T. Kavinkumara,
D. Sastikumarb and
S. Manivannan*a
aCarbon Nanomaterials Laboratory, Department of Physics, National Institute of Technology, Tiruchirappalli 620 015, India. E-mail: ksmani@nitt.edu; ksmaniphysics@yahoo.com; Fax: +91-431-2500133; Tel: +91-431-2503616
bDepartment of Physics, National Institute of Technology, Tiruchirappalli 620 015, India
First published on 7th January 2015
Graphene oxide (GO) was synthesized from graphite through a chemical oxidation process and heat treated at 110 and 220 °C in a vacuum atmosphere. The partial reduction and sp3 to sp2 phase transition of GO was characterized by powder X-ray diffraction, Fourier-transform infrared, micro Raman, ultraviolet-visible-near infrared spectroscopy techniques. Dielectric properties of pristine GO and heat-treated GO were studied in the frequency range 102 to 106 Hz at 27 °C. Hydroxyl, carboxyl functional groups removed GO after 220 °C heat treatment, expressed higher electrical conductivity, dielectric constant and dielectric loss in the order of 10−2 S m−1, 103 and 105 respectively than the pristine GO (10−6 S m−1, 101 and 101). Pristine and heat-treated GO were coated on the partially cladding removed poly-methyl methacrylate optical fiber and used as fiber optic gas sensors. GO and heat treated GO coated fibers were responsive to detect ammonia, ethanol and methanol vapors from 0 to 500 ppm at 27 °C. Sensitivities of GO coated fiber optic sensor were calculated as −0.32, −0.26 and −0.20 counts per ppm for ammonia, ethanol and methanol vapors respectively. The effect of functional groups on dielectric and gas sensing properties of GO was investigated and reported.
Recently, GO and graphene are widely used as active materials to detect the gas molecules. Oxygenated sheets of graphene sensor has very high surface area and develop the better gas detection abilities than its counterpart due to adsorption of gas molecules on the surface of sheets.17,18 Some et al. demonstrated the gas sensors using hydrophilic GO and hydrophobic reduced graphene oxide (rGO) to sense the volatile organic gases. It is reported that the sensitivity of GO is higher than the rGO. This is due to high surface area and presence of different functional groups in GO.19 On the other hand, Huang et al. reported in situ investigation on the transition of electrical properties of GO paper from −40 to 150 °C. In situ dielectric spectroscopy studies revealed the four stage transition of electrical properties of GO with increasing temperature.20 The above reports clearly indicates the potential of GO and rGO towards the gas sensing and flexible electronic device applications.
But, for practical applications, room temperature operation of gas sensor and flexible electronic devices require well defined characteristics of GO and rGO. Hence, it is important to understand the role of functional groups towards the gas sensing and dielectric properties. The present work focus on tailor made synthesis of GO through heat treatment processes. For the first time, the role of hydroxyl, carboxyl functional groups on dielectric and optical gas sensing properties of GO, heat treated GO for ammonia, ethanol and methanol vapors is reported.
| ε*(ω) = ε′(ω) − iε′′(ω) | (1) |
![]() | (2) |
The real part of dielectric constant is derived from the following relation,
![]() | (3) |
![]() | (4) |
![]() | (5) |
The AC conductivity (σ) was calculated by the following equation20
σ = ε0ε′ω tan δ
| (6) |
The multimode step index poly-methyl methacrylate (PMMA) fiber cladding region was removed over 3 cm long at the center of 42 cm long fiber adapting mechanical cleaving method. The cladding removed fiber was then polished and its uniformity was observed through an optical microscope after cleaning with ethanol. Diameter of PMMA fiber was 750 μm (core = 735 μm, cladding = 15 μm, NA = 0.51) with the refractive index of 1.492 and 1.402 for core and cladding respectively. The as prepared GO and heat-treated GO110, GO220 were coated independently over the cladding removed areas of fibers using the dip coating method and then dried in air. The prepared fiber optic gas sensor was inserted into the gas sensing chamber. The schematic diagram of experimental set up for the gas sensing measurement is depicted in Fig. 2. Ammonia, ethanol and methanol vapors prepared at different concentrations (0–500 ppm) were allowed separately to pass into the gas sensing chamber through a gas inlet with the help of a regulator. After 10 min of passing the vapor, variation of output light intensity was continuously measured for each concentration. All the measurements were made at 27 °C.
FT-IR spectra of graphite, GO, GO110 and GO220 were presented in Fig. 4 and the assignment was given in Table 1. No absorption peak from graphite confirmed the absence of functional groups in the commercial graphite powder. The absorption peaks at 1725, 3416, 1054 and 1383 cm−1 in GO evidenced the presence of C
O, O–H, C–O and C–OH functional groups respectively. The removal of functional groups from GO after the heat treatments at 110 and 220 °C was responsible for the reduced peak intensity. Nevertheless, the presence of weak absorption peaks, established the GO was partially reduced after the heat treatment processes. The water molecules were evaporated from GO at 110 °C and hence the reduction in O–H peak intensity of GO110. The vibrations of remaining functional groups attached with GO110 are matching with the functional groups in GO. This result supported that the heat treatment did not disturbed the core structure of GO. Meanwhile, most of the functional groups were disappeared from GO after the 220 °C heating. However, the trace of functional groups in the honeycomb carbon structure introduced the limitations of the chemical reduction of GO.26
| Peak position (cm−1) | Assignment | Origin and nature | |||
|---|---|---|---|---|---|
| Graphite | GO | GO110 | GO220 | ||
| 1054 | C–O stretching | No absorption peak | Strong | Strong | Weak |
| 1383 | C–OH stretching | No absorption peak | Strong | Strong | Weak |
| 1623 | C C stretching |
Weak | Strong | Strong | Weak |
| 1725 | C O stretching |
No absorption peak | Strong | Strong | Weak |
| 3416 | O–H stretching | Weak | Broad | Weak peak compared to GO | Very weak |
The thermal stability of graphite and GO was calculated using TGA (Fig. 5). Graphite showed better thermal stability than GO. Linear relationship was noticed between weight loss and temperature for graphite. Meantime, there were three major steps of weight losses observed for GO. The weight loss was about 20% at around 110 °C, attributed to the removal of loosely bounded, adsorbed water and gas molecules. About 50% of weight loss occurred at around 220 °C, was ascribed to the elimination of O2, CO and CO2. This confirmed that highly reactive oxygen atoms from the GO structure react well to form CO and CO2. Further, weight loss at about 500 °C indicates the further reduction process of GO. Substantial weight loss was observed when GO was heated beyond 500 °C indicating the absence of major oxygen functional groups.
Raman spectroscopy (Fig. 6) evidenced the occurrence of structural changes due to the chemical treatment, from graphite to GO and the thermal treatment of GO to the partially reduced GO. Raman spectrum of graphite showed the low intense D band at 1361 cm−1 and was used to measure the defects, arising from a breathing mode of k-point phonons of A1g symmetry. The intense G band corresponds to the first-order scattering of E2g mode of sp2 carbon and was observed at 1575 cm−1. Presence of functional groups such as hydroxyl, epoxy generated the disorder in GO. Raman spectrum of as prepared GO showed relatively high D band at 1344 cm−1 and the G band was shifted to 1596 cm−1 compared to graphite. Transformation of sp3 to sp2 and degree of disorder were indicated by the intensity ratio of D band and G band (ID/IG). The ID/IG was inversely proportional to the average size of the sp2 domains. Variations of ID/IG will lead to the changes in the electronic conjugation state of the GO during the thermal reduction. The Raman spectrum of GO110 displayed the prominent D and G bands at 1350 and 1578 cm−1 respectively. Meanwhile, the increased intensity ratio as 0.98 for GO110 compared to GO (0.94) was due to elimination of water molecules from the graphene sheets. The ID/IG ratio increased further to 1.01 when the sample was annealed at 220 °C and was higher than GO and GO110. This confirmed the effective removal of defects/restoration of sp2 carbon atoms and reduced sizes of the sp2 domains.
Fig. 7a shows the FESEM image of cross-sectional view of vacuum dried free standing GO paper obtained by filtration of GO dispersion in water. Fig. 7b witnessed that the oxygenated graphene sheets were highly exfoliated having layered structure. This qualitative information reinforced the XRD observation on the expansion of interlayer spacing from 0.338 nm (graphite) to 0.930 nm.
![]() | ||
| Fig. 7 FE-SEM image of (a) side-view of thick pristine GO paper, (b) layered and exfoliated GO sheets. | ||
UV-vis-NIR spectra of GO and heat-treated GO are shown in Fig. 8. GO exhibits the maximum absorbance at 230 nm originated from the π–π* transition of C–C and C
C in sp2 hybrid regions, whereas a small hump at 300 nm was due to n–π* transition of the C
O in sp3 hybrid regions. However, the absorption peak of π–π* transition was shifted to the higher wavelength 238 nm for GO110 and the shoulder peak at 300 nm was also started to disappear, indicating the water containing functional groups on the GO surface were removed partially. The absorption peak of GO220 for π–π* transition of C–C was shifted to 273 nm. This was due to the removal of some oxygen functional groups and restoration of C
C bonds in GO220 sheets. This result supports the FT-IR observation on the presence and absence of oxygen functional groups in GO and GO220 nanosheets respectively.
Fig. 9 shows the frequency dependence of AC electrical conductivity calculated for GO, GO110 and GO220 papers at 27 °C. The GO paper was considered as layered composite consists of insulating sp3 matrix and conductive sp2 clusters.20 The pristine GO paper exhibits frequency dependent electrical conductivity and it increases with increasing frequency. The conductivity showed less variation when the applied frequency was low due to more active grain boundaries. Variation at high frequency regions was due to less active grain boundaries. Intercalated water molecules and additional oxygen functional groups in GO paper played a major role in the electrical conductivity. Conductivity of GO paper was controlled by the presence of insulating sp3 structure within the GO sheet. The functional groups break up the sp2-hybridized structure of the stacked graphene sheets and generating electrically poor conducting behavior. GO110 displayed the ionic conductivity increased with increasing frequency about two orders compared to GO due to the removal of water molecules. This confirms that the increase in sp2 cluster during the thermal reduction. A similar trend was observed for GO220 and was due to the elimination of oxygen functional groups. These results suggest the domination of conductive clusters and the removal of oxygen functional groups which are important for attaining maximum conductivity.
![]() | ||
| Fig. 9 The variation of electrical conductivity against the applied frequency of pristine GO, GO110 and GO220. | ||
Fig. 10a indicates the behaviour of dielectric constant (ε′) with respect to the applied frequency at 27 °C. The dielectric constant is a measure of charges stored inside the sample in the presence of applied electric field and detects the strength of alignment of dipoles in the dielectric. It is observed that the dielectric constant increases as frequency decreases due to interfacial polarization, which arises from space charges formed in the sample. In the presence of an electrical field, these space charges were confined by defects and localized accumulation of charge was formed at the electrode and sample interface. Therefore, ε′ show the large values at low frequencies. Though the conductivity was high at higher frequencies, the electrons cannot follow the alternating electric field and so ε′ of the GO paper was approximately constant and almost independent of frequency.
![]() | ||
| Fig. 10 (a) The variation of dielectric constant, (b) dielectric loss and imaginary part of the electric modulus with the applied frequency for pristine GO. | ||
The behaviour of dielectric loss and imaginary parts of the complex dielectric modulus (M′′) of GO was studied at 27 °C against the function of applied frequency (Fig. 10b). In general, dielectric loss (ε′′) arising due to the loss of energy in a dielectric material through conduction (transport-related loss), slow polarization current (bipolar loss) and other deceptive phenomena like interfacial polarization contribution.22,27,28 At low frequencies, dielectric loss showed the maximum value and was related to water intercalated into the GO. Increase of applied frequency decreases the loss and it approaches the stable value at high frequencies. Electric modulus was used to examine the dipole relaxation process, interfacial polarization and long-range conduction process within the bulk materials. The present study indicates the relaxation process associated with interfacial polarization clearly observed in the M′′ curves of the GO.29,30 Electrical modulus increases with increasing frequency and at certain frequency relaxation peak exist with a higher value. Here the, interfacial polarization appears only at certain frequencies for the GO and the electrical modulus drop to zero for higher frequencies.
GO110 showed the higher ε′ with respect to applied frequency (Fig. 11a) than the GO and was attributed to interfacial polarization at low frequencies. The ε′ and ε′′ decreases with increasing frequency. Removal of adsorbed water molecules played a key role in the increased dielectric constant. The dielectric loss of GO110 was greatly improved due to increases of conductive sp2 clusters. Fig. 11b displays the imaginary part of the complex electrical modulus of GO110 with respect to applied frequency. Relaxation peak shifted towards the higher frequency compared to GO due to elimination of functional groups. Hence, the relaxation time decreases with the peak shifts towards the higher frequency region.
![]() | ||
| Fig. 11 (a) The variation of dielectric constant, (b) dielectric loss and imaginary part of the electric modulus with the applied frequency for 110 °C heat-treated GO. | ||
Meantime, GO220 showed the highest ε′ and ε′′ among the three samples (Fig. 12a and b). This property was arising from the domination of conductive sp2 clusters after the restored sp2 carbon atoms through the thermal reduction. Fig. 12b displays the imaginary part of the complex electrical modulus of GO220 against frequency. Relaxation peak was shifted towards the higher frequency region compared to GO, GO110 and was due to the elimination of functional groups in GO paper.
![]() | ||
| Fig. 12 (a) The variation of dielectric constant, (b) dielectric loss and imaginary part of the electric modulus with the applied frequency for 220 °C heat-treated GO. | ||
According to Huang et al. very high electrical conductivity, ε′ and ε′′ values have been measured in the order of 10−2 S m−1, 105 and 109 respectively at 150 °C for GO. The present measurement on electrical conductivity of GO220, was in the same order of earlier work. However, two and four orders lower ε′ and ε′′ than the literature was observed.20 All the three samples have the strong frequency dependent of ε′, ε′′ and electrical modulus. The large value of ε′ for GO220 was due to majority of conductive sp2 clusters after the removal of insulating sp3 through heat treatment. Moreover, ε′′ occurs in GO220 was due to conduction, dipole and vibrational losses. The electrical conduction loss was due to elimination of functional groups, which in turn increases the dielectric loss.
In an optical fiber, when light crosses an interface between the core and modified cladding with different refractive indices, the light beam partially refracted and partially reflected at the interface depending on the angle of incidence. When the angle of incidence was equivalent to the critical angle, the light was reflected back into the core. The penetrated or refracted light intensity called evanescent field and its intensity decreasing exponentially away from the core–cladding interface. Evanescent waves in the cladding region was used for developing fiber optic gas sensors. Absorption of evanescent field phenomenon in the modified cladding region was studied through the variation of output light intensity to detect gas molecules. The total internal reflection was occurring at the interface between core and modified cladding when the refractive index of core was greater than the cladding of optical fiber.23,24 Thus, a part of the light was penetrating into the cladding material, when the refractive index of modified cladding was higher than the core and enters the modified cladding–air interface. Variation of output light intensity occurs when the refractive indices of core (ncore) and modified cladding material (nmclad) changes. When ncore > nmclad, the total internal reflection occurs at core–cladding interface. The characteristics of gas sensing exhibits when ncore < nmclad and partial reflection of light occurs at the core–cladding interface and the light enters into the modified cladding. Loss of partially refracted light through the optical fiber leads to decrease in output intensity. The characteristic of gas sensing was occurred at the modified cladding–air interface. The total internal reflection occurred at this interface and re-entry of leaked light changed the output intensity.
Absorption of the evanescent field exists in the outer air medium. The light beam undergoes total internal reflection at the modified cladding–air interface and the intensity of output light would vary when the gas molecules interact with modified cladding. The loss of light intensity undergoes in the absence of the gas molecules because of evanescent wave absorption occur at the outer air medium. The evanescent wave absorption may be higher or lower than the reference value in the presence of gases and magnitude of evanescent wave absorption.23 In the present study, as purchased PMMA raw fiber and cladding removed uncoated fiber were exposed to ammonia, ethanol and methanol vapors in the concentration range 0–500 ppm at 27 °C. No change in output intensity was observed in the absence of GO and heat-treated GO. The GO and heat-treated GO coated fibers were exposed to the water vapor to find out the interference of water vapor in the sensing performance. Interestingly, there is no remarkable response obtained during the exposure. The GO, GO110 and GO220 coated fibers were then tested in air atmosphere in the absence of test vapors. No change in output intensity was noticed and that output intensity was taken as reference intensity (reference or zero intensity mentioned by dotted line in Fig. 13) for sensing.
![]() | ||
| Fig. 13 Graph between a change in output intensity and gas concentration (0–500 ppm) of (a) ammonia (b) ethanol and (c) methanol for GO, GO110, GO220. | ||
Variation of output light intensity was measured in the wavelength range 200 to 1000 nm during interaction between the test gases and the sensing material. Concentration of sensing gas was fixed in the range 0–500 ppm and varied in steps of 100 ppm. The maximum change in intensity was observed at a particular wavelength for each vapor and was taken for sensitivity calculation. In the case of pristine GO, GO110 and GO220 coated fibers exposed to NH3 vapor, decrease (for GO and GO110) and increase (GO220) in output intensity (from the reference value) at 672 nm was observed as shown in ESI 1.† Similarly, the maximum wavelength responses for pristine GO, GO110 and GO220 coated fibers for ethanol/methanol vapors were observed as 675/672, 665/672 and 665/672 respectively and are shown in ESI 2–3.† These results indicate that the GO, GO110 and GO220 sheets were sensitive to the vapors.
Change in intensity from the reference value against vapor concentration of NH3, ethanol and methanol for GO, GO110 and GO220 sensors were plotted in Fig. 13. Sensitivity of the fiber optic gas sensor was defined as the ratio of change in output intensity with that of change in concentration of test gas. Thus, the sensitivities of pristine GO, GO110 and GO220 coated sensors were found to be −33, −15 and +7 counts per 100 ppm respectively for NH3 (Fig. 13a). In the case of ethanol vapor, sensitivities were calculated as −31, −15 and +6 counts per 100 ppm for GO, GO110 and GO220 respectively (Fig. 13b). Weak interaction between the polar groups in sensing material and the methanol vapor reflected on sensitivities and were calculated as −21, −9 and +3 counts per 100 ppm for GO, GO110 and GO220 respectively (Fig. 13c). Here the positive and negative signs are representing the increase and decrease in output intensities from the reference value of the fiber optic sensor. Interestingly, the pristine GO and GO110 coated fibers showed the decrease in output light intensity from the reference value against vapor concentration (ESI 1–2†). This was due to more evanescent field absorption at the boundary between GO/GO110 and test vapors. In contrast, GO220 coated sensor showed increased output light intensity from the reference value with respect to vapor concentration (ESI 3†). This was due to the less evanescent field absorption at the boundary between GO220 and the test vapors. Presence of oxygen functional groups in GO made the material hydrophilic. High surface area due to the exfoliation as observed in powder XRD provides the much better vapor sensing property than the heat-treated GO. Interaction between the oxygen functional groups in pristine GO and the test vapors were higher than the heat-treated samples due to the formation of intermolecular polar interactions.31,32 Heat treatment lead to the removal of functional groups confirmed from TGA and FT-IR made the materials became hydrophobic. The hydrophilic and hydrophobic nature of GO and heat-treated GO were further confirmed by dispersing the material in water. The GO dispersion in H2O was stable even after three weeks and will be useful for water based solution processing. In contrast, sedimentation of GO110 and GO220 dispersed in H2O was found after three hours, introduced limitations towards aqueous based processing.
In the present study, output light intensity variation arises predominantly due to physical adsorption of gas molecules on the surface of sensing materials. The refractive index of GO was controlled by the amount of hydroxyl and carboxyl functional groups through the interaction with analyte. In the case of NH3 sensing, the –OH and –COO ions in GO were possibly involved in hydrogen bonding with ammonia. This results in changing the refractive index of the modified cladding and so decrease in output intensity. The pristine GO was hydrophilic and had more number of attached oxygen functional groups than the heat treated GO. Hence, the pristine GO showed better sensing performance than the heat treated GO. Thus, the evanescent wave absorption occurred at the modified cladding–air interface during the interaction of ammonia with GO. Here, the magnitude of absorption was higher than the pure air environment. In the low vapor concentration, ammonia molecules were adsorbed over the surface of the GO sheets till to reach the critical concentration. At the critical concentration and above, the coexistence of vapor and liquid phase of ammonia molecules did not interrupt much on the refractive index of modified cladding. At this juncture, thick layer of adsorbed ammonia was formed over GO sheets resulting minor change in output intensity. This hypothesis was confirmed through the attainment of saturation in output intensity after the critical concentration. Similar trend was observed for ethanol and methanol vapors with lesser sensitivities than the ammonia. Here, the hydrogen bondings between the –OH in alcohol and –OH/–COO ions in GO are responsible for sensing. Lesser electronegativity of alcohol than ammonia was reflected on the sensitivity variation. Thus, the experimental results clearly indicate the role of functional groups in fiber optic gas sensing.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra12766h |
| This journal is © The Royal Society of Chemistry 2015 |