2 mg/mL GO for 2 h to obtain GO hydrosols whose pH was changed, the protonation and deprotonation of the carboxylic
adjusted from 2 to 11 with HCl or KOH aqueous solution [103]. Out- groups could electrostatically charge the GO, thereby shifting the
side the pH range [2–11], the hydrosol became unstable. After the Fermi level.
centrifugation to remove the residual nanosheets, the GO hydrosols In another work, PAA and poly(2-vinylpyridine) (P2VP) were
were heated in a water-thermal bath at about 85 ◦ C for 30–45 min grafted onto the surfaces of CdS/ZnS (PAA-BQDs, blue emission at
and formed thin membranes at the liquid/air interface. These GO 440 nm) and CdSe/ZnS (P2VP-OQDs, orange emission at 580 nm)
membranes (GOMs) had a thickness of between 1.5 and 2 ␮m at QDs, respectively [106]. These two compounds were anchored to
pH 11 and 2, respectively. Although Lv et al. studied these GOMs GO by ␲–␲ stacking interactions between the pyrene groups of PAA
to tune their optical properties, the GOMs has the potential to be and P2VP and the basal plane of GO. The PAA-BQDs and P2VP-OQDs
also used a pH sensor. According to the ultraviolet-visible (UV-vis) were mixed in aqueous media (1 mg/mL) and characterized at dif-
spectra, the highest optical transmittance at 600 nm was about 30% ferent pH conditions with a ␭ex of 365 nm. The ratio of the two
at pH 2, and non-linearly decreased to about 1% at pH 11 (Fig. 4 a). intensities, I580 /I440 , plotted versus pH from 1 to 7 showed a non-
Lv et al. explained that this result was due to the deoxygenation of linear relationship that could be linearly approximated between
the GOMs in basic conditions, which led to a partial reduction of GO. pH 1 and 5 (sensitivity ∼-0.56/pH, Fig. 4d). The PL intensity of the
The transmittance increased with increasing wavelength and was blue peak was almost stable from pH 1 to 4, and increased from
65% at 1000 nm and pH 2. The transmittance of the GOMs at pH 2, 7 pH 4 to 7. On the other hand, the PL intensity of the orange peak
and 11 was also determined in the range [200–3300] nm. Perhaps decreased from pH 1 to 5, and saturated after pH 5. This effect
because of the presence of water molecules, the GOMs transmit- could be explained by the fact that for pH < pKa = 3 the P2VP chains
tances showed a non-linear dependence after 1300 nm, which can swelled (P2VP-OQDs size increased from ∼45 to ∼85 nm), whereas
be thus assumed as the wavelength limit to use these GOMs as pH the PAA chains swelled for pH > pKa = 4.5 (PAA-BQDs size increased
sensors. from ∼30 to ∼70 nm). The swelling reduced the GO quenching and
Yan et al. proposed an optical pH sensor consisting of a film favoured the blue or the orange emission depending on the pH
made up of polyethylenimine (PEI) functionalized lanthanide- value (Fig. 4e). These GO-QDs were also reversible when cyclically
doped (NaYF4 :Yb,Er) upconversion nanoparticles (PEI-UCNPs) and exposed to acid and basic pH. The GO-QDs were only tested in buffer
GO [104]. GO is a pH-dependent quencher for the luminescent PEI- solutions; therefore, the effect of interfering agents in real solutions
UCNPs, and the fluorescent intensity can be used to sense pH. PEI should be further investigated.
has many amine groups that can protonate and interact with the A pH sensing material was obtained by combining DNA probes
negative charges of GO. Thus, the PEI-UCNPs interaction with GO is and GO in a work by Luo et al. [107]. This sensor was fabri-
weak at acidic pH because GO has few negative charges, whereas it cated by incubating DNA probes marked with the fluorescent dye
is strong at basic pH where the GO nanosheets are more negatively tetramethylrhodamine (TAMRA, 100 nM) in 20 ␮L phosphate buffer
charged. This interaction was further confirmed from the enhanced at different pH values (different pH were obtained with 20 mM
binding between pure PEI and GO with increasing enthalpy, thus KH2 PO4 , 20 mM Na2 HPO4 and 100 mM NaCl) for 5 min, then adding
binding depended on the long-range electrostatic interaction. Since ultra-pure water (68 ␮L) and GO (12 ␮L, concentration 100 ␮g/mL).
enthalpy increased with increasing pH, the authors were able to Two DNA probes were designed to have different contents of the
confirm the pH-dependent behaviour. An amount of 0.5 mg/mL of protein Tat, 50% (probe 1) and 80% (probe 2). Tat allowed the DNA
PEI-UCNPs at pH 5 was mixed with 1 mg/mL GO and, after vacuum conformation to change at different pH values, whereas the flu-
filtration and washing with water, a film was obtained leaving this orescence signal of TAMRA was independent of pH. At basic pH,
dispersion to dry in air for two weeks. The GO-PEI-UCNPs film was the DNA probes had a hairpin structure with single stranded tails
dipped in buffer solutions and excited with a 980 nm laser. The flu- and were absorbed onto the GO surface, thus the fluorescence was
orescence intensity at 540 nm linearly decreased (maximum 35%) quenched. At acidic pH, the DNA probes assumed a triplex structure,
when the pH changed in the physiological range from 5 to 8, with a which is more rigid than the single stranded tail conformation. In
response time of about 60 s for unitary changes of pH. The repeata- this case, the probes were not absorbed, and the fluorescence was
bility, stability and reversibility were demonstrated by cycling the not quenched (Fig. 4f). For pH < 5, probe 1 had a weak fluorescence
film between pH 6 and 8 for five hours. This sensor was also tested intensity and increased from pH 5 to 7. For pH > 7, the intensity
with mice urine, which was diluted to obtain different pH condi- saturated. Probe 2 had a similar behaviour with a dynamic range
tions. The response was almost linear, with a maximum intensity between pH 6 and 8. Therefore, a combination of probes 1 and 2
shift of 55% between pH 4.95 and 8.05 (Fig. 4b). The GO-PEI-UCNPs could be used to measure pH from 5 to 8. This DNA probes have
P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991 985

Fig. 4. a) Optical transmittance of the GOMs at 600 nm. The decrease in transmittance depended on the deoxygenation and partial reduction of the GOMs in basic
conditions (Adapted with permission from [103]); b) Response of the GO-PEI-UCNPs in solutions of diluted mice urine. F0 is the fluorescence intensity at 540 nm
at pH 5, whereas Fx is the fluorescence intensity at the pH under test (Adapted from [104] with permission from The Royal Society of Chemistry); c) fluorescence
intensity of the GO single-layer nanosheets at ␭ex of 650 nm. The intensity can be linearly approximated in the pH physiological range [4–8] (Adapted from [105]
with permission from The Royal Society of Chemistry); d) response of the PAA-BQDs and P2VP-OQDs solution reported as the ratio between the orange (I580 ) and
the blue (I440 ) emission intensities versus pH. Inset: linear approximation between pH 1 and 5 (Adapted with permission from [106]. Copyright 2014 American Chemi-
cal Society); e) Top: chemical structure of P2VP-OQDs (left) and PAA-BQDs (right). Bottom: conformational changes in the PAA and P2VP chains at different pH values. At acidic
986 P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991

showed a notable potential as pH sensor, but this potential needs ing, ultra-sonication and centrifugation, a lysozyme-rGO (lys-rGO)
to be confirmed when the probes are exposed to real biological solution was obtained. The lysozymes adsorbed on GO, proba-
samples. bly due to ␲-␲ stacking and electrostatic interactions. They then
acted as a dispersant, leading to a low folding degree and flat lys-
3.2. Electrochemical GO-based pH sensor rGO with a higher carbon-oxygen (C/O) atomic ratio than that of
bare rGO. Furthermore, the lysozymes stabilized the rGO disper-
Melai et al. presented an electrochemical GO pH sensor with sion, and promoted the reduction of the oxygen-based groups of
a gold three-electrode screen-printed set-up [108]. The working GO, thus enhancing the restoration of the sp2 hybridization. The
electrode (WE) was coated by a 400 nm GO layer and the ref- zeta-potential increased in absolute value when pH increased from
erence electrode (RE) was Ag/AgCl (the counter electrode was 2.36 to 12.5, and were assumed to be linear between pH 4 and 10
short-circuited to RE). The open-circuit circuit potential between (Fig. 5a).
WE and RE was linearly related to pH (range [4–10]) in buffer Liu et al. added chitosan to a GO solution before the reduction
solutions, with a sensitivity of about 32 mV/pH. This sensor was with hydrazine [115]. The zeta potential was higher than 30 mV
intended to monitor wound pH, and measured values that only for 2 < pH < 5, probably because of the strong chitosan protonation.
differed less than 0.1 pH unit from a commercial pH meter over a After pH 5, the zeta potential rapidly decreased to about 9 mV at
4-day test in human exudate, i.e. a fluid mainly produced in the pH 6 and was negative for pH ≥ 7. This effect was probably due
inflammatory and proliferative phases of a wound [109]. to the increased ionization of the carboxylic groups in rGO and
Salvo et al. expanded the previous work and used a GO disper- the decreased chitosan protonation. These effects led to the aggre-
sion (4 mg/mL) to fabricate a pH responsive film in the range [4–10] gation of rGO at pH values higher than the pKa of chitosan (6.5)
[110,111]. The resulting sensor (sensitivity ∼40 mV/pH, repeatabil- because the electrostatic repulsive forces became too weak. The
ity ±4 mV and high linearity R2 = 0.99) performed comparably to aggregation of rGO was highly reversible when tested in repet-
a commercial pH meter (error = 0.14 ± 0.09 pH units) when tested itive cycles between pH 4 (dispersion) and 7 (aggregation). The
in human plasma over one month. When tested in human serum same behaviour was observed between pH 7 (aggregation) and 10
over one week, the pH sensor was compared with a commercial (dispersion) (Fig. 5b).
pH meter (glass electrode) and had a mean difference of about Ren et al. described a polyacrylamide (PAM)–rGO composite
0.1 ± 0.1 pH units. The standard deviations for pH meter and the that could be dispersed for pH ≥ 4 [116]. Although the PAM-rGO
GO sensor were 0.05 and 0.1 pH unit, respectively. This GO pH sen- composite showed a pH dependence when characterized by UV-vis,
sor showed a dependence from temperature of about 0.01 pH/◦ C this optical analysis does not seem ready for pH sensing because the
in the range [25–45] ◦ C and had an hysteresis of 0.15 pH unit. The discrimination between different pH values still needs improving.
biocompatibility of the GO film was positively assessed in a broth Yang et al. treated GO with ozone before and after a thermal
medium that included minimum essential medium (MEM) with reduction process to obtain an ozonized rGO quantum dots solu-
10% fetal bovine serum (FBS) and human fibroblast cells (MRC-5 tion (O-rGO-QD) [82]. After the thermal reduction, the rGO sheets
(p33)) cells. were treated with ozone in water solution at pH 2 (O-rGO1) and 13
Liu et al. followed a similar approach using a 3-electrode con- (O-rGO2), and hydrothermally treated at pH < 1. Four other types
figuration and fabricated a cysteic acid/GO composite film [112]. of rGO sheets (O-rGO3 to O-rGO6) were obtained after ozonation
The advantage of the cysteic acid was to add carboxylate and sul- at pH 2, 7, 10 and 13, and a hydrothermal treatment at pH 13.
fonated pH- sensitive groups to the GO structure. The cysteic acid After drying, the O-rGOs were filtered through a microporous mem-
was obtained oxidizing L-Cysteine (CySH) by cyclic voltammetry brane to obtain a solution of O-rGO quantum dots (O-rGO-QDx,
in the range [−0.8 m to +2.2] V at a scanning rate of 200 mV s−1 x = 1.6, diameter 2–5 nm for x = 1 and 2, and 3–5 nm for x = 3.6). The
in 0.04 M HCl solution containing 2.5·10−3 M CySH for 10 consec- re-oxidation of GO through ozone formed carboxylic and ketone
utive cycles. The optimum GO concentration was 0.15 mg mL−1 . groups, i.e. reactive sites for thermal reduction. These reactive sites
The cysteic acid/GO electrochemical sensor had a linear response were also introduced with the ozone treatment after the reduc-
(R2 = 0.997) in the pH range [2–12] with a sensitivity of about tion. The percentage of reactive sites depended on the pH of the
50 mV/pH. The sensor response was less than 1. The stability was hydrothermal treatment, which increased the oxygen percentage
tested at pH 5 for 11 h at 25 ◦ C and the relative standard deviation from about 16% for O-rGO-QD1 and O-rGO-QD 2 to a range of
was 0.8%. The storage life was about 1 month. This promising cys- about [17,27] % for O-rGO-QD3 up to O-rGO-QD6. With an excita-
teic acid/GO sensor was compared with a commercial pH meter to tion wavelength of 254 nm, the fluorescence emission peak shifted
measure the pH of Coca Cola and tap water and showed relative towards shorter wavelengths when pH increased from 2 to 13 dur-
deviations of 1.6% and 0.7%, respectively. ing ozonation. Yang et al. hypothesized that this phenomenon could
be explained by the reaction of rGO with one of the two forms of
4. Reduced graphene oxide-based pH sensors ozone in acid–base aqueous solution, i.e. dissolved ozone or HO.
radicals. In acidic solution, the dissolved ozone is more reactive
Like for GO, the literature reports several attempts to produce than the HO. radicals, whereas the opposite happens at basic pH.
pH-sensitive rGO-based dispersions. Liu et al. proposed pyrene- The O-rGO-QD1 fluorescence emission peak at 404 nm reversibly
terminated poly(2-N-(dimethyl amino ethyl acrylate) (PDMAEA) increased for pH from 1 to 13 pH, with a negligible change in shape
and PAA for polymer-rGO composites, but the pH-dependence of the intensity curve and a strong variation for pH > 7. The authors
of rGO was negligible [113]. In another paper, after the synthe- suggested several theories that could explain this result, such as
sis of GO by a modified Hummers method, Yang et al. added the triplet ground state [117] and the islets theories [118].
10 mg of lysozyme from hen egg white to 20 mL of a GO solution Srinives et al. fabricated an rGO pH sensor using L-ascorbic acid
before reduction with 1 mL of hydrazine hydrate [114]. After heat- (L–AA) as the reducing agent [119]. An rGO layer of unknown thick-

pH, the P2VP chains were extended because of the protonation of amine groups, whereas the PAA chains were collapsed. At basic pH, the P2VP chains are collapsed, whereas the
PAA chains are extended because of the ionization of the carboxylic acid groups. At pKaP2VP < pH < pKaPAA , both PAA and P2VP chains are collapsed (Adapted with permission
from [106]. Copyright 2014 American Chemical Society); f) Conformational changes of the DNA probes at different pH values. At basic pH, the DNA probes were absorbed
onto the GO surface and the fluorescence was quenched. At acidic pH, the DNA probes were more rigid and not absorbed onto the GO surface, thus the fluorescence was
maintained Adapted from [107] with permission from The Royal Society of Chemistry).
P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991 987

Fig. 5. a) pH dependence of a lysozyme-rGO solution. The lysozyme improved the restoration of the sp2 facilitating the reduction of the oxygen-based groups of GO (Adapted
with permission from [114]. Copyright 2010 American Chemical Society); b) Synthesis process and dependence of the chitosan-rGO solution on pH. At acidic pH, the chitosan
was highly protonated and the electrostatic repulsion prevents the aggregation of rGO. In the pH range [6–8], the chitosan deprotonated and the carboxylic groups became
ionized. The electrostatic repulsion became too weak and the rGO started aggregating. At pH ≥ 9, the ionization of the carboxylic groups on the graphene surface is the
predominant effect (Adapted with permission from [115]); c) SEM images of the reverse-pyramid rGO-FET, which had a higher pH sensitivity than the rGO-FET because of the
increased sensitive area; (left) top-view, (middle) cross-sectional view, (right) example of the reverse-pyramid structure with pitch = 70 ␮m. (Adapted from [122]. Copyright
(2016) The Japan Society of Applied Physics).

ness and size bridging two electrodes (the electrodes distance is not was tested during real-time measurements at constant VG = −0.2 V.
reported) was used to fabricate a chemiresistive sensor. The nor- Unlike graphene, the rGO has oxygen functional groups in its struc-
malized resistance (R/R0 ) to the rGO resistance in air (R0 ∼ 6.7 k) ture; thus, it is reasonable to assume that the modulation of the
decreased for unitary changes of pH from 3 to 8 with a sensitivity reduction process could improve the FET performances.
of 0.5 pH−1 , roughly. However, this study does not seem mature Li et al. started with a SiO2 /Si substrate anisotropically etched
enough for pH sensing as stability, repeatability, cycling tests and to obtain reverse pyramids (35 ␮m deep, pitch distances 18, 35, 70,
large variations at low pH should be further characterized. 105 and 140 ␮m) in Si (Fig. 5c) [121,122]. After spraying a com-
mercial rGO solution onto the substrate and heating at 120 ◦ C, the
rGO was annealed in nitrogen at 300 ◦ C for 1 h and treated with
4.1. rGO FETs oxygen plasma for 1 min. The sensing window was 4 × 4 mm2 per
sample. Transmission electron microscopy (TEM) and XPS analy-
Sohn et al. published the first paper on a solution-gated ses showed that the oxygen-plasma-treated (OPT)–rGO FET had
rGO FET for pH detection [120]. The rGO FET had a SiO2 /Si more oxygen functional groups than the rGO FET. In terms of pH
substrate and the SiO2 (100 nm thick) was exposed to a 3% 3- sensitivity, at VG ∈ [0, 4] V and VDS = 0.2 V, the OPT-rGO FET (sensi-
aminopropyltriethoxysilane (APTES) aqueous solution. The APTES tivity ∼53 mV/pH) performed better than the rGO FET on planar Si
positively charged SiO2 with amino functional groups, so that the (45 mV/pH), whereas the OPT-rGO FET with the reverse-pyramid
aggregation of GO nanosheets was prevented when a GO solu- structure (RP-OPT-rGO FET) was even better (57.5 mV/pH for a
tion was dropped on the substrate (nanosheet thickness ∼1 nm, pitch/depth ratio of 0.5) with high linearity (R2 = 0.996) because
diameter 0.5–1.5 mm). The rGO channel (resistance ∼10–20 k) of the increased sensing surface. The optimal power for oxygen
between source and drain (length 400 mm, ratio width/length = 20) plasma treatment was 20 W, since a decrease in sensitivity was
was obtained by reducing GO with hydrazine. The VG was applied observed at higher values. Hysteresis, defined as the voltage off-
through an Ag/AgCl electrode in the range [−0.2 to 0.2] V and con- set between the beginning and the end of the test, was calculated
stant VDS = 0.1 V. The device had a VDirac ∼
= 0 ± 0.01 V at pH 7 and for the pH loop sequence 7, 3, 7, 11 and 7. For rGO, OPT-rGO and
low rGO carrier mobility (0.5 cm2 /(V s)) due to the residual OH RP-OPT-rGO FETs, the offsets were about 19, 13, and 11 mV, respec-
and COOH functional groups on the rGO surface after reduction. tively. Therefore, the RP-OPT-rGO FET outperformed the other
The rGO FET was tested from pH 6 to 9 with unitary increments. configurations.
The response was almost linear (R2 ∼ = 0.98) with a sensitivity, i.e.
VDirac /pH, of about 29 mV/pH. The reversibility seems good
from the reported plot, but no numerical error was given. The 5. Conclusions and outlook
repeatability error increased for increasing pH, with a maximum
of about 20 mV at pH 9, i.e. almost one pH unit. The pH values Developing a graphene-based pH sensor is challenging.
from 9 to 6.11 were correctly discriminated when the rGO FET Although graphene-based materials have pH-dependent elec-
988 P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991

Table 2
Characteristics of the graphene and rGO field effect transistors for measuring pH.

FET Configuration Substrate VG (V) VDS (V) Sensitivity (mV/pH) pH range Ref.

Solution-gated 6H-SiC [−1, 1] −1 99 [2–12] [49]


Gate-free SiO2 /Si Not applicable Not reported 2.13 k/pH [4–10] [50]
Gate-free Paper Not applicable [−1, 1] 30.8 /pH [1–10.5] [51]
Solution-gated SiO2 /Si [−0.05, 0.05] Not reported 30 [4–8.2] [52]
Solution-gated 6H-SiC [−1.2, 0.3] −0.05 19 [3–12] [53]
Solution-gated, suspended graphene SiO2 /Si [−0.05, 0.05] Not reported 17 [6–9] [54]
Solution-gated SiO2 /Si [0, 1] 0.05 −78, −38 and −7 for 0.05 M Na+ , Cl− [6–8] [55]
and K+ solutions, respectively; +69 for
reference buffers
Solution-gated SiO2 /Si [−0.4, 0.2] ≤0.005 At pH 7, the VDirac shift changed by 3 and 7 [56]
−42.7 mV/decade with the buffer
concentration. At pH 3, there was a
sign inversion (+18.9 mV/decade)
Solution-gated SiO2 /Si [0.1, 0.8] [0.01, 0.05] 6 [5–10] [57]
Solution-gated, graphene mesh SiO2 /Si [−0.1, 0.6] 0.05 90 (single use); 7 (after 3 cycles) [6.55–8.25] [58]
Solution-gated, graphene nanoribbons SiO2 /Si [0.07, 0.32] 0.01 23.6 [6–8] [59]
Solution-gated, suspended graphene SiO2 /Si [−1.8, 1.8] 1 25 [5–9] [60]
nanoribbons
Solution-gated, anodized graphene SiC Not reported 51.3 2 and 10 [61]
Solution-gated SiO2 /Si, PEN [−0.1, 0.5] 50 22 (both substrates) [4–8] [66]
Solid gate, HfO2 dielectric SiO2 /Si [0.6, 1.6] Not reported 57.5 [5.3–9.3] [68]
Solid gate, 3D graphene SiO2 /Si [−1.5, 1.5] 0.5 71 ± 7 [3–9] [69–71]
Solution-gated, APTES treatment, rGO SiO2 /Si [−0.2, 0.2] 0.1 29 [6–9] [120]
Solution-gated, reverse-pyramid, Si [1.3, 2.4] 0.2 57 [1–13] [121,122]
oxygen plasma treated, rGO

trochemical properties, current pH sensors are not sufficiently This list includes borohydrides (e.g. NaBH4 ), aluminium hydride,
advanced for new products to enter the market. hydrohalic acid (e.g. HI), sulphur-containing reducing agents (e.g.
In this review, the field effect transistor was the most used pH- thiourea dioxide/NaOH), nitrogen-containing reducing agents (e.g.
sensing architecture. Table 2 summarizes the efforts reported in hydrazine, urea/NH3 , and pyrrole), oxygen-containing reducing
the literature to create a graphene-based pH sensor. Most FETs agents (e.g. methanol, ethanol, L-ascorbic acid, and glucose/NH3 ),
adopt graphene as the conduction channel, whereas there are only sulphur-containing reducing agents (e.g. NaHSO3 and thiourea),
two examples of rGO FETs. The sensitivity of these devices ranges metal-acid (e.g. AlHCl, Fe/HCl and Mg/HCl), metal–alkaline (e.g.
from about 6 to 99 mV/pH and such a wide spectrum suggests Zn/NH3 and Na/NH3 ), amino acid (e.g. L-cysteine and glycine), plant
that researchers are still looking for an optimum solution. Pristine extracts (e.g. green tea and R. damascena), microorganisms (e.g.
graphene is scarcely sensitive to pH, but defects induced by oxy- E. coli and shewanella), proteins (e.g. bovine serum albumin/NaOH),
genated functional groups can modify its chemical structure and and hormones (e.g. Melatonin/NH3 ) [123]. This list should also
turn it into a pH-sensitive material that can rival other existing pH be taken into account for reduced graphene oxide since a further
sensors. reduction will eliminate the defects, therefore decreasing the sen-
However, a standard fabrication process is not yet available and sitivity to pH.
small differences can lead to large variations. For example, Ristein This review has presented proof-of-concept pH sensors usually
et al. found that the VG should be carefully chosen to drive cor- tested in reference buffer solutions. However, some works showed
rectly the gate [53], and the roles of ionic charge screening [55] and that the pH response changes when electrolytes are added to the
radius [56] need to be further investigated. The control of defects buffer solutions. This could limit the applicability of these sensors
is of primary importance to obtain reliable pH sensors, especially in real case studies. Quantum dots based on graphene, GO and
the amount of COOH groups at the graphene edges. It has been rGO are an emerging alternative to FETs to bridge the performance
demonstrated that oxygen plasma treatment or anodization can gap between the FETs and pH sensing (Table 3). For example, the
increase the number of reactive sites to H+ and OH− ions. Another GQDs described in [84] were successfully tested in river, rain and
improvement could result from the increase of the pH sensitive tap water, and passed the biocompatibility tests. However, GQDs
area, as some transistors with suspended and 3D graphene have are still confined to lab research because of limitations, such as
demonstrated. the difficulty to achieve large scale production and high quantum
These approaches can pave the way towards a standardized fab- yield, and to understand correctly the quantum confinement and
rication technology, but they are only one side of the whole picture. the doping mode (e.g. lattice and edge modes) [82,124,125]. The
Unfortunately, the pH response can be affected by impurities and optical pH measurement showed a great potential, but most results
residual resist such as silanol and silylamine groups that modify the were based on tests performed in buffer solutions. The optical
net charge of the substrate surface, typically SiO2 on Si. This prob- response is always associated with a hybrid material that includes
lem could be solved by improving the manufacturing techniques, graphene, therefore the effects of interfering agents that can mod-
but the effect of interfering ions such as Na+ , Cl− and K+ could still ify the chemistry of the hybrid material or mislead the pH readings
be a serious limiting factor. Redox agents can also interfere with the should be discussed in future upgrades of these sensors. Like for
pH measurement. For graphene, Eqs. 1 and 2 can be used to under- the electrochemical sensors that use a reference electrode, a ref-
stand the effects of the electron exchange during the measurement. erence measurement would be essential for the optical sensors to
The redox agents can seriously impair the graphene oxide as the assess the reliability of the results. In his seminal paper, Janata dis-
removal of oxygen groups decreases the sensitivity to pH. Many cussed the problems inherent to the optical pH measurement [126].
redox agents are used to obtain graphene from graphene oxide. Most of the optical sensors usually ignore the contributions of dilute
Thus, when the pH is measured in solutions where these agents aqueous solutions, where the activity coefficients tend to unity and
are present, the result can be incorrect or hardly reproducible. their ratio converges even faster with dilution. When dealing with
P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991 989

Table 3
Characteristics of the graphene-base quantum dots.

Starting Material Size (diameter, nm) Excitation Sensitivity pH range Ref.


wavelength (␭ex , nm)

Nitrogen-doped graphene 2.3 365 −4 (mV/pH) [2–9] [84]


Nitrogen-doped (from L-DOPA) graphene 12.5 346 ∼80 (a.u. of fluorescence [3–8] [85]
intensity)/pH
Graphene (in 5 mL DMSO solution which 10.6 365 ∼5 nm/pH [1–14] [86]
contained 0.01 M TBAP)
P7AC-b-PNIPAAm-grafted graphene 10 365 Not reported [1–11] [87]
Graphite powder 2 and 18 250–390 Not reported [2–12] [89]
PAA-CdS/ZnS and P2VP-CdSe/ZnS nanodots PAA-CdS/ZnS = ∼85 at pH 7; 365 ∼−0.56/pH [1–7] [106]
anchored to graphene oxide P2VP-CdSe/ZnS = ∼70 nm at pH
7
Ozonized reduced graphene oxide 2–5 254 Not reported [1–13] [82]

optical sensors, the effects of ionic strength, the presence of poly- [14] R. Bizzarri, M. Serresi, S. Luin, F. Beltram, Green fluorescent protein based pH
electrolytes, mixed solvents, and adsorption of molecules onto the indicators for in vivo use: a review, Anal. Bioanal. Chem. 393 (2009)
1107–1122.
surface of the pH-sensitive material should be investigated. There- [15] Y. Chen, X. Wang, S. Erramilli, P. Mohanty, A. Kalinowski, Silicon-based
fore, optical sensors should prove to be reliable in a wide spectrum nanoelectronic field-effect pH sensor with local gate control, Appl. Phys.
of conditions before they can be used for practical applications. Lett. 89 (2006) 223512.
[16] A. Das, D.H. Koa, C.H. Chen, L.B. Chang, C.S. Lai, F.C. Chu, et al., Highly
The future of graphene-based pH sensors will probably be sensitive palladium oxide thin film extended gate FETs as pH sensor, Sens.
decided by the ability of researchers to produce effective results Actuators B—Chem. 205 (2014) 199–205.
with FETs and GQDs. Although they can be combined with other [17] P. Duroux, C. Emde, P. Bauerfeind, C. Francis, A. Grisel, L. Thybaud, et al., The
ion sensitive field effect transistor (ISFET) pH electrode: a new sensor for
pH-sensitive materials, graphene, GO and rGO are likely to have
long term ambulatory pH monitoring, Gut 32 (1991) 240–245.
a considerable impact on pH sensing as stand-alone pH sensitive [18] T. Guinovart, G. Valdés-Ramírez, J.R. Windmiller, F.J. Andrade, J. Wang,
materials and this entails a deeper understanding of the chemical Bandage-based wearable potentiometric sensor for monitoring wound pH,
Electroanalysis 26 (2014) 1345–1353.
interactions with hydrogen ions.
[19] O. Korostynska, K. Arshak, E. Gill, A. Arshak, Review on state-of-the-art in
polymer based pH sensors, Sensors (Basel) 7 (2007) 3027–3042.
[20] O. Korostynska, K. Arshak, E. Gill, A. Arshak, Review paper: materials and
Competing financial interests techniques for in vivo pH monitoring, IEEE Sens. J. 8 (2008) 20–28.
[21] A. Richter, G. Paschew, S. Klatt, J. Lienig, K.F. Arndt, H.J.P. Adler, Review on
hydrogel-based pH sensors and microsensors, Sensors (Basel) 8 (2008)
The authors declare no competing financial interests. 561–581.
[22] S. Schreml, R.J. Meier, K.T. Weiß, J. Cattani, D. Flittner, S. Gehmert, et al., A
sprayable luminescent pH sensor and its use for wound imaging in vivo,
Acknowledgements Exp. Dermatol. 21 (2012) 951–953.
[23] B. Schyrr, S. Pasche, E. Scolan, R. Ischer, D. Ferrario, J.A. Porchet, et al.,
Development of a polymer optical fiber pH sensor for on-body monitoring
This work was partially supported from the European Commis- application, Sens. Actuators B—Chem. 194 (2014) 238–248.
sion through the SWAN-iCare project (FP7-ICT-317894). [24] N.F. Sheppard Jr., M.J. Lesho, P. McNally, A. Shaun Francomacaro,
Microfabricated conductimetric pH sensor, Sens. Actuators B—Chem. 28
(1995) 95–102.
References [25] W. Shi, X. Li, H. Ma, A tunable ratiometric pH sensor based on carbon
nanodots for the quantitative measurement of the intracellular pH of whole
cells, Angew. Chem. Int. Ed. Engl. 51 (2012) 6432–6435.
[1] C. Bohnke, H. Duroy, J.L. Fourquet, pH sensors with lithium lanthanum [26] K.K. Shiu, F. Song, H.P. Dai, Potentiometric pH sensor with
titanate sensitive material: applications in food industry, Sens. Actuators anthraquinonesulfonate adsorbed on glassy carbon electrodes,
B—Chem. 89 (2003) 240–247. Electroanalysis 8 (1996) 1160–1164.
[2] U. Forstner, G.T.W. Wittmann, Metal Pollution in the Aquatic Environment, [27] B.C. Thompson, O. Winther-Jensen, B. Winther-Jensen, D.R. MacFarlane, A
2nd ed., Springer-Verlag, Heidelberg, Berlin, 1983. solid-state pH sensor for nonaqueous media including ionic liquids, Anal.
[3] R.J. Gillies, N. Raghunand, M.L. Garcia-Martin, R.A. Gatenby, pH imaging. A Chem. 85 (2013) 3521–3525.
review of pH measurement methods and applications in cancers, IEEE Eng. [28] M. Tian, X. Peng, J. Fan, J. Wang, S. Sun, A fluorescent sensor for pH based on
Med. Biol. 23 (2004) 57–64. rhodamine fluorophore, Dyes Pigm. 95 (2012) 112–115.
[4] R.D. Harter, Effect of soil pH on adsorption of lead, copper zinc, and nickel, [29] W. Vonau, V. Guth, pH Monitoring: a review, J. Solid State Electrochem. 10
Soil Sci. Soc. Am. J. 47 (1982) 47–51. (2006) 746–752.
[5] A. Kurkdjian, J. Guern, Intracellular pH: measurement and importance in cell [30] C. Zhao, S. Nie, M. Tang, S. Sun, Polymeric pH-sensitive membranes—a
activity, Annu. Rev. Plant Biol. 40 (1989) 271–303. review, Prog. Polym. Sci. 36 (2011) 1499–1520.
[6] J.L. Slonczewski, M. Fujisawa, M. Dopson, T.A. Krulwich, Cytoplasmic pH [31] C. Zuliani, G. Matzeu, D. Diamond, A potentiometric disposable sensor strip
measurement and homeostasis in bacteria and archaea, Adv. Microb. for measuring pH in saliva, Electrochim. Acta 132 (2014) 292–296.
Physiol. 55 (2009) 1–79. [32] A.K. Geim, K.S. Novoselov, The rise of graphene, Nat. Mater. 6 (2007)
[7] R.G. Bates, Determination of pH: Theory and Practice, 2nd ed., John Wiley & 183–191.
Sons, London, 1973. [33] D.A.C. Brownson, C.E. Banks, Graphene electrochemistry: an overview of
[8] A.K. Covington, R.G. Bates, R.A. Durst, Definition of pH scales standard potential applications, Analyst 135 (2010) 2768–2778.
reference values, measurement of pH and related terminology, Pure Appl. [34] L. Feng, Z. Liu, Graphene in biomedicine: opportunities and challenges,
Chem. 57 (1985) 531–542. Nanomedicine 6 (2011) 317–324.
[9] R.P. Buck, S. Rondinini, A.K. Covington, F.G.K. Baucke, C.M.A. Brett, M.F. [35] M. Pumera, Electrochemistry of graphene: new horizons for sensing and
Camões, et al., Measurement of pH. Definition standards, and procedures energy storage, Chem. Rec. 9 (2009) 211–223.
(IUPAC recommendations 2002), Pure Appl. Chem. 74 (2002) 2169–2200. [36] Y. Song, Y. Luo, C. Zhu, H. Li, D. Du, Y. Lin, Recent advances in
[10] S. Karastogianni, S. Girousi, S. Sotiropoulos, pH: principles and electrochemical biosensors based on graphene two-dimensional
measurement, Encycl. Food Health 4 (2016) 333–338. nanomaterials, Biosens. Bioelectron. 76 (2016) 195–212.
[11] J.R. Sandifer, J.J. Voycheck, A review of biosensor and industrial applications [37] Y. Zhu, S. Murali, W. Cai, X. Li, J.W. Suk, J.R. Potts, et al., Graphene and
of pH-ISFETs and an evaluation of Honeywell’s DuraFET, Mikrochim. Acta graphene oxide: synthesis, properties, and applications, Adv. Mater. 22
131 (1999) 91–98. (2010) 3906–3924.
[12] J.-H. Ahn, J.Y. Kim, M.L. Seol, D.J. Baek, Z. Guo, C.H. Kim, et al., A pH sensor [38] D.R. Cooper, B. D’Anjou, N. Ghattamaneni, B. Harack, M. Hilke, A. Horth,
with a double-gate silicon nanowire field-effect transistor, Appl. Phys. Lett. et al., ISRN condensed matter physics, Exp. Rev. Graphene 2012 (2012) 1–56.
102 (2013) 083701. [39] H. Gao, H. Duan, 2D and 3D graphene materials: preparation and
[13] F. Baldini, Critical review of pH sensing with optical fibres, Proc. SPIE 3540, bioelectrochemical applications, Biosens. Bioelectron. 65 (2015) 404–419.
Chemical, Biochemical, and Environmental Fiber Sensors X (1998) 2–9.
990 P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991

[40] A.T. Lawal, Synthesis and utilisation of graphene for fabrication of [72] Y. Du, S. Guo, Chemically doped fluorescent carbon and graphene quantum
electrochemical sensors, Talanta 131 (2015) 424–443. dots for bioimaging sensor, catalytic and photoelectronic applications,
[41] S. Abdolhosseinzadeh, H. Asgharzadeh, H.S. Kim, Fast and fully-scalable Nanoscale 8 (2016) 2532–2543.
synthesis of reduced graphene oxide, Sci. Rep. 5 (2015) 10160. [73] J. Gu, M.J. Hu, Q.Q. Guo, Z.F. Ding, X.L. Sun, J. Yang, High-yield synthesis of
[42] Y. Si, E.T. Samulski, Exfoliated graphene separated by platinum graphene quantum dots with strong green photoluminescence, RSC Adv. 4
nanoparticles, Chem. Mater. 20 (21) (2008) 6792–6797. (2014) 50141–50144.
[43] Y. Hernandez, V. Nicolosi, M. Lotya, F.M. Blighe, Z. Sun, S. De, et al., [74] X. Li, M. Rui, J. Song, Z. Shen, H. Zeng, Carbon and graphene quantum dots
High-yield production of graphene by liquid-phase exfoliation of graphite, for optoelectronic and energy devices: a review, Adv. Funct. Mater. 25
Nat. Nanotechnol. 3 (2008) 563–568. (2015) 4929–4947.
[44] M. Wang, S.K. Jang, W.J. Jang, M. Kim, S.Y. Park, S.W. Kim, et al., A platform for [75] H. Sun, L. Wu, W. Wei, X. Qu, Recent advances in graphene quantum dots for
large-scale graphene electronics—CVD growth of single-layer graphene on sensing, Mater. Today 16 (2013) 433–442.
CVD-grown hexagonal boron nitride, Adv. Mater. 25 (19) (2013) 2746–2752. [76] L. Wang, Y. Wang, T. Xu, H. Liao, C. Yao, Y. Liu, et al., Gram-scale synthesis of
[45] A.N. Obraztsov, Chemical vapour deposition: making graphene on a large single-crystalline graphene quantum dots with superior optical properties,
scale, Nat. Nanotechnol. 4 (2009) 212–213. Nat. Commun. 5 (2014) 5357.
[46] R. Muñoz, C. Gómez-Aleixandre, Review of CVD synthesis of graphene, [77] Z. Wang, H. Zeng, L. Sun, Graphene quantum dots: versatile
Chem. Vap. Depos. 19 (10-11-12) (2013) 297–322. photoluminescence for energy, biomedical, and environmental applications,
[47] Y. Zhang, L. Zhang, C. Zhou, Review of chemical vapor deposition of graphene J. Mater. Chem. C 3 (2015) 1157–1165.
and related applications, Acc. Chem. Res. 46 (10) (2013) 2329–2339. [78] M. Bacon, S.J. Bradley, T. Nann, Graphene quantum dots, Part. Part. Syst.
[48] C. Mattevi, H. Kim, M. Chhowalla, A review of chemical vapour deposition of Charact. 31 (2014) 415–428.
graphene on copper, J. Mater. Chem. 21 (2011) 3324–3334. [79] L. Li, G. Wu, G. Yang, J. Peng, J. Zhao, J.J. Zhu, Focusing on luminescent
[49] P.K. Ang, W. Chen, A.T. Wee, K.P. Loh, Solution-gated epitaxial graphene as graphene quantum dots: current status and future perspectives, Nanoscale
pH sensor, J. Am. Chem. Soc. 130 (2008) 14392–14393. 5 (2013) 4015–4039.
[50] N. Lei, P. Li, W. Xue, J. Xu, Simple graphene chemiresistors as pH sensors: [80] R. Liu, D. Wu, X. Feng, K.J. Müllen, Bottom-up fabrication of
fabrication and characterization, Meas. Sci. Technol. 22 (2011) 107002. photoluminescent graphene quantum dots with uniform morphology, Am.
[51] C.Y. Lee, K.F. Lei, S.W. Tsai, N.M. Tsang, Development of graphene-based Chem. Soc. 133 (2011) 15221–15223.
sensors on paper substrate for the measurement of pH value of analyte, [81] J. Shen, Y. Zhu, X. Yang, C. Li, Graphene quantum dots: emergent nanolights
BioChip J. 10 (3) (2016) 182–188. for bioimaging, sensors, catalysis and photovoltaic devices, Chem. Commun.
[52] Y. Ohno, K. Maehashi, Y. Yamashiro, K. Matsumoto, Electrolyte-gated 48 (2012) 3686–3699.
graphene field-effect transistors for detecting pH and protein adsorption, [82] F. Yang, M. Zhao, B. Zheng, D. Xiao, L. Wu, Y. Guo, Influence of pH on the
Nano Lett. 9 (2009) 3318–3322. fluorescence properties of graphene quantum dots using ozonation
[53] J. Ristein, W. Zhang, F. Speck, M. Ostler, L. Ley, T. Seyller, Characteristics of pre-oxide hydrothermal synthesis, J. Mater. Chem. 22 (2012) 25471–25479.
solution gated field effect transistors on the basis of epitaxial graphene on [83] P. Huang, J.J. Shi, M. Zhang, X.H. Jiang, H.X. Zhong, Y.M. Ding, et al.,
silicon carbide, J. Phys. D: Appl. Phys. 43 (2010) 345303. Anomalous light emission and wide photoluminescence spectra in graphene
[54] Z. Cheng, Q. Li, Z. Li, Q. Zhou, Y. Fang, Suspended graphene sensors with quantum dot: quantum confinement from edge microstructure, J. Phys.
improved signal and reduced noise, Nano Lett. 10 (2010) 1864–1868. Chem. Lett. 7 (2016) 2888–2892.
[55] M.H. Lee, B.J. Kim, K.H. Lee, I.S. Shin, W. Huh, J.H. Cho, et al., Apparent pH [84] Z.L. Wu, M.X. Gao, T.T. Wang, X.Y. Wan, L.L. Zheng, C.Z. Huang, A general
sensitivity of solution-gated graphene transistors, Nanoscale 7 (2015) quantitative pH sensor developed with dicyandiamide N-doped high
7540–7544. quantum yield graphene quantum dots, Nanoscale 6 (2014) 3868–3874.
[56] I. Heller, S. Chatoor, J. Männik, M.A. Zevenbergen, C. Dekker, S.G. Lemay, [85] B. Shi, L. Zhanga, C. Lana, J. Zhaoa, Y. Sua, S. Zhaoa, One-pot green synthesis
Influence of electrolyte composition on liquid-gated carbon nanotube and of oxygen-rich nitrogen-doped graphene quantum dots and their potential
graphene transistors, J. Am. Chem. Soc. 132 (2010) 17149–17156. application in pH-sensitive photo-luminescence and detection of
[57] W. Fu, C. Nef, O. Knopfmacher, A. Tarasov, M. Weiss, M. Calame, et al., mercury(II) ions, Talanta 142 (2015) 131–139.
Graphene transistors are insensitive to pH changes in solution, Nano Lett. 11 [86] F. Yuan, L. Ding, Y. Li, X. Li, L. Fan, S. Zhou, et al., Multicolor fluorescent
(2011) 3597–3600. graphene quantum dots colorimetrically responsive to all-pH and wide
[58] S.S. Kwon, J. Yi, W.W. Lee, J.H. Shin, S.H. Kim, S.H. Cho, et al., Reversible and temperature range, Nanoscale 7 (2015) 11727–11733.
irreversible responses of defect-Engineered graphene-based [87] C.H. Park, H. Yang, J. Lee, H.H. Cho, D. Kim, D.C. Lee, et al., Multicolor
electrolyte-gated pH sensors, ACS Appl. Mater. Interfaces 8 (2016) 834–839. emitting block copolymer-Integrated graphene quantum dots for
[59] X. Tan, H.J. Chuang, M.W. Lin, Z. Zhou, M.M.C. Cheng, Edge effects on the pH colorimetric simultaneous sensing of temperature, pH, and metal ions,
response of graphene nanoribbon field effect transistors, J. Phys. Chem. C. Chem. Mater. 27 (2015) 5288–5294.
117 (2013) 27155–27160. [88] S.H. Song, M.H. Jang, J. Chung, S.H. Jin, B.H. Kim, S.H. Hur, et al., Highly
[60] B. Zhang, T. Cui, Suspended graphene nanoribbon ion-Sensitive field-effect efficient light-emitting diode of graphene quantum dots fabricated from
transistors formed by shrink lithography for pH/cancer biomarker sensing, J. graphite intercalation compounds, Adv. Opt. Mater. 2 (2014) 1016–1023.
Microelectromech. Syst. 22 (2013) 1140–1146. [89] S.H. Song, M. Jang, H. Yoon, Y.H. Cho, S. Jeon, B.H. Kim, Size and pH
[61] C.X. Lim, H.Y. Hoh, P.K. Ang, K.P. Loh, Direct voltammetric detection of DNA dependent photoluminescence of graphene quantum dots with low oxygen
and pH sensing on epitaxial graphene: an insight into the role of oxygenated content, RSC Adv. 6 (2016) 97990–97994.
defects, Anal. Chem. 82 (2010) 7387–7393. [90] D. Joseph, N. Tyagi, A. Ghimire, K.E. Geckeler, A direct route towards
[62] A.W. Bott, Electrochemistry of semiconductors, Curr. Sep. 17 (1998) 87–91. preparing pH-sensitive graphene nanosheets with anti-cancer activity, RSC
[63] N.M. Zaifuddin, S. Okamoto, T. Ikuta, Y. Ohno, K. Maehashi, M. Miyake, et al., Adv. 4 (2014) 4085–4093.
pH sensor based on chemical-vapor-deposition-synthesized graphene [91] S. Stankovich, D.A. Dikin, G.H. Dommett, K.M. Kohlhaas, E.J. Zimney, E.A.
transistor array, Jpn. J. Appl. Phys. 52 (2013) 06GK04. Stach, et al., Graphene-based composite materials, Nature 442 (2006)
[64] X. Lin, J. Shi, J. Pang, W. Liu, H. Liu, X. Wang, Graphene channel liquid 282–286.
container field effect transistor as pH sensor, J. Nanomater. 2014 (2014) [92] D.R. Dreyer, S. Park, C.W. Bielawski, R.S. Ruoff, The chemistry of graphene
547139. oxide, Chem. Soc. Rev. 39 (2010) 228–240.
[65] M.J. Kiani, M.T. Ahmadi, H. Karimi Feiz Abadi, M. Rahmani, A. Hashim, F.K. [93] H.Y. He, J. Klinowski, M. Forster, A. Lerf, A new structural model for graphite
Che Harun, Analytical modelling of monolayer graphene-based oxide, Chem. Phys. Lett. 287 (1998) 53–56.
ion-sensitive FET to pH changes, Nanoscale Res. Lett. 8 (2013) 173–181. [94] A. Lerf, H. He, M. Forster, J. Klinowski, Structure of graphite oxide revisited, J.
[66] B. Mailly-Giacchetti, A. Hsu, H. Wang, V. Vinciguerra, F. Pappalardo, L. Phys. Chem. B 102 (1998) 4477–4482.
Occhipinti, et al., pH sensing properties of graphene solution-gated [95] H. Bai, C. Li, X. Wang, A pH-sensitive graphene oxide composite hydrogel,
field-effect transistors, J. Appl. Phys. 114 (2013) 084505. Chem. Commun. 46 (2010) 2376–2378.
[67] Q.H. Wang, Z. Jin, K.K. Kim, A.J. Hilmer, G.L. Paulus, C.J. Shih, et al., [96] H. Wu, J.J. Shao, C. Zhang, M.B. Wu, B.H. Li, Q.H. Yang, pH-dependent size,
Understanding and controlling the substrate effect on graphene surface chemistry and electrochemical properties of graphene oxide, New
electron-transfer chemistry via reactivity imprint lithography, Nat. Chem. 4 Carbon Mater. 28 (2013) 327–335.
(2012) 724–732. [97] C.J. Shih, S. Lin, R. Sharma, M.S. Strano, D. Blankschtein, Understanding the
[68] Y. Zhu, C. Wang, N. Petrone, J. Yu, C. Nuckolls, J. Hone, et al., A solid-gated pH-dependent behavior of graphene oxide aqueous solutions: a
graphene FET sensor for pH measurements, Proc. of the 28th IEEE comparative experimental and molecular dynamics simulation study,
International Conference on Micro Electro Mechanical Systems (MEMS) Langmuir 28 (2012) 235–241.
(2015) 869–872. [98] W.S. Hummers Jr., R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem.
[69] S.K. Ameri, P.K. Singh, S.R. Sonkusale, Liquid gated three dimensional Soc. 80 (1958) 1339.
graphene network transistor, Carbon 79 (2014) 572–577. [99] G. Wang, L.T. Jia, B. Hou, D.B. Li, J.G. Wang, Y.H. Sun, Self-assembled
[70] S.K. Ameri, P.K. Singh, S.R. Sonkusale, Three dimensional monolayer graphene monoliths: properties, structures and their pH-dependent
graphene foam for ultra-sensitive pH sensing, 18th International Conference self-assembly behaviour, New Carbon Mater. 30 (2015) 30–40.
on Solid-State Sensors, Actuators and Microsystems (TRANSDUCERS) (2015) [100] S.Y. Qin, X.L. Liu, R.X. Zhuo, X.Z. Zhang, Microstructure-controllable
1378–1380. graphene oxide hydrogel film based on a pH-responsive graphene oxide
[71] S.K. Ameri, P.K. Singh, S.R. Sonkusale, Three dimensional graphene transistor hydrogel, Macromol. Chem. Phys. 213 (2012) 2044–2051.
for ultra-sensitive pH sensing directly in biological media, Anal. Chim. Acta
934 (2016) 212–217.
P. Salvo et al. / Sensors and Actuators B 256 (2018) 976–991 991

[101] S. Sun, P. Wu, A one-step strategy for thermal- and pH-responsive graphene [126] J. Janata, Do optical sensors really measure pH? Anal. Chem. 59 (1987)
oxide interpenetrating polymer hydrogel networks, J. Mater. Chem. 21 1351–1356.
(2011) 4095–4097.
[102] H. Hosseinzadeh, S. Ramin, Magnetic and pH-responsive
starch-g-poly(acrylic acid-co-acrylamide)/graphene oxide superabsorbent Biographies
nanocomposites: one-pot synthesis, characterization, and swelling
behaviour, Starch/Stärke 68 (2016) 200–212.
[103] W. Lv, C.H. You, S. Wu, B. Li, Z.P. Zhu, M. Wang, et al., pH-Mediated Pietro Salvo received the M.Sc. degree in Electronics Engineering in 2004 and the
fine-tuning of optical properties of graphene oxide membranes, Carbon 50 Ph.D. degree in Bioengineering in 2009 from University of Pisa, Italy. He was at the
(2012) 3233–3239. Institute for the Conservation and Promotion of Cultural Heritage (ICVBC-CNR), Italy,
[104] L. Yan, Y.N. Chang, W. Yin, X. Liu, D. Xiao, G. Xing, et al., Biocompatible and from 2008 until 2009. In 2009–2013, he worked on wearable sensors and microflu-
flexible graphene oxide/upconversion nanoparticle hybrid film for optical idic systems at the Centre for Microsystems Technology (CMST), Ghent University,
pH sensing, Phys. Chem. Chem. Phys. 16 (2014) 1576–1582. Belgium. In 2013–2016, he worked on the monitoring of physiological parameters
[105] J.L. Chen, X.P. Yan, Ionic strength and pH reversible response of visible and at the Department of Chemistry, Pisa, Italy. Presently, he is at the Institute of Clinical
near-infrared fluorescence of graphene oxide nanosheets for monitoring the Physiology (IFC-CNR), Italy, where he is working on breath analysis and (bio)sensors.
extracellular pH, Chem. Commun. 47 (2011) 3135–3137.
Bernardo Melai received the M.Sc. degree in Chemistry in 2004, working on inor-
[106] K. Paek, H. Yang, J. Lee, J. Park, B.J. Kim, Efficient colorimetric pH sensor
ganic compounds, and the Ph.D. degree, working on Ionic Liquids, in Science of
based on responsive polymer—quantum dot integrated graphene oxide, ACS
Drug and Bioactive substances in 2009 from Department of Pharmacy, University of
Nano 8 (2014) 2848–2856.
Pisa, Italy. He was at the Department of Pharmacy, Phytochemical Group from 2010
[107] F. Luo, Q. Xi, J. Jianhui, R. Yu, Graphene oxide based DNA nanoswitches as a
until 2012 working on Essential oil by extraction using ionic liquids. In 2013–2016,
programmable pH-responsive biosensor, Anal. Methods 8 (2016)
he worked on the monitoring of physiological parameters at the Department of
6982–6985.
Chemistry, Pisa, Italy.
[108] B. Melai, P. Salvo, N. Calisi, L. Moni, A. Bonini, C. Paoletti, et al., A graphene
oxide pH sensor for wound monitoring, Proc. of 38th Annual International Nicola Calisi graduated in Chemistry at the University of Pisa, Italy. From
Conference of the IEEE Engineering in Medicine and Biology Society (EMBC) 2012–2015, he was a Ph.D. student in Chemistry and Material Science at the same
(1901) 1898–1901. University where he worked on the development of wearable pH and temperature
[109] P. Salvo, V. Dini, F. Di Francesco, M. Romanelli, The role of biomedical sensors. From 2015, he is a postdoc researcher at the Department of Chemistry, Uni-
sensors in wound healing, Wound Med. 8 (2015) 15–18. versity of Pisa, where he is working on the functionalization of nanocomposites, and
[110] P. Salvo, N. Calisi, B. Melai, V. Dini, C. Paoletti, T. Lomonaco, et al., on 2D materials and quantum dots for biological, clinical and industrial applications.
Temperature and pH sensitive wearable materials for Monitoring Foot
Ulcers, Int. J. Nanomed. 12 (2017) 949–954. Clara Paoletti, (Female) graduated in Chemistry at the University of Pisa, Italy, in
[111] P. Salvo, N. Calisi, B. Melai, B. Cortigiani, M. Mannini, A. Caneschi, et al., 2011. After graduation, she worked for the analysis of organic micropollutants at
Temperature and pH sensors based on graphenic materials, Biosens. the Regional Agency for the Environmental Protection of Tuscany (ARPAT), Italy. In
Bioelectron. 91 (2017) 870–877. 2012, she won a Chemistry scholarship at the Italian National Research Council. In
[112] L. Liu, Y. Huang, J. Zhao, Z. Zhu, H. Zhang, C. Wang, All-Solid-Sate pH sensing 2013, she worked as quality control analyst at Lusochimica S.p.A. Presently, she is
material based on cysteic acid/graphene oxide nanocomposites, in: Proc. of a Ph.D. student in Chemistry and Industrial Chemistry, University of Pisa, Italy. She
the 3rd International Conference on Industrial Application Engineering, is involved in the development of wearable sensors for the monitoring of chronic
Kitakyushu (Japan), 2015, pp. 144–148. wounds and gas sensors based on nanocomposites.
[113] J. Liu, L. Tao, W. Yang, D. Li, C. Boyer, R. Wuhrer, et al., Synthesis,
characterization, and multilayer assembly of pH sensitive Francesca Bellagambi obtained a bachelor in Analytical Toxicology and a master’s
graphene-polymer nanocomposites, Langmuir 26 (2010) 10068–10075. degree in Chemistry at the University of Pisa, Italy. From 2012, she is a Ph.D. can-
[114] F. Yang, Y. Liu, L. Gao, J. Sun, pH-sensitive highly dispersed reduced didate at the Department of Chemistry and Industrial Chemistry, University of Pisa,
graphene oxide solution using lysozyme via an in situ reduction method, J. Italy, where she is involved in new analytical methodologies for the quantification
Phys. Chem. C 114 (2010) 22085–22091. of drugs in biological matrices and analysis of biological fluids.
[115] J. Liu, S. Guo, L. Han, W. Ren, Y. Liu, E. Wang, Multiple pH-responsive
Arno Kirchhain obtained a bachelor in Molecular Science and master’s degree in
graphene composites by non-covalent modification with chitosan, Talanta
Chemistry from the Friedrich-Alexander-Universität Erlangen-Nürnberg, Germany,
101 (2012) 151–156.
in 2011 and 2014, respectively. During his studies, his focus was on molecular mate-
[116] L. Ren, T. Liu, J. Guo, S. Guo, X. Wang, W. Wang, A smart pH responsive
rials and organic synthesis. From 2015, he is a Ph.D. candidate at the Department of
graphene/polyacrylamide complex via noncovalent interaction,
Chemistry and Industrial Chemistry, University of Pisa, Italy. His main research inter-
Nanotechnology 21 (2010) 335701–335706.
est is on the development of opto-electronic sensors for enzyme level assessment
[117] D. Pan, J. Zhang, Z. Li, M. Wu, Hydrothermal route for cutting graphene
in chronic wounds.
sheets into blue-luminescent graphene quantum dots, Adv. Mater. 22
(2010) 734–738. Maria Giovanna Trivella is a Senior Investigator of Italian National Research Coun-
[118] G. Eda, Y.Y. Lin, C. Mattevi, H. Yamaguchi, H.A. Chen, I.S. Chen, et al., Blue cil (CNR), Head of Experimental Laboratory of IFC-CNR and Head of Experimental
photoluminescence from chemically derived graphene oxide, Adv. Mater. 22 Surgery Laboratory, a joint laboratory in cooperation between CNR and University of
(2010) 505–509. Pisa, Italy. From 2015, she is the head of the IFC-CNR Clinical Research Unit in Milan,
[119] S. Srinives, W. Klongthong, K. Selamassakul, N. Peaunbida, S. Jiamjitton, C. Italy. She received a Medical Doctor Degree from the University of Pisa in 1976 and
Phamornnak, et al., A development of a graphene based chemiresistive she majored in Cardiology in 1981. Her research activity is focused on pathophys-
sensor: demonstrations on pH sensing and cell detection, Adv. Mat. Res. iology of myocardial ischemia and infarction, hemodynamics monitoring, breath
1103 (2015) 137–143. and sweat as disease markers, personality and psychological characterizations in
[120] I.Y. Sohn, D.J. Kim, J.H. Jung, O.J. Yoon, T.N. Thanh, T.T. Quang, et al., pH patients waiting for organ transplantation, sensors for biomedical applications.
sensing characteristics and biosensing application of solution-gated reduced
graphene oxide field-effect transistors, Biosens. Bioelectron. 45 (2013) Roger Fuoco is full professor of Analytical Chemistry at the Department of Chemistry
70–76. and Industrial Chemistry, University of Pisa, Italy, since 1994. He has been the co-
[121] Y.R. Li, S.H. Chang, W.L. Tsai, C.T. Chang, K.Y. Wang, P.Y. Yang, et al., Highly ordinator of a research programme on the development of analytical methodologies
sensitive pH sensors of extended-gate field-effect transistor with the funded by the CNR for many years, and the scientific supervisor of research units
oxygen-functionalized reduced graphene oxide films on reverse pyramid granted by the CNR national project on the special project on the “Safeguard of the
substrates, IEEE Electron. Device Lett. 36 (2015) 1189–1191. Venice Lagoon”. Prof. Fuoco is a member of the Executive Committee of the Analytical
[122] Y.R. Li, S.H. Chang, C.T. Chang, W.L. Tsai, Y.K. Chiu, P.Y. Yang, et al., Chemistry Division of the S.C.I. He is the author of about 100 scientific publications.
High-sensitivity extended-gate field-effect transistors as pH sensors with
oxygen-modified reduced graphene oxide films coated on different Fabio Di Francesco graduated in Physics and received a Ph.D. in bioengineering
reverse-pyramid silicon structures as sensing heads, Jpn. J. Appl. Phys. 55 at the Interdepartmental Center “E. Piaggio” – University of Pisa, where he was
(2016) 04EM08. involved in researches concerning multisensor systems for environmental and food
[123] C.K. Chua, M. Pumera, Chemical reduction of graphene oxide: a synthetic applications. In 2000–2006 he worked at the Institute of Clinical Physiology – CNR,
chemistry viewpoint, Chem. Soc. Rev. 43 (2014) 291–312. where he further extended his research on sensors to the biomedical field. Since
[124] D. Jiang, Y. Chen, N. Li, W. Li, Z. Wang, J. Zhu, et al., Synthesis of luminescent 2006, he is researcher at the Department of Chemistry and Industrial Chemistry of
graphene quantum dots with high quantum yield and their toxicity study, the University of Pisa. His current research activity is focused on the development
PLoS One 10 (2015) e0144906. of chemical and physical sensors, wearable sensor systems, and the identification
[125] J. Sun, S. Yang, Z. Wang, H. Shen, T. Xu, L. Sun, et al., Ultra-high quantum of chemical markers in breath, saliva and sweat for the diagnosis of disease and the
yield of graphene quantum dots: aromatic-nitrogen doping and monitoring of health conditions.
photoluminescence mechanism, Part. Part. Syst. Charact. 32 (2015) 434–440.

You might also like