Catalytic oxidation of NO by Au2 dimers: a DFT study

Ramesh Ch. Deka*, Debajyoti Bhattacharjee, Arup Kumar Chakrabartty and Bhupesh Kumar Mishra
Department of Chemical Sciences, Tezpur University, Napaam, Tezpur, Assam 784 028, India. E-mail: ramesh@tezu.ernet.in

Received 7th May 2013 , Accepted 23rd October 2013

First published on 28th October 2013


Abstract

Density functional theory (DFT) calculations are performed to study the mechanistic details of NO oxidation promoted by gold dimer anions. Furthermore, we studied a full catalytic cycle producing two NO2 molecules. The reaction is explored along three possible pathways. Our theoretical results show that anionic gold dimers present catalytic activity towards NO oxidation, as indicated by the calculated low energy barriers and high exothermicities. The present results enrich our understanding of the catalytic oxidation of NO by Au-cluster based catalysts. For the first time, we have presented a systematic study on the structure and energetics of various reaction intermediates involved in NO oxidation by Au2 clusters using density functional theory (DFT).


Introduction

The removal of poisonous NO gas, emitted mainly in automobile exhaust, is a major challenge in catalysis research and has been the subject of many investigations. Thus, understanding the adsorption mechanisms, dissociation of oxygen, and oxidation of NO at the molecular level is of great importance in designing suitable catalysts. As a new kind of catalyst to reduce air pollution, nano-sized gold clusters have attracted considerable interest recently from both the industrial and scientific communities.1–5 Experimental and theoretical studies have proved that free and supported nano-sized gold clusters can efficiently catalyze CO, NO and H2 oxidation reactions at low temperatures.6,7 Nitrogen oxides exhibit rich chemistry on gold surfaces, particularly on oxygen covered gold surfaces.8,9 Endou et al.10 have studied the adsorption and activation properties of precious metals (Ir, Pt and Au) towards NO by using the cluster model. McClure et al.11 studied the adsorption and reaction of NO with atomic oxygen covered Au(111) surfaces, using collision induced desorption and temperature programmed desorption spectra. They suggested that nitric oxide may react with chemisorbed oxygen atoms at low temperatures. Torres et al.12 theoretically studied the adsorption and oxidation of NO molecules with O on Au(111) surfaces at the low-coverage limit. They suggested that adsorbed NO reacts with pre-adsorbed oxygen by the Langmuir–Hinshelwood (LH) mechanism, in which the reaction occurs between chemisorbed oxygen atoms and NO molecules. The main route to nitrogen dioxide (NO2) formation in combustion systems is through the oxidation of nitric oxide (NO). This process was originally investigated in order to explain the high proportion of NO2 found in NOx emissions from the exhaust of some gas turbine engines.11 Ding et al.13 have used the density functional theory to investigate NO molecule adsorption on the gold clusters, observing a strong ability to adsorb NO molecules for the majority of gold clusters. Hao et al.14 studied the catalytic oxidation of NO over TiO2 supported Pt clusters by in situ FTIR spectroscopy and proposed that the direct oxidation of gaseous NO to NO2 is the main pathway for NO2 formation. If one considers that Au2 promotes CO oxidation in the gas phase,15,16 it is both natural and promising to investigate the catalytic activity of Au2 towards NO oxidation. Xu et al.17 reported density functional theory calculations (GGA-PW91) that probed the relationship between particle size, intermediate structures, and the energetics of CO and NO oxidation by molecular and atomic oxygen on Ptx clusters (x = 1–5 and 10). Very recently, Xie et al.18 have investigated the reaction mechanism of NO reduction on the Rh7+ cluster by means of DFT calculations and show that the first NO molecule which dissociates on the Rh7+ cluster has an activation barrier of 39.6 kcal mol−1, which is the rate-determining step of the whole catalytic cycle. In recent years, zeolites have been one of the most studied catalysts for NOx decomposition. Kikuchi et al.19 showed that the NO reduction with methane on In/H-ZSM-5 could be improved by adding a transition metal such as Pt, Ir or Rh for enhancing the oxidation of NO to NO2 at the early stage of the reaction, which is a necessary step for NO reduction by methane. Recently, Morpurgo et al.20 investigated the mechanism of NO decomposition catalyzed by Cu-ZSM-5 by means of computational methods. Metkar et al.21 performed a comprehensive experimental and kinetic modeling study on NO oxidation reactions using Fe-ZSM-5, Cu-ZSM-5 and Cu-chabazite catalysts. Izquierdo et al.22 analyzed NOx decomposition reactions over a Cu-ZSM-5 catalyst by ONIOM-DFT methods. Silver and proton-exchanged zeolites are also promising catalysts for the reduction of NO2 with hydrocarbons.23 It has been stated by Zeng et al.24 that carbonaceous material supported metal oxides can remove NO effectively at room temperature. Since NO is easily oxidized to NO2 at room temperature, and since NO2 is easily removed by water, the catalytic oxidation of NO to NO2 at room temperature is a promising method for the removal of NO. Therefore we restrict our attention to studying the mechanism of the title reaction at room temperature only. Recently, Gao et al.25 systematically studied CO oxidation on subnanometer anionic gold clusters and their neutral counterparts by using ab initio calculations. Their study reveals that compared to the neutral clusters, the extra negative charge on the anionic clusters can enhance not only O2 adsorption but also the strength of the co-adsorption of CO and O2, which leads to a lower reaction barrier in the oxidation step. Thus, it may be concluded that in general, the anionic clusters are more efficient catalysts than their neutral counterparts. The catalytic reduction of NO by CO has also been studied extensively since both are present in exhaust gases.26 Recently, supported Ir particles have shown unusually high reactivity for the reduction of NO by CO in the presence of oxygen.27 NO oxidation has also been catalyzed by Rh and Co clusters.28

To the best of our knowledge, no detailed theoretical or experimental studies have been reported for this reaction. In view of the potential importance and the rather limited information available, we carried out a detailed theoretical study on the potential energy surfaces of the title reaction to (1) explore elaborate oxidation processes; (2) investigate the products of the title reaction to assist in further experimental identification; and (3) give deep insights into the mechanism of the reaction of gold anionic dimers with nitric oxide. Therefore, to get a more complete picture of the oxidation processes of NO on the anionic Au dimers, it is pertinent to perform a comprehensive theoretical study to determine the energetics involved in the NO oxidation by considering a full structure optimization. The three most plausible pathways for the oxidation of NO are investigated during the present study and are given by pathways (1–3), as shown in Scheme 1.

Au2 + O2 + NO → Au2–O + NO2 (pathway 1)

Au2 + 1/2O2 + NO → Au2 + NO2 (pathway 2)

Au2 + O2 + 2NO → Au2 + 2NO2 (pathway 3)


image file: c3ra42240b-s1.tif
Scheme 1 Reaction pathways for the oxidation of NO to NO2 by Au2 dimers.

In recent years, density functional theory (DFT) has become a valuable tool for studying the properties of molecules and materials29,30 and for identifying reaction mechanisms.31,32 In this communication, for the first time, we have presented a systematic study on the structure and energetics of various reaction intermediates involved in NO oxidation by Au2 dimers using density functional theory (DFT).

Computational methods

All calculations have been performed using the GAUSSIAN 09 suite of programs.33 In the framework of density functional theory (DFT), we employ the hybrid B3LYP34,35 functional to explore the stationary points on the potential energy surfaces. Considering the strong relativistic effects of Au, we used the Los Alamos LANL2DZ36,37 Effective Core Pseudopotentials (ECP) and valence double-ζ basis sets for Au atoms. The C, N and O atoms were treated with the 6-311G(d,p) basis sets. No symmetric constraints were imposed during geometry optimization. In order to determine the nature of different stationary points on the potential energy surface, vibrational frequency calculations were also performed using the same level of theory at which the optimization was made. All the stationary points have been characterized as either minima (the number of imaginary frequencies (NIMAG) = 0) or transition structures (NIMAG = 1). To ensure the reliability of the reaction path, the connections between the transition state and the corresponding minima were verified using intrinsic reaction coordinate (IRC) calculations developed by Gonzalez and Schlegel in the mass-weighted internal coordinate system.38 Previous investigations39–41 showed that the B3LYP/LANL2DZ combinations are sufficiently accurate for describing noble-metal systems. To further clarify the reliability of our calculations, we calculated the geometries and binding energies for Au2–O2, Au2–CO, and Au2–NO. Our calculated values for the binding energies of O2, CO and NO over Au2 are 19.29, 7.16 and 16.50 kcal mol−1, respectively. The binding energy of CO is in good agreement with the reported value of Liu et al.42 It is clear from Fig. 1 that the HOMO and LUMO isosurfaces of Au2, O2, CO and NO have matched shapes and symmetries. This further explains the binding orientations of NO with Au2, which are similar to O2 and CO.
image file: c3ra42240b-f1.tif
Fig. 1 The HOMO and LUMO isosurfaces for Au2 dimers and O2, CO and NO molecules (isosurface value = 0.02).

The thermochemical parameters like standard reaction enthalpy (ΔHr) and Gibbs free energy (ΔGr) at a temperature T were estimated from the differences of H and G values of products and reactants at that temperature:

image file: c3ra42240b-t1.tif

image file: c3ra42240b-t2.tif
where E0 is the total electronic energy including ZPE and Hcorr and Gcorr are the factors to be added to E0 to obtain the enthalpy and Gibbs free energy, respectively, at a temperature T for taking into account the contribution of the translational, rotational and vibrational motion of a molecule. The Hcorr and Gcorr are defined as:
Hcorr = Htrans + Hrot + Hvib + RT

Gcorr = HcorrTStotal where entropy; Stotal = Strans + Srot + Svib + Sel

The thermal zero point energy correction (TZPE) is the thermal correction to the internal energy at 298 K and is given as a sum over four components for contributions from electronic, vibrational, rotational and translational degrees of freedom. The zero point vibrational energy ZPVE (or ZPE) results from the vibrational motion of molecular systems even at 0 K and is calculated for a harmonic oscillator model as a sum of the contributions from all vibrational modes of the system. The heat of adsorption (−ΔHads) is defined as:

−ΔHads = (Ecluster + Hadsorbate) − Ecluster/adsorbate,
where Ecluster/adsorbate, is the total energy of the adsorbate on the cluster, Ecluster is the total energy of the bare cluster and Hadsorbate is the enthalpy of the adsorbate.

Results and discussion

The detailed thermodynamic calculations performed at the DFT level of theory for the free energies, reaction enthalpies and entropies with thermal zero-point energy corrections (TZPE) associated with pathways (1–3) are listed in Table 1. The optimized geometries of the reactants, intermediates, transition states and products obtained at the DFT level are shown in Fig. 2. Transition states found through a search on the potential energy surfaces of pathways (1–3) are characterized by TS1, TS2 and TS3, respectively. The search was made along the minimum energy path on a relaxed potential energy surface.
Table 1 Thermochemical data for the oxidation pathways of NO calculated at the DFT level of theory with TZPE corrections
Reaction channels ΔGr (298 K) (kcal mol−1) ΔHr (298 K) (kcal mol−1) ΔS° (298 K) (cal mol−1 K−1)
Pathway 1 18.08 9.90 −33.29
Pathway 2 −24.16 −22.79 −16.20
Pathway 3 −48.32 −45.59 −32.40



image file: c3ra42240b-f2.tif
Fig. 2 Optimized geometries of reactants, intermediates, transition states and products involved in the oxidative pathways of NO at B3LYP level.

Harmonic vibrational frequencies of the stationary points were calculated at the B3LYP level of theory and are given in Table 2. These results show that the reactant and products have stable minima on their potential energy surface characterized by the occurrence of only real and positive vibrational frequencies. On the other hand, transition states are characterized by the occurrence of only one imaginary frequency obtained at 390i, 246i and 534i cm−1 for TS1, TS2 and TS3, respectively. These frequencies were analyzed using the ChemCraft43 visualization program. Visualization of the imaginary frequency gives a qualitative confirmation of the existence of transition states connecting reactants and products. Intrinsic reaction coordinate (IRC) calculations38 have also been performed for each transition state. The IRC plots for the transition states reveal that the transition state structure smoothly connects the reactant and the product sides. The energies of the reactants, transition states and products obtained in the IRC calculations are in excellent agreement with the individually optimized values at the B3LYP/LANL2DZ level of theory. The formation of complexes of Au2 with O2 and NO molecules is expected to be the initial step of NO oxidation. Pathway 1 starts from the adsorption of O2 on Au2 to form the superoxo-like complex Au2–O2. After the adsorption of O2, the incoming NO attacks the adsorbed O2 to form a three species meta-stable complex IM1 which lies below the reactants by 39.31 kcal mol−1. Then, the NO adsorbs the O atom to produce a NO2 molecule via the transition state TS1. The calculated geometrical parameters of TS1 clearly indicate that the O–O bond is weakening and the N–O bond is forming. Visualization of the optimized structure of TS1 further reveals the elongation of the O–O bond (O1–O2) bond length from 1.257 to 1.436 Å (14%) with a simultaneous shrinkage of the N–O (N1–O1) bond from 1.378 to 1.192 Å (13%). In the optimized geometry of TS1, NO abstracts an O atom to produce NO2 and complete the cycle. Thermodynamic calculations reveal that this pathway proceeds with an endothermicity of 9.90 kcal mol−1 along with an energy barrier of 16.18 kcal mol−1. The calculated heat of adsorption for this path is found to be 48.33 kcal mol−1.

Table 2 Harmonic vibrational frequencies of reactants, intermediates, transition states and products
Species Vibrational frequencies (cm−1)
Au2 114
O2 1641
Au2–O 119, 295, 296, 311
Au2–O2 45, 60, 136, 216, 408, 1176
IM1 28, 56, 61, 137, 158, 228, 367, 465, 613, 808, 1006, 1647
IM2 12, 49, 137, 141, 185, 299, 820, 1067, 1557
IM3 42, 45, 73, 123, 132, 171, 172, 260, 379, 509, 1280, 1758
IM4 18, 30, 40, 47, 55, 68, 119, 139, 208, 231, 248, 437, 740, 825, 1329, 1499, 1647
TS1 390i, 27, 60, 88, 138, 191, 245, 435, 502, 728, 846, 1753
TS2 246i, 46, 57, 137, 141, 219, 785, 1223, 1505
TS3 534i, 27, 45, 53, 63, 138, 171, 176, 217, 242, 329, 397, 592, 638, 720, 953, 1453, 1699
NO 1988
NO2 766, 1399, 1706


In the case of pathway 2, the initial complex of Au2 with O and NO is denoted as IM2. The results given in Table 1 reveal that pathway 2 is exergonic (ΔG < 0) and exothermic in nature and thus, thermodynamically facile. The heat of adsorption for this pathway is found to be 128.97 kcal mol−1. Our calculation shows that this complex is more stable (by 58.59 kcal mol−1) than the reactants. The transition state involved along this path is TS2, where the transition vector is related to the imaginary frequency of 246 cm−1 giving rise to the formation of the product-like intermediate IM2. The barrier along this path is calculated to be 11.91 kcal mol−1. The calculated geometrical parameters of TS2 indicate that the Au–O bond is weakening and the O–N bond is forming. A similar analysis made on the optimized structure of TS2 reveals the elongation of the Au–O (Au2–O1) bond length from 2.164 to 2.462 Å, an increase of about 14%.

Considering the strong interaction of Au clusters with NO (with a binding energy of 16.50 kcal mol−1), we explored another reaction channel (pathway 3) which can be described as the O–N–O–O–N–O group involved pathway with the participation of two NO molecules. Since the ΔGr value for this pathway is significantly negative at 298 K, it should be a spontaneous process at lower temperatures. Interestingly, ΔHr and ΔGr values indicate that this should be the more thermodynamically favored pathway. Pathway 3 involves the simultaneous formation of two NO2 molecules with a heat of adsorption value of 164.24 kcal mol−1. Similar to pathway 1, by attaching the first NO molecule to IM1 we locate another meta-stable complex IM3 which lies below the reactants by 53.76 kcal mol−1. By subsequent attachment of a second NO molecule to IM3, we locate IM4. Visualization of IM4 reveals that the N atom binds to Au2 and associates with one O of O2 to form an O–N–O–O–N–O group which lies below the reactants by 43.58 kcal mol−1, with the simultaneous formation of two NO2 molecules via TS3 with a barrier of 8.92 kcal mol−1. A schematic of the potential energy surface of the titled reaction obtained at the B3LYP + ZPE level of theory is plotted in Fig. 3. In the construction of the energy diagram, zero-point corrected total energies of the species are utilized. These energies are plotted with respect to the ground state energy of Au2 + O2/O arbitrarily taken as zero. From Fig. 3, it is clear that path 3 (the O–N–O–O–N–O group) is the dominant path, having a lower barrier height for NO oxidation promoted by the Au2 dimer. Our calculated barrier heights for all the plausible pathways are comparable with the barrier height calculated by Liu et al.42 for CO oxidation promoted by the Au2 dimer. This result strongly suggests that like with CO oxidation, the Au2 dimer shows catalytic activity towards NO oxidation at low temperatures.


image file: c3ra42240b-f3.tif
Fig. 3 Potential energy profile of NO oxidation by the Au2 dimer. The relative energies (in kcal mol−1) were calculated with ZPE corrections at the B3LYP/LANL2DZ level of theory.

Conclusion

We present here the potential energy profile (including geometries, energies and vibrational frequencies of the reactants, intermediates, transition states and products) and thermochemical data for the oxidation of NO to NO2 promoted by gold anionic dimers investigated at the DFT level of theory. Energetic calculations reveal that the most dominant oxidation pathway is path 3 (Au2 + O2 + 2NO) that occurs with a barrier height of 8.92 kcal mol−1. The present theoretical study may give useful information on the reaction mechanism and can provide powerful guidelines for future experimental studies for the title reaction.

Acknowledgements

The authors are thankful to DST, New Delhi for financial support. BKM is thankful to University Grants Commission, New Delhi for providing a Dr D. S. K. Post doctoral fellowship. The financial support in the form of a Junior Research Fellowship to DB from CSIR, New Delhi is also acknowledged.

References

  1. M. Okumura, S. Nakamura, S. Tsubota, T. Nakamura, M. Azuma and M. Haruta, Catal. Lett., 1998, 51, 53 CrossRef CAS.
  2. G. C. Bond and D. Thompson, Catal. Rev. Sci. Eng., 1999, 41, 319 CAS.
  3. M. Valden, X. Lai and D. W. Goodman, Science, 1998, 281, 1647 CrossRef CAS.
  4. M. Haruta, Chem. Rec., 2003, 3, 75 CrossRef CAS PubMed.
  5. M. C. Daniel and D. Astruct, Chem. Rev., 2004, 104, 293 CrossRef CAS PubMed.
  6. N. Lopez and J. K. Norskov, J. Am. Chem. Soc., 2002, 124, 11262 CrossRef CAS PubMed.
  7. H. Hakkinen and U. Landman, J. Am. Chem. Soc., 2001, 123, 9704 CrossRef CAS.
  8. N. Saliba, D. H. Parker and B. E. Koel, Surf. Sci., 1998, 410, 270 CrossRef CAS.
  9. J. Wang and B. E. Koel, J. Phys. Chem. A, 1998, 102, 8573 CrossRef CAS.
  10. A. Endou, N. Ohashi, S. Takami, M. Kubo, A. Miyamoto and E. Broclawik, Top. Catal., 2000, 11/12, 271 CrossRef CAS.
  11. S. M. McClure, T. S. Kim, J. D. Stiehl, P. L. Tanaka and C. B. Mullins, J. Phys. Chem. B, 2004, 108, 17952 CrossRef CAS.
  12. D. Torres, S. Gonzalez, K. M. Neyman and F. Illas, Chem. Phys. Lett., 2006, 422, 412 CrossRef CAS PubMed.
  13. X. L. Ding, Z. Y. Li, J. L. Yang, J. G. Hou and Q. S. Zhu, J. Chem. Phys., 2004, 121, 2558 CrossRef CAS PubMed.
  14. L. Li, Q. Shen, J. Cheng and Z. Hao, Catal. Today, 2010, 158, 361 CrossRef CAS PubMed.
  15. L. D. Socaciu, J. Hagen, T. M. Bernhardt, L. Woste, U. Heiz, H. Hakkinen and U. Landman, J. Am. Chem. Soc., 2003, 125, 10437 CrossRef CAS PubMed.
  16. M. L. Kimble, A. W. Castleman, R. Mitric Jr, C. Burgel and V. B. Koutecky, J. Am. Chem. Soc., 2004, 126, 2526 CrossRef CAS PubMed.
  17. Y. Xu, R. B. Getman, W. A. Shelton and W. F. Schneider, Phys. Chem. Chem. Phys., 2008, 10, 6009 RSC.
  18. H. Xie, M. Ren, Q. Lei and W. Fang, J. Phys. Chem. A, 2011, 115, 14203 CrossRef CAS PubMed.
  19. E. Kikuchi, M. Ogura, N. Aratani, Y. Sugiura, S. Hiromoto and K. Yogo, Catal. Today, 1996, 27, 35 CrossRef CAS.
  20. S. Morpurgo, G. Moretti and M. Bossa, J. Mol. Catal. A: Chem., 2012, 358, 134 CrossRef CAS PubMed.
  21. P. S. Metkar, V. Balakotaiah and M. P. Harold, Catal. Today, 2012, 184, 115 CrossRef CAS PubMed.
  22. R. Izquierdo, L. J. Rodriguez, R. Anez and A. Sierraalta, J. Mol. Catal. A: Chem., 2011, 348, 55 CrossRef CAS PubMed.
  23. J. A. Martens, A. Cauvel, A. Francis, C. Hermans, F. Jayat, M. Remy, M. Keung, J. Lievens and P. A. Jacobs, Angew. Chem., Int. Ed., 1998, 37, 1901 CrossRef CAS.
  24. Z. Zeng, P. Lu, C. Li, L. Mai, Z. Li and Y. Zhang, Catal. Sci. Technol., 2012, 2, 2188 CAS.
  25. Y. Gao, N. Shao, Y. Pei, Z. F. Chen and X. C. Zeng, ACS Nano, 2011, 5, 7818 CrossRef CAS PubMed.
  26. V. I. Parvulescu, P. Grange and B. Delmon, Catal. Today, 1998, 46, 233 CrossRef CAS.
  27. W. Chen, Q. Shen, R. A. Bartynski, P. Kaghazchi and T. Jacob, Langmuir, 2013, 29, 1113 CrossRef CAS PubMed.
  28. M. M. Yung, E. M. Holmgreen and U. S. Ozkan, J. Catal., 2007, 247, 356 CrossRef CAS PubMed.
  29. V. Stamenkovic, B. S. Mun, K. J. J. Mayrhofer, P. N. Ross, N. M. Markovic, J. Rossmeisl, J. Greeley and J. K. Norskov, Angew. Chem., Int. Ed., 2006, 45, 2897 CrossRef CAS PubMed.
  30. B. Kalita and R. C. Deka, J. Am. Chem. Soc., 2009, 131, 13252 CrossRef CAS PubMed.
  31. R. Valero, J. R. B. Gomes, D. G. Truhlar and F. Illas, J. Chem. Phys., 2010, 132, 104701 CrossRef PubMed.
  32. W. Song and E. J. M. Hensen, Catal. Sci. Technol., 2013, 3, 3020 CAS.
  33. M. J. Frisch, et al., Gaussian 09, Revision B. 01, Gaussian, Pittsburgh, PA, 2009 Search PubMed.
  34. A. D. Becke, J. Chem. Phys., 1992, 96, 2155 CrossRef CAS PubMed.
  35. C. Lee, W. Yang and R. G. Parr, Phys. Rev. B, 1988, 37, 785 CrossRef CAS.
  36. P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 270 CrossRef CAS PubMed.
  37. P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 299 CrossRef CAS PubMed.
  38. C. Gonzalez and H. B. Schlegel, J. Chem. Phys., 1989, 90, 2154 CrossRef CAS PubMed.
  39. A. M. Joshi, M. H. Tucker, W. N. Delgass and K. T. Thomson, J. Chem. Phys., 2006, 125, 194707 CrossRef PubMed.
  40. A. M. Joshi, W. N. Delgass and K. T. Thomson, J. Phys. Chem. B, 2006, 110, 23373 CrossRef CAS PubMed.
  41. D. Y. Wu, B. Ren, Y. X. Jiang, X. Xu and Z. Q. Tian, J. Phys. Chem. A, 2002, 106, 9042 CrossRef CAS.
  42. P. Liu, K. Song, D. Zhang and C. Liu, J. Mol. Model., 2012, 18, 1809 CrossRef CAS PubMed.
  43. G. Zhurko and D. Zhurko, ChemCraft 1.6, 2011 Search PubMed.

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.