Carbon dioxide activation and transformation to HCOOH on metal clusters (M = Ni, Pd, Pt, Cu, Ag & Au) anchored on a polyaniline conducting polymer surface – an evaluation study by hybrid density functional theory

Ramasamy Shanmugamad, Arunachalam Thamaraichelvanb, Tharumeya Kuppusamy Ganesanc and Balasubramanian Viswanathan*d
aDepartment of Chemistry, Thiagarajar College, Madurai, Tamilnadu 625 009, India
bFaculty of Allied Health Sciences, Chettinad Hospital & Research Institute, Kelambakkam, Tamilnadu 603 103, India
cDepartment of Chemistry, The American College, Madurai, Tamilnadu 625 002, India
dNational Center for Catalysis Research, Indian Institute of Technology Madras, Chennai, Tamilnadu 600 036, India. E-mail: bvnathan@iitm.ac.in

Received 17th August 2016 , Accepted 17th October 2016

First published on 18th October 2016


Abstract

Developing new efficient catalysts for the electrochemical reduction of carbon dioxide to formic acid is important in the process of mitigating environmental CO2. In the present work, we have designed metal (M) clusters anchored on a polyaniline (PANI) conducting polymer electrode (M@PANI), where, M = Ni, Pd, Pt, Cu, Ag & Au, and evaluated their potential catalytic activity towards CO2 reduction by means of computational hydrogen electrode using hybrid density functional theory methods. The predicted binding energy and electronic properties of M@PANI suggest a thermodynamically feasible reaction which retains its conducting property with enhancement. The modified electrodes favour the formation of HCOOH involving H*COO species via the formate pathway. The computed limiting potentials suggest that Cu@PANI is a suitable electrode material for the CO2 reduction reaction leading to HCOOH.


Introduction

In recent years, renewable energy sources have received immense attention due to their utility in the electrochemical conversion of greenhouse gas, CO2, into value added chemicals1–3 which can potentially decrease the environmental impact of CO2. CO2 can produce a variety of chemicals4–7 among which HCOOH is considered in the present work. Being a highly functionalised molecule, the activated CO2 species has been used as a building block in fine chemical production,8,9 fuel cells10–12 and hydrogen storage.13 General electro-reduction of CO2 has been carried out on electrodes either in a direct or an indirect manner. Although a large number of electrode14–20 materials have been tried in this regard, the low efficiency of the process needs to be improved. Special attention is required in the choice of electrode material for the production of chosen products from CO2. HCOOH production is feasible on certain metal electrodes such as zinc, lead, mercury, thallium, indium, tin, cadmium and bismuth,21,22 though it requires a high negative potential of about −2.0 V vs. SCE making the yield very low. Various electrode materials were attempted in improving the efficiency by decreasing the over potential.23–26

The main problem in electro reduction of CO2 is the stability of chosen electrodes. Certain electrodes are not stable both in acidic and basic media. During the reaction, intermediate ions in the medium can change the pH. Hence, keeping the electrode for a prolonged period under stable condition is desirable. Growing interest in using polymer supported systems has been due to their enhanced activity in CO2 transformation compared to unsupported one.27–29 Polyaniline (PANI) as a conducting material has special characteristics which include easy synthesis, low cost and stability at ordinary conditions.30 It plays a vital role in many potential applications30 including light emitting diodes,31 fuel cells,32 capacitor,33 super capacitor,34 selective electrode,35 electrical conductivity,36 redox potential higher than metal, corrosion-resistance37,38 tuneable solubility and thermal stability.39

Composites of PANI were used in electro based sensors,40 as well as catalytic and photo catalytic reactions.41,42 Recent reports reveal that a multifunctional catalyst exhibits better activity in CO2 fixation reactions.43,44 CO2 reduction involves electron addition in the first step followed by addition of proton, and then a series of proton, electron and a couple of proton and electron. All these reaction steps are competitive and control the product formation and selectivity. Under these conditions, a suitable catalyst must change its role according to the nature of the substrate. However, this does not happen in the most routinely used catalytic systems. Therefore, designing a catalyst with multifunctional activity is desired. PANI can serve as a better candidate in such reactions, since it can either uptake or release electron and proton45 by altering its structural form. In addition, it can provide a good support in stabilising the metal nanoparticles.46 Sulfonated PANI has found application in CO2 sensing.47 Moreover, PANI is used in hydrogenation of N2 and CO2 under electrochemical conditions.48,49 Metals like Ni, Pd, Pt, Cu, Ag, and Au serve an electrodes in most of the electro reduction reactions of CO2.50 Small quantities of metals are preferred due to their availability and cost. Generally, pure forms of Ni, Pd, Pt, Cu, Ag, and Au are not useful in the production of HCOOH due to CO positioning and their tendency to form other hydrogenated products.21 However, suitably modified metal electrodes are performing better activity towards HCOOH formation with lower negative potentials.12,22,23,26,51–53 Ni coated PANI has been used in sensors54 and absorption of microwaves.55 Pd–PANI/carbon nanotube composite has better catalytic activity towards conversion of CO2 to HCOOH.56 Platinum and its based alloys supported on PANI electrode were used in oxidation of glycerol.46 Copper polypyrrole electrode was employed in CO2 reduction under high pressure conditions.57 PANI protected Ag behaves as a good electrode in oxidation of hydrazine.58 Gold–PANI nanocomposite has performed better in electrochemical reduction of oxygen than other counterparts.59 These reports reveal that metal@PANI systems have a significant role in electrode mediated reactions. Hence, designing and evaluating the activity of a suitable electrode with multiple functionality involving metal–PANI system for CO2 reduction reaction is more trustworthy.

The present work aims to probe as catalyst, the electrodes made of metal clusters anchored on PANI as polymer support, towards CO2 reduction leading to HCOOH by employing hybrid density functional theory methods.

Molecular models

The model used for the study was constructed as follows: (i) the structure of metal cluster with M7 unit, where, M = Ni, Pd, Pt, Cu, Ag & Au studied extensively in literature,60–65 was chosen. (ii) An oligomer of PANI, which consists of 3 units of aniline was used to mimic the polymeric site. (iii) Subsequently, the metal M7 unit was placed on the terminal of oligomer at the N site and is referred to as M@PANI. This configuration was considered for further study.

Computational techniques

The geometry optimization of pure M@PANI and those with adsorbates were performed at B3LYP level of hybrid density functional theory using a basis set of LANL2DZ for metals and 6-31G* for other elements. Since, the configurations of M@PANI are present in the ground state, the total spin value of 1 for Ni, Pd & Pd and 2 for Cu, Ag & Au models were assigned respectively. Using the same level of theory, vibrational frequencies were calculated to identify whether the obtained configuration is either a stationary point or transition state on the potential energy surface. The binding energy of M@PANI clusters was obtained using the formula of BEM = EM@PANIEMEPANI where, EM, EM@PANI, and EPANI are total zero point energy (ZPE) corrected electronic energy of metal cluster, metal PANI and pure PANI respectively. For the adsorbates the binding energy was calculated by, BEads = Eadsorbate+M@PANIEM@PANIEadsorbate where, ECuadsorbate+M@PANI, EM@PANI and Eadsorbate are zero point energy corrected electronic energy of Cuadsorbate+M@PANI, M@PANI and adsorbate respectively. Ab initio molecular dynamics (AIMD) simulations were performed by using the final configurations obtained in the geometry optimisation as the initial structure. The ab initio Born–Oppenheimer molecular dynamics at B3LYP/LANL2DZ level was employed to find the stability of the electrode. Natural Bonding Orbital (NBO) analysis was carried out using NBO 3.1 (ref. 66) software implemented in Gaussian 09. Computational hydrogen electrode method67–69 was applied to analyze the free energy of reaction pathway. For an elementary reaction, A* + (H+ + e) → AH*, the free energy is defined as, ΔGA*→AH*(U) = GAH*GA* − 0.5GH2 + ne(U) where, the terms G, ne and U represent free energy, number of electrons and an applied potential respectively. All these quantum mechanical calculations were carried out using Gaussian 09 software package.70 Density of States (DOS) and Partial Density of States (PDOS) were plotted using GaussSum software.71

Results and discussion

Structural models designed for M–PANI

The metal cluster anchored on oligomer model of M–PANI, considered for the examination is depicted in Fig. 1. Studies on M7 (M = Ni, Pd, Pt, Cu, Ag & Au) clusters are extensively available in literature.60–65 The M7 clusters anchored on oligomer containing three units of aniline are represented in Fig. 1a–f respectively.
image file: c6ra20715d-f1.tif
Fig. 1 Optimised geometries of M@PANI, where M is (a) Ni, (b) Pd, (c) Pt, (d) Cu, (e) Ag and (f) Au respectively.

The metal clusters were anchored through metal–nitrogen sigma bond. The structural parameters of M@PANI are presented in Table 1.

Table 1 Binding energy (ΔE) eV, bond angle (θ)°, bond length (l) Å and bond order for M@PANI
M ΔE (eV) θ (°) l (Å) Bond order
C–N–M C–N N–M N–M
Ni −2.25 108.71 1.37 1.86 0.745
Pd −1.48 110.31 1.35 2.08 0.715
Pt −3.83 110.96 1.34 2.03 0.674
Cu −1.83 119.99 1.36 1.95 0.652
Ag −1.07 121.02 1.38 2.16 0.739
Au −1.72 131.26 1.34 2.11 0.478


The observed bond distances of M–N reveal a slight variation compared to those observed in complexes.72–75 This is attributed to the cluster nature of metal instead of being a single atom which causes electronic and steric repulsion with the oligomer unit. Furthermore, the bond order of M–N experiences a gradual decrease from Ni to Au with the exception being Ag, where an increase in observed. Pure clusters of all metals have pentagonal bipyramid structure (ESI S1). On anchoring, the geometry of metal clusters for Ni, Pd and Pt was found to become distorted cube. However, Cu and Ag exhibited pentagonal bipyramid geometry while Au revealed distorted pentagonal bipyramid geometry. This structural distortion reveals the strong interaction of the metal with the conducting polymer. The data for the angle, C–N–M are presented in Table 1, reveal that the cluster is located approximately in a triangular fashion which is seen above the polymer plane. The binding energy between metal and PANI shows a negative value and predict that the anchoring of metal clusters on PANI is a thermodynamically feasible process. The order of binding energy for different metals is found to be Pt > Ni > Cu > Au > Pd > Ag. That all the computed infrared frequencies are positive, predicts that the configurations are at stationary points in the potential energy surface. Furthermore, the stability of metal–PANI is characterised by means of ab initio Born–Oppenheimer molecular dynamics (BOMD) method which is presented in ESI S2.

Frontier molecular orbital (FMO) and electronic structure analysis

The conducting nature of PANI is mainly controlled by its orbital energy density of HOMO and LUMO and the position where it is located. While anchoring of metal cluster on PANI may perturb the frontier molecular orbitals of PANI, it helps to locate at specific sites on M@PANI and control of electron transfer processes. Frontier molecular orbital (HOMO and LUMO) contours of pure and M@PANI are depicted in Fig. 2.
image file: c6ra20715d-f2.tif
Fig. 2 Frontier molecular orbitals of (a) pure PANI and M@PANI where M is (b) Ni, (c) Pd, (d) Pt, (e) Cu, (f) Ag and (g) Au respectively with their energies.

The HOMO and LUMO of pure PANI spread over π bond between C–H and nitrogen, in a symmetric manner. The FMO analysis of M@PANI reveals that, anchoring of metal cluster shifted the HOMO and LUMO orbitals. In Ni, Pd and Pt, the HOMO comprises of orbitals located on metal, the π-CH and nitrogen. Compared to pure PANI, the shift of HOMO was observed in the π-CH which is present adjacent to the metal cluster. The same trend was observed in LUMO also. It is interesting to note that the shift was not similar in Pt and Pd. While the HOMO is distributed all the metal atoms in the cluster, in Pd, it is concentrated on the metal atoms nearer to the aniline molecule of PANI in Pt. In Pt and Pd, the orbitals observed are denser than in Ni and are as dense as in pure PANI. In contrast to this, the HOMO in Cu@PANI is mostly on PANI and LUMO is centred on Cu. On moving to Ag system, the HOMO levels gradually increase their presence on Ag and the LUMO moves gradually towards PANI. On Au, the HOMO was completely shifted to metal site and the LUMO was located on PANI's site. The participation of metal orbitals in the HOMO population is in the order, Pd > Pt > Ni > Au > Ag > Cu. The order appears to be influenced by the valency and size factors of the metals. While higher valence state leads to greater electron population of HOMO, the atom with larger size contributes more among a given group of elements. The perturbation of HOMO and LUMO is due to the electron releasing nature of the metal which is influenced by the arrangement of atoms in the clusters.

In general, individual metal and PANI reveal conducting behaviour. Anchoring of metal on PANI alters the conducting property of the polymer. It is essential to know the electronic properties of M@PANI in a detailed manner for better understanding of its conducting properties. Density of states of pure PANI, pure metal clusters and M@PANI are presented in Fig. 3–5 respectively. The values corresponding to HOMO and LUMO are presented in Fig. 2. In pure PANI (Fig. 3), the DOS is made up of s and p orbitals. The HOMO and LUMO contours in PANI reveal that in Ni, Pd & Pt catalyst same atom contributes to both the HOMO and LUMO. However, in Cu, Ag & Au, the HOMO & LUMO are centred in different regions. Hence, the HOMO and LUMO overlap with a minimum gap leading to metal like conducting behaviour of PANI.76 Typically, metals are good conductors, though the conducting property may change at nano level. The PDOS of pure metal clusters (Fig. 4) suggests that the metallic behaviour may be due to the population of mainly d orbital rather than s and p orbitals. Further, the orbitals in PANI are overlapping with valance and conduction bands of metal clusters from (a) to (f) (Fig. 5). PDOS of M@PANI reveals that there is no energy gap window upon anchoring of metal on PANI near HOMO–LUMO region. This result suggests that the M@PANI still retains it conducting behaviour. Since the energy gap between HOMO and LUMO is inversely proportional to the reactivity of the material, the order of any reaction that depends redox property may be in the order Cu > Au > Ag > Pt > Pd > Ni.


image file: c6ra20715d-f3.tif
Fig. 3 Density of states of pure PANI.

image file: c6ra20715d-f4.tif
Fig. 4 PDOS of pure metal clusters (a) Ni, (b) Pd, (c) Pt, (d) Cu, (e) Ag and (f) Au.

image file: c6ra20715d-f5.tif
Fig. 5 Partial density of states of M@PANI where M is (a) Ni, (b) Pd, (c) Pt, (d) Cu, (e) Ag and (f) Au respectively.

Adsorption of CO2 on PANI

M@PANI is in a heterogeneous environment consisting of a polymer and metal cluster unit, the interaction of CO2 with polyaniline site has been predicted to be predominately physical adsorption.77 Hence, the interaction of CO2 at PANI site was not taken into account and the metal site alone was considered for CO2 interaction. Such an interaction may be through various coordination modes through carbon, oxygen and combined forms are possible.78 Among the different modes, the metal–carbon bond via CO2 bent mode was considered for activation of CO2. The configurations of CO2 interaction with M@PANI for different M atoms after energy minimisation are presented in Fig. 6. The binding energy and structural parameters for the above interactions are presented in Table 2.
image file: c6ra20715d-f6.tif
Fig. 6 Geometry optimized configurations of CO2 on M@PANI where M is (a) Ni, (b) Pd, (c) Pt, (d) Cu, (e) Ag and (f) Au respectively.
Table 2 The calculated adsorption energies (Eads, eV), structural parameters and NBO charge (|q|) of CO2 on M@PANI
  Eads Angle (°) O–C–O Bond order Bond length (Å) |q|
C–O C–O C–O C–O C–N M–N M–N–C C O O
CO2   180 2.000 2.000 1.16 1.16            
Ni −0.70 145.07 1.728 1.437 1.20 1.25 1.37 1.85 106.80 0.798 −0.555 −0.588
Pd −0.20 151.22 1.577 1.823 1.22 1.19 1.35 2.07 110.08 0.848 −0.567 −0.535
Pt −0.27 146.82 1.437 1.840 1.25 1.19 1.34 2.04 111.09 0.469 −0.356 −0.358
Cu −1.23 179.54 1.816 2.031 1.17 1.16 1.36 1.95 118.95 1.082 −0.514 −0.478
Ag −0.19 179.78 1.867 2.037 1.17 1.16 1.35 2.18 120.76 1.072 −0.523 −0.488
Au −0.06 178.87 1.942 1.981 1.17 1.16 1.34 2.11 131.67 1.067 −0.537 −0.499


Data on binding energy reveal that except for copper, all other clusters weakly hold CO2 on their surfaces. On the other hand, the structural features of adsorbed CO2 on Cu@PANI resemble that of pure CO2 though the binding energy reveal that CO2 is adsorbed through chemisorption mode. A similar trend is observed in Pd and Pt too. However Ni, Pd and Pt clusters remarkably play a vital role in altering the bond angle of CO2. As seen in the table, the angle of CO2 is decreased from 180° to about 34.93°, 28.78° and 33.18° on Ni, Pd and Pt clusters respectively. The NBO charge analysis indicates that two types of charge transfer take place, the first being from metal onto CO2 with values of, −0.345q, −0.254q and −0.245q on Ni, Pd and Pt respectively in the M@PANI. The second type of electron transfer from CO2 to M where, a positive charge is observed on CO2 due the weaker adsorption. The corresponding charges are Cu(0.09q), Ag(0.061q) and Au(0.031q) on M@PANI. The negative charge transferred from M@PANI to CO2 alters its bond angle leading to structural changes. Similar structural changes are observed in pure metal clusters also. The angles for O–C–O bond found in Pd and Ni are smaller than those in other M@PANI and for Ni, it is 6.15° lower than those for Pd and Pt indicating that in Ni, the metal–polymer support interaction predominates. Thus, the polymer support helps only to stabilise the cluster in the new geometry rather than in free form geometry. In general, during the interaction, the adsorbate will be able to change the geometry of the adsorbent. The calculated values of M–N bond length and M–N–C angle for the different metals after interaction are presented in Table 2. The values resemble those observed in pure M@PANI indicating that, even after interaction with adsorbate, the configuration of the adsorbent was not changed. Thus, the M@PANI system is stable under the reaction conditions. All the computed frequencies of CO2 adsorbed on M@PANI are positive revealing that the configurations are on the stationary points in the potential energy surface. Selected vibrations of CO2 are presented in Table 3.

Table 3 Calculated vibrational frequencies of CO2 on M@PANI (cm−1)
O–C–O asymm stretching C–O symm stretching
1927 1186
2057 1225
2009 1187
2424 1364
2425 1367
2429 1371


Adsorption of HCOO* species on M@PANI

Subsequent to the interaction of CO2 with M@PANI, a proton and an electron in the medium attach to the species to produce either formate (H*COO) or *COOH species leading to two pathways for HCOOH formation. The further stability of the species is controlled by the strength of adsorption strength thus making available the species for the reaction. The optimised geometries of the formate on M@PANI site are shown in Fig. 7. The binding energy and structural features of the optimised HCOO on M@PANI are given in Table 4. The values for H*COO and *COOH are negative indicating the thermodynamic feasibility of formation of both the species. The fact that the H*COO species has greater binding energy than *COOH reveals that H*COO species is preferably adsorbed compared to *COO and cover the active site predominantly to enter into further reaction.
image file: c6ra20715d-f7.tif
Fig. 7 Geometry optimized configuration of HCOO on M@PANI where, M in (a) Ni, (b) Pd, (c) Pt, (d) Cu, (e) Ag and (f) Au respectively.
Table 4 The calculated adsorption energies (Eads, eV) and structural parameters of HCOO* and COOH* species on M@PANI
M Eads Angle (°) O–C–O Bond length (Å)
HCOO COOH C–H C–O C–O C–N M–N M–N–C
CHOO     113.15 1.10 1.25 1.25
Ni −9.13 −7.13 127.10 1.10 1.26 1.26 1.41 1.98 121.49
Pd −3.09 −3.15 131.08 1.18 1.21 1.28 1.35 2.05 111.00
Pt −3.23 −2.71 121.42 1.09 1.25 1.28 1.33 2.06 111.74
Cu −2.99 −1.23 128.52 1.10 1.26 1.26 1.33 1.96 133.86
Ag −2.61 −1.14 129.94 1.10 1.26 1.26 2.23 1.33 133.37
Au −2.61 −2.14 124.63 1.10 1.27 1.25 1.33 2.08 128.65


The formate species is adsorbed mainly through two oxygen atoms via the bridge bidendate mode. However, in Pd, the HCOO is held via bridge bidendate mode not by two oxygen atoms but connected via H–Pd on one side and through O–Pd on the other side with a distance of 1.93 Å and 2.08 Å respectively. It is interesting to observe this bridge on Ni, Cu and Ag with two metal atoms, while in Pt and Au, single metal site forms the bridge with both oxygens. The bond length of C–H and C–O are more or less similar to those observed in free HCOO species. The data indicate that the two C–O bonds in HCOO exhibit double bond character. Alteration of bond length in the symmetric one requires more energy than that for the asymmetric one. Hence, there is not much change observed in C–O bond. The angle at carbon of O–C–O was greatly influenced, which is attributed to the formation of bridge for the adjustment of coordination between the two metal atoms which are located at a distance greater than that of O–O distance in HCOO. During the interaction, the structural features of interface between the metal and PANI are slightly modified. This reveals that the interface facilitates the interaction with adsorbate in a flexible manner. To verify whether this mode of adsorption of formate is in the stationary point, the vibrational frequencies were calculated. All the values, except for Au, are positive which indicate that except Au, all are located on stationary points and Au is present in the transition state. The characteristic vibrations for the CHOO obtained from the computations are presented in Table 5. The frequency data reveal that there is a small difference among the values of vibrational modes of various M@PANI systems which is metal dependant.

Table 5 Calculated vibrational frequencies of HCOO on M@PANI (cm−1)
M C–H out of plane bending C–O symm str. C–O asymm str. O–C–O scissor C–H in plane bending C–H stretching
Ni 1053 1389 1631 766 1407 3061
Pd 946 1195 1820 763 1328 2057
Pt 1037 1387 1615 798 1324 3076
Cu 1047 1374 1630 767 1408 3018
Ag 1048 1367 1640 756 1412 2961
Au 1047 1380 1634 795 1343 3012


Adsorption of HCOOH on M@PANI site

For the elucidation of the most stable adsorption site for HCOOH on M@PANI, HCOOH was placed on metal and allowed for full relaxation. The final configurations obtained are depicted in Fig. 8. HCOOH on M@PANI (M = Ni, Pd & Pt) is independently located. On Ni, HCOOH is coordinating via bidendate mode through O and H of C[double bond, length as m-dash]O and O–H respectively with two adjacent metal atoms. However, monodentate coordination of HCOOH is observed on Pd through C. In the case of Pt, only a monodentate coordination was observed through O of C[double bond, length as m-dash]O to the metal. In contrast to the above, HCOOH is held on Cu, Ag and Au through monodentate coordination from O of C[double bond, length as m-dash]O perpendicular to the metal cluster. The distance between the HCOOH and M reveals that instead of a typical bonding, a weak interaction operates and hold the molecule onto the metal site. The binding energy and structural features of M@PANI are given in Table 6. That the HCOOH has a tendency to get adsorbed spontaneously on the metal site is revealed by the data. Ni@PANI has the maximum binding energy compared to other M@PANI systems. On adsorption, the M–N and C–N distances remains constant and are comparable to the values in pure form which indicates that the structure is stable. The stability of the cluster was further evaluated by means of frequency calculation. All the vibrational frequencies returned positive values thus predicting that the resultant configuration is present in one of the energy minima in the potential energy profile. Furthermore, selected vibrations related to HCOOH are tabulated in Table 7. The frequencies for Ni, Pd and Pt return lower frequencies than those of Cu, Ag and Au systems.
image file: c6ra20715d-f8.tif
Fig. 8 Geometry optimized configuration of HCOOH on M@PANI where M is (a) Ni, (b) Pd, (c) Pt, (d) Cu, (e) Ag and (f) Au respectively.
Table 6 The calculated adsorption energies (Eads, eV) and structural parameters of HCOOH* on M@PANI
M Eads Bond length (Å) Angle (°)
M–H*COOH C–H O–H C[double bond, length as m-dash]O C–O C–N M–N M–N–C O–C–O
Ni −1.38 1.92 1.09 0.97 1.27 1.33 1.37 1.88 91.53 123.05
Pd −0.54 2.16 1.10 0.97 1.27 1.38 1.35 2.07 111.51 120.06
Pt −0.58 2.19 1.09 0.97 1.27 1.35 1.34 2.04 111.23 121.87
Cu −0.39 2.09 1.09 0.98 1.22 1.32 1.36 1.89 130.44 123.84
Ag −0.46 2.44 1.09 0.98 1.21 1.32 1.36 2.19 118.55 124.23
Au −0.76 2.37 1.09 0.97 1.22 1.32 1.37 2.02 129.81 123.72


Table 7 Calculated vibrational frequencies of HCOOH on M@PANI (cm−1)
M C–H out of plane bending C[double bond, length as m-dash]O str. O–H str. C–H str. C–H + OH in plane bending C–OH str. O–H out of plane bending C–H in plane bending
Ni 867 1554 3670 3149 1299 1096 529 1354
Pd 892 1493 3676 2971 1260 1059 532 1313
Pt 930 1550 3670 3130 1313 1113 547 1347

M C–H out of plane bending C[double bond, length as m-dash]O str. O–H str. C–H str. C–H + OH in plane bending OH in plane bending O–H out of plane bending C–H in plane bending
Cu 1066 1778 3649 3152 1194 1361 721 1408
Ag 1065 1792 3652 3137 1186 1351 720 1416
Au 1066 1769 3646 3155 1194 1361 723 1412


Electrochemical reduction of CO2 to HCOOH

The electrochemical transformation pathway of CO2 to HCOOH proceeds through the following steps:53 (i) the gaseous CO2 is coupled with an electron available in solution to form CO2 species, which is further adsorbed on the metal site. Subsequently, the reaction proceeds via two possible path ways: (ii-a) since the surface species has negative charge on O, the available proton in solution reacts to form COOH* species in the adsorbed state (carboxylate pathway). (ii-b) Contrary to (ii-a), H+ + e in the solution attack the carbon of CO2 to form adsorbed HCOO species (formate pathway). (iii) Further addition of a H+ + e couple to COOH or HCOO leads to HCOOH finally. The above scheme is presented in Fig. 9. As seen from the binding energy data, Ni@PANI exhibits greater binding energy than other M@PANI systems and the values are more negative indicating that the active site is positioned and will not easily be regenerated for further reaction. Hence, metals other than Ni@PANI were included in the electrochemical reduction of CO2. The calculated Gibbs free energy profiles of CO2 to HCOOH conversion through formate and carboxylate pathways are presented in Fig. 10 and 11 respectively.
image file: c6ra20715d-f9.tif
Fig. 9 Schematic representation of electrochemical reduction of CO2 via different reaction pathways to formic acid.

image file: c6ra20715d-f10.tif
Fig. 10 Calculated Gibbs free energy profile without (U-black lines), and with applied potential (UL-red lines) for CO2 electroreduction to HCOOH through formate pathway on M@PANI where M = (a) Pd, (b) Pt, (c) Cu, (d) Ag, and (e) Au.

image file: c6ra20715d-f11.tif
Fig. 11 Calculated Gibbs free energy profile without (U-black lines), and with applied potential (UL-red lines) for CO2 electroreduction to HCOOH through carboxylate pathway on M@PANI where M = (a) Pd, (b) Pt, (c) Cu, (d) Ag, and (e) Au.

Reactions of CO2(g) + H+ + e → HCOO and HCOO + H+ + e → HCOOH, on M@PANI have an uphill step with an energy barrier (Fig. 10) of 0.88 eV, 0.85 eV, 0.48 eV, 0.71 eV and 0.58 eV for the metals of Pd, Pt, Cu, Ag and Au respectively at an applied potential of 0 V. It is interesting to note that the energy barrier is independent of metals which reveal that the reaction step in Pd, Pt and Cu is HCOO to HCOOH and, in Ag and Au, CO2 to COO are the rate determining steps in CO2 to HCOOH transformation.

In carboxylate reaction path, the calculated energy barriers (Fig. 11) for the step CO2(g) + H+ + e → COOH are 1.06 eV, 0.53 eV, 1.72 eV, 1.80 eV and 0.68 eV respectively for Pd, Pt, Cu, Ag and Au on PANI. For Pd and Pt, COOH to HCOOH and, in Cu, Ag and Au, CO2 to COOH are the rate determining steps respectively in the reduction of CO2 to HCOOH.

A comparison of the energy barrier values suggests that metal on PANI plays an important role in controlling the rate determining step and also the reaction pathway. For Pd@PANI, Ag@PANI and Au@PANI, energy barrier for the formate path way is less compared to that of the carboxylate route. In contrast to this for Pt@PANI, the carboxylate route has lower barrier. The reaction barrier can be moved further downhill by means of applying extra potential (limiting potential). The required limiting potentials correspond to the same barrier value with negative sign. By applying this limiting potential, the energy barrier is nullified and the reaction proceeds in the forward direction. The order of over potential to be applied on M@PANI for CO2 to HCOOH transformation is found to be: Cu < Au < Ag < Pt < Pd. This observed activity is mainly controlled by the HOMO–LUMO gap of the M@PANI systems. The HOMO–LUMO gap of the M@PANI follows in the order of Cu > Au > Ag > Pt > Pd. This result indicates that higher the HOMO–LUMO gap favour in lowering of the over potentials.

Conclusion

In the present work, we have explored the activation and transformation of CO2 to HCOOH on the catalytic metal cluster sites anchored on polyaniline support by employing DFT/B3LYP level of theory. Binding energy reveals that anchoring of metal cluster on PANI is thermodynamically feasible. The frontier molecular orbital analysis of the catalyst predicts the order of HOMO–LUMO gaps for different metal atoms to be Cu > Au > Ag > Pt > Pd, which reveals the order of reactivity as Cu < Au < Ag < Pt < Pd. Data on PDOS reveal that M@PANI retains the conducting behaviour of and involvement of d orbitals too. Calculation of vibrational frequencies of the adsorbed configurations indicates their stability due to the presence in energy minima. Geometry optimisation and electrochemical reduction studies for the carboxylate and formate pathway revealed that the formate pathway is favoured on M@PANI. Among the different metals, Cu appears to be most preceded one.

Acknowledgements

We acknowledge Department of Science and Technology (DST), India, for setting up the National Centre for Catalysis Research (NCCR) at India Institute of Technology Madras (IITM). The authors also acknowledge the High Performance Computing Environment (HPCE), IITM for providing computational facilities. The author, R. S, is thankful to Council for Scientific and Industrial Research (CSIR) for providing financial support vide., Ref. No. 08/117(0001)-2013-EMR-I.

References

  1. O. Inganäs and V. Sundström, Ambio, 2015, 45, 15–23 CrossRef PubMed.
  2. D. R. Kauffman, J. Thakkar, R. Siva, C. Matranga, P. R. Ohodnicki, C. Zeng and R. Jin, ACS Appl. Mater. Interfaces, 2015, 7, 15626–15632 CAS.
  3. S. E. Hosseini and M. A. Wahid, Renewable Sustainable Energy Rev., 2016, 57, 850–866 CrossRef CAS.
  4. W.-H. Wang, Y. Himeda, J. T. Muckerman, G. F. Manbeck and E. Fujita, Chem. Rev., 2015, 115, 12936–12973 CrossRef CAS PubMed.
  5. J. Klankermayer and W. Leitner, Philos. Trans. R. Soc., A, 2016, 374, 1–8 CrossRef PubMed.
  6. G. Centi, E. A. Quadrelli and S. Perathoner, Energy Environ. Sci., 2013, 6, 1711–1731 CAS.
  7. S. Moret, P. J. Dyson and G. Laurenczy, Nat. Commun., 2014, 5, 1–7 Search PubMed.
  8. X. Liu, S. Li, Y. Liu and Y. Cao, Chin. J. Catal., 2015, 36, 1461–1475 CrossRef CAS.
  9. W. Leitner, Angew. Chem., Int. Ed. Engl., 1995, 34, 2207–2221 CrossRef CAS.
  10. H. Jeon, B. Jeong, J. Joo and J. Lee, Electrocatalysis, 2014, 6, 20–32 CrossRef.
  11. N. V. Rees and R. G. Compton, J. Solid State Electrochem., 2011, 15, 2095–2100 CrossRef CAS.
  12. K. Jiang, H.-X. Zhang, S. Zou and W.-B. Cai, Phys. Chem. Chem. Phys., 2014, 16, 20360–20376 RSC.
  13. A. K. Singh, S. Singh and A. Kumar, Catal. Sci. Technol., 2016, 6, 12–40 Search PubMed.
  14. O. R. Luca and A. Q. Fenwick, J. Photochem. Photobiol., B, 2015, 152, 26–42 CrossRef CAS PubMed.
  15. I. Ganesh, Renewable Sustainable Energy Rev., 2016, 59, 1269–1297 CrossRef CAS.
  16. A. J. Martin, G. O. Larrazabal and J. Perez-Ramirez, Green Chem., 2015, 17, 5114–5130 RSC.
  17. P. Kang, Z. Chen, M. Brookhart and T. J. Meyer, Top. Catal., 2014, 58, 30–45 CrossRef.
  18. R. J. Lim, M. Xie, M. A. Sk, J.-M. Lee, A. Fisher, X. Wang and K. H. Lim, Catal. Today, 2014, 233, 169–180 CrossRef CAS.
  19. J.-P. Jones, G. K. S. Prakash and G. A. Olah, Isr. J. Chem., 2014, 54, 1451–1466 CrossRef CAS.
  20. F. A. Viva, Adv. Chem. Lett., 2013, 1, 225–236 CrossRef CAS.
  21. R. P. S. Chaplin and A. A. Wragg, J. Appl. Electrochem., 2003, 33, 1107–1123 CrossRef CAS.
  22. S. Lee, H. Ju, R. Machunda, S. Uhm, J. K. Lee, H. J. Lee and J. Lee, J. Mater. Chem. A, 2015, 3, 3029–3034 CAS.
  23. R. Kortlever, C. Balemans, Y. Kwon and M. T. M. Koper, Catal. Today, 2015, 244, 58–62 CrossRef CAS.
  24. Y. Fu, Y. Liu, Y. Li, J. Qiao and X.-D. Zhou, ECS Trans., 2015, 66, 53–59 CrossRef.
  25. S. Y. Choi, S. K. Jeong, H. J. Kim, I.-H. Baek and K. T. Park, ACS Sustainable Chem. Eng., 2016, 4, 1311–1318 CrossRef CAS.
  26. R. Kortlever, I. Peters, S. Koper and M. T. M. Koper, ACS Catal., 2015, 5, 3916–3923 CrossRef CAS.
  27. N. Elgrishi, S. Griveau, M. B. Chambers, F. Bedioui and M. Fontecave, Chem. Commun., 2015, 51, 2995–2998 RSC.
  28. D. H. Won, J. Chung, S. H. Park, E.-H. Kim and S. I. Woo, J. Mater. Chem. A, 2015, 3, 1089–1095 CAS.
  29. S. Zhang, P. Kang, S. Ubnoske, M. K. Brennaman, N. Song, R. L. House, J. T. Glass and T. J. Meyer, J. Am. Chem. Soc., 2014, 136, 7845–7848 CrossRef CAS PubMed.
  30. D. Li, J. Huang and R. B. Kaner, Acc. Chem. Res., 2009, 42, 135–145 CrossRef CAS PubMed.
  31. H. Bejbouj, L. Vignau, J. L. Miane, T. Olinga, G. Wantz, A. Mouhsen, E. M. Oualim and M. Harmouchi, Mater. Sci. Eng., B, 2010, 166, 185–189 CrossRef CAS.
  32. Y. F. Huang, C. S. Chang and C. W. Lin, Int. J. Hydrogen Energy, 2012, 37, 11975–11983 CrossRef CAS.
  33. L. J. Sun and X. X. Liu, Eur. Polym. J., 2008, 44, 219–224 CrossRef CAS.
  34. S. Mondal, U. Rana and S. Malik, Chem. Commun., 2015, 51, 12365–12368 RSC.
  35. T. Lindfors and A. Ivaska, Anal. Chem., 2004, 76, 4387–4394 CrossRef CAS PubMed.
  36. Y. Xia and H. Zhu, Soft Matter, 2011, 7, 9388–9393 RSC.
  37. J. R. Santos Jr, L. H. C. Mattoso and A. J. Motheo, Electrochim. Acta, 1998, 43, 309–313 CrossRef.
  38. A. Thota, R. Arukula, R. Narayan, C. R. K. Rao and K. V. S. N. Raju, RSC Adv., 2015, 5, 106523–106535 RSC.
  39. Z. Qiang, G. Liang, A. Gu and L. Yuan, Mater. Lett., 2014, 115, 159–161 CrossRef CAS.
  40. X. Feng, C. Mao, G. Yang, W. Hou and J.-J. Zhu, Langmuir, 2006, 22, 4384–4389 CrossRef CAS PubMed.
  41. H. H. Zhou, S. Q. Jiao, J. H. Chen, W. Z. Wei and Y. F. Kuang, J. Appl. Electrochem., 2004, 34, 455–459 CrossRef CAS.
  42. R. M. Mohamed and E. S. Aazam, Appl. Catal., A, 2014, 480, 100–107 CrossRef CAS.
  43. M. H. Beyzavi, R. C. Klet, S. Tussupbayev, J. Borycz, N. A. Vermeulen, C. J. Cramer, J. F. Stoddart, J. T. Hupp and O. K. Farha, J. Am. Chem. Soc., 2014, 136, 15861–15864 CrossRef CAS PubMed.
  44. Q.-W. Song, W.-Q. Chen, R. Ma, A. Yu, Q.-Y. Li, Y. Chang and L.-N. He, ChemSusChem, 2015, 8, 821–827 CrossRef CAS PubMed.
  45. K. A. Weerakoon, J. H. Shu, M.-K. Park and B. A. Chin, ECS J. Solid State Sci. Technol., 2012, 1, Q100–Q105 CrossRef CAS.
  46. A. Nirmala Grace and K. Pandian, Electrochem. Commun., 2006, 8, 1340–1348 CrossRef.
  47. T. C. D. Doan, R. Ramaneti, J. Baggerman, J. F. van der Bent, A. T. M. Marcelis, H. D. Tong and C. J. M. van Rijn, Sens. Actuators, B, 2012, 168, 123–130 CrossRef CAS.
  48. F. Köleli and T. Röpke, Appl. Catal., B, 2006, 62, 306–310 CrossRef.
  49. D. B. Kayan and F. Köleli, Appl. Catal., B, 2016, 181, 88–93 CrossRef CAS.
  50. C. Costentin, M. Robert and J.-M. Saveant, Chem. Soc. Rev., 2013, 42, 2423–2436 RSC.
  51. T. N. Huan, E. S. Andreiadis, J. Heidkamp, P. Simon, E. Derat, S. Cobo, G. Royal, A. Bergmann, P. Strasser, H. Dau, V. Artero and M. Fontecave, J. Mater. Chem. A, 2015, 3, 3901–3907 CAS.
  52. W. Lv, J. Zhou, F. Kong, H. Fang and W. Wang, Int. J. Hydrogen Energy, 2016, 41, 1585–1591 CrossRef CAS.
  53. J. S. Yoo, R. Christensen, T. Vegge, J. K. Nørskov and F. Studt, ChemSusChem, 2016, 9, 358–363 CrossRef CAS PubMed.
  54. D. Chowdhury, J. Phys. Chem. C, 2011, 115, 13554–13559 CAS.
  55. X. L. Dong, X. F. Zhang, H. Huang and F. Zuo, Appl. Phys. Lett., 2008, 92, 013127 CrossRef.
  56. C. Zhao, Z. Yin and J. Wang, ChemElectroChem, 2015, 2, 1974–1982 CrossRef CAS.
  57. R. Aydın, H. Ö. Doğan and F. Köleli, Appl. Catal., B, 2013, 140–141, 478–482 CrossRef.
  58. P. Paulraj, N. Janaki, S. Sandhya and K. Pandian, Colloids Surf., A, 2011, 377, 28–34 CrossRef CAS.
  59. U. Bogdanović, I. Pašti, G. Ćirić-Marjanović, M. Mitrić, S. P. Ahrenkiel and V. Vodnik, ACS Appl. Mater. Interfaces, 2015, 7, 28393–28403 Search PubMed.
  60. W. Q. Tian, M. Ge, B. R. Sahu, D. Wang, T. Yamada and S. Mashiko, J. Phys. Chem. A, 2004, 108, 3806–3812 CrossRef CAS.
  61. S. N. Khanna, M. Beltran and P. Jena, Phys. Rev. B: Condens. Matter, 2001, 64, 235419–235422 CrossRef.
  62. S. J. Li, X. Zhou and W. Q. Tian, J. Phys. Chem. A, 2012, 116, 11745–11752 CrossRef CAS PubMed.
  63. S. Gautam, K. Dharamvir and N. Goel, Comput. Theor. Chem., 2013, 1009, 8–16 CrossRef CAS.
  64. S. Xu, Z. Wang, C. Wang, Z. Wang and Y. Cui, New J. Chem., 2015, 39, 3105–3108 RSC.
  65. C. Jia and W. Fan, Phys. Chem. Chem. Phys., 2015, 17, 30736–30743 RSC.
  66. E. D. Glendening, A. E. Reed, J. E. Carpenter and F. Weinhold, NBO Version 3.1 Search PubMed.
  67. J. K. Nørskov, J. Rossmeisl, A. Logadottir, L. Lindqvist, J. R. Kitchin, T. Bligaard and H. Jónsson, J. Phys. Chem. B, 2004, 108, 17886–17892 CrossRef.
  68. A. A. Peterson, F. Abild-Pedersen, F. Studt, J. Rossmeisl and J. K. Norskov, Energy Environ. Sci., 2010, 3, 1311–1315 CAS.
  69. C. Liu, H. He, P. Zapol and L. A. Curtiss, Phys. Chem. Chem. Phys., 2014, 16, 26584–26599 RSC.
  70. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, N. J. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian Revision C.01, 2009 Search PubMed.
  71. N. M. O'Boyle, A. L. Tenderholt and K. M. Langner, J. Comput. Chem., 2008, 29, 839–845 CrossRef PubMed.
  72. S. H. Ahn and H. Lee, Bull. Korean Chem. Soc., 2016, 37, 27–32 CrossRef CAS.
  73. S. Nayab, H.-I. Lee and J. H. Jeong, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2013, 69, m238–m239 CAS.
  74. S.-Y. Lai, T.-W. Lin, Y.-H. Chen, C.-C. Wang, G.-H. Lee, M.-h. Yang, M.-k. Leung and S.-M. Peng, J. Am. Chem. Soc., 1999, 121, 250–251 CrossRef CAS.
  75. Y. Kim and S. K. Kang, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2015, 71, 1058–1060 CAS.
  76. X. P. Chen, J. K. Jiang, Q. H. Liang, N. Yang, H. Y. Ye, M. Cai, L. Shen, D. G. Yang and T. L. Ren, Sci. Rep., 2015, 5, 16907 CrossRef CAS PubMed.
  77. H. Ullah, A.-U.-H. A. Shah, S. Bilal and K. Ayub, J. Phys. Chem. C, 2013, 117, 23701–23711 CAS.
  78. H. J. Freund and M. W. Roberts, Surf. Sci. Rep., 1996, 25, 225–273 CrossRef.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra20715d

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.