You are on page 1of 8

Van der Waals interactions in complex materials

Van der Waals (vdW) interactions are ubiquitous in nature, playing a major role in defining the structure, stability, and function for a wide variety of molecules and materials. An accurate first-principles description of vdW interactions is extremely challenging, since the vdW dispersion energy arises from the correlated motion of electrons and, in principle, must be described by many-electron quantum mechanics. Rapid increase in computer power and advances in theoretical models for vdW interactions have allowed to achieve "chemical accuracy" for binding between small organic molecules. However, the lack of accurate and efficient methods for large and complex systems hinders truly quantitative predictions of properties and functions of technologically relevant materials. Typical applications where vdW interactions play an essential role include the design of novel hybrid inorganic/organic interfaces for photovoltaics, energy storage, and sensor devices [1], understanding the structure and dynamics of drug binding to proteins, as well as the creation of "smart" nanomechanical devices with tunable electronic properties [2].

Within this context our proposed CECAM Workshop "Towards First-Principles Description of van der Waals Interactions in Complex Materials" wants to be an opportunity to (i) bring the major players up to date in the most recent developments in the field, (ii) identify a set of (complex) systems -- beyond small molecules -- that can be used to benchmark newly developed methods, (iii) discuss the implementation and validation of different vdW methods in electronic structure codes, and (iv) identify future major challenges in the area of modeling vdW interactions. This workshop finds its motivation in the realization that, while the vdW interactions areconstantly found to be significantly more important than anticipated for a broad range of materials, there are still few initiatives on bringing together both the developers and users of vdW methods at the same place and time. We therefore propose a focused workshop on vdW interactions in complex materials, with a strong emphasis on the collaboration between developers and users. This workshop is extremely timely as the different methods for vdW interactions have been or are in the process of being implemented in most major electronic structure software projects. Density-functional theory (DFT) is currently the method of choice for the modeling of complex materials. During the last decade there has been a surge of interest in developing new methods for the modeling of vdW interactions in DFT [3, 4, 5, 6]. Loosely speaking, three successful (and somewhat different) ways can be identified: (i) Non-local density functionals, capturing the pairwise (two-body) part of the vdW energy [7, 8, 9, 10]; (ii) Interatomic (pairwise) vdW potentials, added to the DFT total energy in a post-processing fashion [11, 12, 13, 14, 15]; (iii) Computationally expensive DFT functionals on the fifth rung of Perdew's Jacob's ladder, which explicitly include Coulomb screening and many-body vdW energy [16, 17, 18, 19, 20]. Despite significant progress in the field of modeling vdW interactions during the last decade, many questions remain unanswered and much development needs to be done before a truly universally applicable (accurate and efficient) method emerges. For example, the employed approximations and the connections between different ways of modeling vdW interactions [(i), (ii), (iii) above] are not always transparent. Furthermore, the domain of applicability of every method is not clearly defined. For example, interatomic vdW potentials are frequently employed for the modeling of hybrid inorganic/organic interfaces [21, 22, 23, 24], neglecting the rather strong Coulomb screening present within inorganic bulk materials. On the other hand, the popular non-local vdW-DF functionals [7, 8] use a purely local approximation for the polarizability, which is not expected to be accurate for molecules. Nevertheless, the interaction energies between organic molecules turn out to be reasonably accurate. Understanding the physical reasons of why these different approaches "work" outside of their expected domain of applicability is important for the development of more robust approximations.

References
[1] L. Kronik and N. Koch, MRS Bull. 35, 417 (2010). [2] N. Marom, J. Bernstein, J. Garel, A. Tkatchenko, E. Joselevich, L. Kronik, and O. Hod, Phys. Rev. Lett. 105, 046801 (2010). [3] K. E. Riley, M. Pitonak, P. Jurecka, and P. Hobza, Chem. Rev. 110, 5023 (2010). [4] A. Tkatchenko, L. Romaner, O. Ho man, E. Zojer, C. Ambrosch-Draxl, and M. Scheffler, MRS Bull. 35, 435 (2010). [5] S. Grimme, Comput. Mol. Sci. 1, 211 (2011). [6] D. C. Langreth, B. I. Lundqvist, S. D. Chakarova-Kack, V. R. Cooper, M. Dion, P. Hyldgaard, A. Kelkkanen, J. Kleis, L. Kong, S. Li, P. G. Moses, E. Murray, A. Puzder, H. Rydberg, E. Schroder, T. Thonhauser, J. Phys.: Condens. Matter 21, 084203 (2009). [7] M. Dion, H. Rydberg, E. Schroder, D. C. Langreth, and B. I. Lundqvist, Phys. Rev. Lett. 92, 246401 (2004). [8] K. Lee, E. D. Murray, L. Kong, B. I. Lundqvist, and D. C. Langreth, Phys. Rev. B 82, 081101 (2010). [9] O. A. Vydrov and T. van Voorhis, Phys. Rev. Lett. 103, 063004 (2009). [10] J. Klimes, D. R. Bowler, and A. Michaelides, J. Phys.: Condensed Matter 22, 022201

Document: Reading materials for Introduction - Chapter01.doc Author: Kester Wen Jie Wong Save Date: 03/07/2012 Page 1 of 8

(2010). [11] R. Ahlrichs, R. Penco, and G. Scoles, Chem. Phys. 19, 119 (1977). [12] S. Grimme, J. Antony, S. Ehrlich, and H. Krieg, J. Chem. Phys. 132, 154104 (2010). [13] P. L. Silvestrelli, Phys. Rev. Lett. 100, 053002 (2008). [14] A. Tkatchenko and M. Scheffler, Phys. Rev. Lett. 102, 073005 (2009). [15] F. O. Kannemann and A. D. Becke, J. Chem. Theory and Comput., 6, 1081 (2010). [16] J. Harl and G. Kresse, Phys. Rev. Lett. 103, 056401 (2009). [17] D. Lu, Y. Li, D. Rocca, and G. Galli, Phys. Rev. Lett. 102, 206411 (2009). [18] B. G. Janesko, T. M. Henderson, and G. E. Scuseria , J. Chem. Phys. 130, 081105 (2009). [19] S. Lebegue, J. Harl, T. Gould, J. Angyan, G. Kresse, and J. F. Dobson, Phys. Rev. Lett. 105, 196401 (2010). [20] X. Ren, A. Tkatchenko, P. Rinke, and M. Scheffler, Phys. Rev. Lett. 106, 153003 (2011). [21] G. Mercurio, E. R. McNellis, I. Martin, S. Hagen, F. Leyssner, S. Soubatch, J. Meyer, M. Wolf, P. Tegeder, F. S. Tautz, and K. Reuter, Phys. Rev. Lett. 104, 036102 (2010). [22] N. Atodiresei, V. Caciuc, P. Lazic, and S. Bluegel, Phys. Rev. Lett. 102, 136809 (2009). [23] K. Tonigold and A. Gross, J. Chem. Phys. 132, 224701 (2010). [24] D. Stradi, S. Barja, C. Diaz, M. Garnica, B. Borca, J. J. Hinarejos, D. Sanchez-Portal, M. Alcami, A. Arnau, A. L. Vazquez de Parga, R. Miranda, and F. Martin, Phys. Rev. Lett. 106, 186102 (2011).

Spin states in inorganic chemistry


The electrons surrounding the nuclei of all matter around and within us can be in two different states, denoted the electron spin. This effect, although being purely quantum-chemical, has profound implications for real-world, large-scale systems like, for example, living tissue. In most cases, electrons of different spin pair up, effectively canceling the spineffect. In some cases, however, electrons prefer not to pair up, which leads to an excess of one type of electrons in a system. Depending on the exact conditions and surroundings, the number of these unpaired electrons can vary, leading to different spin-states not only for the individual electrons, but for much larger molecular species. In many cases, the spin-state has been found to be a key factor governing the behavior of the system. Elucidating the role and effect of different spin states on the properties of a system, even deciding which spin-state occurs naturally, is presently one of the most challenging endeavors both from an experimental and theoretical point-of-view. Reactivity patterns in organometallic and bioinorganic chemistry often depend critically on the spin state, e.g., on the spin-state preferences of reactants, products, intermediates and transition states.[1-2] For heme-proteins, this has been acknowledged for decades,[3] and recently, spin-state controlled reactions have been proposed to be important also outside the realm of biomolecules, that is, in industrial catalysis.[4] As a specific, intriguing example of this, merely one of a myriad, we consider the catalytic cycle of cytochrome P450cam that catalyzes the hydroxylation of R-camphor to 5-exo camphorol.[5] In the catalytic cycle,[6] a low-spin doublet is observed for the resting state which goes to a high-spin sextet after substrate binding.[7-12] This spin-flip in the first step seems to be determined primarily by the presence or absence of water molecules in the active site, and is vital for the specificity of the reaction taking place. Another key factor is a tyrosine residue that serves as an anchor for the natural substrate (R-camphor) and prepares it maximally for the subsequent hydroxylation.[13-15] For instance, a previous study[13] showed that upon binding to Tyr96, the substrate is found to be distorted, especially at the atom (C5) where the reaction takes place. The replacement of the Tyr96 residue by phenylalanine indeed reduces the regio- and stereo-specificity dramatically.[16] Another intriguing aspect of the catalytic cycle of P450cam is related to the step after the substrate has entered the active site.[5] The cycle continues with an electron transfer from a reducing agent, in the natural system putidaredoxin,[17-18] to give a quintet state after the transfer. This electron transfer takes place only when substrate is present, that is, only for the high spin state. Moreover, it seems that only when the electron transfer has taken place, that dioxygen can enter the active site. In this process of dioxygen binding, there is a second spin-flip present. The paramagnetic dioxygen molecule (with a triplet ground-state) on its own already induces a change of the spin-state, which would have lead to a triplet. Instead, it goes to a singlet state. The subsequent steps to product formation probably follow a rebound mechanism involving compound I, but there are also studies that suggest other mechanisms.[19-23] In order to help understand these subtleties and explain the missing steps in the reaction mechanism that occur too fast for experiments to follow, theoretical chemistry can play an important role. However, theory is not without its own problems (see State of the art section). At present, the situation is such that it is advisable to always include the results from a number of different quantum-chemical methods, to be ensure the results are consistent. In practice, when large molecular systems are of interest, the density functional theory (DFT) approach is the only viable alternative. There are, however, many different functionals that are proposed by different research groups (B3LYP*, TPSSh, OPBE/OLYP, SSB-D, M06-L, B3LYP, B2PLYP, ), without a clear consensus on which should be used (and why). The motivation for proposing this workshop is to be able to get together important players in the field and discuss this unfortunate situation thoroughly. The aim is to reach consensus on the current situation at hand, and more importantly, discuss best practices on how to deal with the problem in future scientific research. The focus of the workshop will not be limited only to theory; many of the assumptions about spin-state preferences are based on experimental data, which may not be not conclusive in each and every case. For instance, different
Document: Reading materials for Introduction - Chapter01.doc Author: Kester Wen Jie Wong Save Date: 03/07/2012 Page 2 of 8

complications may occur in the experiments, such as dimerization, ligand exchange, or disproportionation, which make that the system studied theoretically does not correspond to the one for which experimental data are observed. Likewise, there may be metal impurities present in the sample that might interfere with the observation of magnetic moments. Recent examples of failures of DFT in our research group in Girona turned out to be most likely due to inconsistencies in the experiments (e.g. disproportionation as described in ref. [24-25]). In order to resolve this, and get a clear discussion about these effects, we have also invited several experimentalists (Mayer, Que, Ghosh, Costas) to share their experiences and insights on this. The most popular method for studying (bio)inorganic catalysis is presented by Density Functional Theory (DFT),[26-28] due to its efficiency that enables to treat large systems of up to several hundreds, even thousands of atoms in a reasonable time. Also wave-function based methods can be used (for example, CASPT2), such as done by e.g. de Graaf or Pierloot, with the significant draw-back that only much smaller systems can be treated. Almost all DFT studies in the literature so far have either used the B3LYP[29-30] or BP86[31-32] functional, which for most simple cases both give good results. This is no longer true when spin-states with energy levels in close proximity are involved,[33-34] that is, when a transition metal like chrome, manganese, iron, cobalt or nickel is present. Previous studies[35-37] have shown that both of these functionals are unable to correctly predict the spin ground-state of these transition-metal complexes. Early GGA functionals like BP86 tend to overstabilize the low-spin state, while hybrid functionals like B3LYP tend to overstabilize the high-spin state (due to the inclusion of a portion of Hartree-Fock exchange).[34, 38] Several remedies have been proposed, such as lowering the amount of Hartree-Fock exchange in B3LYP to 15% (B3LYP*),[35] mixing the Becke88 and PW91x[39-40] exchange functionals (XLYP, X3LYP),[41] or introducing a Hubbard U parameter,[42] but neither one of these was really satisfactory for all situations. In fact, it was suggested to always calculate the spin-state energies with a number of functionals. An important step forward was made by the application of Handy and Cohens optimized exchange (OPTX) functional (which is abbreviated as O in combination with other functionals, as in OPBE or OLYP).[43] Previous validation studies have shown the validity of the OPBE[33, 44] (and OLYP)[45-51] functional for the spin-state splittings of iron complexes. While for the vertical spin-state splittings a number of DFT functionals could be trusted to give the correct spin ground-state,[33] this picture changed for relaxed splittings.[25, 34] The experience with the OPBE and PBE functionals for spin-states and (SN2) reaction barriers has led to the development of an improved GGA functional (SSB-D), which like OPBE works well for spin states. Other groups have advocated other functionals such as B3LYP* (Reiher, Siegbahn), OLYP (Ghosh), TPSSh (Jensen), or B2PLYP (Neese). Note also the EuroBic conference that is being held the week before the CECAM workshop in Granada: http://www.eurobic11.com/

References
[1] J. N. Harvey, R. Poli, K. M. Smith: "Understanding the reactivity of transition metal complexes involving multiple spin states", Coord. Chem. Rev., 238-239, 347-361 (2003). [2] J. N. Harvey: "DFT Computation of Relative Spin-State Energetics of Transition Metal Compounds", Structure and Bonding, 112, 151-183 (2004). [3] W. R. Scheidt, C. A. Reed: "Spin-state/stereochemical relationships in iron porphyrins: implications for the hemoproteins", Chem. Rev., 81, 543-555 (1981). [4] M. P. Shaver, L. E. N. Allan, H. S. Rzepa, V. C. Gibson: "Correlation of Metal Spin State with Catalytic Reactivity: Polymerizations Mediated by -DiimineIron Complexes", Angew. Chem. Int. Ed., 45, 1241-1244 (2006). [5] P. R. Ortiz de Montellano, Cytochrome P450 structure, mechanism and biochemistry, Plenum, New York, 1995. [6] C. A. Tyson, J. D. Lipscomb, I. C. Gunsalus: "The roles of putidaredoxin and P450cam in methylene hydroxylation", J. Biol. Chem., 247, 5777-5784 (1972). [7] G. H. Loew, D. L. Harris: "Role of the heme active site and protein environment in structure, spectra, and function of the cytochrome p450s", Chem. Rev., 100, 407-419 (2000). [8] D. Harris, G. Loew: "Determinants of the Spin-State of the Resting State of Cytochrome-P450cam", J. Am. Chem. Soc., 115, 8775-8779 (1993). [9] T. L. Poulos, B. C. Finzel, A. J. Howard: "Crystal structure of substrate-free Pseudomonas putida cytochrome P450", Bioch., 25, 5314-5322 (1986). [10] T. L. Poulos, B. C. Finzel, A. J. Howard: "High-resolution crystal structure of cytochrome P450cam", J. Mol. Biol., 195, 687-700 (1987). [11] M. T. Green: "Role of the axial ligand in determining the spin state of resting cytochrome P450", J. Am. Chem. Soc., 120, 10772-10773 (120). [12] J. H. Dawson, M. Sono: "Cytochrome P450 and chloroperoxidase: thiolate-ligated heme enzyme, spectroscopic determination of their active site structures and mechanistic implications of thiolate ligation", Chem. Rev., 87, 1255-1276 (1987). [13] M. Swart, A. R. Groenhof, A. W. Ehlers, K. Lammertsma: "Substrate binding in the active site of cytochrome P450cam", Chem. Phys. Lett., 403, 35-41 (2005). [14] I. Schlichting, C. Jung, H. Schulz: "Crystal structure of cytochrome P450cam complexed with the (1S)-camphor enantiomer", FEBS Lett., 415, 253-257 (1997). [15] R. Raag, T. L. Poulos: "Crystal structures of cytochrome P450cam complexed with camphane, thiocamphor and adamantane: factors controlling P450 substrate hydroxylation", Bioch., 30, 2674-2684 (1991). [16] C. Di Primo, G. Hui Bon Hoa, P. Douzou, S. G. Sligar: "Mutagenesis of a single hydrogen bond in cytochrome P450 alters cation binding and heme solvation", J. Biol. Chem., 265, 5361-5363 (1990).
Document: Reading materials for Introduction - Chapter01.doc Author: Kester Wen Jie Wong Save Date: 03/07/2012 Page 3 of 8

[17] M. J. Hintz, D. M. Mock, L. L. Peterson, K. Tuttle, J. A. Peterson: "Equilibrium and kinetic studies of the interaction of cytochrome P450cam and putidaredoxin", J. Biol. Chem., 257, 14324-14332 (1982). [18] H. Shimada, S. Nagano, Y. Ariga, M. Unno, T. Egawa, T. Hishiki, Y. Ishimura, F. Masuya, T. Obata, H. Hori: "Putidaredoxin-cytochrome P450cam interaction", J. Biol. Chem., 274, 9363-9369 (1999). [19] J. J. Warren, J. M. Mayer: "Proton-Coupled Electron Transfer Reactions at a Heme-Propionate in an IronProtoporphyrin-IX Model Compound", J. Am. Chem. Soc., 133, 8544-8551 (2011). [20] A. R. Groenhof, A. W. Ehlers, K. Lammertsma: "Proton Assisted Oxygen-Oxygen Bond Splitting in Cytochrome P450", J. Am. Chem. Soc., 129, 6204-6209 (2007). [21] V. Guallar, B. F. Gherman, S. J. Lippard, R. A. Friesner: "Quantum chemical studies of methane monooxygenase: comparision with P450", Curr. Opin. Chem. Biol., 6, 236-242 (2002). [22] V. Guallar, M.-H. Baik, S. J. Lippard, R. A. Friesner: "Peripheral heme substituents control the hydrogen-atom abstraction chemistry in cytochromes P450", Proc. Natl. Acad. Sci. USA, 100, 6998-7002 (2003). [23] R. D. Bach, O. Dmitrenko: "The "Somersault" mechanism for the P450 hydroxylation of hydrocarbons. The intervention of transient inverted metastable hydroperoxides", J. Am. Chem. Soc., 128, 1474-1488 (2006). [24] M. P. Johansson, M. Swart: "Subtle effects control the polymerisation mechanism in -diimine iron catalysts", Dalton Trans., 40, accepted (2011). [25] M. Swart, A. W. Ehlers, K. Lammertsma: "Relaxed spin states of iron complexes: the case for OPBE", (to be) submitted, (2008). [26] R. G. Parr, W. Yang, Density functional theory of atoms and molecules, Oxford University Press, New York, 1989. [27] W. Koch, M. C. Holthausen, A Chemist's Guide to Density Functional Theory, Wiley-VCH, Weinheim, 2000. [28] R. Dreizler, E. Gross, Density Functional Theory, Plenum Press, New York, 1995. [29] A. D. Becke: "Density-functional Thermochemistry. III. The role of exact exchange", J. Chem. Phys., 98, 5648-5652 (1993). [30] P. J. Stephens, F. J. Devlin, C. F. Chabalowski, M. J. Frisch: "Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields", J. Phys. Chem., 45, 11623-11627 (1994). [31] A. D. Becke: "Density-functional exchange-energy approximation with correct asymptotic behavior", Phys. Rev. A, 38, 3098-3100 (1988). [32] J. P. Perdew: "Density-functional approximation for the correlation-energy of the inhomogeneous electron-gas", Phys. Rev. B, 33, 8822-8824. Erratum: Ibid. 8834, 7406 (1986). [33] M. Swart, A. R. Groenhof, A. W. Ehlers, K. Lammertsma: "Validation of exchange-correlation functionals for spin states of iron-complexes", J. Phys. Chem. A, 108, 5479-5483 (2004). [34] M. Swart: "Accurate spin state energies for iron complexes", J. Chem. Theory Comp., 4, 2057-2066 (2008). [35] M. Reiher, O. Salomon, B. A. Hess: "Reparameterization of hybrid functionals based on energy differences of states of different multiplicity", Theor. Chem. Acc., 107, 48-55 (2001). [36] E. M. Sproviero, J. A. Gascon, J. P. McEvoy, G. W. Brudvig, V. S. Batista: "Characterization of synthetic oxomanganese complexes and the inorganic core of the O2-evolving complex in photosystem II: evaluation of the DFT/B3LYP level of theory", J. Inorg. Biochem., 100, 786-800 (2006). [37] M. Bruschi, L. De Gioia, G. Zampella, M. Reiher, P. Fantucci, M. Stein: "A theoretical study of spin states in Ni-S4 complexes and models of the [NiFe] hydrogenase active site", J. Biol. Inorg. Chem., 9, 873-884 (2004). [38] M. Swart: "Metal-ligand bonding in metallocenes: differentiation between spin state, electrostatic and covalent bonding", Inorg. Chim. Acta, 360, 179-189 (2007). [39] J. P. Perdew, "Unified Theory of Exchange and Correlation Beyond the Local Density Approximation", in Electronic structure of Solids 1991 (Eds.: P. Ziesche, H. Eschrig), Akademie, Berlin, 1991, p. 11. [40] J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J. Singh, C. Fiolhais: "PW91", Phys. Rev. B, 46, 6671-6687 Ibid E 6648 (1993) 4978 (1992). [41] X. Xu, W. A. Goddard III: "The X3LYP extended density functional for accurate descriptions of nonbond interactions, spin states, and thermochemical properties", Proc. Natl. Acad. Sci. USA, 101, 2673-2677 (2004). [42] D. A. Scherlis, M. Cococcioni, P. Sit, N. Marzari: "Simulation of heme using DFT + U: a step toward accurate spinstate energies", J. Phys. Chem. B, 111, 7384-7391 (2007). [43] N. C. Handy, A. J. Cohen: "Left-right correlation energy", Molec. Phys., 99, 403-412 (2001). [44] M. Swart, A. W. Ehlers, K. Lammertsma: "The performance of OPBE", Molec. Phys., 102, 2467-2474 (2004). [45] C. Lee, W. Yang, R. G. Parr: "Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density.", Phys. Rev. B, 37, 785-789 (1988). [46] J. Baker, P. Pulay: "Assessment of the Handy-Cohen optimized exchange density functional for organic reactions", J. Chem. Phys., 117, 1441-1449 (2002). [47] J. Baker, P. Pulay: "Assessment of the OLYP and O3LYP density functionals for first- row transition metals", J. Comput. Chem., 24, 1184-1191 (2003). [48] W. M. Hoe, A. J. Cohen, N. C. Handy: "Assessment of a new local exchange functional OPTX", Chem. Phys. Lett., 341, 319-328 (2001). [49] A. P. Bento, M. Sol, F. M. Bickelhaupt: "Ab initio and DFT benchmark study for nucleophilic substitution at carbon (SN2@C) and silicon (SN2@Si)", J. Comput. Chem., 26, 1497-1504 (2005). [50] X. Xu, W. A. Goddard III: "Assessment of Handy-Cohen Optimized Exchange Density Functional (OPTX)", J. Phys. Chem. A, 108, 8495-8504 (2004). [51] J. Conradie, A. Ghosh: "DFT calculations on the spin-crossover complex Fe(salen)(NO): a quest for the best functional", J. Phys. Chem. B, 111, 12621-12624 (2007).

Document: Reading materials for Introduction - Chapter01.doc Author: Kester Wen Jie Wong Save Date: 03/07/2012 Page 4 of 8

Modelling realistic nanostructures: bridging the gap between theory and experiment
The last twenty years have seen a colossal experimental effort to prepare and characterise ever smaller structures of inorganic materials down to as small as one nanometre in size. Such nanostructures have been demonstrated to often have widely different properties to that of the corresponding bulk material. As a result of these unique size-dependent characteristics inorganic nanostructures find application in numerous fields such as catalysis, solar cells, batteries and (bio)medical imaging. The same reduced size that makes these nanostructures interesting, however, also makes it difficult to understand the origins of their unique features fully by experiment alone. This knowledge gap is inherently linked to the difficulty of resolving the detailed atomic structures of nanostructured samples in experiment due to disorder and the associated broadening of diffraction and spectroscopic peaks.

Complementing experimental results with theoretical results is an attractive approach to closing the nanoscale knowledge gap. Here one would use computational chemistry calculations with realistic structural models of the nanostructure to gain atomistic insight into the fundamental processes underlying the size-dependency. Structural models would be obtained through global optimisation, where one aims to find the lowest energy structure for a given nanostructure, together with any information that can be obtained from experiment (e.g. composition, dimensionality, shape). Several groups have been developing such strategies in recent years using different global optimisation strategies (e.g. basin-hoping, genetic algorithms, simulated annealing, data mining) and have published proof of principle papers demonstrating that not only is it possible to reproduce experiment but also to successfully obtain microscopic insight unattainable by any other means. Most computational studies, however, at the moment still employ cuts from bulk structures or manually constructed clusters of unknown pedigree.

In the workshop we plan to bring together for the first time not only people working on global optimisation of nanostructures but also computational chemist/physicists interested in modelling their properties, and experimental chemist and physicists who prepare and characterise nanostructures. This mix of experience, as also reflected in objective 1 (see below), is aimed at promoting the developments to people outside the primary global optimisation community who can directly benefit from these approaches in their research. In the proposed workshop we will also for the first time discuss strategies for how to extend the existing global optimisation methods (previously developed for studying nanostructures in vacuo) to handle nanostructures prepared in the presence of solvents and ligands. We finally hope also to shed light on the underexplored field of structure prediction for structures that are extended in one (e.g. nanotubes) or two dimensions (e.g. thin films) and nanosized in the other dimensions. As discussed above, complementing experimental observations with theoretical predictions is an attractive proposition for closing the knowledge gap that originates from the problems with fully characterising nanostructures by experiment alone. The first computational studies predicting the likely structure of nanostructures from the bottom up focussed primarily on Lennard-Jones clusters; a theoreticians idealised yet surprisingly useful system. However, since then, the various global optimisation approaches (e.g. genetic algorithms, simulated annealing and basin-hopping) have been applied to a very diverse range of nanostructures with importance relevance to experimental nanoscience: e.g. those made out of (mixed) metals [1,2], metal oxides (SiO2 [3], ZnO [4], TiO2 [5,6], In2O3 [7], CeO2-x [8]) and metal chalcogenides (ZnS, ZnSe, CdS, CdSe [9-11]). At the same time top-down global optimisation methods were developed that use continuum surface thermodynamics and the surface energies of bulk crystal faces which predict the likely morphologies of crystalline nanostructures as a function of size, temperature and pressure [12]. Results from both bottom up and top down studies have since been applied in a number of proof of principle studies to explain the optical [13,14], catalytic [6,8,15,16] and toxicity [16] properties of nanostructures. Some very important issues remain, however, that we will address in the workshop. For example, the sheer majority of bottom-up studies focus on pure nanostructures in vacuo while in experiment the surface of such nanostructures is often decorated with organic ligands, hydroxyl groups and/or coordinating solvent molecules. Other issues we intend to address are predicting the likely atomic structures of nanotubes/nanowires [17] and thin films [18] and how to tackle ever larger nanostructures with bottom-up methods. Most bottom-up studies currently stop at nanostructures of 20-40 formula units, i.e. at the bottom of the experimental size range, because of technical problems related to the combinatorial explosion of possible low energy structural configurations. We will also discuss how results from top-down methods that do not yield explicit atomic structures, but rather relative expression of crystal faces, can be used to calculate nanostructure properties. Beside the above method related points, we also intend to discuss interesting experimental questions that need addressing. One such question is the effect of ligands on order/disorder in nanostructures. For instance, for ZnS nanostructures it has been reported that introducing ligands (and subsequently varying their surface binding strength) appears to dramatically change the nanostructure crystallinity as probed by real-space pair distribution function [19]. Similar effects have been also predicted for small gold clusters in the presence of carbon monoxide [20] and been observed in a combined experimental/computational study for thin films of Potassium Bromide after adsorption of specific organic molecules [21]. Another interesting experimental question is for above what size the nanostructures can be expected to be crystalline and to have the bulk crystal structure. For several oxide and chalcogenide nanostructures (e.g. SiO2 [3,22], ZnS [23]) bottom-up approaches show that the lowest energy structures for small size nanostructures are

Document: Reading materials for Introduction - Chapter01.doc Author: Kester Wen Jie Wong Save Date: 03/07/2012 Page 5 of 8

non-crystalline and the transition size is currently not accurately known. Recent work also demonstrated that the experimental presence of broad diffraction peaks is not necessarily a sign of bulk order [14].

References
[1] Z. Li, N.P. Young, M. Di Vece, S. Palomba, R. E. Palmer, A.L. Bleloch, B.C. Curley, R.L. Johnston, J. Jiang, Jun Yuan, Nature, 451, 46, 2008. [2] D.T. Tran, R.L. Johnston, Proc. Roy. Soc. A., 2131, 2004, 2011. [3] S.T. Bromley, E. Flikkema, Phys. Rev. Lett., 95, 185505, 2005 [4] A.A. Al-Sunaidi, A.A. Sokol, C.R.A. Catlow, S.M. Woodley, J. Phys. Chem. C, 112, 18860, 2008. [5] Z.W. Qu, G.J. Kroes, J. Phys. Chem. C, 111, 16808, 2008. [6] M. Calatayud, L. Maldonado, C. Minot, J. Phys. Chem. C, 112, 16087, 2008. [7] A.Walsh, S.M. Woodley, Phys. Chem. Chem. Phys., 12, 8446, 2010. [8] A. Migani, K.M. Neyman, F. Illas, S.T. Bromley, J. Chem. Phys., 131, 064701, 2009. [9] E. Spano, S. Hamad, C.R.A. Catlow, J. Phys. Chem. B, 107, 10337, 2003. [10] A. Burnin, E. Sanville, J.J. BelBruno, J. Phys. Chem. A, 109, 5026, 2005. [11] E. Sanville, A. Burnin, J.J. BelBruno, J. Phys. Chem. A, 110, 2378, 2006. [12] A.S. Barnard, L.A. Curtiss, Nano Lett., 5, 1261, 2005. [13] J.M. Matxain, L.A. Eriksson, J.M. Mercero, J.M. Ugalde, E. Spano, S. Hamad, C.R.A. Catlow, Nanotechnology, 17, 4100, 2006. [14] M.A. Zwijnenburg, Nanoscale, accepted for publication, June 2011. [15] A. Migani, G.N. Vayssilov, S.T. Bromley, F. Illas, K.M. Neyman, Chem. Commun., 46, 5936, 2010. [16] A.S. Barnard, Energy & Environmental Sci., 4, 439, 2011. [17] W. Sangthong, J. Limtrakul, F. Illas F, S.T. Bromley ST, Nanoscale, 2, 72, 2010. [18] J. Goniakowski, C. Noguera, Phys. Rev. B, 79, 155433, 2009. [19] B. Gilbert, F. Huang, Z. Lin, C. Goodell, H. Zhang, J.F. Banfield, Nano Lett., 6, 605, 2006. [20] K.P. McKenna, A.L. Shluger, J. Phys. Chem. C, 111, 18848, 2007. [21] T. Trevethan, B. Such, T. Glatzel, S. Kawai, A.L. Shluger, E. Meyer, P. de Mendoza, A.M. Echavarren, Small, 7, 1264, 2011. [22] E. Flikkema, S.T. Bromley, Phys. Rev. B, 80, 035402, 2009. [23] S. Hamad, C.R.A. Catlow, J. Cryst. Growth., 294, 2, 2006.

Theoretical challenges in electronic structure of nanoparticles


Clusters and nanoparticles have been used for centuries e.g. in artwork (stained glasses and paints), photography, and medicine (colloidal gold). However, it is only in the past three decades that attention has been focused on a fundamental understanding of their properties as a function of size, shape, and composition. Advancement of theoretical techniques and computer hardware and software has made it possible to study the structure-property relationships of matter containing from a few to a few thousand atoms. The accuracy with which theory can predict some of the properties as well as new species has made it an invaluable tool in the design of new materials and in guiding experiments in their focussed discovery. Computational materials science, therefore, has been the third pillar in exploring the new frontiers of nanoscience. In spite of the successes many challenges remain in developing a seamless multi-scale approach that can treat matter from molecules and small clusters to large nanoparticles and modelling the interaction from very weak van der Waals to strong covalent bonding, correlations and finally the excitations. This is particularly important when dealing with material problems in biology, energy science, catalysis, and device applications. This Workshop will bring some of the top experimental and theoretical experts in the area of cluster science and focus on hard questions and challenges that could lead to future developments for the applications of clusters and nanoparticles in catalysis, optical, magnetic, sustainable energy, device science, and biological systems.

We summarize the status of the electronic structure of clusters and nanoparticles in two parts: 1) The calculation of the ground state properties in which structure and bonding play dominant role and 2) optical and other excitation properties that are important for applications.

1) Structure and bonding The most common approach to understand the ground state structure and properties of clusters and nanoparticles is using Kohn-Sham density functional theory approach within local density approximation (LDA) or generalized gradient approximation (GGA) that has many variants to describe exchange and correlations in the system. These calculations
Document: Reading materials for Introduction - Chapter01.doc Author: Kester Wen Jie Wong Save Date: 03/07/2012 Page 6 of 8

have been highly successful in many cases in understanding experimental results such as those obtained from mass abundance spectroscopy, measurements of IP, EA, electric dipole moments, IR and Raman modes, and photoemission experiments. Such studies have not only helped the development of cluster science but in some cases led to the prediction of novel structures such as silicon fullerenes and other polyhedral forms [1] that have been subsequently realized in laboratory [2] as well as the findings of planar and other novel structures of gold clusters [3] that are attracting great interest in catalysis and nanomedicine as well as molecular assemblies. Such predictive ability of the computational methods is going to play a very important role in the development of cluster science. However, such striking findings are few and in some cases such as clusters of transition metals which are very important in catalysis and optical properties, our understanding needs to be improved because recent experiments of IR spectra [4] have been found to agree with one form of exchange-correlation functional while another exchange correlation functional gives very different energies of isomers. These experiments highlight the need for the development of theoretical and computational methods to treat these high-spin transition metal clusters. Such difficulties are also expected for partially occupied f electron systems such as rare earth doped semiconductor quantum dots for use in LED and it is important that good understanding be evolved for clusters and nanoparticles of such systems. Such developments would have a profound effect in understanding the catalytic behavior and magnetic as well as optical applications and in designing new species for cluster assembly. While in most above cases the interactions within clusters and nanoparticles are generally strong, past few years have seen emergence of the applications of clusters and nanoparticles in hydrogen storage [5] and biological systems [6]. However, these involve weak interaction of some molecules such as hydrogen (for hydrogen storage) and organic molecules (for biological applications) on clusters and nanoparticles. Research in these areas is currently evolving. Recently there have been developments of treating van der Waals interaction using the ground state density obtained from the DFT programs based on GGA approach [7] and these are promising. Such studies are also very important for studying clusters and nanostructures on graphene which is currently a very important area of research. Another area in which both weak and strong interactions play a very important role is reactions and the developments of the treatment of weak interactions as well as strong interactions in the case of partially occupied d shells will help to obtain better understanding of the reaction barriers. While reactions have been simulated on clusters for quite a long time, theoretical tools often used in these calculations were not sufficient to treat weak and strong interactions.

2) Optical spectra and other excitation properties Understanding of the optical and other excitation properties is important for many applications of clusters and nanoparticles. The above mentioned theoretical formulations of the ground state properties suffer from the deficiency that even for the weakly correlated systems such as s-p bonded metals and semiconductors, the band gap is underestimated by 20-50% and thus it is difficult to describe the optical and other excitation properties of nanoclusters properly. This problem is more severe for systems with atoms involving open d or f shells where electron correlations become more important. Often different variants of GGA or hybrid exchange-correlation functionals are used within the time dependent (TD) DFT framework for optical absorption calculation. However, their success may be limited to particular systems. What is the correct description is an important question and one needs to find better ways to describe such systems. While Bethe-Saltpeter equations for electron and hole using two particle Green's function is an appropriate approach [8], it is computationally very demanding and often used for small systems and this could be used for benchmaking approximate methods. There are also recent reports of simplifying this method for applications to large systems of the order of 1 nm size [9]. For sp bonded systems, the so-called GW method has been very successful and has been the method of choice for quantitative description of quasi-particle excitations. However, its applications to d and f systems are much fewer and these are not as successful as anticipated. Recently GW method has been used together with LDA+U [10 ] to treat localised d and f electrons and it is referred to GW@LDA+U, but this is just the beginning and much more work would be required to develop computational approaches to treat correlations and excitations in clusters and nanoparticles of such systems. There are efforts to treat correlations for the calculation of electronic and optical spectra of small systems [11] and treat excitations in metal-molecular interfaces [12]. An important application of nanoparticles such as those of CdTe, PbSe, PbTe is in solar energy area which is currently a very important area for sustainable energy solutions. This would require the calculation of excitonic properties such as exciton binding energies and optical absorption spectra. Such efforts are going on in some laboratories. For increased efficiency, important developments are taking place on the ways to have multiple exciton formation and breaking of exciton in to electron and hole to drive current. In order to obtain vital insight into the electronic processes involving excited states, efforts are now going on to extend the powerful techniques of electronic structure theory to include electron dynamics in clusters and nanparticles to describe the excited electron motion in such systems [13]. This is an emerging area for theoretical studies.

References
[1] V. Kumar, Y. Kawazoe, Phys. Rev. Lett. 87, 045503 (2001); V. Kumar, Compt. Mater. Sci. 36, 1 (2006); Nanosilicon, V. Kumar Ed., Elsevier, Oxford 2008. [2] K. Koyasu, M. Akutsu, M. Masaaki, A. Nakajima, J. Am. Chem. Soc. 127, 4995 (2005); K. Koyasu,
Document: Reading materials for Introduction - Chapter01.doc Author: Kester Wen Jie Wong Save Date: 03/07/2012 Page 7 of 8

J. Atobe, M. Akutsu, M. Mitsui, A. Nakajima, J. Phys. Chem. A 111, 42 (2007); S. Neukermans, X. Wang, N. Veldeman, E. Janssens, R.E. Silverans, P. Lievens, Int. J. Mass Spec. 252, 145 (2006). [3] R. Ferrando, G. Barcaro, A. Fortunelli, Phys. Rev. Lett. 102, 206102 (2009); T. Ishida, M. Haruta, Angew. Chem. Int. Ed. 46, 7154 (2007); P. Gruene, D.M. Rayner, B. Redlich, A.F.G. van der Meer, J.T. Lyon, G. Meijer, A. Fielicke, Science 321, 674 (2008); M. Homberger, U. Simon, Phil. Trans. R. Soc. A 368, 1405 (2010); P.D. Jadzinsky, G. Calero, C.J. Ackerson, D.A. Bushnell, R.D. Kornberg, Science 318, 430 (2007); M. Walter, J. Akola, O. Lopez-Acevedo, P.D. Jadzinsky, G. Calero, C.J. Ackerson, R.L. Whetten, H. Grnbeck, H. Hkkinen, Proc. Natl. Acad. Sci. U.S.A. 105, 9157 (2008); O. Lopez-Acevedo, J. Akola, R.L. Whetten, H. Grnbeck, H. Hkkinen, J. Phys. Chem. C 113, 5035 (2009); E. Hulkko, O. Lopez-Acevedo, J. Koivisto, Y. Levi-Kalisman, R.D. Kornberg, M. Pettersson, H. Hkkinen, J. Am. Chem. Soc. 133, 3752 (2011). [4] D.J. Harding, P. Gruene, M. Haertelt, G. Meijer, A. Fielicke, S.M. Hamilton, W.S. Hopkins, S.R. Mackenzie, S. Neville, T.R. Walsh, J. Chem. Phys. 133, 214304 (2010); D.J. Harding, T.R. Walsh, S.M. Hamilton, W.S. Hopkins, S.R. Mackenzie, P. Gruene, M. Haertelt, G. Meijer, A. Fielicke, J. Chem. Phys. 132, 011101 (2010). [5] Q. Sun, Q. Wang, P. Jena, Y. Kawazoe, J. Am. Chem. Soc. 127, 14582 (2005). [6] D.A. Giljohann, D.S. Seferos, W.L. Daniel, M.D. Massich, P.C. Patel, C.A. Mirkin, Angew. Chem. Int. Ed. 49, 3280 (2010). [7] A. Tkatchenko, M. Scheffler, Phys. Rev. Lett. 102, 073005 (2009). [8] J.C. Grossman, M. Rohlfing, L. Mitas, S.G. Louie, M.L. Cohen, Phys. Rev. Lett. 86, 472 (2001); Y. Noguchi, K. Ohno, Phys. Rev. A 81, 045201 (2010). [9] D. Rocca, D. Lu, G. Galli, J. Chem. Phys. 133, 164109 (2010). [10] H. Jiang, R.I. Gomez-Abal, P. Rinke, M. Scheffler, Phys. Rev. B 82, 045108 (2010). [11] M. Palummo, C. Hogan, P. Bagal, F. Sottile, A. Rubio, J. Chem. Phys. 131, 84101 (2009). [12] K.S. Thygesen, A. Rubio, Phys. Rev. Lett. 102, 46802 (2009). [13] J.A. Alonso, A. Castro, A. Rubio, Eds A.K. Ray, American Scientific Publishers 3, 79, USA (2010).

Document: Reading materials for Introduction - Chapter01.doc Author: Kester Wen Jie Wong Save Date: 03/07/2012 Page 8 of 8

You might also like