Next Article in Journal
Evaluation of the Shallow Geothermal Potential for Heating and Cooling and Its Integration in the Socioeconomic Environment: A Case Study in the Region of Murcia, Spain
Next Article in Special Issue
A Review on Advanced Manufacturing for Hydrogen Storage Applications
Previous Article in Journal
Experimental Investigation of Frost Formation Influence on an Air Source Heat Pump Evaporator
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Palladium-Functionalized Graphene for Hydrogen Sensing Performance: Theoretical Studies

1
Department of Mechanical Engineering, Indian Institute of Technology, Jodhpur 342037, Rajasthan, India
2
Membrane Science and Separation Technology, CSIR Central Salt and Marine Chemicals Research Institute, Bhavnagar 364002, Gujarat, India
3
Materials Science and Engineering Department, Idaho National Laboratory, Idaho Falls, ID 83415, USA
4
Department of Civil & Environmental Engineering, Idaho State University, Pocatello, ID 83209, USA
*
Authors to whom correspondence should be addressed.
Equal Contribution.
Energies 2021, 14(18), 5738; https://doi.org/10.3390/en14185738
Submission received: 18 August 2021 / Revised: 7 September 2021 / Accepted: 8 September 2021 / Published: 12 September 2021

Abstract

:
The adsorption characteristics of H2 molecules on the surface of Pd-doped and Pd-decorated graphene (G) have been investigated using density functional theory (DFT) calculations to explore the sensing capabilities of Pd-doped/decorated graphene. In this analysis, electrostatic potential, atomic charge distribution, 2D and 3D electron density contouring, and electron localization function projection, were investigated. Studies have demonstrated the sensing potential of both Pd-doped and Pd-decorated graphene to H2 molecules and have found that the gap between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO), i.e., the HOMO-LUMO gap (HLG), decreases to 0.488 eV and 0.477eV for Pd-doped and Pd-decorated graphene, respectively. When H2 is adsorbed on these structures, electrical conductivity increases for both conditions. Furthermore, chemical activity and electrical conductivity are higher for Pd-decorated G than Pd-doped G, whereas the charge transfer of Pd-doped graphene is far better than that of Pd-decorated graphene. Also, studies have shown that the adsorption energy of Pd-doped graphene (−4.3 eV) is lower than that of Pd-decorated graphene (−0.44 eV); a finding attributable to the fact that the recovery time for Pd-decorated graphene is lower compared to Pd-doped graphene. Therefore, the present analysis confirms that Pd-decorated graphene has a better H2 gas sensing platform than Pd-doped graphene and, as such, may assist the development of nanosensors in the future.

1. Introduction

Understanding the various kinds of interactions between metallic molecules and graphene is of utmost importance if a number of scientific problems, associated with the enhancement of sensing performance, are to be solved. Over the last few decades, semiconducting materials such as hydrogen sensing films have been extensively explored [1,2]. Carbon-based nanomaterials and their derivatives, viz., graphene, reduced graphene, reduced graphene oxide, carbon nanotubes, and fullerene, play a vital role in gas sensing, adsorption, and storage performance due to their interesting and beneficial chemical and physical properties [3,4,5,6]. Carbon-based nanomaterials possess the largest mechanical strength, high specific surface area, good conductivity, compatibility with surface modification, and high electron mobility [7,8,9,10]. As such, they have found widespread applications in the fields of electronics, chemistry, and medicine; as catalysts in gas sensors; and as components in the pursuit of energy conversion, production, and storage [11,12,13,14,15,16,17,18,19]. Graphene and its oxide derivatives have numerous properties to choose from and, for this reason, are an ideal gas sensor material. Some of these properties include: (1) large surface area, which provides better interaction with gas molecules; (2) charge transfer capabilities; and (3) hybridization, which results in strong interaction between the electronic state of graphene and gas molecules [20,21]. Despite these advantages, the lack of bandgap in graphene limits its use in electronic devices. Graphene nanoflakes and graphene nanoribbons are H2-passivated, graphene-based, materials of finite width [22]. Over recent decades, many transition materials have been explored to enhance sensing capabilities, adsorbance, and storage properties. Additionally, the ability of graphene and its oxides to sense various gases has been investigated. Previously, researchers used transition metals to decorate graphene, thus enhancing its surface reactivity and facilitating the separation of targeted gases. Cu [23], Au [24], Pd [25], Fe [26], Ru [27], and Mn [28] are only some of the transition metals frequently reported on in the literature. The reason we selected Pd in this work was because of Pd’s greater affinity towards hydrogen gas [29,30,31]. Alfano et al. [32] developed a fully eco-friendly, Pd-decorated, graphene device for fast and selective detection of H2 within a 30 s timeframe and with a 26% response value for exposure of 1% H2. Singla et al. [33] studied the hydrogen sensing mechanism of a copper-decorated, N-doped, defective graphene through DFT calculation and found that it increased the binding energy of Cu atoms due to the deficiency of electrons in the N-doped site preventing the aggregation of the metal in their graphene sheets. It was also found that hydrogen adsorption intensity was much higher for the copper-decorated complex system. In other work, Bakshi et al. [34] studied the hydrogen storage capabilities of these new adsorbents by introducing Pd into co-doped, pristine and defective, graphene sheets containing N2 and boron as adatoms. They reported that Pd enhanced the hydrogen adsorption energy of the pristine graphene, whereas the defective sheets prevented clustering, resulting in a reduction to hydrogen adsorption. Thus, hydrogen adsorption can be enhanced by doping graphene sheets with N2 and boron adatoms. Finally, results concluded that Pd-decorated, N-B co-doped, graphene sheets behave as a novel material for hydrogen adsorption. Zhou et al. [35] studied the adsorption of TCDF on vacancy-defective graphene doped with Mn and Fe via DFT calculations and reported enhanced sensitivity of the graphene towards the TCDF molecule due to the presence of vacancy and dopants. From this investigation, Zhou et al. found that the hydrogen adsorption of Ni4 clusters embedded in multi-vacancy defective graphene absorbed up to four hydrogen molecules in circumstances conducive to hydrogen storage [36]. In another study, the hydrogen storage of a graphene-Pd(T) graphene sandwich structure was studied, with findings suggesting that the binding energy of the Pd atoms at the T-sites of the bilayer graphene is higher than the experimental cohesion energy of the Pd atoms; this resulted in the elimination of a metal cluster effect on the adsorption property, producing an outstanding H2 storage material [37]. Gecim et al.’s [38] work focused on the methanol sensing capabilities of Ga- and Ge-doped graphene using DFT calculations and found that, during adsorption, Ga- and Ge-doped graphene increased the positive charge of methanol transfer to graphene as well as electrical conductivity. At the same time, the gap between HOMO and LUMO decreases for the doped transition metals. It was also observed that Ge-doped graphene had higher electrical conductivity, a shorter recovery time, and increased sensitivity for methanol compared to Ga-doped graphene. Zheng et al. [39] studied the hydrogen storage performance of pure graphene nanoribbon and Li-decorated graphene using a DFT calculation study. The results revealed that the Li atom contains high interaction energy towards pure graphene ribbon with a high adsorption energy, and has little effect on geometry. At the same time, H2 molecules interact strongly with Li-decorated graphene, suggesting that Li-decorated graphene nanoribbon exhibits excellent hydrogen storage capability. Ostad et al. [40] investigated the adsorption of transition metal clusters over a graphene layer through DFT and identified the site between the two neighboring carbon atoms as the most stable for Pt, Pd, and Ir adatoms. These authors determined that, as the number of atoms increases, the binding energy of the cluster also increases. By contrast, the adsorption energy of the cluster attached to the graphene lowered when the lowest energy configuration for the graphene-transition metal was applied. The adsorption energy of the adatoms increases due to the Van der Waals correction applied in the DFT calculation. The charge transmitted to the carbon atom via transition metal atoms was also investigated, showing that the d orbital of the adatoms introduced a peak at the dirac point of the graphene. In another study, Montejo-Alvaro et al. [41] investigated the properties and ground-state structure of the Pdn cluster (n = 3–10) using DFT calculation and found it was supported on pristine, B-doped, and defective graphene quantum dots. They also reported that reactivity towards oxygen reduction was partially improved. In another work, hydrogen adsorption on Ni (111) surfaces, Ni13–graphene, and isolated Ni13 clusters through DFT calculation, were studied. Through pseudo charge density difference calculation, a build-up of charge density in the region of the Ni-graphene interfacial was observed. An activation barrier that stabilizes Ni13-graphene was observed using hydrogen adsorption [42]. Cui et al. [43] conducted a DFT calculation study to assess the sensing behavior of SOF2 and SO2F2 molecules using the surface of Pt/Pd-decorated CNT and Pt2Pd2-decorated CNT. The study found that decorated CNT had better adsorption ability for SOF2 compared to SO2F2, and that large charge transfer numbers were generated by enlarging the bandgap after adsorption and redistribution of the total electrons in the adsorption system due to the strong interaction between the surface and SOF2 molecule. Cheng et al. [44] investigated the sensitivity enhancement of a Pd-decorated ZnO nanocluster for H2 sensing via DFT calculation. They found that Pd released 42.1 kcal/mol energy and that it was located above the center of the tetragonal ring. Furthermore, adsorption of the O2 molecule on the Pd atom weakened the O-O bond, resulting in a good site for H2 attachment requiring 1.3 kcal/mol as an energy barrier at 325 K temperature. Response time also increased from 0.11 to 123.53 for the Pd-decorated ZnO nanocluster, and the prediction of recovery time was 0.7 ms for H2O desorption in the case of the Pd-decorated ZnO nanocluster. Khodadadi et al. [45] evaluated the sensing characteristics of H2S through doped and decorated metals attached to graphene and found that doping of transition metals (Ni, Cu, and Zn) enhanced the molecular adsorption of H2S among all tested metals. Cu showed the best adsorption system due to low adsorption energy and a short distance. While, for metal decoration, Ni showed the lowest distance between the metal and carbon atom best suited for electron transfer from H2S to metal-decorated graphene.
In this work, the sensing characteristics of H2 on both Pd-doped and Pd-decorated graphene were investigated using DFT calculations to analyze the effect of Pd-doping and Pd-decoration upon graphene sensitivity towards these molecules.

2. Computational Method

The adsorption process of the H2 molecule on Pd-doped and Pd-decorated graphene was investigated via DFT calculations [46], using the Gaussian09 software [47]. The graphene structures were optimized by employing the B3PW9 method with a basis set of 6–31 g*. To plot the density of states (DOS) the GaussSum package was used [48] and the remaining 3D plots were plotted through Multiwfn. Adsorption energies and optimized geometry were based on Equilibrium Geometry calculations.
E = Eelectronic + Zero-point energy (ZPE) + Evibrational +Erotational + Etranslational
All energy values were calculated at room temperature (298 K) and atmospheric pressure, including zero-point energy and single-point energy. The Gaussian09 software was used to obtain the total thermal energy, vibrational frequency, enthalpy, and thermal free energy [49].
H = E + RT
where E stands for thermal energy, H for thermal enthalpy, R for ideal gas constant, and T for temperature (298 K).
The electronic properties of the H2, adsorbed on the Pd-doped and Pd-decorated graphene, were investigated and compared. To evaluate the interaction between the H2 molecule and the Pd-doped graphene, their adsorption energy (Eads) was calculated with Equation (3).
Adsorption Energy; (∆Eads) = Esystem − (Ecluster+ Eadsorptive)
Where Esystem stands for the total energy of the H2 on Pd-doped graphene geometry, Ecluster stands for the energy of optimized Pd-doped graphene geometry, and Eadsorptive stands for the energy of optimized H2 geometry. To evaluate the interaction between the H2 molecule and the Pd-decorated graphene, their adsorption energy (Eads) was calculated with Equation (3), similar to the calculation for Pd-doped graphene.
Adsorption Energy (∆Eads) = Esystem − (Ecluster+ Eadsorptive)
where Esystem stands for the total energy of the H2 on Pd-decorated graphene geometry, Ecluster stands for the energy of optimized Pd-decorated graphene geometry, and Eadsorptive stands for the energy of optimized H2 geometry. As per the results, a negative value for the adsorption energy was obtained, indicating a stable structure and the occurrence of an exothermic process. The difference between the energy of HOMO and LUMO (HLG) was used to evaluate the chemical activity of Pd-doped and Pd-decorated graphene structures. Low HLG values imply that chemical activity is high for that atomic condition, and vice versa [50,51]. Furthermore, chemical hardness, electronegativity, electrophilicity, and chemical potential were calculated [52,53,54] to investigate the activity of the clusters. These values are based on Koopman’s approach [55,56].
Chemical   Potential   ( μ ) = I + A 2
Chemical   Hardness   ( η ) = I A 2
Electrophilicity   ( ω ) = μ 2 2 η
Electronegativity (λ) = − μ
where I = −ε HOMO and A= −ε LUMO, εHOMO-. To estimate net charge transfers and atomic charges; natural bond orbital charges, i.e., NBO population analysis, was applied [57]. The minimum electrophilicity principle [58] and maximum hardness principle [59] determined that the chemical stability of a system has a positive correlation with hardness and a negative correlation with electrophilicity.

3. Results and Discussion

A DFT optimized geometry of Pd-doped/decorated graphene before, and after, adsorption of H2 molecules is shown in Figure 1.
In this process, a carbon atom of pristine graphene was substituted with a dopant atom (Pd) for the doped condition. The Pd atom was attached to the carbon atom by forming an external bond without affecting the electron density of the Pd-decorated graphene surface. This demonstrates that, after doping/decorating the Pd atom to the graphene, the intermolecular bond between the C-C atom increases in comparison to pristine graphene. The estimated lengths between the three carbon atoms and the Pd atom were 1.80 Å, 1.81 Å, and 1.82 Å for the doped condition (Figure 1a). For the decorated state, the length between the Pd atom and the carbon atom was 2.13 Å (Figure 1c). In addition, after adsorption of H2 gas, the lengths between the three carbon atoms and Pd increased to 1.97 Å, 2.0 Å, and 2.0 Å for the doped condition, whereas molecular distance between the H2 and Pd molecules measured up to 2.34 Å (Figure 1b). There was no change in the bond length of the carbon and Pd atoms for the decorated condition; however, the H2 molecule attached to the Pd atom directly at a bond length of 1.85 Å (Figure 1d).
The electrostatic potential behavior of Pd-doped and Pd-decorated graphene clusters before, and after, adsorption of H2 is shown in Figure 2.
It is clearly demonstrated that the charge accumulation in the doped region is less in comparison to Pd-decorated graphene, and that charge accumulation increases for both conditions following adsorption of the H2 molecules. The charge accumulation regions are depicted in red.
To determine the stability of configurations, the hydrogen molecule was located above the Pd-doped/decorated graphene cluster. To determine the characteristic behaviour of hydrogen adsorption over the Pd-doped/decorated graphene cluster, the calculation was computed for adsorption energies and adsorption enthalpies. This is shown in Table 1 and Table 2.
The bond length between Pd-C atoms increases after the adsorption of H2 molecules for both the Pd-doped/decorated graphene clusters. The new estimated bond lengths following H2 adsorption were 1.97 Å, 2.0 Å, and 2.0 Å for the Pd-doped graphene clusters. No change in bond length between Pd-C for the Pd-decorated graphene occurred, as shown in Figure 1. The adsorption of hydrogen on both Pd-doped and Pd-decorated graphene structures is an exothermic process. Adsorption energies for the Pd-doped, and Pd-decorated, graphene clusters were calculated as −4.3 eV and −0.44 eV, respectively. The H-H bond length for the adsorbed hydrogen molecule was calculated as 0.759 Å for the Pd-doped graphene, and as 0.803 Å for the Pd-decorated graphene clusters. Both values were more significant than that of the free hydrogen molecule (0.74 Å).
To estimate the atomic charge and total charge transfer, the natural bond orbital population was applied [57], as shown in the Figure 3.
While performing calculations for Pd-doped graphene and Pd-decorated graphene clusters, H2 was adsorbed on both Pd-doped and Pd-decorated graphene clusters. The highest positive charges were observed for Pd atoms in both conditions (+1.214e for Pd-doped graphene and +0.599e for Pd-decorated graphene) and charges after H2 adsorption (+1.006e for Pd-doped graphene and +0.484e for Pd-decorated graphene). To determine the electronic behavior of both clusters, the net charge transfers of adsorbed H2 onto both Pd-doped graphene and Pd-decorated graphene clusters, were examined. Prior to adsorption, H2 molecules had no charge; the net charge transferred to adsorbed H2 molecules on Pd-doped and Pd-decorated graphene was calculated as (0, +0.067) and (+0.023, +0.027), respectively. The results suggest that the charge was transferred from H2 to Pd-doped/decorated-graphene. With respect to the charge transfer comparisons, results suggest that the charge transferred from H2 to Pd-doped graphene far quicker than to Pd-decorated graphene. This clearly demonstrates the difference in the adsorption energy for Pd-doped and Pd-decorated graphene clusters.
The molecular orbital (HOMO and LUMO) representation for both Pd-doped and Pd-decorated graphene is shown in Figure 4.
Figure 4a shows HOMO and LUMO representations of α and βMOs for the optimized structure of Pd-doped graphene. Figure 4c shows optimised structure representations for Pd-decorated graphene. The term α MOs is used for the spin-up, and βMOs for spin-down. The HLG is used to determine the chemical activity of Pd-doped and Pd-decorated graphene clusters HLG   α 1 Chemical   Activity ; the lower the value of HLG, the higher the chemical activity [50,51]. The value of electrophilicity, electronegativity, HLG, HOMO, LUMO, chemical hardness, and chemical potential for Pd-doped and Pd-decorated graphene, is given in Table 1 and Table 2. Adsorption energy is a dependent term for chemical potential; if the chemical potential is low, then the adsorption energy for H2 is also low [60,61]. HLG for Pd-doped graphene (0.56) is comparatively higher than Pd-decorated graphene (0.509), i.e., chemical activity of Pd-decorated graphene is higher than Pd-doped graphene. While, after the hydrogen adsorption, HLG for Pd-doped graphene decreased up to 0.488, meaning that chemical activity increased for Pd-decorated HLG value and decreases up to 0.477, the results demonstrate that chemical activity is higher for Pd-decorated graphene in comparison to Pd-doped graphene when all these conditions are met. For Pd-doped graphene, chemical hardness and electronegativity decreased following H2 adsorption, whereas the amount of chemical potential and electrophilicity increased. For Pd-decorated graphene, chemical hardness and electronegativity decreases as the amount of chemical potential and electrophilicity increases. In comparison to Pd-doped and Pd-decorated graphene, the H2 adsorption value of chemical hardness, electronegativity, and electrophilicity decreased as the value of chemical potential increased.
Figure 5 represents the DOS plots for Pd-doped and Pd-decorated graphene before and after Hydrogen adsorption, resulting in both Pd-doped and Pd-decorated graphene enhancing electrical conductivity due to decreased HLG following the adsorption of the Hydrogen molecule.
The Pd-decorated graphene shows a much lower HLG value in comparison to Pd-doped graphene given the changes in DOS plot gaps suggesting greater electrical conductivity for Pd-decorated graphene than Pd-doped graphene. The electrical conductivity depends on temperature and the value of HLG as shown in Equation (8) [62,63,64].
σ = AT 3 / 2 exp ( E g 2 kT )
As the value of HLG (Eg) increases, the value of electrical conductivity (σ) decreases due to the negative exponential in the expression and vice versa. In both Pd-doped and Pd-decorated graphene conditions, the value of HLG decreased. However, a comparison between Pd-doped and Pd-decorated graphene following hydrogen adsorption showed that the value of HLG decreased from 0.488 to 0.477 for Pd-doped and Pd-decorated graphene, respectively. Thus, as per Equation (8), electrical conductivity was higher for the Pd-decorated graphene. This shows that the Pd-decorated graphene is a better sensor for H2 detection. Also, recovery time appears to play an essential role in the detection of gases and indicates the time consumption for the desorption of gases following adsorption. It is the desirable short recovery time for better gas sensor capabilities which can be obtained through the lowest value of the adsorption energy, as mentioned in the Equation (9) [62,65].
τ = υ 0 1 exp ( E ads kT )
where τ stands for the recovery time, Eads for the adsorption energy, υ0 for the attempt frequency, T for temperature, and k for Boltzmann constant. As per the equation, the lower the value of adsorption energy, the lower the recovery time will be; this is one of the sensor’s desirable attributes. As per the results, adsorption energy for the Pd-doped graphene was −4.3 eV and, for the Pd-decorated graphene, it was −0.44 eV. This is far lower than Pd-doped graphene, meaning that recovery time for the Pd-decorated graphene was much lower than Pd-decorated graphene. We can conclude, therefore, that Pd-decorated graphene is a better gas sensor than Pd-decorated graphene.
The distribution map for electron density with the projection of Pd-doped graphene and Pd-decorated graphene is shown in Figure 6a and Figure 6b,c, respectively.
These maps showed that, for both Pd-doped graphene and Pd-decorated graphene, conditions electron density concentrated at the Pd atom. Pd-decorated graphene, however, showed a concentration of electron density at Pd (above the surface of graphene) when visualized using only 3D projection.
In addition to that, Figure 6d–f shows a projection map with the distribution of an electronic localisation function shown later [66,67,68,69] for both Pd-doped and Pd-decorated graphene. To determine the location of electron pairs, electron localisation function is a vital phenomenon [69]. As shown in the ELF graph, Figure 6d shows the location of electron pairs for Pd-doped conditions. It shows that, for Pd-doped graphene, electron pairs localised at the place where the Pd atom attached while, for Pd-decorated graphene, electron pairs are evenly distributed at the corner atoms, as depicted in Figure 6.
The condition for electron density contour plots, and electron localisation function plots with the projection of Pd-doped/decorated graphene after adsorption H2 molecule, is shown in Figure 7.
Figure 7 shows that, for Pd-doped graphene after H2 adsorption, the concentration of electron density shifted from the Pd atom to the H2 molecule (Figure 7a,b), while there was no effect of electron density concentration over Pd-decorated graphene (Figure 7c,d). Because Pd atoms, unlike the doped condition, are not on the surface of the graphene in the decorated condition, the Pd atom localised at the surface of the graphene. Additionally, Figure 7e–h shows electron localisation function for Pd-doped/decorated graphene after H2 adsorption. It shows that, following H2 adsorption, the concentrated electron pairs in Pd-doped graphene are evenly distributed at the atom’s corners. It also shows that, at the point where the Pd atom was doped, no charge localisation occurred (Figure 7e,f). In contrast, for Pd-decorated graphene after H2 adsorption, electron pairs were concentrated on the corner, at the location where H2 was attached to a Pd atom (Figure 7g,h).

4. Conclusions

In the present work, the sensing capabilities of Pd-doped graphene and Pd-decorated graphene were investigated through DFT calculations of hydrogen molecules. Properties of these clusters were analysed by using the value of bond distance, adsorption energy, and NBO population analysis. During the adsorption of the H2 molecule, charge transfer from the H2 molecule to the Pd-doped and Pd-decorated graphene clusters occurred. This shows that Pd-doped/decorated graphene behaves as an electron acceptor. Further results show that decreases to the HLG values of Pd-doped and Pd-decorated graphene increases the electrical conductivity for both Pd-doped and Pd-decorated graphene following adsorption of H2. While analysing a comparison between Pd-doped and Pd-decorated graphene for the adsorption of H2 molecule, the chemical activity and electrical conductivity are higher for Pd-decorated graphene in comparison to Pd-doped graphene. In contrast, the charge transfer of Pd-doped graphene is better than Pd-decorated graphene. Furthermore, recovery time is much lower for Pd-decorated graphene compared to Pd-doped graphene. It can be concluded, therefore, that the Pd-decorated graphene has much better sensing capabilities compared to Pd-doped graphene and, therefore, the analysed sensing film can be envisioned as a potential nanosensor platform.

Author Contributions

Conceptualization, A.G. and A.Y.; resources, A.G. and A.Y.; writing—original draft preparation, V.K. and A.Y.; writing—review and editing, V.K., A.Y.; supervision and project administration, K.M. and A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

Authors gratefully acknowledge Start Research Grant (SRG/2020/001895) provided by Science and Engineering Research Board, Department of Science and Technology, India. Kunal Mondal gratefully acknowledges the Energy & Environment S & T at the Idaho National Laboratory, the USA for their support.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Ji, H.; Zeng, W.; Li, Y. Gas sensing mechanisms of metal oxide semiconductors: A focus review. Nanoscale 2019, 11, 22664–22684. [Google Scholar] [CrossRef] [PubMed]
  2. Gupta, A.; Pandey, S.S.; Nayak, M.; Maity, A.; Majumder, S.B.; Bhattacharya, S. Hydrogen sensing based on nanoporous silica-embedded ultra dense ZnO nanobundles. RSC Adv. 2014, 4, 7476–7482. [Google Scholar] [CrossRef]
  3. Kaskhedikar, N.A.; Maier, J. Lithium Storage in carbon nanostructures. Adv. Mater. 2009, 21, 2664–2680. [Google Scholar] [CrossRef]
  4. Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J.W.; Potts, J.R.; Ruoff, R.S. Graphene and graphene oxide: Synthesis, properties, and applications. Adv. Mater. 2010, 22, 3906–3924. [Google Scholar] [CrossRef]
  5. Wang, Y.; Li, Z.; Wang, J.; Li, J.; Lin, Y. Graphene and graphene oxide: Biofunctionalization and applications in biotechnology. Trends Biotechnol. 2011, 29, 205–212. [Google Scholar] [CrossRef]
  6. Esrafili, M.D.; Dinparast, L. The selective adsorption of formaldehyde and methanol over Al- or Si-decorated graphene oxide: A DFT study. J. Mol. Graph. Model. 2018, 80, 25–31. [Google Scholar] [CrossRef] [PubMed]
  7. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [Green Version]
  8. Li, X.; Wang, X.; Zhang, L.; Lee, S.; Dai, H. Chemically derived, ultrasmooth graphene nanoribbon semiconductors. Science 2008, 319, 1229–1232. [Google Scholar] [CrossRef]
  9. Neto, A.H.C.; Guinea, F.; Peres, N.M.R.; Novoselov, K.; Geim, A.K. The electronic properties of graphene. Rev. Mod. Phys. 2009, 81, 109–162. [Google Scholar] [CrossRef] [Green Version]
  10. Lahiri, J.; Lin, Y.; Bozkurt, P.; Oleynik, I.; Batzill, M. An extended defect in graphene as a metallic wire. Nat. Nanotechnol. 2010, 5, 326–329. [Google Scholar] [CrossRef]
  11. Choi, H.J.; Jung, S.-M.; Seo, J.-M.; Chang, D.W.; Dai, L.; Baek, J.-B. Graphene for energy conversion and storage in fuel cells and supercapacitors. Nano Energy 2012, 1, 534–551. [Google Scholar] [CrossRef]
  12. Bonaccorso, F.; Colombo, L.; Yu, G.; Stoller, M.; Tozzini, V.; Ferrari, A.C.; Ruoff, R.S.; Pellegrini, V. Graphene, related two-dimensional crystals, and hybrid systems for energy conversion and storage. Science 2015, 347, 1246501. [Google Scholar] [CrossRef] [PubMed]
  13. Kamat, P.V. Graphene-based nanoassemblies for energy conversion. J. Phys. Chem. Lett. 2011, 2, 242–251. [Google Scholar] [CrossRef]
  14. Yoon, H.J.; Jun, D.H.; Yang, J.H.; Zhou, Z.; Yang, S.S.; Cheng, M.M.-C. Carbon dioxide gas sensor using a graphene sheet. Sens. Actuators B Chem. 2011, 157, 310–313. [Google Scholar] [CrossRef]
  15. Zhang, B.; Cui, T. An ultrasensitive and low-cost graphene sensor based on layer-by-layer nano self-assembly. Appl. Phys. Lett. 2011, 98, 073116. [Google Scholar] [CrossRef]
  16. Rad, A.S. Density functional theory study of the adsorption of MeOH and EtOH on the surface of Pt-decorated graphene. Phys. E Low-Dimens. Syst. Nanostruct. 2016, 83, 135–140. [Google Scholar] [CrossRef]
  17. Hughes, Z.E.; Walsh, T.R. Computational chemistry for graphene-based energy applications: Progress and challenges. Nanoscale 2015, 7, 6883–6908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Perreault, F.; Faria, A.F.; Elimelech, M. Environmental applications of graphene-based nanomaterials. Chem. Soc. Rev. 2015, 44, 5861–5896. [Google Scholar] [CrossRef] [PubMed]
  19. Prezhdo, O.V.; Kamat, P.V.; Schatz, G.C. Virtual issue: Graphene and functionalized graphene. J. Phys. Chem. C 2011, 115, 3195–3197. [Google Scholar] [CrossRef]
  20. Wanno, B.; Tabtimsai, C. A DFT investigation of CO adsorption on VIIIB transition metal-doped graphene sheets. Superlattices Microstruct. 2014, 67, 110–117. [Google Scholar] [CrossRef]
  21. Sun, Y.; Chen, L.; Zhang, F.; Li, D.; Pan, H.; Ye, J. First-principles studies of HF molecule adsorption on intrinsic graphene and Al-doped graphene. Solid State Commun. 2010, 150, 1906–1910. [Google Scholar] [CrossRef]
  22. Ri, N.-C.; Wi, J.-H.; Kim, N.-H.; Ri, S.-I. First principles study on the structural, electronic, and transport properties of the Armchair Graphane, fluorographene, fluorographane/graphene heterostructure nanoribbons terminated by H and F atoms. Phys. E Low-Dimens. Syst. Nanostruct. 2019, 108, 226–232. [Google Scholar] [CrossRef]
  23. Song, E.H.; Wen, Z.; Jiang, Q. CO Catalytic oxidation on copper-embedded graphene. J. Phys. Chem. C 2011, 115, 3678–3683. [Google Scholar] [CrossRef]
  24. Zhang, T.; Xue, Q.; Shan, M.; Jiao, Z.; Zhou, X.; Ling, C.; Yan, Z. Adsorption and catalytic activation of O2 molecule on the surface of Au-doped graphene under an external electric field. J. Phys. Chem. C 2012, 116, 19918–19924. [Google Scholar] [CrossRef]
  25. Ma, L.; Zhang, J.-M.; Xu, K.-W.; Ji, V. A first-principles study on gas sensing properties of graphene and Pd-doped graphene. Appl. Surf. Sci. 2015, 343, 121–127. [Google Scholar] [CrossRef]
  26. Li, Y.; Zhou, Z.; Yu, G.; Chen, W.; Chen, Z. CO catalytic oxidation on iron-embedded graphene: Computational quest for low-cost nanocatalysts. J. Phys. Chem. C 2010, 114, 6250–6254. [Google Scholar] [CrossRef]
  27. Huang, C.; Ye, X.; Chen, C.; Lin, S.; Xie, D. A computational investigation of CO oxidation on ruthenium-embedded hexagonal boron nitride nanosheet. Comput. Theor. Chem. 2013, 1011, 5–10. [Google Scholar] [CrossRef]
  28. Wu, M.; Cao, C.; Jiang, J.Z. Electronic structure of substitutionally Mn-doped graphene. New J. Phys. 2010, 12, 063020. [Google Scholar] [CrossRef]
  29. Gupta, A.; Gangopadhyay, S.; Gangopadhyay, K.; Bhattacharya, S. Palladium-functionalized nanostructured platforms for enhanced hydrogen sensing. Nanomater. Nanotechnol. 2016, 6, 40. [Google Scholar] [CrossRef] [Green Version]
  30. Gupta, A.; Srivastava, A.; Mathai, C.J.; Gangopadhyay, K.; Gangopadhyay, S.; Bhattacharya, S. Nano porous palladium sensor for sensitive and rapid detection of hydrogen. Sens. Lett. 2014, 12, 1279–1285. [Google Scholar] [CrossRef]
  31. Gupta, A.; Parida, P.K.; Pal, P. Functional Films for Gas Sensing Applications: A Review; Springer: Singapore, 2019. [Google Scholar]
  32. Alfano, B.; Polichetti, T.; Miglietta, M.L.; Massera, E.; Schiattarella, C.; Ricciardella, F.; Di Francia, G. Fully eco-friendly H 2 sensing device based on Pd-decorated graphene. Sens. Actuators B Chem. 2017, 239, 1144–1152. [Google Scholar] [CrossRef]
  33. Singla, M.; Jaggi, N. Enhanced hydrogen sensing properties in copper decorated nitrogen doped defective graphene nanoribbons: DFT study. Phys. E Low-Dimens. Syst. Nanostruct. 2021, 131, 114756. [Google Scholar] [CrossRef]
  34. Bakhshi, F.; Farhadian, N. Co-doped graphene sheets as a novel adsorbent for hydrogen storage: DFT and DFT-D3 correction dispersion study. Int. J. Hydrogen Energy 2018, 43, 8355–8364. [Google Scholar] [CrossRef]
  35. Zhou, Q.; Yong, Y.; Ju, W.; Su, X.; Li, X.; Wang, C.; Fu, Z. DFT study of the adsorption of 2, 3, 7, 8-tetrachlorodibenzofuran (TCDF) on vacancy-defected graphene doped with Mn and Fe. Curr. Appl. Phys. 2018, 18, 61–67. [Google Scholar] [CrossRef]
  36. Orazi, V.; Ambrusi, R.; Marchetti, J.; Pronsato, M. DFT study of the hydrogen adsorption and storage on Ni4 cluster embedded in multivacancy graphene. Int. J. Hydrogen Energy 2020, 45, 30805–30817. [Google Scholar] [CrossRef]
  37. Tian, W.; Zhang, Y.; Wang, Y.; Liu, T.; Cui, H. A study on the hydrogen storage performance of graphene–Pd(T)–graphene structure. Int. J. Hydrogen Energy 2020, 45, 12376–12383. [Google Scholar] [CrossRef]
  38. Gecim, G.; Ozekmekci, M.; Fellah, M. Ga and Ge-doped graphene structures: A DFT study of sensor applications for methanol. Comput. Theor. Chem. 2020, 1180, 112828. [Google Scholar] [CrossRef]
  39. Zheng, N.; Yang, S.; Xu, H.; Lan, Z.; Wang, Z.; Gu, H. A DFT study of the enhanced hydrogen storage performance of the Li-decorated graphene nanoribbons. Vacuum 2020, 171, 109011. [Google Scholar] [CrossRef]
  40. Ostad, F.Z.; Ghazi, M.; Javan, M.; Izadifard, M. DFT study of Ptn, Pdn, and Irn (n = 5, 6) clusters adsorbed on graphene: Structural and electronic properties. Phys. B Condens. Matter 2019, 575, 411678. [Google Scholar] [CrossRef]
  41. Montejo-Alvaro, F.; Rojas-Chávez, H.; Román-Doval, R.; Mtz-Enriquez, A.; Cruz-Martínez, H.; Medina, D. Stability of Pd clusters supported on pristine, B-doped, and defective graphene quantum dots, and their reactivity toward oxygen adsorption: A DFT analysis. Solid State Sci. 2019, 93, 55–61. [Google Scholar] [CrossRef]
  42. Amaya-Roncancio, S.; Blanco, A.G.; Linares, D.; Sapag, K. DFT study of hydrogen adsorption on Ni/graphene. Appl. Surf. Sci. 2018, 447, 254–260. [Google Scholar] [CrossRef]
  43. Cui, H.; Zhang, X.; Chen, D.; Tang, J. Pt & Pd decorated CNT as a workable media for SOF2 sensing: A DFT study. Appl. Surf. Sci. 2018, 471, 335–341. [Google Scholar] [CrossRef]
  44. Cheng, J.; Hu, D.; Yao, A.; Gao, Y.; Asadi, H. A computational study on the Pd-decorated ZnO nanocluster for H2 gas sensing: A comparison with experimental results. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 124, 114237. [Google Scholar] [CrossRef]
  45. Khodadadi, Z. Evaluation of H2S sensing characteristics of metals–doped graphene and metals-decorated graphene: Insights from DFT study. Phys. E Low-Dimens. Syst. Nanostruct. 2018, 99, 261–268. [Google Scholar] [CrossRef]
  46. Kohn, W.; Sham, L. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133. [Google Scholar] [CrossRef] [Green Version]
  47. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09, Revision d. 01; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  48. O’Boyle, N.M.; Tenderholt, A.L.; Langner, K. cclib: A library for package-independent computational chemistry algorithms. J. Comput. Chem. 2008, 29, 839–845. [Google Scholar] [CrossRef]
  49. Foresman, J.B. Exploring Chemistry with Electronic Structure Methods: A Guide to Using Gaussian, 2nd ed.; Gaussian Inc.: Pittsburgh, PA, USA, 1996. [Google Scholar]
  50. Kulkarni, B.S.; Krishnamurty, S.; Pal, S. Probing Lewis acidity and reactivity of Sn- and Ti-beta zeolite using industrially important moieties: A periodic density functional study. J. Mol. Catal. A Chem. 2010, 329, 36–43. [Google Scholar] [CrossRef]
  51. Modi, C.K.; Trivedi, P.M.; Chudasama, J.A.; Nakum, H.D.; Parmar, D.K.; Gupta, S.; Jha, P.K. Zeolite-Y entrapped bivalent transition metal complexes as hybrid nanocatalysts: Density functional theory investigation and catalytic aspects. Green Chem. Lett. Rev. 2014, 7, 278–287. [Google Scholar] [CrossRef] [Green Version]
  52. Geerlings, P.; De Proft, A.F.; Langenaeker, W. Conceptual density functional theory. Chem. Rev. 2003, 103, 1793–1874. [Google Scholar] [CrossRef]
  53. Parr, R.G.; Szentpály, L.V.; Liu, S. Electrophilicity index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  54. Janak, J.F. Proof that∂E/∂ni = ε in density-functional theory. Phys. Rev. B 1978, 18, 7165–7168. [Google Scholar] [CrossRef]
  55. Pearson, R.G. Chemical hardness and density functional theory. J. Chem. Sci. 2005, 117, 369–377. [Google Scholar] [CrossRef]
  56. Pearson, R.G. The electronic chemical potential and chemical hardness. J. Mol. Struct. TheoChem. 1992, 255, 261–270. [Google Scholar] [CrossRef]
  57. Nobel, N.; Bamba, K.; Patrice, O.; Ziao, N. NBO population analysis and electronic calculation of four azopyridine ruthenium complexes by DFT method. Comput. Chem. 2017, 5, 51–64. [Google Scholar] [CrossRef] [Green Version]
  58. Pan, S.; Solà, M.; Chattaraj, P.K. On the validity of the maximum hardness principle and the minimum electrophilicity principle during chemical reactions. J. Phys. Chem. A 2013, 117, 1843–1852. [Google Scholar] [CrossRef] [PubMed]
  59. Pearson, R.G. The principle of maximum hardness. Accounts Chem. Res. 1993, 26, 250–255. [Google Scholar] [CrossRef]
  60. Mao, S.; Zhou, H.; Wu, S.; Yang, J.; Li, Z.; Wei, X.; Wang, X.; Wang, Z.; Li, J. High performance hydrogen sensor based on Pd/TiO2 composite film. Int. J. Hydrogen Energy 2018, 43, 22727–22732. [Google Scholar] [CrossRef]
  61. Raghu, S.; Ramaprabhu, S. Nanostructured palladium modified graphitic carbon nitride–High performance room temperature hydrogen sensor. Int. J. Hydrogen Energy 2016, 41, 20779–20786. [Google Scholar] [CrossRef]
  62. Hadipour, N.L.; Peyghan, A.A.; Soleymanabadi, H. Theoretical study on the Al-doped ZnO nanoclusters for CO chemical sensors. J. Phys. Chem. C 2015, 119, 6398–6404. [Google Scholar] [CrossRef]
  63. Yong, Y.; Jiang, H.; Lv, S.; Cao, J.; Li, X. The cluster-assembled nanowires based on M12N12 (M = Al and Ga) clusters as potential gas sensors for CO, NO, and NO2 detection. Phys. Chem. Chem. Phys. 2016, 18, 21431–21441. [Google Scholar] [CrossRef] [PubMed]
  64. Bhuvaneswari, R.; Nagarajan, V.; Chandiramouli, R. First-principles insights on the electronic and field emission properties of Ga and Al doped germanium nanocones. J. Electron Spectrosc. Relat. Phenom. 2018, 227, 15–22. [Google Scholar] [CrossRef]
  65. Beheshtian, J.; Peyghan, A.A.; Bagheri, Z. Detection of phosgene by Sc-doped BN nanotubes: A DFT study. Sens. Actuators B Chem. 2012, 171–172, 846–852. [Google Scholar] [CrossRef]
  66. Savin, A.; Becke, A.D.; Flad, J.; Nesper, R.; Preuss, H.; Von Schnering, H.G. A new look at electron localization. Angew. Chem. Int. Ed. 1991, 30, 409–412. [Google Scholar] [CrossRef]
  67. Becke, A.D.; Edgecombe, E.K. A simple measure of electron localization in atomic and molecular systems. J. Chem. Phys. 1990, 92, 5397–5403. [Google Scholar] [CrossRef]
  68. Silvi, B.; Savin, A. Classification of chemical bonds based on topological analysis of electron localization functions. Nature 1994, 371, 683–686. [Google Scholar] [CrossRef]
  69. Fuentealba, P.; Chamorro, E.; Santos, J.C. Chapter 5 Understanding and using the electron localization function. In Theoretical Aspects of Chemical Reactivity; 2007; Elsevier: Amsterdam, The Netherlands; Volume 19, pp. 57–85. [Google Scholar] [CrossRef]
Figure 1. DFT optimized geometries of (a) Pd-doped graphene, (b) Pd-doped graphene-H2, (c) Pd-decorated graphene, (d) Pd-decorated graphene-H2.
Figure 1. DFT optimized geometries of (a) Pd-doped graphene, (b) Pd-doped graphene-H2, (c) Pd-decorated graphene, (d) Pd-decorated graphene-H2.
Energies 14 05738 g001
Figure 2. Electrostatic potential of (a) Pd-doped graphene, (b) Pd-doped graphene-H2, (c) Pd-decorated graphene, (d) Pd-decorated graphene-H2. The charge accumulation regions are in red.
Figure 2. Electrostatic potential of (a) Pd-doped graphene, (b) Pd-doped graphene-H2, (c) Pd-decorated graphene, (d) Pd-decorated graphene-H2. The charge accumulation regions are in red.
Energies 14 05738 g002
Figure 3. NBO atomic charge distributions for the optimised structures of (a) Pd-doped graphene, (b) Pd-decorated graphene, (c) Pd-doped graphene-H2, and (d) Pd-decorated graphene-H2.
Figure 3. NBO atomic charge distributions for the optimised structures of (a) Pd-doped graphene, (b) Pd-decorated graphene, (c) Pd-doped graphene-H2, and (d) Pd-decorated graphene-H2.
Energies 14 05738 g003
Figure 4. HOMO and LUMO representations of α and βMOs for the optimised structures of (a) Pd-doped graphene, (b) Pd-doped graphene-H2, (c) Pd-decorated graphene, and (d) Pd-decorated graphene-H2.
Figure 4. HOMO and LUMO representations of α and βMOs for the optimised structures of (a) Pd-doped graphene, (b) Pd-doped graphene-H2, (c) Pd-decorated graphene, and (d) Pd-decorated graphene-H2.
Energies 14 05738 g004
Figure 5. DOS plots shown in (a,c) for Pd doped graphene and (b,d) for Pd decorated graphene before and after adsorption of H2. Blue color indicates the DOS spectrum, green color indicates occupied orbitals and red color indicates virtual orbitals.
Figure 5. DOS plots shown in (a,c) for Pd doped graphene and (b,d) for Pd decorated graphene before and after adsorption of H2. Blue color indicates the DOS spectrum, green color indicates occupied orbitals and red color indicates virtual orbitals.
Energies 14 05738 g005
Figure 6. 2D and 3D Electron density contour plots of (a) Pd-doped graphene (b,c) Pd-decorated graphene and Electron localization function projections of (d) Pd-doped graphene (e,f) Pd-decorated graphene. The charge accumulation regions are in red, while the charge depletion regions are in blue.
Figure 6. 2D and 3D Electron density contour plots of (a) Pd-doped graphene (b,c) Pd-decorated graphene and Electron localization function projections of (d) Pd-doped graphene (e,f) Pd-decorated graphene. The charge accumulation regions are in red, while the charge depletion regions are in blue.
Energies 14 05738 g006aEnergies 14 05738 g006b
Figure 7. 2D and 3D Electron density contour plots and projections of: (a,b) Pd-doped graphene-H2, (c,d) Pd-decorated graphene (G)-H2, 2D and 3D Electron localization function plots, and projections of (e,f) Pd-decorated graphene (G)-H2 and (g,h) Pd-decorated graphene-H2. The charge accumulation regions are in red, while the charge depletion regions are in blue.
Figure 7. 2D and 3D Electron density contour plots and projections of: (a,b) Pd-doped graphene-H2, (c,d) Pd-decorated graphene (G)-H2, 2D and 3D Electron localization function plots, and projections of (e,f) Pd-decorated graphene (G)-H2 and (g,h) Pd-decorated graphene-H2. The charge accumulation regions are in red, while the charge depletion regions are in blue.
Energies 14 05738 g007aEnergies 14 05738 g007b
Table 1. Properties of the Pd-doped graphene cluster before and after H2 interaction.
Table 1. Properties of the Pd-doped graphene cluster before and after H2 interaction.
Properties Pd-Doped GrapheneH2 Adsorbed on Pd-Doped Graphene
Distances (Å)Pd-C 1.80, 1.81,1.82Å1.97 Å, 2.0 Å, 2.0 Å
Pd-H 2.34 Å
H-H0.740.759
NBO charges (e)Pd atom1.2141.006
C atoms
(bonded with Pd)
−0.274, −0.274, −0.274−0.215, −0.174, −0.191
Adsorbed H2 molecule----0, 0.067
Adsorption energies (eV)∆E----−4.3
α MOs (eV)HOMO−4.171−3.999
LUMO−3.611−3.511
HLG0.560.488
Chemical Hardness0.280.244
Chemical Potential−3.891−3.755
Electronegativity3.8913.755
Electrophilicity27.03528.893
Table 2. Properties of the Pd-decorated graphene cluster before and after H2 interaction.
Table 2. Properties of the Pd-decorated graphene cluster before and after H2 interaction.
Properties Pd-Decorated Graphene H2 Adsorbed on Pd-Decorated Graphene
Distances (Å)Pd-C 2.13 2.13
Pd-H --1.85
H-H0.740.803
NBO charges (e)Pd atom0.5990.484
C atoms (bonded with Pd)0.006−0.073
Absorbed H2 molecule-0.023, 0.027
Adsorption energies (eV)∆E −0.44
α MOs (eV)HOMO−4.008 −3.937
LUMO−3.499 −3.414
HLG0.5090.477
Chemical Hardness0.2540.238
Chemical Potential−3.754−3.676
Electronegativity3.7543.676
Electrophilicity27.74128.388
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kishnani, V.; Yadav, A.; Mondal, K.; Gupta, A. Palladium-Functionalized Graphene for Hydrogen Sensing Performance: Theoretical Studies. Energies 2021, 14, 5738. https://doi.org/10.3390/en14185738

AMA Style

Kishnani V, Yadav A, Mondal K, Gupta A. Palladium-Functionalized Graphene for Hydrogen Sensing Performance: Theoretical Studies. Energies. 2021; 14(18):5738. https://doi.org/10.3390/en14185738

Chicago/Turabian Style

Kishnani, Vinay, Anshul Yadav, Kunal Mondal, and Ankur Gupta. 2021. "Palladium-Functionalized Graphene for Hydrogen Sensing Performance: Theoretical Studies" Energies 14, no. 18: 5738. https://doi.org/10.3390/en14185738

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop