Ganaka G. Chandrakumaraa,
Jin Shangd,
Ling Qiub,
Xi-Ya Fangc,
Frank Antolasice,
Christopher D. Eastonf,
Jingchao Songb,
Tuncay Alan*a,
Dan Li*b and
Jefferson Zhe Liu*a
aDepartment of Mechanical and Aerospace Engineering, Monash University, Clayton, Victoria 3800, Australia. E-mail: zhe.liu@monash.edu; tuncay.alan@monash.edu
bDepartment of Materials Engineering, Monash University, Clayton, Victoria 3800, Australia. E-mail: dan.li2@monash.edu
cCenter for Electron Microscopy, Monash University, Clayton, Victoria 3800, Australia
dDepartment of Chemical and Biomolecular Engineering, The University of Melbourne, Victoria 3010, Australia
eSchool of Applied Chemistry, RMIT University, Victoria 3000, Australia
fCSIRO Manufacturing Flagship, VIC 3168, Australia
First published on 24th July 2015
The superior mechanical flexibility, mechanical strength, electrical conductivity, high specific surface area, and a special two-dimensional crystalline structure make graphene a very promising building block material for flexible electromechanical actuators. However, graphene papers have exhibited limited electromechanical actuation strain in aqueous electrolyte solution. In this paper, we show an easy approach to significantly improve the electromechanical actuation of reduced graphene oxide (rGO) papers via fine tuning the oxygen functional groups in rGO sheets, which was achieved by careful control of quantity of the reduction agent used in the chemical reduction process of graphene oxide. The actuation strains are enhanced up to 0.2% at an applied voltage of −1 V, which is more than a 2 fold increase compared to the regular pristine rGO paper. Further theoretical and experimental analyses reveal that the change of the capacitance and the stiffness of the rGO papers are two key factors responsible for the observed improvement.
Recent studies show that carbon nanotube and graphene based materials are promising actuators in artificial muscles and micro/nano-electromechanical systems.10–13 Particularly, the actuation performance of graphene has been anxiously studied with expectations of revolutionising the advanced actuator systems along different actuation schemes.14–20 The studies conducted up to date show promising results. Liang et al. showed that a graphene paper (composed of well aligned multilayers of rGO sheets) was capable of achieving electromechanical strain values close to 0.064% at a low operation voltage (−1 V) due to the formation of the electrical double layer (EDL) in an aqueous electrolyte.14 They also found that introduction of nanoparticles increased interlayer spacing of the graphene sheets and allowed a better development of the EDLs, hence enhancing the actuation strain to 0.1%.14 Note that among the available techniques to maintain the interlayer distances, controlling the hydration of the corrugated rGO sheets stands out as it can realise a tunable interlayer spacing without the introduction of any alien particles nor additional synthesis steps. This technique has been developed and successfully applied to synthesis high performance graphene hydrogel papers in supercapacitors.21
In addition to the efforts of increasing the interlayer spacing of rGO sheets in the papers, another route to enhance the electromechanical strain is chemical modification. Xie et al.22,23 modified the surface of rGO papers using oxygen plasma treatment and observed an electromechanical strain value close to 0.4% in aqueous electrolyte. Ab intio simulations reported that some rGO crystal structures had a far greater actuation performance than what has been discovered. A specific ordered rGO crystal structure is predicted to have a strain of up to ∼10%6 and an irreversible strain value of 28%.24 A complete understanding of the involved mechanisms and identifying the extent of the influence by the key parameters involved would help us optimize the rGO structure to reach these predicted performance levels. However, the failure to identify the involved mechanisms and parameters has brought the development of rGO actuators and their implementation in MEMS/NEMS to an abrupt stop.
The recent advance of rGO synthesis techniques show that the oxygen content in rGO sheets can be well controlled by using different amounts of reduction agent in the synthesis process without any additional post treatment steps.25,26 He et al. found that at a weight ratio of hydrazine over graphene oxide (GO) Rw > 0.2, fine tuning of the rGO molecular structures can be achieved without causing any major disruptions to the integrity of the rGO sheets. This method can promote uniformity and integrity of the obtained rGO papers in comparison with those fabricated using the post treatment processes.
In this paper, we used three different weight ratios of hydrazine over GO in the chemical reduction process, i.e., Rw = 1.72, 0.52, and 0.34 to study the effect of oxygen functionality on electromechanical actuation in an aqueous electrolyte. The obtained products are referred to as Rw 1.72 rGO, Rw 0.52 rGO, and Rw 0.34 rGO, respectively. Using these obtained rGO solutions, three types of rGO hydrogel papers were fabricated. A dry graphene paper at Rw = 1.72 was also fabricated as a reference. We observe that electromechanical actuation of the Rw 0.52 and Rw 0.34 rGO hydrogel papers exhibit significantly enhanced actuation strains in comparison with the Rw 1.72 rGO hydrogel and dry papers. An in-depth theoretical analysis and experimental results reveal a relationship among the electromechanical strain, the electrochemical capacitance and mechanical stiffness of the papers. The knowledge gained in this study will be useful for designing of rGO paper actuators to achieve optimal electromechanical actuation performances.
To characterise the chemical functional groups in the rGO sheets, five different methods were employed, i.e., contact angle measurement, Fourier transform of infrared spectroscopy (FTIR), X-ray photoelectron spectroscopy (XPS), Raman spectroscopy, and energy-dispersive X-ray spectroscopy (EDX).
Fig. 2 shows the contact angle of a 10 μl water droplet on the surfaces of as prepared rGO papers. The Rw 1.72 dry paper has a contact angle around 90°, indicating a hydrophobic nature of graphene, which is consistent with previous studies. For the Rw 1.72 hydrogel paper, the contact angle value is reduced to 68°. The excess water in the hydrogel papers should be the reason for this hydrophilic behaviour. For the Rw 0.52 and Rw 0.34 hydrogel papers, the contact angle reduces further to 54° and 51°, respectively. This is reasonable because it is well known that more oxygen functional groups in a graphene sheet can enhance its wettability.22
![]() | ||
| Fig. 2 The contact angles of a 10 μl water droplet on (a) Rw 1.72 dry paper, (b) Rw 1.72, (c) Rw 0.52 and (d) Rw 0.34 hydrogel papers. | ||
A comparison of FTIR spectra of the as-prepared hydrogel papers are depicted in Fig. 3(a). Differences in the FTIR spectra of the three types of hydrogel papers are presented in order to detect the differences in functional groups more clearly. The spectrum within the region from 1226 to 1277 cm−1 (highlighted using the dash box) corresponds to the vibrations of epoxy C–O bond.31–33 Compared with the Rw 1.72 case, the stronger spectrum intensities of the Rw 0.52 and Rw 0.34 hydrogel papers in this region indicate the existence of more C–O groups in the rGO sheets. This comparison also shows that the Rw 0.34 rGO hydrogel paper has more C–O groups than the Rw 0.52 hydrogel paper. This is as expected because more hydrazine reduction agent was used to produce the Rw 0.52 rGO sheets.
Table 1 summarises the quantitative information of atomic percentage of the rGO sheets, which were obtained from the XPS, EDX, and Raman spectroscopy analyses.
| Material | XPS | EDX | Raman | |||||
|---|---|---|---|---|---|---|---|---|
| C (%) | O (%) | N (%) | C–O/hydrocarbon ratio | C (%) | O (%) | N (%) | D/G ratio | |
| Rw 1.72 | 80.92 | 14.17 | 3.62 | 0.400 | 86.62 | 9.03 | 4.31 | 1.32 |
| Rw 0.52 | 79.02 | 16.54 | 3.78 | 0.531 | 75.79 | 16.06 | 8.14 | 1.29 |
| Rw 0.34 | 74.09 | 20.83 | 3.78 | 0.706 | 76.61 | 14.13 | 7.83 | 1.27 |
The elemental quantification obtained from XPS survey scans presented in Table 1 indicate a trend of decreasing O content with increasing weight ratio of hydrazine, as expected. High resolution C 1s spectral overlay of the three hydrogel paper samples are presented in Fig. 3(b). The overall peak shape is characteristic for partially reduced GO, with the main peak at 284.4 eV as expected for graphitic hydrocarbon and a shoulder at approximately 286 eV associated with C–O (and potentially C–N) groups. At higher binding energies we observe another less intense shoulder at approximately 288 eV associated with (N–)C
O and O–C–O and a tail in the spectra at binging energy >289 eV characteristic of rGO. It is apparent from the spectra overlay that the intensity of the shoulder associated with C–O decreases with increasing weight ratio of hydrazine, as expected and consistent with the elemental quantification results. In an attempt to quantify this trend, the high resolution C 1s spectra were fitted with a number of standard Gaussian–Lorentzian based components and the ratios of the components associated with C–O and hydrocarbon groups are presented in Table 1. As expected, sample Rw 0.34 has the largest ratio indicating it has the largest amount of C–O groups relative to hydrocarbon within the parameter space explored. Note that as only standard Gaussian–Lorentzian based components were used, the component for C–O groups is likely overestimated in all cases as the component related to sp2 carbon is lacking a tail that would extend to high binding energy, however the ratios still provide a useful method for comparing the hydrogel paper samples.
The Raman spectra for the three levels of reductions are presented in Fig. 3(c). These spectra display a clear D and G peaks at 1350 cm−1 and 1590 cm−1 respectively for all three samples. This is consistent with the standard spectrum seen for rGO.34 However, as reported in Table 1, slight variations of the D/G intensity ratios can be observed for the studied reduction levels. The slight perceptible decrease of D/G ratios indicate a decrease of sp2/sp3 bond ratios in the carbon structure due to the introduction of extra oxygen molecules to the rGO molecular structure.34–37 This indicates that the Rw 0.52 and Rw 0.34 has higher amount of carbon and oxygen interactions compared to that of Rw 1.72. This analysis further confirms that the Rw 0.34 samples have slightly more carbon oxygen interactions compared to Rw 0.52.
As summarised in Table 1, the semi-quantitative results obtained from EDX mapping shows a carbon, oxygen and nitrogen content of 86.7%, 9% and 4.3% respectively in Rw 1.72 hydrogel papers. This is consistent with the EDX results of graphene papers reported before.30 Although hydrazine hydrate can remove C–O–C, C
O and O–C
O functional groups efficiently during the reduction process, clear traces of C–OH groups are still found even at high reduction levels such as the case of Rw 1.72.25,32 The less reduction agent used in the reduction process left more functional groups in the rGO sheets as seen in Fig. 3(c), increasing the oxygen content in Rw 0.34 and Rw 0.52 to 14.13% and 16.06%, respectively. A slightly higher oxygen content in the Rw 0.52 paper (than that in the Rw 0.34 paper) appears contrary to the contact angle measurements as well as the FTIR, XPS, and Raman analysis results, and the fact that more reduction agent was used to produce the Rw 0.52 rGO sheets. We believe that such a discrepancy can be attributed to the limitations of the semi-quantitative EDX analysis. Nonetheless our EDX analysis yields two conclusions: (1) the Rw 0.52 and Rw 0.34 hydrogel papers have more functional groups than the Rw 1.72 papers; (2) there should be a minor difference of the content of oxygen functional groups in the Rw 0.52 and Rw 0.34 hydrogel papers.
Next, to determine the electromechanical response of the graphene papers and to demonstrate their use as potential MEMS actuators, we fabricated a series of bimorph devices using SU8 as a polymer substrate as shown in Fig. 4. In the bimorph design, via tuning the thickness of rGO papers and substrates, the deflection can be enlarged to ensure an easy and accurate measurement. The surface dimensions of the device is 2 mm × 15 mm and the thickness is ∼90 μm for the dry paper devices and ∼220 μm for hydrogel paper devices. The bimorph device was mounted on a holder and immersed in 1 M NaCl aqueous solution. It was connected to an electrochemical station (Bio Logic VMP 300) as the working electrode using a platinum sheet and wires, meanwhile a platinum mesh and a saturated calomel electrode were used as the counter and reference electrodes, respectively. The chronoamperometry technique was employed to apply the square wave voltage starting from ±0.1 V up to ±1 V at increments of 0.1 V. Three charge/discharge cycles (with a period of 100 s) were conducted at a given voltage value. It should be noted that in this study the actuation characterisation was conducted at a low charging frequency of 0.01 Hz, providing sufficient time to achieve saturated actuation strain of the individual papers. Therefore, the electrical conductivity of rGO layers, which influences the actuation at high frequency, should play a minor role in our case. The deflection of the bimorph cantilevers were captured using a CCD camera (Watec; WAT-231S) mounted to a C-mount microscope (Infinity; InfiniVar CFM-2/S). The tip displacement of the bimorph cantilevers were measured using image processing techniques on the captured footage. For each type of rGO papers, the actuation experiments were repeated for 10 times using three to five bimorph devices.
The electromechanical actuation strain, ε of the active rGO papers can be calculated using the measured tip displacement of the bimorph cantilever, δ in the following equation,38
![]() | (1) |
In eqn (1), t1 and t2 are the thickness of the rGO paper layer and the SU8+epoxy layer, respectively, E1 and E2 are their Young's modulus, and L0 (15 mm) is the initial length of the cantilever. The value of t1 is 5 μm for the dry rGO papers and 150 μm for the hydrogel papers while their t2 values are ∼85 μm and ∼70 μm respectively. The Young's modulus of SU8 is 2 GPa (MICROCHEM, SU8 2000) and that of epoxy is 500 MPa (ref. 39) (Selleys, Aqua Repair). The E2 is calculated as the averaged Young's modulus of SU8 and epoxy layer based on their thickness: ∼1 GPa. Note that eqn (1) is widely used in micro/nano devices community, e.g., Al/SiO2, W/SiO2, Al/Si, Ni/Ni-diamond, SiO2/Ti bimorph beams.8,38,40
Fig. 5(a) shows the calculated actuation strain of the rGO papers as a function of an applied periodic square wave voltage at a frequency of 0.01 Hz (ESI†). A nearly quadratic strain–voltage relation was observed for all devices. The Rw 1.72 dry papers have an averaged maximum actuation strain of 0.037% at −1 V. The actuation strain is improved to ∼0.10% at −1 V for the Rw 1.72 rGO hydrogel papers. The 1.7 fold increase indicates that increasing interlayer distances between the rGO sheets in the papers is very important to improve actuation performance. Our Rw 1.72 hydrogel papers have a similar actuation strain to graphene papers from Liang et al.14 It should be emphasised that in our experiments no alien nanoparticles were used to keep the interlayer distances of our hydrogel papers.
![]() | ||
| Fig. 5 The electromechanical actuation strain vs. (a) applied voltage and (b) injected charges for the rGO papers. | ||
While following the similar quadratic behaviour against the applied potential, the Rw 0.52 and Rw 0.34 hydrogel papers unveil an averaged maximum strain performance of 0.20% and 0.13%, respectively. This is a profound enhancement in comparison with the highly reduced Rw 1.72 papers. The enhancement is about 100% for Rw 0.52 papers and about 40% for the Rw 0.34 papers compared to the actuation strain of Rw 1.72 hydrogel paper. This evidently proves that controlling the reduction level is indeed a very convenient and effective way to improve the electromechanical actuation performance of rGO papers.
Previous studies have shown that the amount of injected charges dominate the electromechanical strain arising from the EDL effect.18,41 Hence, we integrated the measured current with respect to time to estimate the charges stored in the papers. Fig. 5(b) illustrates the strain against the respective charge per atom. At a given voltage, the Rw 1.72 dry papers store much less charge than the hydrogel papers do. This can be attributed to the much less accessible surfaces in the dry papers. The more charge accumulation in hydrogel papers enhance the coulomb forces in the EDL, generating a large driving force for mechanical deformation. At −1 V, the Rw 0.52 and Rw 0.34 hydrogel papers store similar charges: −0.072 vs. −0.069 electrons per atom. In spite of a small ∼4.3% difference in the charge values, there is a considerable ∼54% difference in the actuation strain results. This evidence indicates that the amount of injected charges (per atom) is not the only factor that governs the electromechanical actuation of our rGO papers.
The ion intercalation, electric double layer effect and the quantum mechanical effect are reported as the mechanisms of the electromechanical actuation of carbon based materials.10,18,42 The graphene paper is composed of face-to-face aligned atomically thin graphene layers. Thus the ion intercalation between graphene layers would change the dimension perpendicular to the basal plane.42,43 The change in the basal plane arising from ion intercalation should be minor. Theoretical studies showed that the quantum mechanical effect should lead to extension upon electron doping and contraction upon hole doping and the overall strain–charge relation exhibited a highly non-parabola shape,10,18,43 which are clearly different from our experiments. On the contrary, theoretically the EDL effect endows a quadratic strain–charge relation.18,44 In addition, our previous ab initio simulations indicated that for graphene layers immersed in electrolyte, the EDL effect is much more significant than the quantum mechanical effect. The EDL effect, therefore, is highly likely to be the dominant mechanism in our system.
In an attempt to identify the key physical factors that determine the actuation arising from EDL effect, a theoretical model will be developed in the following. Following Riemenschneider et al.,41 the energy of the system can be expressed as,
![]() | (2) |
![]() | (3) |
![]() | (4) |
Assuming that the change of capacitance C in comparison with the initial value C0 solely arises from the deformation S, we can make some simplifications for α. The electric capacitance can be estimated as
and
, in which A0 or A are the accessible surface area at the initial or final states, εr is the dielectric constant, and d is the thickness of the EDL. The surface area A can be expressed by using A0 and the electromechanical strain ε of the rGO papers as A0 = A(1 + 2ε). Thus the coefficient α is approximately α ≈ 2ε/C0S.
Given that at static equilibrium the internal forces (F) reduce to zero, from eqn (3) we obtain,
![]() | (5) |
To examine the obtained relation eqn (5), we measured the capacitance and the stiffness of our rGO papers. Fig. 6(a) shows the current vs. voltage results of the rGO papers from the CV analysis. The calculated device capacitance are shown in Fig. 6(b). The hydrogel papers always have a higher electrochemical capacitance than the dry paper. This explains why the rGO dry paper bimorph device accumulated less charges at −1 V than those hydrogel papers (Fig. 5). Interestingly the Rw 0.52 rGO hydrogel paper has a slightly higher capacitance than both the Rw 1.72 and the Rw 0.34 hydrogel papers.
Fig. 7 depicts the obtained stress vs. strain relations in the tensile test and summarises the Young's modulus and stiffness results that were fitted using the tensile experimental data in Fig. 7(a) with strain values below 0.2%. The Rw 1.72 dry paper has the highest Young's modulus and stiffness, i.e., 2654 MPa and 1770 N m−1. Overall the hydrogel papers have a comparable Young's modulus and stiffness, in which the Rw 1.72 hydrogel paper has the highest values of modulus and stiffness and the Rw 0.52 rGO paper has the lowest ones. The low stiffness implies the lack of hindrance against the in-plane structural deformation, which should benefit the electro-chemical actuation strain output.
![]() | ||
| Fig. 7 (a) The stress–strain relation of the rGO papers under a tensile loading. (b) The calculated Young's modulus and stiffness results of the rGO papers. | ||
With the obtained capacitance C and stiffness K results, we plot the calculated electromechanical strain results at −1 V (Fig. 5) against the C/K values in Fig. 8. The results of the all four materials exhibit a very good linear relation, which is consistent with the derived relation eqn (5). Hence, we can conclude that indeed EDL is the dominant actuation mechanism for rGO papers and the capacitance and stiffness are two key parameters in determining the electromechanical actuation performance.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra09743f |
| This journal is © The Royal Society of Chemistry 2015 |