Electrochemical Reduction of Graphene Oxide and Its in Situ

download Electrochemical Reduction of Graphene Oxide and Its in Situ

of 7

Transcript of Electrochemical Reduction of Graphene Oxide and Its in Situ

  • 7/30/2019 Electrochemical Reduction of Graphene Oxide and Its in Situ

    1/7

    This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys.

    Cite this: DOI: 10.1039/c2cp42253k

    Electrochemical reduction of graphene oxide and its in situ

    spectroelectrochemical characterization

    Antti Viinikanoja,*a Zhijuan Wang,b Jussi Kauppilaac and Carita Kvarnstro m*a

    Received 3rd July 2012, Accepted 20th August 2012

    DOI: 10.1039/c2cp42253k

    The electrochemical properties of self-assembled films of graphene oxide (GO) on

    mercaptoethylamine (MEA) modified rough Au-surfaces were studied. The film deposition

    process on MEA primed gold was followed by surface plasmon resonance measurements and the

    film morphology on 3-aminopropyltriethoxysilane primed Si(100)-surface was studied by atomic

    force microscopy. The deposited few layer thick GO films on gold were electrochemically reduced

    by cyclic voltammetry simultaneously as the structural changes in the film were recorded byin situ vibrational spectroscopies. In situ surface enhanced infrared spectroscopy results indicate

    that the effect of the applied potential on the GO structure could be divided into two parts where

    the changes occurring at moderate negative potentials are mainly related to changes in the double

    layer at the filmelectrolyte interface and to hydrogen bonding of intercalated water between the

    GO sheets. At potentials more negative than 0.8 V vs. Ag/AgCl the reduction of GO starts to

    take place with concomitant conversion of the different functional groups of the film.

    Introduction

    A pristine graphene sheet possesses high electron conductivity1

    and fast heterogeneous electron transfer rate at the graphene

    sheet edges,2 properties that are crucial for application in

    electrochemical devices in energy storage or production or inelectrochemical sensors of different kinds.35 Since the revival

    of the investigation of graphene in 20046 chemical exfoliation

    of bulk graphite by strong oxidants or strong acid intercalation

    has been developed with the expectation of obtaining graphene

    like materials on a large scale and at low cost.3,4 The harsh

    treatment of graphite in the production of bulk graphite or

    graphene oxide (GO) excludes or depresses the properties found

    in pristine graphene sheets. As a result, the graphite oxide or

    GO is a highly hygroscopic, electrically insulating material,

    and its stoichiometry is shown to depend on the method of

    preparation and on the kind of graphite used.7 Subsequently,

    the chemically oxidized graphite must be reduced by chemical,

    thermal, or electrochemical means to retain at least some of

    the graphene structure.8,9

    The electrochemical reduction was introduced as a fast

    and clean way (i.e. no contamination from reducing agents)

    for producing a partly recovered sp2 lattice.10 Especially in

    manufacturing flexible reduced graphene oxide (rGO) films

    with complex patterns of a specific thickness electrochemical

    reduction has shown its potential.11 Graphite or GO can be

    electrodeposited and simultaneously reduced as thin conduct-

    ing films from buffered aqueous solutions in a wide pH

    range12,13 or from ionic liquids.14 The ionic surface groups

    on the GO sheets enable a thin film preparation by the

    electrostatic layer-by-layer self-assembly technique.15 The

    method of fabrication and the way of reduction of the GO

    film influence the final rGO product and its electrochemistry.

    In comparison to thermally reduced graphene oxide that

    contains a large amount of structural defects the chemically or

    electrochemically reduced GO contains a higher amount of

    oxygen containing groups.16 Electrochemical reduction has

    been reported to only partially remove the oxygen containing

    functionalities leaving groups mainly in the basal plane due to

    a faster electron transfer rate around the edges of the GO

    sheets.17 While electrochemistry alone cannot give details

    about structural changes in thin films it is often applied in

    combination with a spectroscopic technique.18,19 We have

    previously used surface sensitive IR techniques to follow the

    formation of self-assembled mono- and multilayers to probe

    the structure, organization, chemical composition, and inter-

    actions within these thin films.20,21 Previously, the thermal

    reduction of single and multilayer GO was studied by these IR

    techniques after the reduction process and the effect of inter-

    calated water between the GO sheets on the final reduced

    a Turku University Centre of Materials and Surfaces (MatSurf),Laboratory of Materials Chemistry and Chemical Analysis,University of Turku, Vatselankatu 2, 20014 Turku, Finland.E-mail: [email protected], [email protected];Fax: +358-2-333 6700; Tel: +358-2-333 6714, +358-2-333 6729

    b School of Materials Science and Engineering, NanyangTechnological University, 50 Nanyang Avenue, Singapore 639798,Singapore

    c The Finnish National Graduate School in Nanoscience(NGS-NANO), Nanoscience Center, P. O. Box 35, 40014,University of Jyvaskyla, Finland

    PCCP Dynamic Article Links

    www.rsc.org/pccp PAPER

    View Online / Journal Homepage

    http://dx.doi.org/10.1039/c2cp42253khttp://pubs.rsc.org/en/journals/journal/CPhttp://dx.doi.org/10.1039/c2cp42253khttp://dx.doi.org/10.1039/c2cp42253khttp://dx.doi.org/10.1039/c2cp42253k
  • 7/30/2019 Electrochemical Reduction of Graphene Oxide and Its in Situ

    2/7

    Phys. Chem. Chem. Phys. This journal is c the Owner Societies 2012

    structure was discussed. The intercalated water was found to

    affect the defect formation and carbonyl formation in multi-

    layered graphene oxide films.22,23 By combining infrared

    absorption spectroscopy and electrochemistry, the evolution

    of oxygen containing defects can be followed in situ during the

    electrochemical reduction of GO.

    In this work in situ Surface Enhanced Infrared Spectroscopy

    (SEIRAS)24 was used with quasi-single crystalline metal films

    deposited on high refractive materials in attenuated total

    reflection (ATR) geometry together with cyclic voltammetry.

    SEIRAS enhances the signal-to-noise sensitivity by a factor of

    1050 higher than that of IR reflection absorption spectro-

    scopy which makes it a superior technique in studying reaction

    processes on surfaces. We focus on the electrochemical

    reduction in the aqueous media of a few layers of GO

    deposited by self-assembly on the roughened gold plated

    reflection element. Using the ATR geometry i.e. entering the

    film from the backside through the substrate we can minimize

    the interference from electrolyte contribution to the spectra

    and get the main signal from intercalated water and the

    potential dependent change in functional groups in the GO.

    In situ surface enhanced Raman spectroscopy (SERS) and

    atomic force microscopy (AFM) were further applied for rGO

    characterization.

    Experimental methods

    Chemicals

    3-Aminopropyltriethoxysilane (Fluka AG, 96%) (APTES), and

    2-mercaptoethylamine hydrochloride (Aldrich, 98%) (MEA) were

    used for the preparation of self-assembled monolayers. KCl

    (FF-Chemicals, 99.5%), and NaF (J. T. Baker, Baker analyzed)

    were used with Millipore water as electrolytes in electrochemical

    and spectroelectrochemical measurements. The chemical reduction

    of GO was done with hydrazine monohydrate (Aldrich, 98%).

    HF (J. T. Baker, 48%), NH4F (Merck, p.a.), potassium tetra-

    chloroaurate (Aldrich, 98%), sodium sulfite Na2SO3 (Merck, p.a.),

    sodium thiosulfate Na2S2O35 H2O (FF-Chemicals, p.a.), and

    NH4Cl (Acros Organics, 99%) were used in the gold deposition

    process. 1,1,2-Trichloroethylene (Merck, 99%), MeOH, EtOH

    were of analytical grade or higher and were used as received for

    the cleaning of the Si-hemicylinder.

    Synthesis of GO

    Natural graphite (SP-1, Bay Carbon) was used for synthesizing

    the graphite oxide by the modified Hummers method and was

    further dispersed in water to form GO as previously reported.25

    In brief, graphite flakes (0.3 g) were chemically oxidized using a

    mixture of concentrated H2SO4 (2.4 mL), K2S2O8 (0.5 g), and

    P2O5 (0.5 g) and kept at 80 1C for 4.5 h. This preoxidised

    product was washed with distilled water to remove all traces of

    acid and dried. The dried product was dropped into cold (0 1C)

    concentrated H2SO4 (12 mL). Then, KMnO4 (1.5 g) was slowly

    added and the temperature of the mixture was kept below

    20 1C. After the addition, the solution temperature was kept

    at 35 1C for 2 h and was diluted with distilled water (25 mL).

    The mixture was stirred for 2 h; an additional 70 mL of H2O

    was added to the solution. Shortly after the dilution 2 mL of

    30% H2O2 was added to the mixture resulting in a bright yellow

    coloured solution. The resulting mixture was washed with HCl

    and distilled water after which graphite oxide was obtained.

    The obtained graphite oxide was dispersed in water at a

    certain concentration and subsequently sonicated to obtain

    GO. UV/Vis (GO): 231(pp*), B300 nm (np*); rGO,

    reduced with hydrazine: 270 (pp*), >300 nm (increase in

    absorbance).26,27 IR (graphite oxide): 3615 (sh, COH), 3410

    (COH), 3210 (H2O), 1740 (CQO), 1620 (CQC, H2O), 1049

    (CO), 14001300 (OH, COH), 12001000 (ArOH), 986

    (sh, COC), 848 (COOC) cm1.28

    Preparation of the substrates

    An Au-minielectrode (CHI101, 2 mm dia., CH Instruments,

    Inc.) was mechanically polished with an EtOH wetted polishing

    cloth (Struers, DP-Nap) using diamond paste of gradually

    finer grades (Struers, 3, 1, and 0.25 mm) with careful wash by

    EtOH in between. The mechanically polished electrodes were

    electrochemically cleaned in 0.5 M H2SO4 (Aldrich, Suprapur)

    using cyclic voltammetry recording 50 potential cycles between

    0.0 and 1.2 V at 100 mV s

    1 vs. a Hg/HgSO4 referenceelectrode. In SERS measurements, the Au-minielectrode was

    additionally roughened electrochemically in 0.1 M KCl before

    the electrochemical cleaning step based on the previously

    reported procedure.29 Microscope glass slides and Si(100)

    wafers (Okmetic, Finland) for polarization modulation infrared

    reflection-absorption spectroscopy (PM-IRRAS) and AFM

    measurements, respectively, were cleaned in fresh piranha

    solution (concentrated H2SO4/30% H2O2 (3 : 1)), rinsed

    thoroughly with Millipore water and dried. A gold film

    (ca. 100 nm) was evaporated on microscope slides using the

    Edwards E306A coating system. The microscope slides were

    previously covered by an evaporated chromium layer (ca. 5 nm)

    for improving the adhesion of Au. The reflecting plane of thesilicon hemicylinder (Harrick Scientific) was cleaned by

    soaking in hot 1,1,2-trichloroethylene, acetone, and MeOH for

    10 min in each, thereafter etching by 1 min treatment in argon

    saturated 10% HF (aq.). The oxide layer on the surface was

    removed by soaking in argon saturated 40% NH4F (aq.) for

    6 min.30 A thin gold film was deposited chemically on

    the cleaned silicon hemicylinder according to a previous

    report by H. Miyake et al.31 In brief, the reflecting plane of

    the silicon hemicylinder was put into contact with a mixture of

    2% HF and a plating solution of 0.015 M KAuCl 4 + 0.15 M

    Na2SO3 + 0.05 M Na2S2O35 H2O + 0.05 M NH4Cl

    (aq.) (2 : 1 v/v) for 15 minutes at room temperature.

    The hemicylinder was rinsed with Millipore water to finishthe deposition.

    Deposition of the GO film

    Positive charge on substrates, required for GO adsorption,

    was accomplished by immersing cleaned silicon or gold

    substrates in 1% (v/v) APTES in EtOH for 30 min32 or in

    1 mM MEA in H2O for one hour, respectively. Self-assembled

    thin films were deposited on the primed substrates from

    dilute aqueous GO solutions. For this purpose solid graphite

    oxide was dispersed in Millipore water (0.51 mg mL1)

    with the aid of mild ultrasound treatment for 30 min.

    View Online

    http://dx.doi.org/10.1039/c2cp42253k
  • 7/30/2019 Electrochemical Reduction of Graphene Oxide and Its in Situ

    3/7

    This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys.

    Additionally, the resulting yellow suspension was centrifuged

    at 4500 rpm (3939g) for 30 min. The pH of the supernatant was

    set at 2.5 with H2SO4 and used for the thin film deposition. The

    adsorption time of 240 min and pH 2.5 of the GO solution were

    based on a previously reported procedure and on surface

    plasmon resonance (SPR) measurements (vide infra).

    Instrumentation

    The UV-vis spectra of the GO and chemically reduced GO

    solutions were recorded using a Hewlett-Packard 8453 spec-

    trophotometer. The self-assembly process of GO from dilute

    aqueous solution was followed by a Texas Instruments Spreeta

    integrated surface plasmon resonance affinity sensor.33 The

    electrochemical reduction of the thin GO films on gold

    electrodes was performed by cyclic voltammetry in deaerated

    0.1 M NaF aqueous solutions at a scan rate of 10 mV s1 using

    a standard calomel reference electrode (+43 mV vs. Ag/AgCl/

    3.5 M KCl reference system) and a platinum wire as an

    auxiliary electrode. PM-IRRAS differential reflectance spectra

    and SERS spectra of the thin self-assembled films on primed

    gold were measured using a dry-air-purged Nexus 870 FTIR

    spectrometer (Nicolet) equipped with a Tabletop Optical

    Module (TOMt) for Nicolett FT-IR spectrometers for

    polarization modulation experiments and FT-Raman

    accessory. For each PM-IRRAS spectrum, 4096 scans at

    4 cm1 spectral resolution were collected. SEIRAS spectra

    were measured using the Kretschmann attenuated total reflec-

    tion (ATR) configuration24 in the same TOM configuration as

    the PM-IRRAS spectra without the polarization modulation

    of the IR-signal. The p-polarized IR-beam passed at an angle

    of incidence of 701 relative to the surface normal through a

    silicon hemicylinder mounted on a homemade spectroelectro-

    chemical cell. The total reflected beam from the thin film was

    detected using an HgCdTe detector. All IR spectra shown are

    differential spectra where the previous spectrum has been used

    as reference for the following one in the in situ recorded

    spectral array. SERS spectra were obtained from the thin

    films deposited on roughened and primed gold electrodes

    placed in a homemade spectroelectrochemical cell equipped

    with a glass window using a Nd:YAG laser at 1064 nm in 1801

    backscattering geometry and a germanium detector. For each

    SEIRAS and SERS spectrum, 1024 scans were collected

    at constant potential using 4 cm1 spectral resolution. An

    Ag/AgCl-minireference-electrode (Cypress Systems, Inc., leak-

    free) and a platinum wire served as reference and counter

    electrodes. The potentiostat was an EG & G 283 (Princeton

    Applied Research). Argon purged 0.1 M NaF aqueous

    solutions were used as an electrolyte. The surface morphology

    of MEA/GO thin films on APTES modified silicon was

    investigated using a Veeco diCaliber AFM instrument (Veeco

    Instruments Inc.) under ambient conditions in the tapping

    mode. Rectangular silicon cantilevers (TESPA) of 125 mm

    length having a resonance frequency of approximately

    320 kHz were used. The tip height was 1015 mm and the

    nominal radius of curvature 8 nm. The raw data obtained were

    processed by flattening and plane fitting with the SPMLab-

    Analysis 7.00 (Veeco Instruments, Inc.) and Gwyddion

    2.19 data visualization and analysis software.

    Results and discussion

    The adsorption of GO on a MEA primed gold surface was

    followed by SPR measurements in order to optimize the condi-

    tions for the self-assembly process, namely the adsorption time,

    concentration and pH of the GO deposition solution. The pKavalue of the NH2-group in a MEA monomer is close to 7.6,

    3436

    suggesting that the amine group is mainly protonated during the

    GO adsorption process at pH 6 and 2.5 making the surfacepositively charged. Attracted by the opposite charge, GO sheets

    adsorb on top of the MEA layer attached to the gold layer. This

    self-assembly process is terminated when enough of negatively

    charged GO platelets are adsorbed on the surface to reverse

    the surface charge. The change in pH from 6 to 2.5 has an

    effect on this equilibrium. At pH 2.5, the protonation degree of

    the NH2-groups of MEA is higher increasing the positive

    surface charge. On the other hand, the carboxylic acid groups

    responsible for the negative charge of the GO sheets (surface

    charge density is 0.4 units per 100 A 2)15 are mostly protonated

    at this pH, which lowers the negative surface charge density

    of the platelets. As a result, more GO sheets are required

    to reverse the surface charge at lower pH which is seen asa greater change in the refractive index of the SPR curves at

    pH 2.5 in Fig. 1.

    Additionally, at lower concentrations (but the same pH) less

    material is adsorbed on the electrode at the same time indicating a

    diffusion limited mass transport process. SPR data revealed that

    the growth of the GO film levels-off after ca. 80 min (ca. 15 min at

    lower concentration). A small drift in the SPR signal after the

    plateau is comparable to that observed in experiments done

    with pure solvent and is related to instrumental and other

    e.g. temperature and diffusion effects. Based on this and

    previously reported results an adsorption time of 240 min,

    cGO = 0.51 mg mL1 and a pH of 2.5 was chosen.34

    Fig. 2a and b show the AFM images of a few layers thickfilm of GO on an APTES modified silicon surface after short

    (30 min) and long (240 minutes) deposition times. The line in

    the figures indicates the position from where the height profiles

    below the images are obtained. After 30 minutes, the surface is

    unevenly covered with relatively thin particles whose heights

    are below 4.0 nm and a few bigger particles with height rising

    up to 14.4 nm. After 240 minutes, the surface is more evenly

    covered with GO particles whoseheights are below 3.4 nm although

    Fig. 1 SPR curves obtained from a MEA modified gold surface upon

    interaction with aqueous 0.01 mg mL1 GO suspension at pH 6.0 (&),

    pH 2.5 (B), and with 1 mg mL1 GO suspension at pH 2.5 (J).

    View Online

    http://dx.doi.org/10.1039/c2cp42253k
  • 7/30/2019 Electrochemical Reduction of Graphene Oxide and Its in Situ

    4/7

    Phys. Chem. Chem. Phys. This journal is c the Owner Societies 2012

    some regions with higher height profiles are visible as well.

    Both profiles indicate films that are thicker than a single

    GO layer. The widths of the largest particles are about

    150300 nm. The thickness of a single graphene sheet is

    approx. 0.8 nm26,37 and a GO sheet that contains oxygen-rich

    defects has a thickness of ca. 1.2 nm.32,3841 According to this

    the highest spot within a smooth region in our sample would

    consist of three layers of GO sheets if one considered a sheet

    thickness of 1.21.3 nm that is in agreement with the above

    and with the step height of 1.01.3 nm that can be observed in

    some GO sheets. Generally, the film is up to ca. 4 nm thick

    with occasional multilayer GO structures. Importantly, the

    AFM results show a film that consists mainly of a few layer

    GO sheets and occasional multilayer GO particles. It has been

    reported that intercalated water has a significant impact on the

    course of reduction processes in similar kinds of films,22 which

    is also confirmed by SEIRAS results (vide infra).

    Vibrational spectral characterization

    The PM-IRRAS results of the MEA-primed gold substrate in

    Fig. 3 show that the asymmetric and symmetric methylene

    stretching vibrations belonging to a short ethylene chain in the

    monolayer appear at 2925 and 2856 cm1, respectively. Upon

    the adsorption of a GO layer the relative intensity of the

    aforementioned bands increases probably due to the orienta-

    tion effect of the monolayer. The bulk spectrum of graphite

    oxide (data not shown) did not reveal any absorption bands

    at these wavenumbers. An additional broad band around

    ca. 3300 cm1 is assigned to OH stretching vibrations related

    to adsorbed water molecules and structural OH groups. In

    the lower wavenumber range the spectrum of the MEA

    monolayer is featureless whereas after GO adsorption several

    peaks appear below 1800 cm1. The vibration at 1739 cm1

    can be related to CQO stretching from carboxylic acid groups.

    Vibrations belonging to the carboxylate group are seen at

    1541 and 1465 cm1. The relative intensity of the CQO

    absorbance (vs. bulk) has decreased too indicating the ioniza-

    tion of the acidic groups. This is expected because some of the

    surface groups are to be negatively charged in order to enable

    a successful adsorption of the GO sheets on the positively

    charged electrode surface. The band around 1662 cm1 is

    usually assigned to either the CQC vibrations of the basal

    plane of graphitic regions or oxygen containing carbon species

    other than carboxylic acids.42,43 Additionally, the possibility

    for traces of water remaining in the film structure, giving rise

    to absorptions from HOH bending (ca. 1620 cm1), has to

    be taken into account. Hydroxyl or ether (CO) related

    vibrations contribute to the weak peak at 1087 cm1.28

    In situ spectroelectrochemical characterization of the GO

    reduction

    Surface enhanced Raman spectroscopy. Raman spectroscopyrepresents one frequently used technique for the characteriza-

    tion of carbon-based materials and graphene structures in

    particular.44,45 The Raman spectra of GO at different potentials

    are shown in Fig. 4a. All Raman spectra exhibit two prominent

    features, the G-band (B1599 cm1), corresponding to the sp2

    lattice, and the D-band (B1305 cm1), which is related to the

    defects in the sp2 lattice. For a better comparison, all signal

    intensities are normalized to the G-band (1599 cm1) intensity.

    From the presented Raman spectra, it is clear that the intensity

    of the D-band, which is related to disorder structures in the sp2

    lattice, is high in graphene oxide (D/G ratio of 1.41 in Fig. 4b)

    as it exhibits the large abundance of oxygenated functionalities

    that disrupt the planar sp2 structure. In Fig. 4b, the electro-chemical reduction of GO products exhibits D/G ratios between

    1.59 (at 0.4 V) and 1.29 (at 1.0 V), suggesting that the order

    of graphene is partly recovered during the electrochemical

    Fig. 2 The AFM images of Si/APTES/GO (a) after 30, and

    (b) 240 minutes adsorption time from 0.5 mg mL1 aqueous solution.

    A line in an AFM image indicates the position of the corresponding

    height profile.

    Fig. 3 The PM-IRRAS spectrum of Au/MEA (- - -) and Au/MEA/

    GO (___) before the reduction.

    Fig. 4 (a) The normalized Raman spectra of the Au/MEA/GO film

    as a function of applied potential. The spectra are shifted for clarity.

    (b) The D/G-intensity ratio as a function of applied potential. (c) The

    absolute Raman spectra of the Au/MEA/GO film at 0.0 V before (- - -)

    and after (___) the reduction. (d) Cyclic voltammogram of Au/MEA/

    GO in 0.1 M NaF, a scan rate of 10 mV s1. The arrow indicates the

    direction of the potential sweep.

    View Online

    http://dx.doi.org/10.1039/c2cp42253k
  • 7/30/2019 Electrochemical Reduction of Graphene Oxide and Its in Situ

    5/7

    This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys.

    reduction that is consistent with previous reports.46,47 Based

    on Raman spectra, it is possible to calculate the average

    crystallite size (La) from the observed D/G ratio.48 The D/G

    ratio is inversely proportional to the crystallite size and for

    graphene oxide and electrochemically reduced GO the sizes

    16 nm and 15 nm, respectively, have been reported.16 Similarly,

    an increase in the D/G ratio up to 0.4 V indicates a degrada-

    tion of the crystallite size after which the size of the graphitic

    regions grows slightly at higher negative potentials as indicated

    by the decrease in the D/G ratio. The G band position was

    shifted by 8 cm1 from 1599 to 1591 cm1 with increasing

    applied potential. This shift in the G band is in line with the

    reduction of GO.49 The D band also shows a similar variation

    in its position from 1305 to 1295 cm1 in Fig. 4c. This is in

    consistence with the data reported by Ramesha and Sampath

    from studies in the identical potential range.34

    Surface enhanced IR-reflection absorption spectroscopy. The

    reduction of GO to rGO in aqueous solutions was followed

    spectroelectrochemically during stepwise increase in applied

    potential towards more negative values between 0 and 1.0 V.

    Assuming that the spectral treatment and the use of an

    internal reflection set-up compensate the contribution from

    the environment and the electrolyte in the cell upward extending

    peaks show the gain and downward extending peaks the loss in

    the structure of the film during electrochemical reduction shown

    in Fig. 5a. The dashed line in Fig. 5a shows the total change in

    structure induced during the potential cycle. The vibrations

    related to structural OH groups in GO are overlapped with the

    vibrations of intercalated and/or free water molecules. A lot of

    efforts have been made to distinguish between the COH modes

    originating from a wide range of functional groups on the GO

    sheets all overlapping in the range of 36003200 cm1.23,50,51

    Within the band above 3200 cm1 at least four main vibrations

    at around 3200 (overtones of scissor vibrations of adsorbed

    water molecules), 3370 (closely neighbouring hydroxyl groups),

    3500 (carboxyl groups of GO), and 3625 cm1 (COH stretch

    associated with five-membered-ring lactols and hydroxyl groups

    of GO from both the basal plane and from the sheet edges) can

    be distinguished.23,50 The plot in Fig. 5b shows the potential

    induced change in absorbance of the aforementioned bands at

    specific potential related to the previous applied potential. The

    potential induced change in absorbance can be divided into two

    sections, the first covering the potential range from 0 to 0.8 V

    where only a minor increase in the intensity of all the vibration

    modes takes place. Within this potential range mainly changes

    in the double layer at the interface between the GO layer and

    the electrolyte are detected together with the enhancement of the

    absorbance due to hydrogen bonding interactions (hydrogen

    bridges) with COH and COOH groups on the GO sheets.

    These modes are known to be more enhanced in a multilayer

    film due to the contribution of intercalated water.22,28,50,52,53The

    multilayer nature of the investigated GO films is also supported

    by the AFM results in Fig. 2. The reduction of GO involving the

    conversion of the functional groups starts to take place at

    potentials more negative than 0.8 V. This is seen as an abrupt

    change observed above 3200 cm1 where both the width and

    the intensity of the absorption bands strongly increase. This

    coincides well with the significant reduction current increase

    observed in the cyclic voltammogram of the GO film at around

    0.8 V vs. Ag/AgCl in Fig. 4d. At 1.0 V finally, a part of the

    broad band develops a broad negative band with its centre at

    3650 cm1 while the rest of the band remains positive.

    In the lower wavenumber region (Fig. 5a) the same potential

    dependent increase in absorbance is also seen at 1640 cm1

    where a rather broad band is growing. This is the range for the

    asymmetric vibrational stretching of sp2-hybridized CQC,

    non-oxidized or reconstructed graphitic domains.10,22 Within

    this same range also scissor modes of water that are broadened

    through hydrogen interaction with COH are visible. Additionally

    also vibrations from edge carbonyl groups might be visible in

    this region. An additional weak feature arising at 1444 cm1

    can be assigned to the deformation vibrations of carboxyl and

    COH groups (seen only in the spectrum measured upon

    completion of the potential cycle in Fig. 5a).23 Simultaneously

    with the abovementioned increasing vibrations, a negative band

    from vanishing structures appears at around 1220 cm1. The

    asymmetric shape of the peak at 1224 cm1 indicates that

    the band consists of several oxygen related vibrations. This is

    the region for epoxides COC but also for peroxides, ethers,

    ketones, lactols and benzoquinones.22,28

    The response from the intercalated water between the GO

    sheets resulting in further overlapping in the high wavenumber

    region in the IR spectra can be compared with the results

    reported from water in polymer membranes.54 In a LB film of

    a polymer, monomeric non-hydrogen-bonded water shows up

    at 3616 cm1, hydrogen-bonded dimers and small hydrogen-

    bonded clusters have their vibration modes at 3536 cm1 and

    3424 cm1. A fourth mode at 3246 cm1 contains a wide

    Fig. 5 (a) Differential SEIRA spectra of the Au/MEA/GO film in

    0.1 M NaF solution; the reference and sample potentials were 0.0 and

    0.2 (1), 0.2 and 0.4 (2), 0.4 and 0.6 (3), 0.6 and 0.8 (4),

    0.8 and 1.0 V (5) vs. Ag/AgCl, respectively. Curve (6) shows the

    difference in absorbance at 0.0 V vs. Ag/AgCl before and after the reduction

    cycle. (b) The potential dependence of GO (3500 and 3625 cm1) and H2O

    (3200 and 3370 cm1) related vibrational bands.

    View Online

    http://dx.doi.org/10.1039/c2cp42253k
  • 7/30/2019 Electrochemical Reduction of Graphene Oxide and Its in Situ

    6/7

    Phys. Chem. Chem. Phys. This journal is c the Owner Societies 2012

    distribution of hydrogen-stretch vibrations associated with

    water chains. The effect of hydrogen bonding causing

    enhancement of the IR intensity as well as the broadening of

    the OH stretch band is as strongest in the lower wavenumber

    region in the 36003200 cm1 band, i.e. 34003200 cm1.

    The electrochemical reduction of epoxides, carboxylates,

    ethers, etc. has been mostly studied in organic solvents and the

    reduction potentials reported for these species lie close to each

    other and are therefore hard to identify in a mixture.55,56 In

    aqueous solution, the potential window is narrow and the order

    of easiness of the electrochemical reductions of the different

    functional groups in the GO layers might be due to the access

    of electrons and the electrolyte to these sites. According to the

    LerfKlinowski model7 the major components in the basal plane

    of the GO sheets are epoxide and hydroxyl groups.57,58 The

    epoxide group can be considered as the most easily reduced

    functional group in the GO film.59 The reduction potential of

    aromatic epoxide groups has been reported to occur around

    0.75 to 1.5 V vs. SCE in aqueous solutions.6062 The electro-

    chemical reduction of epoxides CHOCH leads mainly to

    CHQCH, CH2CH2 or CH2CHOH slightly depending

    on the medium and its pH where the reduction is made.55

    Aromatic carboxylic acids are converted to the corresponding

    carboxylate anion, aldehyde, alcohol, or hydrocarbon depending

    on the pH and the number of electrons consumed upon electro-

    chemical reduction.56,63,64 Moreover, in aqueous solutions the

    formed aldehydes may exist predominantly as the hydrate. The

    formed hydroxyl groups COH originating from the reduction

    of the epoxides and carboxyl groups can explain the increase in

    absorbance intensity at 36003640 cm1 at the beginning of the

    reduction. The simultaneous reduction of intercalated water

    contributes to the increase in OH species.

    ConclusionsBased on the experimental results obtained from thin GO films

    prepared by the self-assembly technique utilizing electrostatic inter-

    actions an overview of the potential dependent electrochemical

    reduction process could be presented. Due to the fact that the redox

    reactions of the different oxygen containing functional groups lie

    very close to each other a complete identification of each group by

    potential was not possible by electrochemistry in aqueous solutions.

    The in situ spectroelectrochemical measurements, however, give a

    possibility to correlate the specific structural changes with the

    reduction potential. The role of intercalated water in multilayer

    GO samples is taken into account in the spectral interpretation.

    Spectroelectrochemical studies in organic solvents are in progress.

    Acknowledgements

    Financial support from Academy of Finland (Grant No.

    128535) and the Graduate School of Materials Research

    (JK) is gratefully acknowledged.

    Notes and references

    1 X. Du, I. Skachko, A. Barker and E. Y. Andrei, Nat. Nanotechnol.,2008, 3, 491.

    2 T. J. Davies, M. E. Hyde and R. G. Compton, Angew. Chem., Int.Ed., 2005, 117, 5251.

    3 X. Huang, X. Qi, F. Boey and H. Zhang, Chem. Soc. Rev., 2012,41, 666.

    4 X. Huang, Z. Y. Yin, S. X. Wu, X. Y. Qi, Q. Y. He, Q. C. Zhang,Q. Y. Yan, F. Boey and H. Zhang, Small, 2011, 7, 1876.

    5 J. Hou, Y. Shao, M. W. Ellis, R. B. Moore and B. Yi, Phys. Chem.Chem. Phys., 2011, 13, 15384.

    6 K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang,S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004,306, 666.

    7 A. Lerf, H. He, M. Forster and J. Klinowski, J. Phys. Chem. B,1998, 102, 4477.

    8 D. R. Dreyer, S. Park, C. W. Bielawski and R. S. Ruoff, Chem.Soc. Rev., 2010, 39, 228.

    9 S. Pei and H.-M. Cheng, Carbon, 2012, 50, 3210.10 H.-L. Guo, X.-F. Wang, Q.-Y. Qian, F.-B. Wang and X.-H. Xia,

    ACS Nano, 2009, 3, 2653.11 M. Zhou, Y. Wang, Y. Zhai, J. Zhai, W. Ren, F. Wang and

    S. Dong, Chem.Eur. J., 2009, 15, 6116.12 L. Chen, Y. Tang, K. Wang, C. Liu and S. Luo, Electrochem.

    Commun., 2011, 13, 133.13 M. Hilder, B. Winther-Jensen, D. Li, M. Forsyth and

    D. R. MacFarlane, Phys. Chem. Chem. Phys., 2011, 13, 9187.14 C. Fu, Y. Kuang, Z. Huang, X. Wang, N. Du, J. Chen and

    H. Zhou, Chem. Phys. Lett., 2010, 499, 250.15 N. A. Kotov, I. De ka ny and J. H. Fendler, Adv. Mater., 1996, 8, 637.16 A. Ambrosi, A. Bonanni, Z. Sofer, J. S. Cross and M. Pumera,

    Chem.Eur. J., 2011, 17, 10763.17 P. M. Hallam and C. E. Banks, Electrochem. Commun., 2011, 13, 8.18 C. Kvarnstro m, H. Neugebauer and A. Ivaska, in Advanced

    Functional Molecules and Polymers; Processing and Spectroscopy ,ed. H. S. Nalwa, Gordon and Breach Publishers, Amsterdam,2001, vol. 2, p. 139.

    19 A. O sterholm, P. Damlin, C. Kvarnstro m and A. Ivaska, Phys.Chem. Chem. Phys., 2011, 13, 11254.

    20 A. Viinikanoja, J. Lukkari, T. A a ritalo, T. Laiho and J. Kankare,Langmuir, 2003, 19, 2768.

    21 A. Viinikanoja, S. Areva, N. Kocharova, T. A a ritalo,M. Vuorinen, A. Savunen, J. Kankare and J. Lukkari, Langmuir,2006, 22, 6078.

    22 M. Acik, C. Mattevi, C. Gong, G. Lee, K. Cho, M. Chhowalla andY. J. Chabal, ACS Nano, 2010, 4, 5861.

    23 M. Acik, G. Lee, C. Mattevi, A. Pirkle, R. M. Wallace, M. Chhowalla,K. Cho and Y. Chabal, J. Phys. Chem. C, 2011, 115, 19761.

    24 M. Osawa, Bull. Chem. Soc. Jpn., 1997, 70, 2861.25 Z. Wang, X. Zhou, J. Zhang, F. Boey and H. Zhang, J. Phys.

    Chem. C, 2009, 113, 14071.26 J. I. Paredes, S. Villar-Rodil, P. Sols-Ferna ndez, A. Martnez-

    Alonso and J. M. D. Tasco n, Langmuir, 2009, 25, 5957.27 D. Li, M. B. Mu ller, S. Gilje, R. B. Kaner and G. G. Wallace, Nat.

    Nanotechnol., 2008, 3, 101.28 T. Szabo , O. Berkesi and I. De ka ny, Carbon, 2005, 43, 3186.29 P. Gao, D. Gosztola, L.-W. H. Leung and M. J. Weaver,

    J. Electroanal. Chem., 1987, 233, 211.30 C. H. de Villeneuve, J. Pinson, M. C. Bernard and P. Allongue,

    J. Phys. Chem. B, 1997, 101, 2415.31 H. Miyake, S. Ye and M. Osawa, Electrochem. Commun., 2002,

    4, 973.32 X. Zhou, X. Huang, X. Qi, S. Wu, C. Xue, F. Y. C. Boey, Q. Yan,

    P. Chen and H. Zhang, J. Phys. Chem. C, 2009, 113, 10842.

    33 T. M. Chinowsky, J. G. Quinn, D. U. Bartholomew, R. Kaiser andJ. L. Elkind, Sens. Actuators, B, 2003, 91, 266.34 G. K. Ramesha and S. Sampath, J. Phys. Chem. C, 2009,

    113, 7985.35 C. P. Smith and H. S. White, Langmuir, 1993, 9, 1.36 M. A. Bryant and R. M. Crooks, Langmuir, 1993, 9, 385.37 X. Fan, W. Peng, Y. Li, X. Li, S. Wang, G. Zhang and F. Zhang,

    Adv. Mater., 2008, 20, 4490.38 Y.-K. Kim and D.-H. Min, Langmuir, 2009, 25, 11302.39 P. Nemes-Incze, Z. Osva th, K. Kamara s and L. P. Biro , Carbon,

    2008, 46, 1435.40 J. Shen, Y. Hu, C. Li, C. Qin, M. Shi and M. Ye, Langmuir, 2009,

    25, 6122.41 Y. Wang, Z. Shi, Y. Huang, Y. Ma, C. Wang, M. Chen and

    Y. Chen, J. Phys. Chem. C, 2009, 113, 13103.42 P. E. Fanning and A. Vannice, Carbon, 1993, 31, 721.

    View Online

    http://dx.doi.org/10.1039/c2cp42253k
  • 7/30/2019 Electrochemical Reduction of Graphene Oxide and Its in Situ

    7/7

    This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys.

    43 S. J. An, Y. Zhu, S. H. Lee, M. D. Stoller, T. Emilsson, S. Park,A. Velamakanni, J. An and R. S. Ruoff, J. Phys. Chem. Lett., 2010,1, 1259.

    44 A. C. Ferrari, Solid State Commun., 2007, 143, 47.45 L. M. Ma lard, M. A. Pi menta, G. Dr ess elhaus and

    M. S. Dresselhaus, Phys. Rep., 2009, 473, 51.46 S. Liu, J. Wang, J. Zeng, J. Ou, Z. Li, X. Liu and S. Yang, J. Power

    Sources, 2010, 195, 4628.47 H. Kang, A. Kulkarni, S. Stankovich, R. S. Ruoff and S. Baik,

    Carbon, 2009, 47, 1520.48 M. R. Baldan, E. C. Almeida, A. F. Azevedo, E. S. Goncalves,

    M. C. Rezende and N. G. Ferreira, Appl. Surf. Sci., 2007, 254, 600.49 K. N. Kudin, B. Ozbas, H. C. Schniepp, R. K. Prudhomme,

    I. A. Aksay and R. Car, Nano Lett., 2008, 8, 36.50 C. Hontoria-Lucas, A. J. Lo pez-Peinado, J. D. D. Lo pez-Gonza lez,

    M. L. Rojas-Cervantes and R. M. Martn-Aranda, Carbon, 1995,33, 1585.

    51 N. Karousis, A. S. D. Sandanayaka, T. Hasobe, S. P. Economopoulos,E. Sarantopoulou and N. Tagmatarchis, J. Mater. Chem., 2011,21, 109.

    52 G. I. Titelman, V. Gelman, S. Bron, R. L. Khalfin, Y. Cohen andH. Bianco-Peled, Carbon, 2005, 43, 641.

    53 Hydration and Intermolecular Interaction, ed. G. Zundel, AcademicPress Inc, New York, 1969, p. 28.

    54 P. Sutandar, D. J. Ahn and E. I. Franses, Macromolecules, 1994,27, 7316.

    55 K. Boujlel and J. Simonet, Electrochim. Acta, 1979, 24, 481.56 J. H. Wagenknecht, J. Org. Chem., 1972, 37, 1513.57 X. Gao, J. Jang and S. Nagese, J. Phys. Chem. C, 2010, 114, 832.58 S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas,

    A. Kleinhammes, Y. Jia, Y. Wu, S. T. Nguyen and R. S. Ruoff,Carbon, 2007, 45, 1558.

    59 S. Kim, S. Zhou, Y. Hu, M. Acik, Y. J. Chabal, C. Berger,W. D. Heer, A. Bongiorno and E. Riedo, Nat. Mater., 2012, 11, 544.

    60 L. Lee, M. M. Villalba, R. B. Smith and J. Davis, Electrochem.Commun., 2009, 11, 1555.

    61 L. Xiao, G. G. Wildgoose, A. Crossley and R. G. Compton, Sens.Actuators, B, 2009, 138, 397.

    62 A. Bonanni, A. Ambrosi and M. Pumera, Chem.Eur. J., 2012, 18, 4541.63 M. J. Allen, Organic electrode processes, Reinhold, New York,

    1958, p. 71.64 J. H. Wagenknecht, L. Eberson and J. H. P. Utley, in Organic

    Electrochemistry, ed. H. Lund and O. Hammerich, Marcel Dekker,New York, 4th edn 2001, p. 454.

    View Online

    http://dx.doi.org/10.1039/c2cp42253k