Sunil P. Lonkar,
Vishnu Pillai,
Ahmed Abdala and
Vikas Mittal*
Department of Chemical Engineering, The Petroleum Institute, P.O. box 2533, Abu Dhabi, UAE. E-mail: vmittal@pi.ac.ae; Tel: +971-26075491
First published on 22nd August 2016
Nanostructured composites of graphene and highly dispersed sub-20 nm sized ZnO nanoparticles (TRGZ) were prepared via a novel method combining freeze-drying and thermal annealing. A direct synthesis implies thermal reduction of graphite oxide and in situ ZnO formation which acted as mutual spacers preventing restacking of the graphene layers and eventual aggregation of the nanoparticles at moderate temperatures. A series of compositions with different weight ratios of ZnO nanoparticles were prepared and used as a reactive sorbent in low temperature hydrogen sulfide (H2S) removal from natural gas. The composite sorbent having a ZnO mass ratio of 45.1 wt% showed a significantly greater H2S adsorption capacity (3.46 mmol g−1) than that of pure ZnO (1.06 mmol g−1), indicating that hybridization of ZnO with grpahene significantly improved the H2S removal ability. Such enhancement was mainly attributed to the higher surface area, greater pore volume and unique morphology at the nanoscale in the graphene–ZnO hybrid.
Hydrogen sulphide (H2S) is one of the common and undesirable sulphur component often found in natural gas, syngas, biogas and other industrial gases.19,20 Due of its highly toxicity, offensive odour and acidic nature, H2S can pose serious threats to health and environment. Moreover, trace amounts of H2S can cause catalyst poisoning and pipeline corrosion.21 Its oxidation in the atmosphere to SO2 results in an acid rain formation. Therefore, development of efficient adsorbent materials with excellent desulfurization performance is highly essential. Owing to its favourable thermodynamics in sulphidation reaction, zinc oxide (ZnO) is considered as a very effective sorbent for removal of H2S from various gas streams and has also attracted significant attention due to its thermal stability and non-pyrophoric property.22–24 ZnO reacts with H2S to form insoluble zinc sulphide.25 The mechanism involves dissociation of H2S into H+ and HS−, followed by adsorption of HS− on ionic solid such as ZnO which later converts into ZnS by proton transfer from H2S to the chemisorbed OH groups on the Zn–O surface.26 As this reaction is exothermic, a lower temperature favours the forward reaction, resulting in high efficiency of desulfurization.27 The kinetics ZnO sulphidation reaction with H2S was predicted that the ZnO–H2S reaction proceeds with surface reaction, ion migration and completed by sulphidation penetration.28 In the further advancement of the field, the hybrid nanostructures involving ZnO are of special interest in the development of solid sorbents for H2S removal as they are capable of offering high sorption capacity.29–31 The hybrid materials based on graphene and nanoscale ZnO are promising in desulfurization because of excellent surface properties of graphene and catalytic properties of ZnO nanoparticles, which can be integrated together to enhance the overall performance of the sorbent.18,32,33 These studies concluded that the degree of oxygen functional groups and surface area of graphene in conjunction with particle size and distribution of ZnO are determining factors for low temperature H2S adsorption. However, most of these studies report graphene/ZnO composite synthesis that involves chemical reduction of GO using toxic reducing agents and conditions not suitable on industrial scales. Moderate to severe synthetic conditions including use of different solvents, microwaves, etc. under variable pH conditions, temperatures and pressures were employed. Moreover, ZnO particle size in the composites is occasionally found to be ≥100 nm with non-uniform particle dispersion. An in situ growth of ZnO nanoparticles with simultaneous graphene oxide reduction to obtain graphene/ZnO nanohybrids can offer scalable and facile route for the straight forward synthesis of nanostructured graphene/ZnO composites suitable for low temperature H2S adsorption.
In this work, preparation of graphene/ZnO nanostructured composites is reported via in situ process under thermal conditions using zinc hydroxyacetate as a Zn precursor. This simple and environment friendly method can be easily scaled up to obtain high quality graphene/ZnO composite materials suitable for wide range of applications. Finally, the performance of these hybrid materials as low temperature H2S sorbent was evaluated. A series of graphene/ZnO composites were prepared and their effectiveness towards H2S adsorption was assessed and activities were compared with pristine ZnO sorbent.
![]() | ||
Scheme 1 Schematic representation of interaction between GO and zinc hydroxyacetate and TRG/ZnO nanohybrid synthesis. |
The morphology of the resulting nanohybrids was elucidated using TEM and SEM analysis. The typical TEM images of ZnO nanoparticles and TRGZ composites are shown in Fig. 2. It can be seen that the thermal annealing of zinc hydroxyacetate under controlled conditions resulted into hexagon shaped ZnO nanoparticles with average size of ∼100 nm. A significant change in morphology from hexagonal to spherical was observed after hybridizing with graphene. Also, a substantial decrease in the size of ZnO particles was noted in presence of graphene, thus, confirming its control on the size and dispersion of the ZnO seeds. From the TEM images, it could be observed that a large number of uniformly dispersed ZnO nanoparticles with average size below 20 nm were formed on the surface of graphene. For TRGZ-1, the average size of ZnO particles was below 10 nm with more isolated particulates which can be attributed to the lower Zn salt concentration.
![]() | ||
Fig. 2 TEM images of (a) pristine ZnO nanoparticles, (b) TRGZ-1, (c) TRGZ-2, and (d) TRGZ-3 graphene/ZnO nanohybrids. |
For higher ZnO loaded composites namely, TRGZ-2 and TRGZ-3, a slight increase in ZnO size as well as number of particles was observed. At higher Zn salt concentrations, though the size controlling effect of graphene layers was found to reduce, however, the distribution and average size were still uniform. Overall, the TEM images clearly revealed that the nanostructured TRGZ composites with a uniform ZnO dispersion were successfully prepared. Further, the average crystalline size of the pristine ZnO nanoparticles and in situ formed graphene supported ZnO nanoparticles was estimated from Scherrer's formula.38 It was found that the in situ generated ZnO nanoparticles in TRGZ composites forms smaller crystallites in compared to the pristine ZnO (23.4 nm). Amongst, TRGZ composites, TRGZ-1 possesses smallest ZnO crystallites size (10.9 nm) and increases with the ZnO content for TRGZ-2 (14.4 nm) and TRGZ-3 (15.8 nm), respectively. These results are in good agreement with above mentioned TEM results.
As shown in Fig. 3, SEM analyses also confirmed the nanostructured morphology of the TRGZ composites. Uniform nano-assemblies of ZnO were visible in case of highly loaded TRGZ composites. The Zn elemental mapping (Fig. 3 inset) clearly indicated the uniform dispersion of the zinc species i.e. ZnO, along with increase in density for higher ZnO loadings. Overall, the interactions between functional groups on the graphene surface and the ZnO seeds significantly contributed to the formation of the hierarchically structured TRG/ZnO hybrids.
![]() | ||
Fig. 3 SEM images of from left to right, TRGZ-1, TRGZ-2 and TRGZ-3 composites (inset Zn element mapping). |
![]() | ||
Fig. 4 XRD diffraction patterns of (A) ZnO, TRGZ-1, TRGZ-2 and TRGZ-3 (B) TRGZ-1 and (C) GO and TRG samples. |
XPS was used to monitor the chemical states after thermal reduction in GO and its interaction with in situ formed ZnO nano-assemblies. Fig. 5A shows the C 1s and O 1s chemical states of GO and TRGZ hybrids.
![]() | ||
Fig. 5 XPS spectra (A) C 1s region of GO and TRZ-1, (B) survey spectra of TRGZ-1, and (C) high resolution spectra of Zn 2p region, for TRGZ-1 composite. |
It can be seen that XPS spectra of GO can be fitted into three peaks, corresponding to carbon atoms in different functional groups: sp2 carbon (C–C, 284.9 eV), carbon in C–O bonds (287.2 eV), and carbonyl carbon (CO, 288 eV), respectively.41 For TRGZ hybrids, the peak intensity of the C–C bond was significantly increased while oxygen-bonded carbons (C–O and C
O) were considerably reduced, indicating a sufficient reduction of graphite oxide into graphene during the thermo-annealing process.42 Further, the survey spectra of TRG/ZnO nanohybrid (Fig. 5B) shows the characteristic peaks of Zn 2p1/2, Zn 2p3/2, O 1s, and C 1s only and no peaks of other elements were observed which ensured the single phase, contamination free in situ growth of ZnO nano-assemblies on TRG sheets. The high-resolution scan of Zn 2p, shown in Fig. 5C, identified the exact peak location of Zn 2p3/2 at 1022.6 eV. Hence, XPS findings confirmed the successful formation ZnO/TRG nanohybrids under one-step thermo-annealed process.
Raman spectroscopy is a powerful and non-destructive technique to extract useful information about graphene and its related materials.43 Generally, graphitic materials show the characteristic D and G bands corresponding to k-point phonons of A1g symmetry and E2g phonon of sp2 carbon which are assigned to local defects and disorder especially at the edges of graphene and graphite platelets.44 Fig. 6 shows the Raman spectra of the GO, TRG and the corresponding TRGZ hybrids with different ZnO loading. The pristine GO exhibited G and D bands at Raman shifts of 1584 and 1333 cm−1, respectively, with an intensity ratio, ID/IG = 1.42. These bands shifted to 1349 and 1589 cm−1 after the thermal treatment of GO. Also, trivial increase in ID/IG was observed which indicated a decrease in the size of the in-plane sp2 domains, which was mainly attributed to the removal of the oxygen functional group in GO during thermal reduction process.42 For TRG/ZnO composites, further increase in ID/IG ratio was noted which signified the simultaneous GO reduction and ZnO nanoparticle formation implying successful synthesis of TRG/ZnO hybrids. Fig. 7 also shows the FT-IR spectra of the as-prepared GO, TRG, ZnO and TRGZ samples. GO exhibited characteristic IR peaks at 1620 cm−1 corresponding to the remaining sp2 character, the absorption peaks at 1722 cm−1 and 1392 cm−1 were ascribed to CO stretching of COOH groups, tertiary C–OH groups vibrations and epoxy symmetrical ring deformation vibrations, respectively.45 Furthermore, the broad peak at 3330 cm−1 was attributed to the hydroxyl stretching vibrations of the C–OH groups, and the band at 1050 cm−1 was assigned to C–O stretching vibrations mixed with C–OH bending.
In TRG, the characteristic features of GO almost disappeared, indicating the reduction of GO to graphene under thermal treatment. For ZnO, the absorption band at 440 cm−1 resulted due to stretching modes of Zn–O.34 In the FT-IR spectrum of TRGZ composites, only characteristic bands corresponding to the TRG and ZnO were reflected indicating the formation of TRG/ZnO nanohybrids. Further, absence of characteristic absorbance of CH3COO− assigned to raw material zinc hydroxyl acetate confirmed the successful preparation of TRGZ composites. The increment in the intensity of Zn–O peak with the content of ZnO was also noted for these composites.
TGA was used to determine ZnO loading and thermal stability of the resulting TRG/ZnO nanostructured composites under oxidative environment. As shown in Fig. 8, all the composites exhibited weight loss weight around 400 °C which was attributed to the removal of oxygen-containing groups and the decomposition of carbon framework of graphene from the composites. The thermal stability of graphene in case of TRGZ-2 was improved which is attributed to the higher loading and better dispersion of nanosized ZnO. Based on the residues and considering weight loss in pristine ZnO, the weight percentage of ZnO in composites was calculated to be TRGZ-1 (8.89%), TRGZ-2 (25.34%) and TRGZ-3 (45.13%), respectively.
N2 adsorption–desorption isotherms were used to investigate microstructure and exposed surface area in the resulting TRGZ composite materials. As shown in Fig. 9, ZnO and TRGZ composites yielded a type-IV isotherm with a hysteresis loop in the relative pressure range of 0.8–1.0, indicating the presence of inhomogeneous mesopores. The pore size distribution curve calculated from the adsorption branch by the BJH method displayed relatively small pores with a typical size of 1–5 nm. Also, it was observed that the TRGZ-3 composite had a BET surface area of 119.3 m2 g−1, which is almost 10 times higher than that of the pure ZnO nanoparticles (12.4 m2 g−1). The higher surface area for TRGZ composite could be attributed to the presence of high surface area TRG support, leading to more surface active sites and the increased adsorption of reactants during the composite formation. The surface area and pore properties for other composites are tabulated in Table 1.
Nanohybrids | BET surface area (m2 g−1) | Pore volume (cm3 g−1) | BJH pore diameter (Å) |
---|---|---|---|
ZnO | 12.4 | 0.083 | 31.25 |
TRG | 223.3 | 0.93 | 39.50 |
TRGZ-1 | 256.7 | 1.23 | 39.25 |
TRGZ-2 | 207.9 | 1.35 | 37.80 |
TRGZ-3 | 119.3 | 0.80 | 38.05 |
![]() | ||
Fig. 10 H2S breakthrough curves for TRG/ZnO composites treated at ambient conditions with different ZnO ratio and CH4 breakthrough curves for TRGZ-3 composite. |
Sample | H2S Cap (BT) (mmol per g-sorb) | H2S Cap (BT) (mg H2S per g-sorb) |
---|---|---|
TRG | 0.09 | 3 |
TRGZ-1 | 1.06 | 35 |
TRGZ-2 | 2.68 | 88 |
TRGZ-3 | 3.46 | 114 |
ZnO | 1.63 | 53 |
This indicated that CH4 adsorption on the adsorbent was negligible. A remarkable difference in H2S breakthrough curves was observed for the TRGZ hybrid composites as compared to pristine ZnO nanoparticles (Fig. 10). TRGZ-3 and TRGZ-2 sorbents exhibited highest H2S breakthrough capacity of 114 mg per g-sorbent and 88 mg per g-sorbent, respectively, as compared to 53 mg per g-sorbent for pristine ZnO. It was noteworthy that TRGZ composites contained a lesser amount of the active component, ZnO. Amongst TRGZ composites, TRGZ-1 had the lowest H2S sorption performance, as its breakthrough capacity was measured to be 35 mg per g-sorbent. As expected, pure TRG showed negligible breakthrough capacity at 3 mg per g-sorbent due to absence of any active site for H2S adsorption. The high H2S sorption capacities for TRGZ-2 and TRGZ-3 sorbents can be attributed to the sub-20 nm ZnO particles which enhanced the reactivity with H2S at low temperature in conjunction with the high surface area of TRG which provided more active sites. Also, the residual oxygen-containing functional groups on the basal planes of TRG (as confirmed by XPS and FT-IR results) play a critical role in promoting oxygen activation by accelerating the electron transfer, thereby promoting the activity of the terminal groups in surface reaction.46 In addition, these functional groups help the distribution of active ZnO particles on the surface. So the possible mechanism involves the initial physisorption of H2S molecules by oxygenated functional groups of graphene (Fig. 11I) which later can reach to the finely dispersed active ZnO nanoparticles and get chemisorbed through reactive adsorption (Fig. 11II).47,48
In present work, the higher H2S sorption by TRGZ-3 (mg per g sorbent) over pristine ZnO is attributed to the considerably large surface area (190.3 m2 g−1), which was almost 15 times that of ZnO (Table 1). Moreover, the pore volume of TRGZ-3 (0.80 cm3 g−1) was higher than that of ZnO (0.083 cm3 g−1). Whereas, TRGZ-2 possess higher surface properties over TRGZ-3 but showed lower H2S capacity (88 mg per g sorbent).
This difference in adsorption behaviour further indicate that not only high surface area, but also the content of nanosized ZnO active component is determining factor in low temperature H2S adsorption by using graphene/ZnO nanostructured adsorbents. The H2S breakthrough curve for ZnO was significantly broad and might have high saturation capacity as compared to the TRGZ composites due to the differences in pore structure. As the sorption of sulphur compound by metal oxides is a process of volume expansion, the sorbent becomes more compact and even pore closure occurs in some severe cases during the reaction.49 This prevents gaseous reactive molecules from reaching the interior fresh active sites, leading to a very low uptake and utilization of sorbent, especially at ambient temperature. In contrast, TRGZ sorbents due to larger surface areas and smaller size ZnO nanoparticles along with high pore volumes did not suffer from this and exhibited improved H2S adsorption. Overall, the higher amount of the ZnO active component in conjunction with high surface area and good dispersion due to the graphene layers are the critical factors for enhanced low temperature H2S adsorption.
Further, the EDX spectra of H2S treated TRGZ composites were recorded to monitor the changes in elemental composition, as shown in Fig. 12 for TRGZ-3 hybrid. The EDX spectrum exhibited a strong peak for sulphur element indicating the reactive adsorption of H2S on the sorbent. The quantitative results of the S/Zn ratio were calculated from the area of the corresponding spectral K lines and the amount of S in the composites was observed to be around 10 wt% which was in agreement with the breakthrough calculations. The sulphur element mapping image (inset in Fig. 12) also conformed the reactivity of the uniformly dispersed ZnO. However, to further elaborate this study, it would be interesting to investigate the reactivity of these adsorbents at higher temperatures which is under investigation.
This journal is © The Royal Society of Chemistry 2016 |