01.04.2014 Views

Poster Session 3 - RSC - Australian National University

Poster Session 3 - RSC - Australian National University

Poster Session 3 - RSC - Australian National University

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Poster</strong><br />

<strong>Session</strong> 3


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP351<br />

Regioselectivity in the Heck Reaction<br />

Carina Bäcktorp, Signe Teuber Henriksen, Per-Ola Norrby<br />

1 <strong>University</strong> of Gothenburg, Department of Organic Chemistry, Gothenburg, Sweden, 2 Technical<br />

<strong>University</strong> of Denmark, Lyngby, Denmark<br />

Palladium-catalyzed arylation and vinylation of olefins has been used for more than 30 years in<br />

organic synthesis. This methodology is known as the Heck reaction and its basic steps are<br />

schematically depicted below.<br />

Ar<br />

Ar<br />

Ar<br />

R<br />

Ar<br />

R<br />

R<br />

Ar<br />

H<br />

H<br />

Pd<br />

H<br />

Pd<br />

HX<br />

X<br />

Pd<br />

R H<br />

R<br />

X<br />

X<br />

Pd<br />

X Ar R<br />

XPd<br />

Ar<br />

Pd(0)<br />

R<br />

Ar<br />

R<br />

H<br />

Ar<br />

Ar<br />

ArX<br />

Pd<br />

Pd<br />

X<br />

X<br />

R<br />

R<br />

PP352<br />

Air Oxidation of Ethoxylated Surfactants – Computational Estimations of Energies and<br />

Reaction Behaviors<br />

Carina Bäcktorp, Anna Börje, J. Lars G. Nilsson, Ann-Therese Karlberg, Per-Ola Norrby, Gunnar<br />

Nyman<br />

1 Department of Chemistry, <strong>University</strong> of Gothenburg, Gothenburg, Sweden, 2 Göteborg Science Centre<br />

for Molecular Skin Research, Faculty of Science, <strong>University</strong> of Gothenburg, Gothenburg, Sweden<br />

Ethoxylated surfactants are widely used in household and industrial cleaners, in topical<br />

pharmaceuticals, cosmetics and laundry products. Surfactants are known to be skin irritants. Most<br />

cases of occupational dermatitis are considered to be irritant contact dermatitis, caused by work with<br />

water and surfactants. It is also well-known that polyethers, e. g. ethoxylated surfactants are oxidized<br />

by atmospheric oxygen.<br />

Our pervious experimental studies have shown that autoxidation of nonionic alcohol ethoxylates<br />

generates products that are skin allergens. Specific oxidation products, including hydroperoxides,<br />

have been identified in an autoxidation mixture of the pure ethoxylated alcohol pentaethylene glycol<br />

mono-n-dodecyl ether (C 12 E 5 ).<br />

Here we present an investigation where the computational method, DFT (B3LYP), has been used to<br />

calculate reaction barriers and energies for the pathways for formation of previously observed<br />

autoxidation products of the ethoxylated surfactant. In addition to the established radical chain<br />

reaction, several possibilities for intramolecular fragmentation of the intermediate radicals have been<br />

characterized. The results rationalize the formation of the identified autoxidation products, including<br />

several, which have been implicated as allergenic.<br />

H<br />

PdX<br />

Much empirical knowledge about the performance for different ligands, substrates, and reaction<br />

conditions has been built up over the years. However, the complex nature of the reaction still leaves<br />

many question marks to be answered. Only very recently has it become possible to address this<br />

complex mechanism computationally without resorting to small model systems.<br />

We have used DFT methods to investigate the source of selectivities seen in recent experimental<br />

studies. In the initial carbopalladation step, the effect of ligand on the branching ratio can be well<br />

reproduced by the calculations. For the next step, β-hydride elimination, the general assumption has<br />

been that the more reactive hydride is eliminated selectively, but very surprisingly our models instead<br />

show the selectivity to be determined by a restriction in bond rotation in the alkyl-palladium<br />

intermediate. In general, the agreement with experimental results is excellent.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP353<br />

The Amino Group in Adenine. Is it Co-planar with the Molecular Rings or Not?<br />

Wiktor Zierkiewicz 1 , Danuta Michalska 1 , Ludwik Komorowski 1 , Jiri Cerny 2 , Pavel Hobza 2<br />

1 Wroclaw <strong>University</strong> of Technology, Faculty of Chemistry, Wroclaw, Poland, 2 Academy of Sciences of<br />

the Czech Republic, Institute of Organic Chemistry and Biochemistry, Prague, Czech Republic<br />

What is the structure of adenine in the gas phase? Is the amino group in adenine coplanar with the<br />

molecular rings? These questions seem to be still open, in spite of numerous studies. The planarity of<br />

the C-NH 2 group is of particularly great importance for investigation of the molecular recognition<br />

phenomena involving nucleic acids and other systems containing the adenine residues. Ab initio MP2<br />

and density functional B3LYP methods with various basis sets have been used to calculate the<br />

optimized structures and the infrared spectra of the N9-H tautomer of adenine. MP2 calculations<br />

predict the non-planar structure, however, larger basis sets tend to decrease the degree of<br />

pyramidalization of the C-NH 2 group, while the B3LYP method consistently indicates the planar or<br />

nearly planar structure. The planarization barrier of adenine calculated with the MP2 complete basis<br />

set (CBS) limit approach is negligible, 0.015 kcal/mol, which is in agreement with the MP2-predicted<br />

barrier of 0.020 kcal/mol, reported by S. Wang and H. F. Schaefer III [1]. Thus, it can be concluded<br />

that the structure of adenine is extremely flexible with a very small degree of non-planarity of the<br />

amino group, and a low energy barrier to planarization. These results do not support the conclusions<br />

made by F. Dong and R. E. Miller [2] that the amino group in adenine is tilted about 20° out-of-plane.<br />

The anharmonic frequencies of adenine have been calculated by the B3LYP method. The theoretical<br />

results show excellent agreement with the available experimental data. The revised assignment of the<br />

infrared spectrum of adenine in an Ar matrix has been made.<br />

PP354<br />

ChemIME: An Input Method Engine for Chemists<br />

Haruka Tkeuchi, Xu Yang, Shin-Ya Takane<br />

Department of Information Systmes Engineering, Osaka Sangyo <strong>University</strong>, Daito,Osaka, Japan<br />

In the field of chemistry, which mainly treats various compounds, it is often needed to convert a<br />

compound name into corresponding molecular equations, 2D or 3D molecular structures. For such<br />

purpose, there exist some useful commercially available tools to generate the molecular structure or<br />

rational formula from an IUPAC name, for example, ChemDraw and Chemistry 4D-Draw. These<br />

software packages, however, are separate applications for each platform and cannot be invoked on<br />

the spot from other applications. To our knowledge, there is no application that has the function of inline<br />

conversion for chemical compound names. Our idea is that using an input method mechanism,<br />

which allows us to input characters other than Roman alphabets with a standard keyboard, it is<br />

possible to implement such a tool for chemists smoothly and applicably.<br />

In this paper we present a design and a prototype implementation of the input method engine that can<br />

convert various compound names (given as an IUPAC name) into corresponding molecular equations<br />

by using Java and the Java Input Method Framework. We also extend to give various 2D or 3D<br />

coordinate formats (MDL mol, CML, and Gaussian input file) with the help of CDK (Chemistry<br />

Development Kit) and some external programs.<br />

[1] Wang, S.; Schaefer III, H. F. J. Chem. Phys. 2006, 124, 044303.<br />

[2] Dong, F.; Miller, R. E. Science, 2002, 298, 1227.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP355<br />

Ab Initio Calculations of the Zero-Field Splitting Tensors of Organic Open-Shell Molecules<br />

Kenji Sugisaki, Kazuo Toyota, Kazunobu Sato, Daisuke Shiomi, Takeji Takui<br />

Departments of Chemistry and Materials Science, Graduate School of Science, Osaka City <strong>University</strong>,<br />

3-3-138 Sugimoto, Sumiyoshi-ku, Osaka, Japan<br />

Zero-field splitting (ZFS) is the separation of multiplet sublevels in the absence of an external<br />

magnetic field. Its origin is spin-spin (SS) and spin-orbit (SO) couplings. Theoretical calculations of<br />

ZFS tensor (D tensor) have been the subject of intense interest, because the D tensor contains<br />

essentially important information on the spatial distributions of unpaired electrons. In ZFS of organic<br />

open-shell molecules, SS is in general more important than SO, but when (π,π) states and (n,π) states<br />

are energetically close to each other, strong spin-orbit coupling can be expected from the El-Sayed’s<br />

rule [1]. In this work, D SS and D SO of organic molecules having (n,n)-, (n,π)-, and (π,π)-types of spin<br />

configuration are investigated.<br />

Spin-spin contributions to the D tensor can be calculated as follows [2].<br />

D<br />

SS<br />

ab<br />

4 2<br />

( 2S<br />

+ 1)<br />

( d ) ( 2ˆ s sˆ<br />

− sˆ<br />

sˆ<br />

− sˆ<br />

sˆ<br />

)Ψ<br />

= α<br />

Ψ ∑<br />

(1),<br />

ij ab iz jz ix jx iy jy<br />

S<br />

( d )<br />

ij ab<br />

i< j<br />

r δ<br />

=<br />

2<br />

ij ab<br />

− 3<br />

( r ) ( r )<br />

ij a ij b<br />

5<br />

rij<br />

In the past decade, calculations of D SS using the CASSCF wavefunction have been reported [3]. To<br />

investigate D SS of relatively large systems, the SAC-CI method [4] is useful, because computational<br />

cost of the CASSCF method strongly depends on the size of active space. In this work, we will report<br />

preliminary results of the D SS calculations based on McWeeny and Mizuno’s equation (eq. 3) [5], which<br />

is exact when the wavefunction Ψ consists of a single determinant, with the SAC-CI spin density.<br />

D<br />

1<br />

( ) ∑ ρ<br />

2S<br />

+ 1<br />

(2)<br />

α −β<br />

α −β<br />

α −β<br />

α −β<br />

* *<br />

(<br />

µν<br />

ρκλ<br />

− ρ<br />

µλ<br />

ρκν<br />

) ∫ µ ( r1<br />

) κ ( r2<br />

)( d12<br />

) ν ( r1<br />

) λ( r2<br />

)<br />

ab<br />

dr dr<br />

SS<br />

ab<br />

=<br />

1 2<br />

S<br />

µνκλ<br />

Spin-orbit contributions to the D tensor in the spherical representation can be calculated in terms of<br />

the second-order perturbation theory with the sum-over-state formula.<br />

H<br />

SO<br />

D<br />

SO<br />

ij<br />

=<br />

∑<br />

3<br />

i<br />

Ψ H<br />

0<br />

SO λ<br />

k λ k<br />

Ψn<br />

Ψn<br />

H<br />

λ 3<br />

E − E<br />

n, λ,<br />

k<br />

n 0<br />

1 ⎡<br />

⎢<br />

2m<br />

c ⎢⎣<br />

Z e<br />

r<br />

A<br />

=<br />

2 2 ∑ ∑<br />

2<br />

SO 3<br />

Ψ<br />

j<br />

0<br />

(4),<br />

2<br />

e<br />

⎤<br />

I ˆ<br />

iA<br />

⋅ sˆ<br />

i<br />

− Iˆ<br />

ij<br />

⋅ ( sˆ<br />

i<br />

+ 2ˆ s<br />

j<br />

)<br />

3<br />

⎥ (5)<br />

i j rij<br />

⎥⎦<br />

i, A iA<br />

,<br />

Here, Ψ is the n’th eigenfunction of the non-relativistic Schrödinger equation with λ = 2S+1 and k = M S .<br />

The calculations of D SO based on eq. 4 at the CASSCF level have been reported [3a,6]. However,<br />

CASSCF energy difference is not directly comparable to the experimental ∆E. For example, a T 1 state<br />

of pyrazine is calculated to be ππ* at the CASSCF(10,8)/cc-pVDZ level, rather than nπ*. Incorrect<br />

energy difference may seriously affect the accuracy of the calculated D SO . To avoid this difficulty, we<br />

have replaced the CASSCF energy difference to MRMP2 one.<br />

[1] El-Sayed, M. A. J. Chem. Phys. 1963, 38, 2834–2838.<br />

[2] Harriman, J. E. Theoretical Foundations of Electron Spin Resonance; Academic Press: New York, 1978.<br />

[3] (a) Havlas, Z.; Kývala, M.; Michl, J. Mol. Phys. 2005, 103, 407–411., and references therein. (b) Loboda, H., et<br />

al. Chem. Phys. 2003, 286, 127–137., and reference therein.<br />

[4] (a) Nakatsuji, H. Chem. Phys. Lett. 1978, 59, 362–364. (b) Nakatsuji, H.; Hirao, K. J. Chem. Phys. 1978, 68,<br />

2053–2065. (c) Nakatsuji, H. Chem. Phys. Lett. 1979, 67, 329–333.<br />

[5] McWeeny, R.; Mizuno, Y. Proc. R. Soc. London 1961, 259, 554–571.<br />

[6] Rubio-Pons, Ò.; Minaev, B.; Loboda, O.; Ågren, H. Theor. Chem. Acc. 2005, 113, 15–27., and references<br />

therein.<br />

(3)<br />

PP356<br />

A Chemically Reasonable Model of Various Phosphine Ligands: Application of CCSD(T)<br />

Calculation to Large Transition Metal Complexes<br />

Yu-ya Ohnishi, Mayu Nakaoka, Yoshihide Nakao, Hirofumi Sato, Shigeyoshi Sakaki<br />

Department of Molecular Engineering, Graduate School of Engineering, Kyoto <strong>University</strong>, Kyoto,<br />

Kyoto, Japan<br />

Chemically reasonable models of PR 3 (R = Me, Et, i Pr, t Bu, Cy, or Ph) were constructed to apply the<br />

CCSD(T) method to large transition metal complexes. We elucidated that the dominant role of PR 3<br />

listed above is sigma-donation to metal centre with its lone pair orbital. In out new model, R is<br />

replaced with the one electron pseudo carbon atom including the frontier orbital consistent quantum<br />

capping potential (FOC-QCP) (eq. 1) [1] which reproduces the frontier orbital energy of PR 3 . The<br />

steric effect is incorporated well by the new procedure named steric repulsion correction (SRC), which<br />

was not incorporated in the usual QM/MM methods.<br />

FOC-QCP<br />

2<br />

( −ζ<br />

)<br />

U ( r)<br />

= U ( r)<br />

− 3 r exp r (1)<br />

l<br />

l<br />

To examine the performance of this FOC-QCP method with the SRC, the activation barriers and<br />

reaction energies of the reductive elimination reactions of C 2 H 6 and H 2 from M(R 1 ) 2 (PR 2 3) 2 (M = Ni, Pd,<br />

or Pt; R 1 = Me for R 2 = Me, Et, or i Pr, or R 1 = H for R 2 = t Bu) were evaluated with the DFT[B3PW91],<br />

MP4(SDQ), and CCSD(T) methods. The FOC-QCP method reproduced very well the DFT[B3PW91]-<br />

and MP4(SDQ)-calculated energy changes of the real complexes with PMe 3 . For more bulky<br />

phosphine, the SRC is crucially important to present correct energy change, in which the MP2 method<br />

presents reliable steric repulsion correction like the CCSD(T) method because the systems calculated<br />

in the SRC do not include transition metal element.<br />

Also, the coordination energies of CO, H 2 , N 2 , and C 2 H 4 with a large dinuclear complex [RhCl(P i Pr 3 ) 2 ] 2<br />

were theoretically calculated by the CCSD(T) method combined with the FOC-QCP and the SRC. The<br />

CCSD(T)-calculated energies agree well with the experimental ones, as shown in Figure. On the<br />

other hand, the DFT[B3PW91]-calculated energies of the real complexes considerably deviate from<br />

the experimental ones.<br />

We also calculated the reaction energy of CO coordination with CoCl 2 (PR 3 ) 2 (R = Et, Cy, or Ph) (eq. 2)<br />

and decarbonylation from (PPh 3 ) 2 (Cl)PtC(O)C 6 H 5 (eq. 3) to compare with the experimental results.<br />

The details of methods and results will be shown in poster session.<br />

CoCl 2 (PR 3 ) 2 + CO CoCl 2 (CO)(PR 3 ) 2 (2)<br />

(PPh 3 ) 2 (Cl)PtC(O)C 6 H 5 (PPh 3 ) 2 (Cl)PtC 6 H 5 + CO (3)<br />

B3PW91 (in vacuo)<br />

B3PW91 (in toluene)<br />

B3PW91/FOC-QCP + SRC<br />

CCSD(T)/FOC-QCP + SRC<br />

Figure. The error of the coordination energies (kcal/mol) of CO, H 2 , N 2 -end-on, and<br />

C 2 H 4 to [RhCl(P i Pr 3 ) 2 ] 2 from the experimental values.<br />

[1] Ohnishi, Y.-y.; Nakao, Y.; Sato, H.; Sakaki, S. J. Phys. Chem. A 2008, 112, 1946-1955.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP357<br />

A DFT Study on a Natural Diels-Alder Reaction<br />

Sadra Kashefolgheta, Mehdi Irani, Mohammad Reza Gholami<br />

Department of Chemistry-Sharif <strong>University</strong> of Technology, Tehran, Iran, Islamic Republic of<br />

Macrophomate synthase (MPS), which promotes the conversion of 2-pyrones to benzoates, has<br />

garnered considerable attention as a possible natural Diels-Alder reaction [1]. There are two possible<br />

mechanisms for the MPS-Catalyzed Formation of Macrophomate (i) Michel addition followed by aldol<br />

condensation and (ii) Diels-Alder reaction [Figure 1].<br />

The activation and TS energies of both pathways have been investigated by density functional<br />

methods in order to find the preference of the mechanisms. The two possible mechanisms, have been<br />

studied at the B3LYP/6-311++G (d,p) level of theory. In addition, aromatization energy of products has<br />

been determined consequently in the same method.The catalytic characteristic of MPS has been<br />

investigated further by changing metal ion and performing calculation with Ca +2 and Ni +2 . The effects<br />

of different substituents on 2-pyron have been also studied using the DFT method.<br />

PP358<br />

Structure and Vibrational Spectra of Hydrogen-Bonded Clusters by the Anharmonic Downward<br />

Distortion Following Method<br />

Satoshi Maeda, Yu Watanabe, Yi Luo, Koichi Ohno<br />

Department of Chemistry, Graduate School of Science, Tohoku <strong>University</strong>, Sendai, Japan<br />

Vibrational spectroscopy is a powerful tool to investigate hydrogen-bond (H-bond) clusters. Except for<br />

very simple cases, calculations are necessary to reveal structures from experimental spectra. At first,<br />

one needs to explore the potential energy surface (PES) to list up stable isomers. Then, vibrational<br />

analyses will be made for each structural candidate to find out isomers which correspond to observed<br />

vibrational spectra. In this study, we show that the anharmonic downward distortion following (ADD-<br />

Following) method [1] can be a powerful tool for both PES exploration [2] and anharmonic vibrational<br />

analyses [3], with some examples of applications to H-bond cluster systems [4-6].<br />

The ADD-Following [1] is an idea to make uphill walking along reaction routes from a minimum (EQ:<br />

EQuilibrium structure) to (first-order) saddle points (TS: Transition State structure). Since a real<br />

(anharmonic) PES distorts downward from a harmonic PES in direction leading to another EQ, the<br />

ADD can be a symptom of chemical reactions leading to another EQ. ADDs can be found as energy<br />

minima on an iso-energy hypersurface of harmonic potential centered at a starting EQ, and such<br />

energy minima can be detected and then traced by the scaled hypersphere search (SHS) method [1].<br />

The Full-ADD-Following dealing with all ADDs has been employed in global reaction route mapping on<br />

PESs of small systems [1]. The Large-ADD-Following (LADD-Following) was proposed for fast<br />

exploration of low energy parts on PESs by tracing only important large ADDs [2], and it has been<br />

applied to H-bond clusters such as (H 2 O) 8 [2], H + (H 2 O) n (n = 5-8) [4], and H 2 S(H 2 O) n (n = 5-7) [5]. We<br />

further developed the SHS based polynomial fitting (SHS-PF) technique for very efficient construction<br />

of potential energy function (PEF) in the form of sixth-order polynomial [3]. The SHS-PF method can<br />

dramatically reduce numbers of ab initio data for obtaining a sixth-order-PEF from the sixth power to<br />

the second power, by selecting sampling points based on ADD data by the SHS method. Anharmonic<br />

analyses of (H 2 O) n (n = 2-5) were performed by using PEFs constructed by the SHS-PF method, and<br />

experimental intramolecular-mode frequencies were reproduced with errors of about only 10 cm -1 [6].<br />

Figure 1<br />

[1] Jörg M. Serafimov; Thomas Westfeld; Beat H. Meier; Donald Hilvert. J. Am. Chem. Soc. 2007, 129, 9580.<br />

[1] Ohno, K.; Maeda, S. Chem. Phys. Lett. 2004, 384, 277. Maeda, S.; Ohno, K. J. Phys. Chem. A 2005, 109,<br />

5742. Ohno, K.; Maeda, S. J. Phys. Chem. A 2006, 110, 8933.<br />

[2] Maeda, S.; Ohno, K. J. Phys. Chem. A 2007, 111, 4527.<br />

[3] Maeda, S.; Watanabe, Y.; Ohno, K. J. Chem. Phys. 2008, 128, 144111.<br />

[4] Luo, Y.; Maeda, S.; Ohno, K. J. Phys. Chem. A 2007, 111, 10732.; Luo, Y.; Maeda, S.; Ohno, K. (submitted).<br />

[5] Maeda, S.; Ohno, K. J. Phys. Chem. A 2008, 112, 2962.<br />

[6] Watanabe, Y.; Maeda, S.; Ohno, K. (submitted).


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP359<br />

Analysis of an Electronic Spectrum using the Ab Initio Path Integral Molecular Dynamics<br />

Method<br />

Masataka Sugimoto 1 , Motoyuki Shiga 2 , Masanori Tachikawa 1<br />

1 Graduated School of Integaretd Science, Yokohama City <strong>University</strong>, Yokohama, Japan, 2 Japan<br />

Atomic Energy Agency, Tokyo, Japan<br />

In the conventional molecular orbital (MO) calculation, the allowed or forbidden electronic transitions<br />

are determined by geometrical symmetry of molecules. However, to reproduce the experimental<br />

electronic spectrum, it is indispensable to take account of “geometrical fluctuation” to molecular<br />

structure in theoretical study, since the molecules in real system always keep changing its structure<br />

with vibrational motion due to thermal and quantum effects. Although there are some reports that insist<br />

importance of these effects on electronic transition by using ab initio path integral simulation [1, 2],<br />

these reports seem to be not enough because of inconsistency of electronic structure evaluation<br />

between ground and excited states, and the number of beads on path integral calculation. In the<br />

present paper, we have analyzed electronic spectrum of ethylene molecule by ab initio path integral<br />

molecular dynamics (PIMD) calculation with consistent MO level between ground and excited states.<br />

Figure 1 shows (a) calculated and (b) experimental [3] electronic spectra of ethylene at 300K.<br />

Geometrical structure of ground state was obtained with MP2 / 6-31+G* level of calculation, and<br />

electronic structure of excited state was estimated with CIS(D)/6-31+G*. Since the equilibrium<br />

structure of ethylene has D 2h symmetry, allowed transitions are only B 1u<br />

(8.14eV, 10.9eV, 11.1eV), B 2u<br />

(10.6eV), and B 3u (7.76eV, 9.52eV) as shown in vertical lines in Figure 1(a). The electronic spectrum<br />

of PIMD is broader than that of classical MD, because thermal and quantum effects cause distortion of<br />

molecular structure and break its symmetry.<br />

(a)<br />

(b)<br />

200<br />

1<br />

70<br />

180<br />

PIMD<br />

0.9<br />

exp.<br />

Classical<br />

60<br />

160<br />

SP calc.<br />

0.8<br />

140<br />

0.7<br />

50<br />

120<br />

0.6<br />

40<br />

100<br />

0.5<br />

80<br />

0.4<br />

30<br />

60<br />

0.3<br />

20<br />

40<br />

0.2<br />

10<br />

20<br />

0.1<br />

0<br />

0<br />

0<br />

7 8 9 10 11 12<br />

7 8 9 10 11 12<br />

Energy / eV<br />

Energy / eV<br />

Figure 1. Electronic spectrum of ethylene; (a) our works, (b) experimental data.<br />

Cross section / Mb<br />

Oscillator strength<br />

Cross section / Mb<br />

PP360<br />

Computational Study of the Sonogashira Cross-Coupling Reaction in the Gas Phase<br />

Peeter Burk 1 , Jaana Tammiku-Taul 1 , Lauri Sikk 1 , Andras Kotschy 2<br />

1 Institute of Chemistry, <strong>University</strong> of Tartu, Tartu, Estonia, 2 Institute of Chemistry, Eötvös Loránd<br />

<strong>University</strong>, Budapest, Hungary<br />

Many chemical reactions proceed in the presence of catalysts. Cheaper and more effective catalytic<br />

systems are searched to accelerate the reactions and rise their selectivity. The application of transition<br />

metals in organic synthesis has become an accepted and valued tool. Cross-coupling reactions are by<br />

now a widely accepted tool of synthetic chemists. They include a bunch of carbon-carbon bond<br />

forming reactions [1]. Cross-coupling reactions are usually catalyzed by transition metals and<br />

distinguished on the basis of the used transmetalating agent. Organocopper reagent is used in case of<br />

the Sonogashira cross-coupling [2].<br />

The aim of current study was to investigate the reaction mechanism of the Sonogashira coupling in the<br />

gas phase and to find the factors, which influence its kinetics.<br />

The Sonogashira cross-coupling reaction between bromobenzene and phenylacetylene was modelled<br />

using density functional theory B3LYP/cc-pVDZ method. In the case of palladium and copper<br />

Stuttgart-Dresden effective core potential with the accompanying basis sets were used.<br />

The cross-coupling reaction begins with the oxidative addition of bromobenzene onto the palladium<br />

catalyst Pd(PH 3 ) 2 , which is followed by the formation of cis-Pd(PH 3 ) 2 BrPh complex. The subsequent<br />

step is isomerisation leading to trans-Pd(PH 3 ) 2 BrPh complex. The oxidative addition is the reaction<br />

rate-limiting step. Copper(I) bromide as a co-catalyst reacts with phenylacetylene in the presence of a<br />

base (trimethylamine) and copper phenylacetylenide is formed, which in a transmetalation step reacts<br />

with trans-Pd(PH 3 ) 2 BrPh complex. The product of transmetalation step, cis-Pd(PH 3 ) 2 (Ph)C≡CPh,<br />

decomposes into palladium diphosphine, Pd(PH 3 ) 2 , and diphenylacetylene, Ph-C≡C-Ph, during<br />

reductive elimination. The regenerated palladium catalyst is ready to enter a next cycle. The complete<br />

catalytic cycle is exothermic and has a negative Gibbs’ free energy change (∆H = -39.2 kcal/mol, ∆G =<br />

-38.0 kcal/mol). Our calculated catalytic cycle is in agreement with reaction schemes in literature [2,3].<br />

[1]. Kotschy, A.; Timįri, G. Heterocycles from Transition Metal Catalysis, Springer, 2005.<br />

[2]. Sonogashira, K. J. Organomet. Chem. 2002, 653, 46.<br />

[3]. Chinchilla, R; Nàjera, C. Chem. Rev. 2007, 107, 874–922.<br />

[1] Schulte, J; Ramírez, R; Böhm, M. C. Chem. Phys. Let. 2000, 322, 527-535.<br />

[2] Sala, F. D; Rousseau, R; Görling, A; Marx, D. Phys. Rev. Let. 2004, 92, 183401-1-183401-4.<br />

[3] Lu, H-C; Chen, H-K; Cheng, B-M. Anal. Chem. Phys. 2004, 76, 5965-5967.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP361<br />

Identification of an Aryloxenium Ion by Time Resolved Resonance (TR 3 ) Spectroscopy and<br />

Density Functional Theory: First Vibrational Spectrum of an Oxenium Ion<br />

Stephen Glover 1 , Michael Novak 2 , Yue-Ting Wang 2 , David Phillips 3 , Jiadan Xue 3<br />

1 <strong>University</strong> of New England, School of Science and Technology, Armidale, New South Wales,<br />

Australia, 2 Miami <strong>University</strong>, Department of Chemistry and Biochemistry, Oxford, Ohio, United States,<br />

3 Hong Kong <strong>University</strong>, Department of Chemistry, Hong Kong, China<br />

Aryloxenium ions 2 from solvolysis of 1 have previously been generated and trapped by water and<br />

azide, and product studies [1-3] as well as DF calculations [4] point to<br />

their existence in resonance form II rather than I.<br />

O<br />

O<br />

O<br />

Laser flash photolysis studies on 1 (R=C 6 H 4 Me-4) generated 4'-<br />

methylbiphenylyloxenium ion 3 with a UV absorption maximum of 460<br />

nm at pH 7.1 and a lifetime in aqueous solution of 170 ns, similar to<br />

the150-180 ns established earlier by azide trapping methods [2].<br />

Oxenium ion 3 has recently been observed by TR 3 , which with a probe pulse at 460 nm, provides<br />

resonance enhancement of both IR and Raman vibrational bands. The spectrum (Fig. 1) was<br />

simulated at the B3LYP/6-31G* level and the spectral lines and intensities correspond to vibrational<br />

modes of resonance forms 3-II and III. The oxenium ion charge is strongly delocalised into the 4-<br />

methylphenyl ring and the ion, which better resembles a resonance-delocalised 1-oxo-2,5-<br />

cyclohexadien-4-yl cation (3-II), possesses a normal cross conjugated carbonyl at 1635 cm -1 . Analysis<br />

of molecular orbitals indicates that the Raman probe radiation at 460 nm resonates with the carbonyl<br />

and C=C stretch modes of 3-II and III.<br />

O<br />

O<br />

O<br />

∗<br />

∗<br />

Experimental and Theoretical Vibrational Frequencies of<br />

4'-Methylbiphenylyloxenium ion<br />

IR1<br />

R3<br />

R2<br />

∗ ∗<br />

R1<br />

R7<br />

IR4<br />

R6<br />

IR3<br />

R5<br />

IR2<br />

AcO<br />

R4<br />

1<br />

R<br />

R<br />

I<br />

Raman Exp.<br />

IR Calc.<br />

Raman Calc.<br />

2<br />

R<br />

II<br />

PP362<br />

Molecular Dynamic Simulations Can Complement Experiments to Probe Antibiotics Diffusion<br />

through Bacterial Porins<br />

Eric Hajjar, Amit Kumar, Enrico Spiga, Francesca Collu, Paolo Ruggerone, Matteo Ceccarelli<br />

Department of Physics, <strong>University</strong> of Cagliari., Cagliari, Sardinia., Italy<br />

The increase in number and complexity of bacterial resistance to antibiotics severely affects the<br />

treatment of infectious diseases. This calls for a fast development of new antibiotics. As antibiotics<br />

targets are located inside bacteria, the first step of antibacterial activity is the uptake of antibiotics. The<br />

process is a severe bottleneck particularly in Gram-negative bacteria since the presence of the outer<br />

membrane (OM) restricts the entry of any molecule. General diffusion porins, for example outer<br />

membrane protein F (OmpF) are present in large abundance in the OM and constitute the main<br />

pathway for different classes of antibiotics, such as beta-lactams and fourth generation of<br />

floroquinolones.<br />

Bacteria can resist to antibiotics, either by under expressing porins in the OM or by mutations. The<br />

understanding of how antibiotics diffuse through porins can provide information on how to design<br />

novel antibiotics with improved permeation properties, to solve at least partially the problem of<br />

resistance. One of the ways to tackle the problem of antibiotic resistance is to investigate the<br />

permeation properties and what governs the translocation process. Recently, experiments combined<br />

with MD simulations pointed out how the diffusion of antibiotics is a molecular-based process, i.e.<br />

small differences in the antibiotic structure can affect its flux through porins.<br />

The high resolution of MD simulations can be particularly useful to the design of antibiotics with<br />

improved permeation properties. The process of antibiotic diffusion through the porins occurs on<br />

microsecond time scale. To simulate such process we built a molecular system from the X-ray<br />

structure of OmpF porin. OmpF was embedded in biological membrane environment. As we want to<br />

study various antibiotics as well as OmpF porin and its variants, this is not feasible with standard MD<br />

simulations. Instead we use a recent algorithm, metadynamics, that allows to overcome the timescale<br />

problem by adding history dependent potential that accelerates the evolution of the system. Each<br />

antibiotic diffusion process was analysed by (i) constructing the associated free energy map, (ii)<br />

estimating the free energy barrier, (iii) identifying the binding sites inside the porin, and (iv)<br />

characterizing the nature of the interactions (h-bonded and hydrophobic) between antibiotics and<br />

OmpF in the putative binding sites.<br />

Me<br />

I<br />

Me<br />

II<br />

3<br />

Me<br />

III<br />

600 800 1000 1200 1400 1600 1800 2000<br />

Vibrational Frequency (cm-1)<br />

Figure 1<br />

Our MD simulation results for different antibiotics were then compared to the experimental data<br />

available in the literature and provided by our collaborators; this allowed discussing the mechanism of<br />

antibiotic penetration. For example, in electrophysiology experiments interruptions in ionic current<br />

(blockage) are related to antibiotic translocation. Our simulation results show good agreement with<br />

such electrophysiology experiments that report blockage. However, there are cases where we observe<br />

translocation and electrophysiology do not report blockage. To conclude we discuss the following<br />

questions: does blockage always imply translocation? Can theory suggest new experiments or help to<br />

design new measurement strategy? This points out the importance of combining properly theory and<br />

experiments in order to benefit the design of antibiotics with improved permeation properties.<br />

[1] Novak, M.; Glover, S. A. J. Am. Chem. Soc. 2004, 126, 7748; 2005, 127, 8090.<br />

[2] Novak, M.; Poturalski, M. J.; Johnson, W. L.; Jones, M. P.; Wang, Y.; Glover, S. A. J. Org. Chem. 2006, 71,<br />

3778.<br />

[3] Novak, M.; Brinster, A. M.; Dickhoff, J. N.; Erb, J. M.; Jones, M. P.; Leopold, S. H.; Vollman, A. T.; Wang, Y.;<br />

Glover, S. A. J. Org. Chem. 2007, 72, 9954.<br />

[4] Glover, S. A.; Novak, M. Can. J. Chem. 2005, 83, 1372.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP363<br />

Avoiding Heisenberg with Certainty<br />

Yves A. Bernard, Peter M. W. Gill<br />

Research School of Chemistry, <strong>Australian</strong> <strong>National</strong> <strong>University</strong>, Canberra, Australia<br />

The Heisenberg Uncertainty Principle states that we cannot simultaneously measure the position r<br />

and the momentum p of a particle. However, recent electron correlation models based on intracules<br />

require information about both the positions and the momenta of the electrons and the limitations of<br />

the Uncertainty Principle are therefore very inconvenient. To get around this problem, in 1932, Wigner<br />

proposed a phase space distribution W 1 (r,p) for a particle which gives the correct position and<br />

momentum space density when integrated over p or r respectively. However, because W 1 (r,p) is not<br />

necessarily positive, it cannot be strictly interpreted as a joint probability density of r and p.<br />

In this poster, we show that information about r and p can be obtained rigorously by considering a new<br />

variable s = r⋅p, the position-momentum dot product. This “posmom” variable, s, is a quantum<br />

mechanical observable and its density D(s) can be computed efficiently. We also show that the<br />

posmom quasi-density W(s) that is derived from the W 1 (r,p) is an approximation of D(s). We present<br />

results for simple systems such as the harmonic oscillator and the hydrogen atom, as well as for<br />

molecular systems.<br />

PP364<br />

Parallel Implementation of RI-MP2 Energy Calculations of Large Molecules<br />

Michio Katouda, Shigeru Nagase<br />

Department of Theoretical and Computational Molecular Science, Institute for Molecular Science,<br />

Okazaki, Japan<br />

The resolution-of-the-identity (RI) approximation has been developed for second-order Moller-Plesset<br />

perturbation theory (MP2) to reduce the computational cost [1]. To make the RI-MP2 method<br />

applicable to large molecules, the parallel implementation is important [2, 3]. In this study, we present<br />

a parallel implementation of RI-MP2 energy calculations to assess its performance.<br />

The most computationally demanding step in RI-MP2 energy calculations is the construction of two<br />

electron integrals in MO basis with simple matrix-matrix multiplications of three-center two-electron<br />

integral matrices. To make the parallel algorithm efficient, uniform distribution of the construction of<br />

two electron integrals on each processor is needed. Parallelization of the construction of two electron<br />

integrals with efficient load balancing is archived by dividing occupied orbitals into batches whose<br />

sizes are almost equal and distributing these batches on each processor.<br />

Benchmark calculations were performed for the taxol molecule (C 47 H 51 NO 14 ) with the 6-311G** basis<br />

set and the Ahlrichs’s SVP auxiliary basis set (1484 functions, 4175 auxiliary functions, and 164<br />

correlated orbitals). Calculations were performed on a Linux cluster of 3.2 GHz Pentium4 PCs with 4<br />

GB memory and 400 GB hard disk. These PCs are connected by Gigabit-Ethernet network. Table 1<br />

shows the elapsed times and speed-up of parallel RI-MP2 calculations. For 32 processors, the speedup<br />

is 29.6 and the elapsed time is only 25 minutes. Benchmark calculations were also performed for<br />

(C 96 H 24 ) 2 , a cluster model of two layer graphene sheets [4], with the 6-311G** basis set and the<br />

Ahlrichs’s SVP auxiliary basis set (3936 functions, 11280 auxiliary functions, and 408 correlated<br />

orbitals). By using 32 processors, calculation was finished within 43 hours. These results indicate that<br />

RI-MP2 energy calculations of large molecules can be performed in modest time using low-cost PC<br />

clusters.<br />

Table 1. Elapsed times and speed-up of parallel<br />

RI-MP2 energy calculations of taxol.<br />

Processors Time [m] Speed-up<br />

1 741.7 1.0<br />

2 344.9 2.2<br />

4 168.1 4.4<br />

8 86.8 8.5<br />

16 45.5 16.3<br />

32 25.1 29.6<br />

[1] Feyereisen, M.; Fitzgerald, G.; Komornicki, A. Chem. Phys. Lett. 1993, 208, 359.<br />

[2] Bernholdt, D. E.; Harrison, R. J. Chem. Phys. Lett. 1996, 250, 477.<br />

[3] Hättig, C.; Hellweg, A.; Köhn, A. Phys. Chem. Chem. Phys. 2006, 8, 1159.<br />

[4] Grimme, S.; Mück-Lichtenfeld, C.; Antony, J. J. Phys. Chem. C 2007, 111, 11199.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP365<br />

Development of an Efficient Computational Scheme for Relativistic GMC-QDPT and its<br />

Application to Molecular Systems<br />

Ryo Ebisuzaki, Yoshihiro Watanabe, Haruyuki Nakano<br />

Department of Molecular Chemistry, Graduate School of Sciences, Kyushu <strong>University</strong>, Fukuoka,<br />

Japan<br />

In electronic structure calculations for systems that contain heavy elements, treatment of both the<br />

electron correlation and relativistic effects is essential for high accuracy. However, quantum chemical<br />

theory including relativistic effects is much complicated compared to non-relativistic theory, and<br />

therefore even today it is not an easy task to perform relativistic calculations. In addition, calculations<br />

of heavy-atom compounds require much higher computational cost than those of the systems that<br />

consist of first- and second-row atoms such as hydrocarbons. Thus, due to its complexity and<br />

computational cost, highly accurate quantum chemical calculations including both the electron<br />

correlation and relativistic effects are restricted to small sized molecules consisting of few atoms.<br />

Recently, our group has developed the general multiconfiguration quasi-degenerate perturbation<br />

theory (GMC-QDPT) [1] and its relativistic generalized version [2]. GMC-QDPT is a multistate<br />

multireference perturbation theory that is applicable to any multiconfigurational reference wave<br />

functions. This method enjoys both high computational accuracy and efficiency. However, even with<br />

this relativistic GMC-QDPT, still it is not easy to efficiently calculate heavy-atom compounds with many<br />

electrons to be treated.<br />

In this presentation, we report the following of our recent development for improving computational<br />

efficiency of the relativistic GMC-QDPT,<br />

(1) A novel computational scheme based on the matrix elements reference functions and ionized<br />

determinants, which is more efficient than our previous computational scheme based on Goldstone<br />

diagrams [3],<br />

PP366<br />

Ab Initio Benchmark Calculations on Monoligand Ca(II) Complexes and Comparison with<br />

Density Functional Theory Methodologies<br />

Víctor M Rayón 1 , Haydee Valdés 2 , Natalia Díaz 2 , Dimas Suárez 2<br />

1 <strong>University</strong> of Valladolid, Department of Physical Chemistry and Inorganic Chemistry, Valladolid,<br />

Spain, 2 <strong>University</strong> of Oviedo, Department of Physical Chemistry and Analytical Chemistry, Oviedo,<br />

Spain<br />

Calcium is an essential element to life and as such plays a crucial role in many structural and reactive<br />

biochemical processes [1]. The understanding of these processes at the molecular level is of course of<br />

particular importance. To achieve this end, the study of biomimetics, i.e. model systems that only<br />

include the metal and chelating ligands, is usually proposed as a first step.<br />

We have carried out a theoretical study employing both ab initio correlated wave function and density<br />

functional methods on a set of 24 monoligand Ca(II) complexes. Several basis sets ranging from<br />

double to quintuple-zeta quality have been used, including all-electron correlation consistent basis<br />

sets. We can therefore provide a reliable high-level benchmark ab initio database for a set of 1:1<br />

complexes of Ca(II). The performance of four different functionals, namely, PW91, PBE, B3LYP, and<br />

TPSS in the prediction of binding energies, metal-ligand bond distances and proton affinities has been<br />

also assessed. Additionally, an analysis of the metal-ligand interaction has been carried out by means<br />

of an energy decomposition method.This analysis provides information on the relative importance of<br />

the electrostatic, induction and dispersive contributions to the metal-ligand interaction energy and<br />

might be useful for the assessment of ‘non-bonded’ molecular mechanics potentials that are no the<br />

basis of many biomolecular simulations [2].<br />

[1] Berg, J. M.; Tymoczko, J. L.; Stryer, L. Biochemistry, 5 th Ed. Freeman & Co., New York, 2003.<br />

[2] Rayón, V. M.; Valdés, H.; Díaz, N.; Suárez, D. J. Chem. Theory Comput. 2008, 4, 243-256.<br />

(2) The Kramers-restricted formalism corresponding to spin-adapted formalism of non-relativistic<br />

theory,<br />

(3) Parallel algorithm.<br />

As an application, we also report the d-d excitation energy spectra of platinum complexes [PtX n ] 2- (X =<br />

halogen atom) with the new computational scheme. The calculated values were in good agreement<br />

with experimental data.<br />

[1] (a) H. Nakano, R. Uchiyama, and K. Hirao, J. Comput. Chem. 2002, 23, 1166-1175.<br />

(b) H. Nakano, J. Chem. Phys, 1993, 99, 7983–7992.<br />

[2] M. Miyajima, Y. Watanabe, and H. Nakano, J. Chem. Phys. 2006, 124, 044101<br />

[3] R. Ebisuzaki, Y. Watanabe, and H. Nakano, Chem. Phys. Lett. 2007, 442, 164-169.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP367<br />

Aldehyde Dehydrogenase Enzymatic Chemistry: Insights from Hybrid QM/MM Calculations<br />

Troy Wymore, James Keener, Shawn Brown<br />

Pittsburgh Supercomputing Center, <strong>National</strong> Resource for Biomedical Supercomputing, Pittsburgh,<br />

PA, United States<br />

Aldehyde Dehydrogenases (ALDHs) oxidize aldehydes to their corresponding carboxylic acids and<br />

require a cofactor, usually nicotinamide adenine dinucleotide (NAD). The catalytic cycle is initiated by<br />

nucleophilic attack by a cysteine residue on aldehyde substrates to form a thiohemiacetal<br />

intermediate. The traditional mechanism for this step fails to mention even the possibility of proton<br />

transfer to this intermediate. Our past, published results employing hybrid PM3/OPLS umbrellasampling<br />

simulations suggest a novel enzyme mechanism where a proton transfers from a main chain<br />

amide to the oxyanion thiohemiacetal. We hypothesize that proton transfer stabilizes the intermediate<br />

protecting it from forming a “dead-end” cysteine-NAD complex. Our prediction of cysteine-NAD<br />

complex formation in ALDH was confirmed by re-examination of already deposited electron density<br />

maps.<br />

In this presentation, we will describe our strategy for re-parameterizing the semiempirical molecular<br />

orbital method AM1 to calculate, with considerably improved accuracy, various intermolecular<br />

interactions and reactions important for simulating Aldehyde Dehydrogenase (ALDH) chemistry. The<br />

“fitness” of a particular AM1 parameter set was evaluated by comparison to properties collected from<br />

either published experimental results or high-level Quantum Mechanical (QM) calculations. This<br />

reference data includes geometries, dipole moments, ionization potentials, heats of formation and<br />

reaction energies. Various non-linear search algorithms were then employed to search for parameter<br />

sets that best reproduce the reference data using a modified version of the DYNAMO library<br />

(www.pdynamo.org). The use of these parameters in QM/MM simulations to calculate the free energy<br />

profiles/surfaces of reactions in ALDH facilitates examination of the hydride transfer step. The<br />

simulations will test the hypothesis that two mutations in two separate but related ALDHs affect the<br />

essential protonation of the thiohemiacetal intermediate. These two mutations are responsible for two<br />

metabolic diseases, Sjorgren-Larsson syndrome and Hyperprolinemia Type II.<br />

PP368<br />

Simulation Studies of the Folding and Aggregation of Model Amyloid Peptides in Solution and<br />

at an Interface<br />

Volker Knecht, Madeleine Kittner, Reinhard Lipowsky<br />

Max Planck Institute of Colloids and Interfaces, Theory & Bio-Systems Department, Potsdam,<br />

Germany<br />

The development of therapeutic treatments against amyloid diseases requires an understanding of the<br />

(mis)folding and aggregation of fibrillogenic species at a microscopic level. To study these species in<br />

atomic detail experimentally is difficult due to their tendency to aggregate and the transient nature of<br />

early oligomers. Therefore, an indispensable tool to study these systems is provided by computer<br />

simulations. The most accurate treatment consists of a full atomistic description of peptides and<br />

solvent environment. We present molecular dynamics (MD) simulations of the folding or aggregation of<br />

various model amyloid peptides in explicit water and at a water/vapor interface. As shown in Fig. 1, the<br />

18-residue amyloid peptide B18 derived from the sea urchin fertilization protein Bindin was observed<br />

to undergo a transition from beta-sheet/coil conformations in water to partially alpha-helical<br />

conformations at a water/vapor interface. Possible pathways for alpha-beta transformations in solution<br />

and beta-alpha transformations at the interface were suggested [1].<br />

In contrast, a 12-residue amyloid peptide derived from a viral protein and denoted as LSFD was found<br />

to adopt the same type II’ beta-hairpin conformation in equilibrium with more disordered but also U-<br />

shaped conformations both in water and at a water/vapor interface [2]. This suggests that beta-sheet<br />

conformations suggested from previous spectroscopic measurements on LSFD immediately after<br />

preparation of the peptide solution may not arise from the presence of protofilaments as speculated<br />

previously, but, rather, indicate a property of monomers. In addition, combined with previous results<br />

from x-ray scattering, our findings suggest that interfacial aggregation of LSFD implies a transition<br />

from U-shaped to extended peptide conformations. We directly observed such U-shaped to extended<br />

transitions during replica exchange simulations of the dimerization of the (25-35) fragment of the<br />

Alzheimer A beta peptide in water. Simulations of a monolayer consisting of peptides with sequence<br />

G(VT)5 adopting extended intermolecular beta-sheets revealed a strong bending of the beta-sheets at<br />

the termini. Taken together, our results give insights to the properties of fibrillogenic or beta-sheet<br />

forming peptides at an atomic level important from a biophysical, medical, and nanotechnological point<br />

of view.<br />

FIG. 1: Conformational polymorphism of<br />

model amyloid peptide B18 in different<br />

environments from all-atom simulations<br />

[1]. The initial (left) and typical<br />

configurations in explicit water (middle) or<br />

a water/vapor interface (right) are shown.<br />

[1] Knecht, V., Möhwald, H., and Lipowsky, R., J. Phys. Chem. B 2007, 111, 4161-4170.<br />

[2] Knecht V., J. Phys. Chem. B , accepted.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP369<br />

Electrophoretic Mobility Does Not always Reflect the Charge on a Particle<br />

Volker Knecht, H. Jelger Risselada, Alan E. Mark, Siewert-Jan Marrink<br />

1 Max Planck Institute of Colloids and Interfaces, Theory & Bio-Systems, Potsdam, Germany,<br />

2 Groningen Biomolecular Sciences and Biotechnology Institute, and Zernike Institute for Advanced<br />

Materials, <strong>University</strong> of Groningen, Groningen, Netherlands, 3 School of Molecular and Microbial<br />

Sciences, <strong>University</strong> of Queensland, Brisbane, Australia, 4 Groningen Biomolecular Sciences and<br />

Biotechnology Institute, and Zernike Institute for Advanced Materials, <strong>University</strong> of Groningen,<br />

Groningen, Netherlands<br />

Electrophoresis is widely used to determine the electrostatic potential of colloidal particles.<br />

Hydrophobic particles such as oil droplets or air bubbles in water above pH 3 show negative<br />

electrophoretic mobilities. This is commonly attributed to a negative surface charge due to the<br />

adsorption of hydroxide ions. This explanation, however, has been challenged recently by<br />

spectroscopic experiments and atomistic simulations indicating that hydrophobic surfaces, rather,<br />

adsorb hydronium but (weakly) repel hydroxide, thus being positively charged. Alternative<br />

explanations for negative electrophoretic mobilities of hydrophobic particles are anionic impurities at<br />

the interface. Here we present molecular dynamics simulations of oil droplets in water in the presence<br />

of an external electric field but in the absence of any ions [1]. The simulations reproduce the negative<br />

sign and the order of magnitude of the oil droplet mobilities at the point of zero charge in experiment.<br />

The electrostatic potential in the oil with respect to the water phase, induced by anisotropic dipole<br />

orientation at the interface, is positive. Our results suggest that, for hydrophobic interfaces, the effect<br />

of hydronium causing positive electrophoretic mobility is overcompensated by negative mobility arising<br />

from dipolar ordering. More generally, our findings suggest that electrophoretic mobility does not<br />

always reflect the net charge or electrostatic potential of colloidal particles, and that a fundamental<br />

understanding of electrokinetic phenomena requires a molecular description.<br />

PP370<br />

A Multilevel Sidechain Representation Library for Protein Structure Prediction and Docking.<br />

Quentin Kaas<br />

Institute for Molecular Bioscience, <strong>University</strong> of Queensland, Brisbane, Queensland, Australia<br />

Protein-protein docking and ab-initio structure prediction are nowadays the more challenging area of<br />

research in structural bioinformatics. The major problem of protein docking is to account for protein<br />

flexibility [1]. It is often done by using the soft docking method which allows a certain level of<br />

entanglement between the two docked structures [2]. On the other hand ab-initio structure prediction<br />

(loop or whole protein modelling) often uses simplified sidechain representations, with a reduced<br />

number of interaction centres, for faster computation and to comply with low resolution modelling [3].<br />

I have developed a new method called SCHISMo that combines and extends those approaches<br />

thanks to a multilevel library of sidechain representations. The first level represents sidechains with<br />

two interaction centres, the second level uses interaction centres corresponding to a single atom or to<br />

a group of atoms and the third level corresponds to all the heavy atoms. The levels are implemented<br />

hierarchically which allows to pass from a simple to a more complex representation, or the reverse.<br />

The flexibility of the sidechains are modelled by mapping the positions of the interaction centres on a<br />

circle or on a sphere. The sidechain conformations are predicted by considering conformation<br />

frequency in a set of experimental structures and by considering charge and hydrophobic interactions.<br />

A C library allows an easy implementation of SCHISMo in other modeling programs that will be able to<br />

progressively complexify sidechain representations as a model is refined. Finally SCHISMo can build<br />

new simplified representation libraries with different centre definitions and train this library on a<br />

different set of experimental structures, for example corresponding to a specific class of proteins.<br />

FIG. 1: Electrophoresis of oil droplets in water in the absence of ions in molecular dynamics simulations. Left:<br />

(Periodic) simulation box (black frame) with direction of the external electric field. Middle: Heptane droplet of<br />

diameter 9 nm (dark gray) in water (white) moving in negative field direction. Right top: Heptane-water interface.<br />

Right bottom: Single heptane and water molecule. CH 3 and CH 2 groups in heptane were treated as compound<br />

atoms. For heptane, covalent bonds are depicted as white lines, and one of the bond torsion angles is indicated.<br />

For water, partial charges on oxygen and hydrogen atoms are indicated.<br />

[1] Lensink, M. F.; Méndez, R.; Wodak, S.J. Proteins 2007 69, 704-718<br />

[2] Fernandez-Recio, J.; Totrov, M.; Abagyan, R.; Protein Science 2002 11, 280-291<br />

[3] Maupetit, J.; Tuffery, P.; Derreumaux, P.; Proteins 2007 69, 394-408<br />

[1] Knecht, V; Rissela, H.J.; Mark, A.E.; and Marrink, S.J; J. Col. Int. Sc. 2008, 318, 477-486.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP371<br />

In Quest of an Efficient and Accurate Modelling of the Photochemistry of Biological<br />

Photoreceptors: A QM/MM Approach.<br />

Pedro B. Coto, Israel González-Ramírez, Gloria Olaso-González, Daniel Roca-Sanjuán, Juan José<br />

Serrano-Pérez, Manuela Merchán<br />

Instituto de Ciencia Molecular (ICMOL), <strong>University</strong> of Valencia, Valencia, Spain<br />

Light-driven chemical reactions play a key role in different biological processes fundamental for life.<br />

Photoreceptors such as rhodopsins, phytochromes and xanthopsins use the energy of<br />

electromagnetic radiation to unleash complex biochemical processes like vision [1], photosynthesis<br />

[2], and phototaxis [3]. The use of hybrid Quantum Mechanics/Molecular Mechanics (QM/MM) [4]<br />

methods allows tackling the treatment of large biological systems in an efficient and accurate way. In<br />

particular, the CASPT2//CASSCF/Forcefield protocol is able to give a balanced treatment of the<br />

different electronic states involved in the photochemical process by including both dynamic and static<br />

correlation effects. We have applied this scheme to the analysis of the photochemistry of Rhodopsin<br />

[5,6] and PYP [7]. The different factors controlling the wavelength absorption maxima are discussed.<br />

PP372<br />

Efficient Extrapolation of Triple Excitations to the Complete Basis Set Limit<br />

Ericka Barnes, George Petersson<br />

Wesleyan <strong>University</strong>, Chemistry Department, Middletown, CT, United States<br />

Compound model chemistry methods (e.g. G3, CBS-QB3, and W1) designed to approximate<br />

CCSD(T)/CBS energies are limited by the cost of the non-iterative perturbative contribution of triple<br />

excitations. This component scales as O 3 V 4 and thus will ultimately determine the range of<br />

applicability of any such method. It is therefore imperative to make every effort to minimize the basis<br />

sets used for this component. There is little to be gained from reducing the error in the approximation<br />

of the CCSD(T)/CBS limit much below the error in the CCSD(T) energy relative to the FCI energy,<br />

which defines the limiting accuracy possible within this framework. We therefore sought the smallest<br />

basis sets for which the rms deviation in our extrapolated triple excitation component falls below this<br />

inherent error in the CCSD(T) energy (0.5 mE h for our test set). We find that a linear extrapolation of<br />

3s2p (double-ζ without diffuse or polarization functions) and 5s4p1d (triple-ζ augmented with diffuse<br />

valence functions and a single set of polarization functions for first-row atoms) satisfies our criterion.<br />

This extrapolation reduces the rms error (compared to the CBS limit of the perturbative triple excitation<br />

component) below 0.4 mE h for our test set of 57 bond dissociation energies, ionization potentials, and<br />

electron affinities. The combination of this (T) extrapolation with CCSD extrapolations employing larger<br />

basis sets is comparable to the overall accuracy of W1 theory (with triples extrapolated from 4s3p2d<br />

and 5s4p3d2f basis sets for first-row atoms) with up to two orders-of-magnitude advantage in speed.<br />

[1] Wald G. Nobel Lecture 1967.<br />

[2] Quail, P. H.; Boylan M. T.; Parks, B. M.; Xu, Y.; Wagner, D. Science 1995, 268, 675-680.<br />

[3] Sprenger, W. W.; Hoff, W. D.; Armitage, J. P.; Hellingwerf, K. J. J. Bacteriol. 1993, 175, 3096-3104.<br />

[4] Warshel, A.; Levitt, M. J. Mol. Biol. 1976, 103, 227-249.<br />

[5] Coto, P. B.; Strambi, A.; Ferré, N.; Olivucci, M. Proc. Natl. Acad. Sci. (USA) 2006, 103, 17154-17159.<br />

[6] Strambi, A.; Coto, P. B.; Frutos, L. M.; Ferré, N.; Olivucci, M. J. Am. Chem. Soc. 2008, 130, 3382-3388.<br />

[7] Coto, P. B.; Martí, S.; Oliva, M.; Olivucci, M.; Merchán, M.; Andrés, J. J. Phys. Chem. B. doi:<br />

10.1021/jp711396b.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP373<br />

Decomposition of 6,7,8-Trioxybicyclo[3.2.2]nonane Prompted by Fe(II). A Model to Study the<br />

Fisrt Stages of the Decomposition Mechanism of Artemisinin<br />

Pamela Moles, V. Sixte Safont, Mónica Oliva<br />

Universitat Jaume I, Departament de Química Física i Analítica, Castelló de la Plana, Spain<br />

Artemisinin is the most effective drug against malaria at hand nowadays. The endoperoxide group that<br />

Artemisinin possesses has been proven to be critical for its antimalarial activity.<br />

By using the 6,7,8-trioxybicyclo[3.2.2]nonane as Artemisinin model, and the Fe(II)[H 2 O] 2 [OH - ] 2 as<br />

activator agent, we present a theoretical investigation of the decomposition mechanism of Artemisinin.<br />

The first step involves the reductive cleavage of the peroxidic bond prompted by Fe (II), leading to the<br />

formation of oxygen-centered radicals which, in turn, generate carbon-centered radicals. These can<br />

participate in a series of chemical reactions and form other kind of intermediates, one or more of which<br />

could kill the malaria parasite. The molecular structures of the free radicals and the neutral species<br />

that take part in the reactions have been calculated with the B3LYP method at 6-311+G** level. In<br />

addition an electron localization function (ELF) based study has been carried out to have a topological<br />

description of the bonds being broken and formed in the reported steps.<br />

H 3 C<br />

O<br />

O<br />

O<br />

H<br />

O<br />

H 3 C<br />

O<br />

H<br />

CH 3<br />

O<br />

O<br />

O<br />

PP374<br />

Solvation of Platinum Chloro-Complexes in 1,3-Dialkylimidazolium Ionic Liquids<br />

Gerhard A Venter, Kevin J Naidoo<br />

Department of Chemistry, <strong>University</strong> of Cape Town, Rondebosch, Western Cape, South Africa<br />

Room Temperature Ionic Liquids (RTILs) are fast becoming synonymous with "Green Chemistry".<br />

Their qualities of possessing near-negligible vapour pressure and high thermal stability are key in this<br />

regard—but it is their degree of tunability, afforded by mixing cation/anion combinations and varying<br />

individual molecular structure, that designates them "designer solvents" [1].<br />

The solvation behaviour and structure of transition metal systems in ionic liquids are of direct concern<br />

in determining catalytic reaction pathways, even more so where charged precursors or intermediate<br />

species are present. The effect of strong ionic association on reaction rates has been illustrated [1].<br />

The most popular class of ionic liquids are those consisting of imidazolium cations, in particular, the 1-<br />

alkyl-3-methylimidazolium cations. We have simulated the solvation of hexachloroplatinate(IV) in<br />

cations of this type, using a variety of different anions. In doing so we have developed a transferable<br />

CHARMM-type forcefield for the 1,3-dialkylimidazolium cations that is also consistent with our platinum<br />

group metal forcefield [2]. Whereas other popular CHARMM-type RTIL forcefields depend on charges<br />

being calculated as needed for each new ion under investigation [3], we opted for a more general<br />

description favouring transferability [4]. The parameterization strategy outlined by Mackerell and<br />

Foloppe was used as a guide [5].<br />

Other than comparison with experimental properties such as densities, diffusion coefficients and pair<br />

distribution functions (where available) as validation of the forcefield, we also present spatial<br />

distribution functions (SDF) of the solvation structure of [PtCl 6 ] 2- in 1,3-dialkylimidazolium chloride.<br />

2-<br />

Cl<br />

Cl Cl<br />

CH 2<br />

CH 2<br />

Pt<br />

H 3 C m N N n<br />

Cl<br />

CH 3<br />

Cl Cl<br />

Cl<br />

Figure. Artemisinin and 6,7,8-trioxybicyclo[3.2.2]nonane<br />

[1] J. D. Gu, K. X. Chen, H. L. Jiang, J. Leszczynski, Journal of Physical Chemistry A. 1999, 103, 9364.<br />

[2] J. Gu, K. Chen, H. Jiang, J. Leszczynski, Journal of Molecular Structure-Theochem 1999, 491, 57.<br />

[3] P. L. Olliaro, R. K. Haynes, B. Meunier, Y. Yuthavong, Trends in Parasitology 2001, 17, 122.<br />

[4] M. G. B. Drew, J. Metcalfe, M. J. Dascombe, F. M. D. Ismail, Journal of Medicinal Chemistry 2006, 49, 6065.<br />

[5] M. G. B. Drew, J. Metcalfe, F. M. D. Ismail, Journal of Molecular Structure-Theochem 2005, 756, 87.<br />

[6] S. Tonmunphean, V. Parasuk, S. Kokpol, Journal of Molecular Structure-Theochem 2005, 724, 99.<br />

[7] J. Queiroz, J. Walkimar, M. Teixeira, F. H. Andrade, A. Gutterres, Bioorganic and Medicinal Chemistry 2008,<br />

16, 5021.<br />

[1] (a) Welton, T. Coord. Chem. Rev. 2004, 248, 2459–2477. (b) Welton, T. Chem. Rev. 1999, 99, 2071–2083.<br />

[2] Lienke, A.; Klatt, G.; Robinson, D. J.; Koch, K. R.; Naidoo, K. J. Inorg. Chem. 2001, 40, 2352–2357<br />

[3] Maginn, E. J. Acc. Chem. Res. 2007, 40, 1200–1207.<br />

[4] Pádua, A. A. H.; Costa Gomes, M. F.; Canongia Lopes, J. N. A. Acc. Chem. Res. 2007, 40, 1087-1096.<br />

[5] Foloppe, N.; MacKerell, Jr., A. D. J. Comp. Chem. 2000, 21, 86–104.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP375<br />

Mechanistic Analysis of Intermolecular Arene C-H Activation<br />

Stuart A. Macgregor, David L. Davies, Amalia I. Poblador-Bahamonde<br />

Heriot-Watt <strong>University</strong>, Engineering and Physical Sciences (EPS), Edinburgh, United Kingdom<br />

The intermolecular C-H activation of benzene by the model complex [IrCp(PH 3 )AcO] + has been<br />

studied by density functional calculations. Three possible mechanisms have been considered:<br />

oxidative addition; σ–bond methathesis; and electrophilic activation [1]. Our results indicate<br />

electrophilic activation to be the most promising process.<br />

H 3 P<br />

Ir<br />

O<br />

O C CH3<br />

H 3 P<br />

Ir<br />

O<br />

O<br />

C CH3<br />

H 3 P<br />

Ir<br />

O<br />

OH<br />

C CH3<br />

PP376<br />

DFT Study on Charge Conductivity of DNA-Wrapped Carbon Nanotubes<br />

Noriyuki Kurita 1 , Ikuyo Komura 1 , Takayuki Tsukamoto 1 , Yasuyuki Ishikawa 2<br />

1 Toyohashi <strong>University</strong> of Technology, Toyohashi, Aichi, Japan, 2 <strong>University</strong> of Puerto Rico, San Juan,<br />

PR, United States<br />

1. Introduction<br />

Recently, Zheng et al. [1] reported that single-walled carbon nanotubes (SWNTs) can be wrapped by<br />

single-stranded DNAs (ssDNA), and that the ssDNA-SWNT complexes formed can be used in<br />

chromatographic techniques to allow the separation of SWNTs by length, conductivity type and<br />

diameter. In the present study, we investigated the stable structures, electronic and charge-conductive<br />

properties of the ssDNA-wrapped SWNTs by molecular simulations based on classical molecular<br />

mechanics (MM) and ab initio molecular orbital (MO) methods. Because the previous experiment [1]<br />

indicated that the (GT) 14 ssDNA wraps SWNTs, we employed the ssDNAs with the sequence 5’-<br />

d(GTGTGT)-3’ and 5’-d(GGCC)-3’, while the semiconducting (10,0) and metallic (6,6) SWNTs with 30<br />

Å in length were employed, in order to elucidate the difference in the effect of ssDNA wrapping on the<br />

electronic properties between the semiconducting and metallic SWNTs.<br />

base displacement H-transfer<br />

Electrophilic activation is computed to proceed by a two-step mechanism, where the initial<br />

displacement of acetate is rate determining. Replacing acetate with a more weakly coordinating<br />

species such as triflate therefore reduces the barrier to C-H activation. Comparisons with other<br />

systems show that the presence of an intramolecular chelating base greatly facilitates C-H activation.<br />

2. Method of calculations<br />

A standard B-form ssDNA was docked to SWNT by using the automated ligand-docking program<br />

AutoDock3. A total of 100 candidate structures for ssDNA-SWNT were produced and fully optimized<br />

by the AMBER-MM method. In density-functional theory calculations using the Dmol³ program, the<br />

electronic properties of these ssDNA-SWNTs were examined to elucidate the effect of ssDNA<br />

wrapping on the energy levels and spatial distributions of MOs, including the HOMO and LUMO.<br />

Based on the electronic properties, charge conductive properties of the ssDNA-SWNTs as well as<br />

SWNTs were probed via the charge-conductive analysis algorithm developed by Meunier et al. [2].<br />

[1] Davies D. L., Donald S.M.A., Al-Duaij O., Macgregor S. A., Pölleth M., J. Am. Chem. Soc., 2006, 128, 4210.<br />

3. Results and discussion<br />

The GGCC ssDNA can wrap both the (10,0) and (6,6) SWNTs, whereas the (GT) 3 ssDNA wraps only<br />

the (6,6) SWNT. In the GGCC-ssDNA+SWNT complexes, the backbone of ssDNA wraps the SWNTs<br />

in a helical form, while the bases of ssDNA are stacked onto the surface of the SWNTs, as suggested<br />

by the previous molecular modeling study [1]. Therefore, it seems that the bases of the GGCC ssDNA<br />

contribute to the wrapping of SWNTs.<br />

For the most stable structures of the (GT) 3 ssDNA+SWNTs complexes, we investigated the binding<br />

energies between ssDNA and SWNT to find that the (GT) 3 ssDNA can bind more strongly to the<br />

metallic (6,6) SWNT than (10,0) SWNT. In the (GT) 3 ssDNA+(6,6) SWNT complex, some stacking<br />

interactions between the bases of ssDNA and the surface of (6,6) SWNT stabilize the complex. On the<br />

other hand, the (GT) 3 ssDNA cannot wrap the (10,0) SWNT, resulting in the weak binding between<br />

(GT) 3 ssDNA and (10,0) SWNT.<br />

The HOMO and LUMO of the (GT) 3 ssDNA+(10,0) SWNT are localized on both edges of the (10,0)<br />

SWNT, whereas they are distributed over the (6,6) SWNT in the (GT) 3 ssDNA+(6,6) SWNT. Therefore,<br />

the effects of the (GT) 3 ssDNA-wrapping on the electronic properties of the (10,0) and (6,6) SWNTs<br />

differ significantly. The results on the charge conductive properties of ssDNA+SWNTs will be shown in<br />

the poster.<br />

[1] Zheng, M. et al. Nature Materials 2003, 2, 338. [2] Meunier, V. et al. J. Chem. Phys. 2005, 123, 024705.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP377<br />

Specific Interactions between Thermolysin and Dipeptide Ligands Obtained by Fragment<br />

Molecular Orbital Calculations<br />

Noriyuki Kurita 1 , Kenichi Dedachi 1 , Mahmud T. H. Khan 2 , Ingebrigt Sylte 2<br />

1 Toyohashi <strong>University</strong> of Technology, Toyohashi, Aichi, Japan, 2 <strong>University</strong> of Tromsø, Tromsø,<br />

Norway<br />

1. Introduction<br />

Thermolysin (TMN) is one of the widely studied zinc metalloproteases from Bacillus<br />

stearothermophilus. Especially, the hydrolysis mechanism of TMN has been theoretically studied for<br />

over a decade. Recent experimental studies indicated that the activity of TMN is largely dependent on<br />

the types of ligand (dipeptide) molecules. In the present study, we investigated the specific<br />

interactions between TMN and some kinds of dipeptides by classical molecular mechanics (MM) and<br />

multi-layer fragment molecular orbital (MLFMO) methods.<br />

2. Method of calculations<br />

We have obtained the X-ray structures of the complexes with TMN and IY (Ile-Tyr) or LW (Leu-Trp).<br />

These X-ray structures have about 350 crystal water molecules, although the positions of hydrogen<br />

atoms are not determined. We then added hydrogen atoms to these structures and optimized the<br />

positions of only hydrogen atoms by the MM method based on AMBER force field, with considering<br />

the crystal water molecules explicitly.<br />

PP378<br />

Characterization of Weak Interactions between Aromatic Amino Acids and the Natural<br />

Nucleobases<br />

Lesley Rutledge, Holly Durst, Stacey Wetmore<br />

<strong>University</strong> of Lethbridge, Department of Chemistry and Biochemistry, Lethbridge, Alberta, Canada<br />

Interactions between DNA and protein building blocks have been proven to be very important in<br />

nature, where they dictate biomolecular structure and are essential in many biological processes. For<br />

example, DNA replication and transcription rely on protein–DNA interactions. ‘Weak’ noncovalent<br />

interactions, such as hydrogen bonding and stacking have been studied extensively. However, crystal<br />

structures reveal that there are also T-shaped interactions in enzyme active sites, which were<br />

previously believed to be even weaker than stacking interactions. This study uses computational<br />

chemistry to characterize the gas-phase stacking and T-shaped interactions between aromatic amino<br />

acids (HIS, PHE, TYR, TRP) and the natural nucleobases (A, G, C, T). The MP2/6-31G*(0.25)<br />

potential energy surfaces of the nucleobase–amino acid dimers were calculated as a function of<br />

several geometrical variables and higher-level calculations were performed to determine the most<br />

accurate binding strengths possible. Our calculations provide valuable information about the nature<br />

and magnitude of these weak interactions, and our results suggest that these interactions play a more<br />

important role in biochemistry than initially suspected.<br />

The electronic properties of the optimized TMN+dipeptide structures with crystal water molecules<br />

were investigated by using the MLFMO method [1,2]. In this calculation, dipeptide, Zn and amino acids<br />

of TMN existing within a 5.0 Å distance from the dipeptide and Zn were treated by MP2/6-31G(d,p),<br />

while the other amino acids of TMN were treated by HF/6-31G(d,p). The crystal water molecules were<br />

treated in the same way. It is noted that Zn ion was included in the same fragment of Glu166, which<br />

exists near to Zn ion. From the comparison of interaction energies between the amino acids or water<br />

molecules and dipeptide, we attempted to elucidate which amino acids or water molecules are<br />

important for the specific interactions between TMN and dipeptide in electronic level.<br />

3. Results and discussion<br />

The obtained interaction energies between the amino acids of TMN or water molecules and IY indicate<br />

that Asn112, Arg203 and Glu143 amino acids have larger interaction energies. It is thus expected that<br />

these amino acids of TMN mainly contribute to the specific interactions between TMN and IY. In<br />

addition, it is elucidated that one unique water molecule has large (-13.23 kcal/mol) interaction energy<br />

with IY. This water molecule is directly hydrogen bonded to the NH 2 terminal of Ile of IY peptide and<br />

bridges between IY, Zn and Glu166 of TMN. Therefore, it can be concluded that this water molecule<br />

can contribute significantly to the specific interactions between TMN, Zn and IY. On the other hand,<br />

there is no such water molecule in the complex of TMN+LW, so that the binding energy between TMN<br />

and IY is smaller than that between TMN and LW. At the conference, we will also present the results<br />

for the complexes with TMN and other dipeptides.<br />

[1] Fedorov, D. G. et al., J. Phys. Chem. A 2005, 109, 2638.<br />

[2] Mochizuki, Y. et al., Chem. Phys. Lett. 2005, 410, 247.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP379<br />

DFT Study of the Manganese Containing Ribonucleotide Reductase in Chlamydia Trachomatis<br />

Katarina Roos, Per E.M. Siegbahn<br />

Department of Physics, Stockholm <strong>University</strong>, Stockholm, Sweden<br />

Ribonucleotide reductase (RNR) catalyses the reduction of ribonucleotides to deoxyribonucleotides,<br />

the building blocks for DNA synthesis. In the pre-activating step a stable tyrosyl radical needed for<br />

catalysis is formed by oxygen cleavage at a di-iron site. The human pathogen Chlamydia trachomatis<br />

(Ct) RNR lacks the tyrosine crucial for activity in conventional RNR. Instead the active cofactor is a<br />

manganese(IV)/iron(III) metal center. An explanation is suggested to why the enzyme needs this<br />

manganese instead of a second iron to perform the same chemistry as with the tyrosyl radical. In<br />

normal class I RNR (e.g. from Escherichia coli) compound X has a similar oxidation state with<br />

iron(IV)/iron(III). This state precedes the tyrosyl radical and has not yet been structurally determined.<br />

In this study DFT is used to show why it is crucial to have manganese(IV) in the active center of Ct R2<br />

and that switching to iron(IV) would probably generate an inactive complex. The quantum chemical<br />

calculations performed explain experimental observations of Ct RNR and give guidance for future<br />

experiments.<br />

PP380<br />

Computational Studies of the Isomerisation of Nido- and Closo- 12-vertex Carboranes<br />

Stuart A. Macgregor, David McKay, Alan J. Welch<br />

Heriot-Watt <strong>University</strong>, School of Engineering and Physical Sciences, Edinburgh, United Kingdom<br />

The production of supraicosahedral carboranes requires the reduction of a closo 12-vertex species to<br />

a nido fragment followed by capitation. We have used computational methods to study the reduction<br />

of closo-C 2 B 10 H 12 species. The reduction of para-carborane is particularly complex and produces five<br />

different nido dianionic minima. We present the full characterisation of the initial products of reduction<br />

and the subsequent isomerisation pathways to all five nido species.<br />

C<br />

C<br />

para-carborane<br />

REDUCTION<br />

CAPITATION<br />

{ML n } 2+<br />

C<br />

ML n<br />

C<br />

13-vertex cl oso ruthenacarborane<br />

C<br />

C<br />

para-carborane<br />

RED<br />

2-<br />

2-<br />

2-<br />

2-<br />

2-<br />

C<br />

C<br />

C<br />

C<br />

C C<br />

C<br />

+<br />

+ +<br />

+<br />

C<br />

C<br />

C<br />

[1,7-nido-C 2 B 10 H 12 ] 2- [3,7-nido-C 2 B 10 H 12 ] 2- [4,7-nido-C 2 B 10 H 12 ] 2- [7,9-nido-C 2 B 10 H 12 ] 2- [7,10-nido-C 2 B 10 H 12 ] 2-<br />

These results agree with experimental studies which have shown that reduction and capitation of<br />

para-carborane produces 13-vertex ruthenacarboranes derived from these nido intermediates [1].<br />

Additional computational studies show the 2e reduction of ortho- and meta-carborane produces<br />

[7,9-nido-C 2 B 10 H 12 ] 2− , in agreement with experimental studies [2]. Reoxidation then gives a new<br />

species shown to be an intermediate in a new low energy route for isomerisation of ortho- to metacarborane.<br />

C<br />

C<br />

C<br />

C<br />

ortho-carborane<br />

meta-carborane<br />

R<br />

O<br />

C<br />

C<br />

R<br />

[7,9-nido-C 2 B 10 H 12 ] 2-<br />

[1] Zlatogorsky, S.; Edie, M. J.; Ellis, D.; Erhardt, S.; Lopez, M. E.; Macgregor, S. A.; Rosair, G. M.; Welch, A. J.<br />

Angew. Chem. Int. Ed. 2007, 46, 6706.<br />

[2] (a) Dunks, G. B.; Wiersema, R. J.; Hawthorne, M. F. J. Am. Chem. Soc. 1973, 95, 3174. (b) Zakharkin, L. I.;<br />

Kalinin, V. N.; Podvisotskaya, L. S. Bull. Acad. Sci. USSR, Div. Chem. Sci. 1966, 1444.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP381<br />

Short Intramolecular Hydrogen Bonds: Derivatives of Malonaldehyde with Symmetrical<br />

Substituents<br />

Jacqueline Hargis, Francesco Evangelista, Justin Ingels, Henry Schaefer<br />

Center of Computational Chemistry, <strong>University</strong> of Georgia, Athens, GA, United States<br />

A systematic study of various derivatives of malonaldehyde has been carried out to explore very short<br />

hydrogen bonds (r OO < 2.450 Å). Various electron withdrawing groups, including CN, NO 2 , and BH 2<br />

have been attached to the central carbon atom, C 2 . To C 1 and C 3 , strong electron donors and/or steric<br />

hindered substituents were used to strengthen the intramolecular hydrogen bond, including but not<br />

limited to NH 2 , N(CH 3 ) 2 , and C(CH 3 ) 3 . Six molecules (Figure 2) were found to have extremely short<br />

intramolecular hydrogen bonds.<br />

The chemical systems investigated are intriguing due to their low energetic barrier for the<br />

intramolecular hydrogen atom transfers. Energy barriers were predicted using correlated methods<br />

including second-order perturbation theory and coupled cluster theory in conjunction with the Dunning<br />

hierarchy of correlation consistent basis sets, cc-pVXZ (X=D, T, Q, 5). Focal point analyse allowed for<br />

the barriers to be evaluated at the CBS limit including core correlation and zero-point vibrational<br />

energy corrections.<br />

B3LYP energies are benchmarked against highly accurate correlated energies for intramolecular<br />

hydrogen bonded systems. The focal point extrapolated value including coupled cluster full triple<br />

excitation contributions give a hydrogen transfer barrier for malonaldehyde of 3.9 kcal mol -1 . We also<br />

describe two compounds with extremely low barriers; nitromalonamide (0.44 kcal mol -1 ) and<br />

2-borylmalonamide (0.62 kcal mol -1 ). An empirical relationship was drawn between the underestimated<br />

B3LYP energetic barriers and the predicted barriers at the CBS limit. By relating these two quantities,<br />

barrier heights for larger systems may be estimated for systems that possess too many atoms to use<br />

highly correlated electronic structure methods.<br />

PP382<br />

Structural Effects of DNA Modification at the C8 site of Purine Nucleobases<br />

Andrea Millen 1 , Cassandra Churchill 1 , Lex Navarro-Whyte 1 , Jenny Shim 1 , Katie Schlitt 2 , Chris<br />

McLaughlin 2 , Richard Manderville 2 , Stacey Wetmore 1<br />

1 Department of Chemistry & Biochemistry, <strong>University</strong> of Lethbridge, Lethbridge, Alberta, Canada,<br />

2 Departments of Chemistry & Toxicology, <strong>University</strong> of Guelph, Guelph, Ontario, Canada<br />

Modification at the C8 site of the purines can occur when molecules such as phenols are oxidized into<br />

phenoxyl radicals by peroxidase enzymes or redox-active transition metals. These C-bonded DNA<br />

adducts can lead to abasic site formation, and thus increase the vulnerability of the cell to the<br />

carcinogenic process [1,2]. Conversely, the unique properties of DNA modified in this way can be<br />

exploited, where for example the structures can be used as bioprobes [3]. Our research uses<br />

computational methods to understand the structure and reactivity of these modified nucleobases [4].<br />

R O<br />

2<br />

R<br />

12 11 7 6<br />

2<br />

N<br />

5<br />

NH<br />

N<br />

13<br />

R 1 1 R<br />

10 8<br />

1<br />

14 θ N 4<br />

15<br />

2<br />

9 N NH N<br />

HO<br />

3 2<br />

HO<br />

5′ χ<br />

O<br />

O<br />

4′<br />

3′<br />

OH<br />

1′<br />

2′<br />

OH<br />

NH 2<br />

N<br />

N<br />

O<br />

O<br />

1a 7 6<br />

N 5<br />

2a 7 6<br />

X<br />

8 NH<br />

N 5<br />

5a 2a<br />

1a X 3a<br />

8 NH<br />

1<br />

1<br />

4a<br />

9<br />

5a<br />

9<br />

HO 3a θ N 4 N 2<br />

4a<br />

NH 2 HO θ N 4 N 2<br />

5'<br />

NH 2<br />

5' 3<br />

O χ<br />

3<br />

O χ<br />

4'<br />

1'<br />

4'<br />

1'<br />

3' 2'<br />

3' 2'<br />

OH X=O,NH,orS OH<br />

Scheme 1 Scheme 2<br />

Experimental work has proposed that the preferred conformation of ortho phenoxyl purine adducts<br />

(Scheme 1, R 2 =OH, R 1 =H) fluctuates between twisted and planar structures depending on the solvent,<br />

while the para adducts (Scheme 1, R 2 =H, R 1 =OH) adopt only twisted conformations. To better<br />

understand the structures of these adducts, the gas-phase potential energy surfaces of the<br />

deoxyguanosine and deoxyadenosine ortho and para carbon-bonded adducts (Scheme 1) were<br />

systematically studied using DFT by varying the orientation about the glycosidic bond (χ) and phenoxyl<br />

substituent (θ) for two orientations of the C5′-hydroxyl group. The effects of various phenyl<br />

substituents on the adduct structure and the barrier for glycosidic bond cleavage have also been<br />

performed for comparison to experimental deglycosylation studies. Additionally, the structures of a<br />

series of five-membered ring adducts (Scheme 2) with interesting fluorescent properties that could be<br />

used as bioprobes have been studied for their conformational properties. This poster will summarize<br />

some of our important findings about the structures of these important DNA nucleosides and the<br />

implications of these findings for their biological activity and use as bioprobes.<br />

[1] Manderville, R. A. Can. J. Chem.-Rev. Can. Chim. 2005, 83, 1261-1267.<br />

[2] McLaughlin, C. K.; Lantero, D. R.; Manderville, R. A. J. Phys. Chem. A 2006, 110, 6224-6230.<br />

[3] Sun, K. M.; McLaughlin, C. K.; Lantero, D. R.; Manderville, R. A. J. Am. Chem. Soc. 2007, 129, 1894-1895.<br />

[4] Millen, A. L.; McLaughlin, C. K.; Sun, K. M.; Manderville, R. A.; Wetmore, S. D. J. Phys. Chem. A 2008, 112,<br />

3742-3753.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP383<br />

Chlorine-π Interactions: New Methods for Old Problems<br />

Anna K. Croft, Helen M. Howard-Jones, Chris C. Wood<br />

School of Chemistry, <strong>University</strong> of Wales Bangor, Bangor, Gwynedd LL57 2UW, United Kingdom<br />

PP384<br />

The Importance of Solvent Reorganisation in Reactions Performed in Ionic Liquids<br />

Hon Man Yau 2 , Susan A. Barnes 1 , James M. Hook 2 , Tristan G. A. Youngs 3 , Jason B. Harper 2 , Anna K.<br />

Croft 1<br />

1 School of Chemistry, <strong>University</strong> of Wales Bangor, Bangor, Gwynedd LL57 2UW, United Kingdom,<br />

2 School of Chemistry, <strong>University</strong> of New South Wales, Sydney, NSW 2052, Australia, 3 Atomistic<br />

Simulation Centre, Queen's <strong>University</strong> Belfast, Belfast, BT7 1NN, United Kingdom<br />

The structure of the chlorine-atom benzene complex has been a topic of significant controversy for<br />

more than 50 years. We have re-examined the structure of this and related complexes with new<br />

density functional methods especially designed for non-covalent complexes [1], and compared the<br />

structures and energetics to those obtained using standard DFT and high accuracy composite<br />

methods [2]. We find that the popular B3LYP functional fails to identify stationary points revealed by<br />

other functionals [3], and that the η 1 -σ complex appears to be more stable than the η 1 -π complex,<br />

highlighting the careful selection of methods required in non-covalent radical systems.<br />

[1] Zhao, Y. and Truhlar D. G., J. Phys. Chem. A, 2004, 108, 6908-6918.<br />

[2] (a) Henry, D. J., Sullivan, M. B. and Radom, L., J. Chem. Phys., 2003, 118, 4849-4860. (b) Mayer, P. M.,<br />

Parkinson, C. J., Smith D. M., and Radom, L., J. Chem. Phys., 1998, 108, 604-615.<br />

[3] Croft, A. K. and Howard-Jones, H. M., Phys. Chem. Chem. Phys., 2007, 9, 5649-5655.<br />

Ionic liquids have been touted as potential alternatives to environmentally damaging volatile organic<br />

solvents [1]. These salts, typically made up of a bulky organic cation and a charge diffuse anion, are<br />

attractive as alternative solvents due to their negligible vapour pressures and the ability to 'tune' the<br />

properties of the solvent based on the modification of the component ions. Temperature dependent<br />

rate studies demonstrate an enthalpic benefit and an entropic cost associated with the change in the<br />

rate of unimolecular substitution process on addition of a high proportion of an ionic liquid [2].<br />

Molecular dynamics simulations have supported this latter effect by showing a large change in the<br />

degree of organisation around reaction intermediates, compared with that seen for the starting<br />

materials.<br />

[1] Rogers, R. D., and Seddon, K. R., Science, 2003, 302, 792.<br />

[2] Yau, H. M., Barnes, S. A., Hook, J. M., Youngs, T. G. A., Croft, A. K and Harper, J. B., Chem. Commun., 2008,<br />

in press.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP385<br />

Quantum Chemical Study on the Promotion Effect of H 2 in the Selective Catalytic Reduction of<br />

NOx over Ag-MFI Zeolite<br />

Kyoichi Sawabe, Ken-ichi Shimizu, Atsushi Satsuma<br />

Department of Molecular Design and Engineering, Nagoya <strong>University</strong>, Chikusa-ku, Nagoya, Japan<br />

The selective catalytic reduction of NOx to N 2 with hydrocarbons (HC-SCR) is a promising technique<br />

for removing NOx from lean-burn and diesel exhausts. Recently, it was reported that the addition of<br />

hydrogen drastically increases the catalytic activity of Ag/Al 2 O 3 and Ag-exchanged zeolites, such as<br />

Ag/MFI for the HC-SCR [1-3]. Although numerous spectroscopic studies have focused on the<br />

hydrogen promotion effect, the role of hydrogen is still under debate. We have proposed that the Ag n<br />

cluster formed by the hydrogen addition is a key to improve the catalytic activity [2]. On the other<br />

hand, several spectroscopic studies by other groups have indicated no linkage between the high NOx<br />

conversion and the formation of Ag n cluster [3]. In order to sheds light on the controversial reaction<br />

mechanisms ever reported, we have carried out the density functional calculation on the promotion<br />

effect of H 2 addition in the HC-SCR reaction over the Ag-MFI system.<br />

PP386<br />

Towards a 32-electron Principle: Pu@Pb 12 and Related Systems<br />

Jean-Pierre Dognon, Carine Clavaguéra, Pekka Pyykkö<br />

1 CEA/Saclay, DSM/IRAMIS/SCM, Gif sur Yvette, France, 2 CNRS/Ecole Polytechnique, Palaiseau,<br />

France, 3 <strong>University</strong> of Helsinki, Helsinki, Finland<br />

The 18-electron principle goes back to Langmuir [1]. Formally it would correspond to fully occupying at<br />

a central atom its ns, np and (n-1)d orbitals. For early 5f-elements the f-shell becomes chemically<br />

available and remains so until about Am. Theoretically it could be filled with 14 further electrons,<br />

bringing the total to 32, a theoretical possibility already evoked by Langmuir. How far towards that limit<br />

can one go? Thorocene, Th(C 8 H 8 ) 2 , was classified as a ’20e’ case. In the ’metalloactinyl’ compounds,<br />

like the linear IrThIr 2- , one could potentially reach ’24e’ [2].<br />

We now find that the 6p valence band of the recently discovered icosahedral [Pb 12 ] 2- shells forms a<br />

perfect partner for the 5f shell of an enclosed actinide atom, like plutonium [3]. Detailed DFT<br />

calculations suggest that the system is viable. It could be on good grounds characterised as a ’32e’<br />

system.<br />

The calculated molecular geometries, an orbital analysis and a bonding energy analysis in term of<br />

Morokuma-type decomposition will be presented for [Pb 12 ] 2- and [M@Pb 12 ] x- with M=Yb, Th, U, Np, Pu,<br />

Am, Cm. The orbital-energy spectra, the densities of states and the ELF distribution for [An@Pb 12 ] x-<br />

(An=Pu, Am, Cm) will be given. Finally, we will propose computed electronic and spectroscopic<br />

properties for some systems.<br />

We perform now more studies to extend this concept to the search of new 5f-element clusters in a<br />

nanoscience context (biology or materials) [4].<br />

NBO analysis shows that an Ag atom on the MFI is completely ionic. Therefore, the Ag diffusion on<br />

the MFI for the cluster formation requires the neutralization of Ag + . Since hydrogen molecule<br />

heterolytically dissociates between Ag + and oxygen of the MFI with a low activation barrier of 10.7<br />

kcal/mol, one of the roles of the H 2 addition is to promote the Ag diffusion through the formation of<br />

AgH. With this process assumed, a HHAg 4 /MFI system<br />

(Fig.1a) should be an intermediate to form a Ag 4 /MFI<br />

system (Fig1.b). TD-DFT calculation predicts that three<br />

strong absorption bands are observed for the Ag 4 /MFI but<br />

the weak UV absorption for the HHAg 4 /MFI (Fig.2). This<br />

result indicates that the UV-vis measurements are not<br />

adequate for the study of the reaction mechanisms if the<br />

HHAg 4 cluster plays an important role in the HC-SCR<br />

reaction.<br />

Several experiments suggest that oxygen should be<br />

activated for the HC-SCR reaction. No stable structure of<br />

the activated oxygen on the Ag 4 /MFI system has been<br />

found in the geometry optimization. On the other hand, the<br />

HOO - species is exothermically formed from the<br />

HHAg 4 /MFI + O 2 system (Fig.1c). Since the catalytic cycle<br />

of the HC-SCR reaction leads to the formation of Ag 4<br />

cluster, hydrogen addition is essential to reproduce the<br />

HHAg 4 cluster. This conclusion explains the discrepancies<br />

ever reported.<br />

[1]. Langmuir, I. Science 1921, 54, 59-67, this paper mentions the 8, 18 and 32-electron closed shells and uses<br />

on pp. 65-66 Fe(CO)5, Ni(CO)4 and Mo(CO)6 as examples on 18e.<br />

[2]. Hrobárik, P.; Straka, M.; Pyykkö, P. Chem. Phys. Lett. 2006, 431, 6-12.<br />

[3]. Dognon, J. P.; Clavaguéra C.; Pyykkö P. Angew. Chem. Int. Ed. 2007, 46, 1-5<br />

[4]. Dognon, J. P.; Clavaguéra C.; Pyykkö P. in preparation.<br />

[1] Satokawa, S. Chem. Lett. 2000, 29, 294-295.<br />

[2] Shimizu, K; Satsuma, A. Phys. Chem. Chem. Phys. 2006, 8, 2677-2695 and references there in.<br />

[3] Breen, J.P.; Burch, R. Top. Catal. 2006, 39, 53-58 and references there in.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP387<br />

Spin-Orbit Effects on Structures and Vibrational Frequencies of Haloiodomethane Cations and<br />

Halogentriflorides<br />

Hyoseok Kim, Yoon Sup Lee<br />

KAIST, Daejeon, Korea, Republic of<br />

We have been working on ab initio molecular orbital methods that take spin-orbit and other relativistic<br />

effects into account without too much computational effort by using two-component relativistic effective<br />

core potentials (RECPs). The inclusion of spin-orbit effects in the calculations may be important for<br />

molecules containing halogen atoms, especially for cations containing heavy halogen atoms.<br />

Calculated structures and vibrational frequencies of haloiodomethane cations CH 2 XI + (X=Cl and Br)<br />

from conventional density functional theory (DFT) methods without spin-orbit interactions are not in<br />

agreement with experimental ones which have been determined from MATI experiments by Kim et al.<br />

[1,2,3]. While, calculated structures and vibrational frequencies of haloiodomethane cations from DFT<br />

methods with spin-orbit interactions are qualitatively agree well with experimental ones. DFT<br />

calculations using various functionals with various RECPs and the corresponding basis sets were<br />

computed to achieve more accurate calculated results compared to experimental results. Other<br />

halogen series of haloiodomethane cations CH 2 XI + (X=F, I, and At) were also calculated to compare<br />

the extent of spin-orbit effects on cations containing iodine atoms.<br />

The effects of spin-orbit interactions and electron correlations on molecular structures and vibrational<br />

frequencies of halogenflurorides XF 3 (X=I, At, and element 117) were described. Considering spinorbit<br />

effects and other relativistic effects as well as electron correlations, the structure of IF 3 is found to<br />

be a C 2v symmetry and the structure of (117)F 3 is found to be a planar D 3h symmetry, which are in<br />

good agreement with the available estimated data[4,5]. The calculated stable structure of AtF 3 is<br />

critically dependent upon the basis set and electron correlations applied. AtF 3 is a borderline case<br />

between the VSEPR structure of IF 3 and the non-VSEPR structure of (117)F 3 . AtF 3 has a local<br />

minimum in the D 3h symmetry when scalar relativistic effects, spin-orbit interactions, and electron<br />

correlations are considered together.<br />

PP388<br />

Interference Partitioning of the Energy for Generalized Product Functions: N 2 as a Test Case.<br />

Thiago Messias Cardozo, Marco Antonio Chaer Nascimento<br />

Instituto de Química - UFRJ, Rio de Janeiro, RJ, Brazil<br />

The role of quantum mechanical effects in the formation of the chemical bond was first made clear by<br />

recognizing that interference among one-electron states was the driving force for the bonding<br />

phenomena [1]. Although the problem of partitioning the energy to obtain the interference contributions<br />

has been thoroughly discussed [1a], the actual application of the method has been limited to a few<br />

cases. This has to do with the difficulty in partitioning the density of complex molecules in a chemically<br />

intuitive and physically meaningful way.<br />

Here we propose the use of Generalized Product Functions (GPFs) [2] for such an analysis. GPFs are<br />

a natural choice, since their first order reduced density matrices are blocked by electron group and<br />

their second order reduced density matrices are neatly partitioned in intergroup and intragroup blocks.<br />

In particular, by describing valence electron groups with modern Valence Bond methods (GVB, SC,<br />

etc.), the arbitrariness of the choice for atomic orbitals inherent to the interference partition is removed.<br />

We derived the conservation relations for the reduced density matrices and the equations for the<br />

interference partitioning for this type of wavefunction allowing, for instance, in the scope of the σ-π<br />

separation approximation, the separate analysis of interference contributions of σ and π electrons to<br />

the molecular energy. We present the interference energy partition for the N 2 molecule as a test case.<br />

The results were obtained for a wavefunction describing the core and lone-pair electrons at the HF<br />

level, and the bonding electrons at the GVB-PP level, with a cc-pVTZ basis.<br />

The interference contribution to the kinetic energy (KX) has a negative sign for both the σ and π<br />

electron groups, and is the dominant quantum-mechanical term, while the interference potential<br />

energy (VX) is positive, in agreement with previous calculations of simpler molecules. At equilibrium<br />

distance, both σ and π electrons are responsible for stabilizing the bond, with the σ electrons total<br />

interference contribution (EX) being greater than the π electrons EX in module by approximately<br />

0.07hartree.<br />

[1] Lee, M.; Kim, H.; Lee, Y. S.; Kim, M. S. Angew. Chem. Int. Ed. 2005, 49, 2929-2931.<br />

[2] Lee, M.; Kim, H.; Lee, Y. S.; Kim, M. S. J. Chem. Phys. 2005, 122, 244319-1-9.<br />

[3] Lee, M.; Kim, H.; Lee, Y. S.; Kim, M. S. J. Chem. Phys. 2005, 123, 024310-1-9.<br />

[4] Schwerdtfeger, P., J. Phys. Chem. 1996, 103, 2968-2973.<br />

[5] Bae, C.; Han, Y. K.; Lee, Y. S. J. Phys. Chem. A, 2003, 107, 852-858.<br />

0,1<br />

0,4<br />

0,0<br />

0,2<br />

A20<br />

-0,1<br />

-0,2<br />

-0,3<br />

EX(pi bond)<br />

EX(sigma bond)<br />

Energy(adj.)<br />

Energy(hartree)<br />

0,0<br />

-0,2<br />

-0,4<br />

VX(sigma)<br />

VX(pi)<br />

Energy(adj)<br />

KX(sigma)<br />

KX(pi)<br />

-0,4<br />

1 2 3 4 5<br />

-0,6<br />

1 2 3 4 5<br />

A1<br />

Internuclear distance (Angstrom)<br />

[1] (a) Ruedenberg, K. Rev. Mod. Phys. 1962, 34, 326-376. (b) Wilson, C. W.; Goddard III, W. A. Theoret. Chim.<br />

Acta 1972, 26, 195-210<br />

[2] (a) McWeeny, R. Proc. Roy. Soc. Lond. 1959, A253, 242-259. (b) Li, J.; McWeeny, R. Int. J. Quantum Chem.<br />

2002, 89, 208-216


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP389<br />

Canonical Transformation Theory: Review and Application to Transition Metal Oxides<br />

Eric Neuscamman, Takeshi Yanai, Garnet Chan<br />

1 Cornell Department of Chemistry and Chemical Biology, Ithaca, NY, United States, 2 Institute for<br />

Molecular Science, Nishigo-Naka, Myodaiji, Okazaki, Japan, 3 Cornell Department of Chemistry and<br />

Chemical Biology, Ithaca, NY, United States<br />

Canonical Transformation (CT) Theory [1,2] is a new approach for treating dynamic correlation in<br />

multireference systems. Based on a unitary, multireference, exponential ansatz, its accuracy for<br />

calculations on water, nitrogen, and metal-oxides is competitive with state-of-the-art multireference<br />

methods, while its cost is proportional to second order multireference perturbation theory. CT theory’s<br />

central approach is to approximate three and higher body operators produced by the Baker-Campbell-<br />

Hausdorff expansion using one and two body operators. Through the framework of extended normal<br />

ordering [3], this approximation is similar to neglecting the effect of three and higher body connected<br />

excitations. The result is an energy expression that contains no reduced density matrices higher than<br />

second order. Natural connections exist between CT and state-specific multireference coupled cluster<br />

theory [3] as well as recent formulations of the anti-hermitian contracted Schroedinger equation [4].<br />

Absolute energies for TiO, MnO, FeO, and NiO are comparable to those produced by the MRCI+Q,<br />

ACPF, and AQCC methods, much more so than energies computed by CASPT2 and CASPT3.<br />

Current work focuses on developing a straightforward solution condition to overcome the numerical<br />

difficulties presented by a highly linearly dependent first order interaction space.<br />

[1] Yanai, T. and Chan, G. J. Chem. Phys. 2006, 124, 194106<br />

[2] Yanai, T. and Chan, G. J. Chem. Phys. 2007, 127, 104107<br />

[3] Kutzelnigg, W. and Mukherjee D. J. Chem. Phys. 1997, 107, 432-449<br />

[4] Mazziotti, D. Phys. Rev. Lett. 2006, 97, 143002<br />

PP390<br />

Temperature and Isotope Effects on Water Cluster Ions with Path Integral Molecular Dynamics<br />

Suzuki Kimichi 1 , Shiga Motoyuki 2 , Tachikawa Masanori 1<br />

1 Quantum Chemistry Division, Graduate School of Science, Yokohama-City <strong>University</strong>, Seto 22-2,<br />

Kanazawa-ku, Yokohama, Japan, 2 CCSE, Japan Atomic Energy Agency, Higashi-Ueno 6-9-3, Taitoku,<br />

Tokyo, Japan<br />

To analyze the temperature dependency of geometrical isotope effect (GIE) on water cluster ions, we<br />

have combined path integral molecular dynamics (PIMD) and 4th order correction of Trotter expansion<br />

[1,2]. Although quiet huge number of beads is required at low temperature in the conventional PIMD<br />

method, our approach can provide accurate description of the quantum system with less number of<br />

-<br />

+<br />

beads. In this paper, we show the results of temperature dependency of GIE on H 3 O 2 and H 5 O 2<br />

species.<br />

Figure 1 shows schematic illustration of H 3 O - 2 . Figure 2<br />

δ OH* = r 1 – r 2<br />

-<br />

shows one-dimensional distribution on H 3 O 2 with<br />

R<br />

respect to the δ OH* at 50 and 400K, where δ OH* is<br />

OO<br />

defined as the difference of r 1 and r 2 . Apart from the<br />

r 1 r<br />

“quantum simulation” by PIMD calculation, the<br />

H*<br />

2<br />

“classical simulation” has been also performed with the<br />

Figure 1. Schematic illustration of H 3 O 2- .<br />

condition of 1 bead. For convenience, proton and triton substitutions are labeled as “H-species” and<br />

“T-species” in quantum simulation, respectively. In Figure 2 a), the distribution of classical simulation<br />

has double peaks, since it is difficult for the central proton to go over the potential barrier at δ OH* =0. As<br />

the temperature becomes higher, the distribution at δ OH* =0 is gradually increased as shown Figure 2<br />

b).<br />

On the other hand, in the quantum simulation we can find that<br />

distributions of H and T species have single peak at 50K, since<br />

these species completely go over the potential barrier at δ OH* =0<br />

by both thermal and nuclear quantum effects. Figure 2 b) shows<br />

distribution of T-species is double peaks at 400K, while that of<br />

H-species is still single peak. This result means that hydrogen<br />

bonded triton is located at closer to either oxygen atoms, since<br />

the potential barrier height at δ OH* =0 becomes higher as the<br />

elongation of bond length between two oxygen atoms (R OO ).<br />

Actually, GIE with respect to the R OO is clearly different between<br />

low and high temperatures. Average bond length between two<br />

oxygen atoms () of H-species is longer than that of T-<br />

species at low temperature, while the inverse relation is<br />

observed at high temperature. We have found that GIE at low<br />

temperatures is due to the difference of zero point vibration,<br />

while multidimensional effect which is coupled between proton<br />

transferring and oxygen-oxygen stretching motions should be<br />

indispensable at high temperatures. Further detail will be<br />

presented on my poster session.<br />

Distribution<br />

a) 50K<br />

b) 400K<br />

H-species<br />

T-species<br />

classical<br />

-1.2 -0.6 0.0 0.6 1.2<br />

δ OH* (Å)<br />

Figure 2. 1D distributions of δ OH*<br />

in H 3 O 2- at a) 50K and b) 400K.<br />

[1] Takahashi, M.; Imada, M. J. Phys. Soc. Jpn., 1984, 53, 3765.<br />

[2] Li, X-P.; Broughton, J. Q. J. Chem. Phys., 1987, 86, 5094.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP391<br />

Molecular Dynamics Simulation of Photodesorption Process of CO Ice<br />

Junko Takahashi 1 , Marc van Hemert 2<br />

1 Meiji Gakuin <strong>University</strong>, Tokyo, Japan, 2 Leiden Institute of Chemistry, <strong>University</strong> of Leiden, Leiden,<br />

Netherlands<br />

At such a low density and temperature in interstellar space, gaseous CO is still detected, although all<br />

molecules other than H 2 should be kept frozen on dust grains within the timescale shorter than the<br />

cloud lifetime. Since thermal desorption is negligible, photodesorption has been suggested as a<br />

possible mechanism to explain the abundance of CO in interstellar gas phase. In the laboratory, this<br />

process has been simulated by depositing CO on a 10 K surface and following CO desorption upon<br />

irradiation of a D 2 discharge lamp [1].<br />

In order to analyze this process theoretically, we have performed a classical molecular dynamics<br />

simulation on CO photodesorption from an amorphous cluster of a few hundred CO molecules as a<br />

model of a piece of interstellar dust grain. All simulations were performed on the non-periodic systems.<br />

The equations of motion were integrated according to the velocity-Verlet algorithm. The system<br />

temperature was set at 10-40 K, the typical temperatures in interstellar clouds.<br />

We have constructed a force field by performing ab initio coupled cluster calculations for the CO dimer<br />

on a large grid of orientations and inter- and intra- molecular distances. In order to get a potential<br />

energy surface in a useful form for MD simulation, we fit about 15000 data points to a potential energy<br />

function consisting of Lennard-Jones parameters and site point charges. We also fit the intramolecular<br />

potential of the individual CO molecule to a Morse-type function.<br />

At the first step of MD simulation, we made amorphous CO clusters. The initial loose geometries were<br />

generated on a one by one basis by minimizing the interaction energy of the newly arriving CO with<br />

the CO molecules already present with a simplex method. After about 50 ps MD simulation, we<br />

obtained stable amorphous CO clusters.<br />

At the second step, we excited a selected CO molecule in a CO<br />

cluster with a 9-10 eV photon. This energy is insufficient to break the<br />

C-O bond, but there are several electronically excited states in this<br />

energy region. The dominant excitation is X 1 Σ + to A 1 Π. We translated<br />

the photon energy into ground or excited state vibration by extending<br />

the intramolecular distance of the CO molecule to a value<br />

corresponding to the outer turning point of the Morse oscillator.<br />

We assumed the following two mechanisms; (1) Instantaneous internal conversion; (2) Internal<br />

conversion with a relaxation period. In the first mechanism, the excited CO molecule does not<br />

propagate on the excited state potential energy surface, and there were no resulting trajectories of CO<br />

desorption. On the other hand, in the second mechanism, the excited CO propagates on the excited<br />

state potential energy surface, and we observed several trajectories of CO desorption where the<br />

excited CO or a neighbouring CO molecule was kicked out from the CO cluster. The latter case<br />

severely depends on the pair potential for the ground state CO and the excited state CO. We are still<br />

developing this potential energy function. In the poster, we will show the latest results of the potential<br />

energy functions and the MD simulation.<br />

[1] Öberg, K. I.; Fuchs, G. W.; Awad, Z.; Fraser, H. J.; Schlemmer, S.; van Dishoeck, E. F.; Linnartz, H.<br />

Astrophys. J., 2007, 662, L23-L26.<br />

PP392<br />

HERON Reaction of N-Acyloxy-N-alkoxyamides — Theoretical and Experimental Study<br />

Stephen Glover<br />

School of Science and Technology, <strong>University</strong> of New England, Armidale, N.S.W., Australia<br />

Anomeric amides 1 are amides bearing two heteroatoms at the amide nitrogen [1-3]. They are<br />

pyramidal at nitrogen and behave as N-acylamines in all respects. They undergo the HERON<br />

rearrangement, which is anomerically driven by a lone pair destabilisation of the N-Y bond by the less<br />

electronegative atom X [4, 5].<br />

R<br />

O<br />

C N<br />

1<br />

X<br />

Y<br />

R C O<br />

X<br />

N Y N Y<br />

Figure 1. HERON reaction of an anomeric amide.<br />

HERON has mainly been observed for NNO, systems. However, ONO systems in the form of N-<br />

acyloxy-N-alkoxyamides (1, Y=OR, X=Oacyl) have recently been found to undergo the rearrangement<br />

affording anhydrides and alkoxynitrenes. 6<br />

The reaction has been modelled at the B3LYP/6-31G* level using N-formyloxy-N-methoxyformamide 2<br />

and DFT calculations predict the rearrangement to favour migration of the formyloxy group over the<br />

methoxyl group. The energetics of reaction pathways have been deduced and are in line with the<br />

significant energy requirement of the reaction, which with N-acyloxy-N-alkoxyamides is found to occur<br />

above 90°C in non-polar toluene.<br />

H<br />

O<br />

E<br />

O<br />

A =39 kcal mol -1<br />

O<br />

O<br />

O<br />

N OMe<br />

H<br />

H H<br />

H<br />

O<br />

N<br />

O C<br />

C H<br />

O<br />

MeO<br />

N OMe<br />

O<br />

(2) (3) (4)<br />

[1] Glover, S. A.; Rauk, A. J. Org. Chem. 1996, 61, 2337.<br />

[2] Glover, S. A. Tetrahedron 1998, 54, 7229.<br />

[3]Glover, S. A.; Rauk, A. J. Org. Chem. 1999, 64, 2340.<br />

[4] Glover, S. A.; Rauk, A.; Buccigross, J. M.; Campbell, J. J.; Hammond, G. P.; Mo, G.; Andrews, L. E.; Gillson,<br />

A.-M. E. Can. J. Chem. 2005, 83, 1492.<br />

[5] Glover, S. A., HERON Rearrangement. In Merck Index, Organic Name Reactions ONR-43, 14 ed.; O'Neil, M.<br />

J., Ed. Merck & Co., Inc.: Whitehouse Station, N.J., 2006.<br />

[6] Glover, S. A., N-Acyloxy-N-alkoxyamides —Structure, Properties, Reactivity and Biological Activity. In Adv.<br />

Phys. Org. Chem., Richard, J., Ed. Elsevier: London, 2007; Vol. 42, p35.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP393<br />

A Linear-Scaling Spectral-Element Method for Computing Electrostatic Potentials<br />

Mark A. Watson, Kimihiko Hirao<br />

Department of Applied Chemistry, The <strong>University</strong> of Tokyo, Tokyo, Japan<br />

A new linear-scaling method for the fast numerical evaluation of the electronic Coulomb potential is<br />

presented. Our work is an extension of the low-scaling algorithm introduced by Sundholm<br />

[J.Chem.Phys.122, 194107 (2005), J.Chem.Phys.126, 094101 (2007)] which obtains the electrostatic<br />

potential of an arbitrary density by direct summation of differential contributions avoiding the solution<br />

of Poisson's equation. The current work describes three main improvements. Firstly, we replace the<br />

uniform finite-element (FE) representation with a high-order spectral expansion in a tensorial basis of<br />

Chebyshev polynomials. We show that the new basis can converge the Coulomb energy to an<br />

accuracy 1000 times higher than the FE representation for the same number of expansion<br />

coefficients, or, alternatively, for a target accuracy of one part per million, 8 times fewer coefficients<br />

are required, which implies a 16-fold reduction in CPU time when using the basic O(N 4/3 ) algorithm.<br />

Secondly, we reduce the scaling to the true O(N) regime by combining the numerical algorithm with<br />

the fast multipole method for distant interactions. Thirdly, we explore further ways to increase the<br />

efficiency using adaptive resolution for different regions of space. Finally, benchmark calculations to<br />

demonstrate the above improvements are reported.<br />

PP394<br />

Automatized Derivation and String-Based Evaluation of Explicitly Correlated Wavefunctions<br />

Andreas Köhn, Gareth Richings<br />

<strong>University</strong> of Mainz, Institute of Physical Chemistry, D-55099 Mainz, Germany, Germany<br />

The inclusion of terms in the wavefunction ansatz that explicitly take into account the interelectronic<br />

cusp condition has been shown to greatly improve the basis-set convergence of the correlation<br />

energy, for a review see e.g. [1]. One draw-back of such methods, however, is the increasing<br />

complexity of the final expressions. This makes it difficult to implement new wavefunction models “by<br />

hand” in a computer code which is both sufficiently fast and error-free.<br />

Here, we present the program GeCCo, which provides tools to set up and manipulate the explicit<br />

expressions that arise from wavefunction models that can be treated in second quantization. In<br />

particular, the program features a compact, yet considerably efficient kernel for the numerical<br />

evaluation of these expressions which uses the string-based approach [2,3] for addressing arbitrarily<br />

indexed quantities with high permutational symmetry.<br />

As applications, we will discuss the implementation of explicitly correlated higher-order coupled-cluster<br />

methods and the development of explicitly correlated excited-state methods.<br />

[1] W. Klopper, F. R. Manby, S. Ten-no, and E. F. Valeev, Int. Rev. Phys. Chem. 2006, 3, 427<br />

[2] W. Duch, J. Phys. A. 1985, 18, 3283<br />

[3] (a) M. Kállay and P. R. Surján, J. Chem. Phys. 2001, 115, 2945; (b) A. Köhn and J. Olsen, J. Chem. Phys.<br />

2006, 125, 17411


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP395<br />

Conical for Stepwise, Glancing for Concerted: The Role of Excited State Topology in Threebody<br />

Dissociation of sym-Triazine<br />

Vadim Mozhayskiy, Anna I. Krylov<br />

<strong>University</strong> of Southern California, Chemistry Department, Los Angeles, CA, United States<br />

We investigated the highly debated three-body dissociation of sym-triazine to three HCN products.<br />

Two dissociation channels are present in the experiment: stepwise and concerted. Calculated state<br />

energies and electronic couplings suggest that sym-triazine is produced in the 3s Rydberg and π*←n<br />

and manifolds. Analysis of the topology of these manifolds along with momentum correlation in the<br />

dissociation products suggest that the 3s Rydberg manifold is characterized by a conical intersection<br />

of two potential energy surfaces and leads to stepwise dissociation, while the π*←n manifold consists<br />

of a four-fold glancing intersection that leads to a symmetric concerted reaction.<br />

PP396<br />

Enzymic H-Tunnelling – A Role for Promoting Vibrations?<br />

Linus O. Johannissen 1 , Micheal J. Sutcliffe 1 , Nigel S. Scrutton 2<br />

1 School of Chemical Engineering and Analytical Science, Manchester Interdisciplinary Biocentre,<br />

<strong>University</strong> of Manchester, Manchester, United Kingdom, 2 Faculty of Life Sciences, Manchester<br />

Interdisciplinary Biocentre, <strong>University</strong> of Manchester, Manchester, United Kingdom<br />

Our understanding of enzyme catalysis has evolved tremendously since the early lock-and-key model.<br />

It is now well established that enzymes are inherently dynamic molecules, undergoing a wide range of<br />

internal motions – from sub-picosecond atomic fluctuations to nanosecond domain motions and even<br />

millisecond conformational rearrangements. Nevertheless, the current text-book picture of enzyme<br />

catalysis – transition state complementarity – is still very much static, and the role of such motions in<br />

the catalytic process is currently the centre of heated debates. In particular, it has come to light over<br />

the past decade or so that enzymically-catalysed H-transfers, which involve nuclear quantum<br />

tunnelling to varying degrees, are influenced by thermally activated vibrations, but the nature and role<br />

of these vibrations are not well established. According to one school of thought, collective, thermally<br />

equilibrated motions are required to attain quantum degeneracy between the reactant and product<br />

states, while faster “promoting” vibrations can enhance the probability of tunneling by compressing the<br />

donor-acceptor distance once such a state has been achieved. The concept of promoting vibrations is<br />

contentious, however, and their involvement in the catalytic effect unclear. For example, donoracceptor<br />

compressions might be caused by stochastic fluctuations rather than a specific vibrational<br />

mode; additionally, similar compressions might occur in solution-based reactions. We present results<br />

from molecular modelling studies of the role of promoting vibrations, using the H-transfer steps from<br />

the reactions catalysed by aromatic amine dehydrogenase and morphinone reductase as examples. In<br />

particular, molecular dynamics simulations are used to observe how enzyme motions impact the active<br />

site, and hybrid molecular mechanical / quantum mechanical methods are used to analyse how these<br />

motions affect the H-transfer barrier.<br />

[1]. Johannissen, L.O., Scrutton, N.S., Sutcliffe, M.J. The enzyme aromatic amine dehydrogenase induces a<br />

substrate conformation crucial for a promoting vibration that significantly reduces the effective potential energy<br />

barrier to proton transfer. J. Roy. Soc. Interface. In press.<br />

[2]. Johannissen, L.O., Hay, S., Scrutton, N.S., Sutcliffe, M.J. Proton tunneling in aromatic amine dehydrogenase<br />

is driven by a short-range sub-picosecond promoting vibration: Consistency of simulation and theory with<br />

experiment. J. Phys. Chem. B 2007, 111, 2631-2638.<br />

[3]. Masgrau L., Roujeinikova A., Johannissen L.O., Hothi P., Basran J., Ranaghan K.E., Mulholland A.J., Sutcliffe<br />

M.J., Scrutton N.S., Leys D. Atomic description of an enzyme reaction dominated by proton tunneling. Science<br />

2006, 312, 237-241.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP397<br />

Calculation of the Effective Chemical Shielding Anisotropy in L-alanyl-L-alanine,<br />

Conformational and Charge dependence study<br />

Ladislav Benda 1 , Petr Bouř 1 , Norbert Müller 2 , Vladimír Sychrovský 1<br />

1 Institute of Organic Chemistry and Biochemistry, Molecular Spectroscopy Group, Praha, Czech<br />

Republic, 2 Johannes Kepler <strong>University</strong>, Institute of Organic Chemistry, Linz, Austria<br />

The high-resolution nuclear magnetic resonance (NMR) spectroscopy is a powerful tool for the<br />

molecular structure determination. Theoretical modeling of NMR parameters can complement the<br />

NMR experiment and allows accurate predictions of structural parameters and dynamical behavior of<br />

biomolecular compounds. Recently we determined the structure of the L-alanyl-L-alanine di-peptide<br />

(A) on the basis of complete assignment of isotropic chemical shifts and scalar coupling constants [1].<br />

There is, however, additional structural information contained in the anisotropy of the chemical<br />

shielding tensor (CSA) that can be accessed through the measurement of NMR relaxation<br />

parameters. The cross-correlated relaxation rates (CCRR) are important NMR relaxation parameters<br />

that depend on both CSA and local geometry. A reliable structural interpretation of the CCRRs still<br />

requires theoretical modeling [2].<br />

In this work we correlated the effective CSAs calculated for atoms along the peptide backbone with its<br />

major descriptors, the torsion angles phi, psi (A, B). Further we investigated the dependence of the<br />

effective CSAs on total charge of the di-peptide (anion, zwitterion, cation) that can be experimentally<br />

accessed at different pH (~12, ~7, ~2, respectively). The geometries were optimized at the BPW91 / 6-<br />

311++G** level employing the PCM solvent model for all three forms of the di-peptide. The NMR<br />

parameters were calculated using the B3LYP / IGLO-III / PCM approach. The calculated surfaces of<br />

effective CSA can improve interpretation of the cross-correlated relaxation rates in peptides.<br />

PP398<br />

Insights into the Structural Basis of N2 and O6 Substituted Guanine Derivatives as Cyclin-<br />

Dependent Kinase 2 (CDK2) Inhibitors: Prediction of the Binding Modes and Potency by<br />

Docking and ONIOM Calculations<br />

Jans Alzate-Morales, Julio Caballero Ruiz, Ariela Vergara, Fernando Danilo González Nilo<br />

Bioinformatics and Molecular Simulation Centre, <strong>University</strong> of Talca, Talca, Region VII, Chile<br />

The cyclin dependent kinases (CDKs) are a class of enzymes which play a fundamental role in cell<br />

cycle regulation [1]. The aberrant CDK control and consequent loss of cell cycle checkpoint function<br />

have been directly linked to the molecular pathology of cancer [2]. One of the medical strategies to<br />

face this pathology is using synthetic inhibitors. For example, N2 and O6 substituted guanine<br />

derivatives are well known as potent and selective CDK2 inhibitors [3]. The ability of molecular<br />

docking, using the program AutoDock3 [4], and the hybrid method ONIOM [5] to obtain some quantum<br />

chemical descriptors, with the aim to successfully rank these inhibitors (80 compounds), was<br />

assessed. The quantum chemical descriptors were used to explain the affinity, of the series studied,<br />

by a model of the CDK2 binding site. The initial structures were obtained from docking studies and<br />

the ONIOM method was applied with only a single point energy calculation on the protein-ligand<br />

structure. We obtained a good correlation model between the ONIOM derived quantum chemical<br />

descriptor “H-bond Energy” and the experimental biological activity, with a correlation coefficient value<br />

of R = 0.80 for 75 compounds. To the best of our knowledge, this is the first time that both<br />

methodologies are used in conjunction in order to obtain a correlation model. The model suggests that<br />

electrostatic interactions are the principal driving forces in this protein-ligand interaction. Overall, the<br />

approach was successful for the cases considered and suggests that this may be useful for the design<br />

of inhibitors in the lead optimization phase of drug discovery.<br />

(A) (B)<br />

9<br />

R = 0.80<br />

N = 75<br />

Figure could<br />

not be<br />

loaded<br />

Log (10 6 /IC 50<br />

)<br />

8<br />

7<br />

6<br />

5<br />

4<br />

-10 -15 -20 -25 -30 -35 -40 -45<br />

ONIOM[(High, A)+(Mid, AB-A)] Energy (kcal/mol)<br />

[1] (a) Bouř, P.; Buděšínský, M.; Špirko, V.; Kapitán, J.; Šebestík, J.; Sychrovský, V. J. Am. Chem. Soc. 2005,<br />

127, 17079-17089. (b) Sychrovský, V.; Buděšínský, M.; Benda, L.; Špirko, V.; Vokáčová, Z.; Šebestík, J.; Bouř, P.<br />

J. Phys. Chem. B 2008, 112, 1796-1805.<br />

[2] (a) Reif, B.; Diener, A.; Hennig, M.; Maurer, M.; Griesinger, C. J. Magn. Reson. 2000, 143, 45–68. (b)<br />

Sychrovský, V.; Müller, N.; Schneider, B.; Smrečki, V.; Špirko, V.; Šponer, J; Trantírek, L. J. Am. Chem. Soc.<br />

2005, 127, 14663-14667.<br />

[1] Morgan, D. O. Nature 1995, 374, 131-134.<br />

[2] Hall, M.; Peters, G. Adv Cancer Res 1996, 68, 67-108.<br />

[3] Hardcastle, I. R.; Arris, C. E.; Bentley, J.; Boyle, F. T.; Chen, Y. et al. J Med Chem 2004, 47, 3710-3722.<br />

[4] Garrett M. Morris, D. S. G., Robert S. Halliday, Ruth Huey, William E. Hart, Richard K. Belew, Arthur J. Olson.<br />

Journal of Computational Chemistry 1998, 19, 1639-1662.<br />

[5] Dapprich, S.; Komaromi, I.; Byun, K. S.; Morokuma, K.; Frisch, M. J. Journal of Molecular Structure:<br />

THEOCHEM 1999, 461-462, 1-21.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP399<br />

New Implementation and Parallelization of DMRG: Towards Large-Scale Multireference<br />

Electronic-Structure Calculations<br />

Yuki Kurashige, Takeshi Yanai<br />

Institute for Molecular Science, Okazaki, Aichi, Japan<br />

Recently, Density Matrix Renormalization Group (DMRG) algorithm, which was invented to solve<br />

quantum lattice problems, has been applied to solution of the electronic Schrodinger equation with ab<br />

initio Hamiltonian [1]. Ab initio DMRG is especially expected to be a robust alternative to CASCI<br />

method for large-scale multireference problems that cannot be handled by the conventional CI. As can<br />

be seen in past ab initio DMRG studies by Chan's group [2], the DMRG approach is successful in<br />

dealing with large-scale multireference correlation, particularly, in long chain π-conjugated molecules<br />

that are spatially quasi-one-dimensional. For the π-conjugated systems, the one-dimensional<br />

correlation in π space is easily described by a small number of the renormalized correlation basis, the<br />

number of which interestingly remains small (e.g. 300-500) for yielding the certain accuracy no matter<br />

how long the system grows. Targeting the large-scale multireference problems for general molecular<br />

systems, say the polyatomic complexes containing transition metals, however, it seems to be still an<br />

open question whether ab initio DMRG can accurately describe the complicated strong electron<br />

correlation in such systems where the one-dimensionality of the correlation structure is absent while<br />

the exponentially-growing number of necessary renormalized correlation basis is marked contrast to<br />

one-dimensional situation. To investigate and extend the applicability of ab initio DMRG to such<br />

problems, we have developed a highly efficient algorithm which can take full advantage of point-group<br />

symmetry and have exploited a new parallelism that allows for maintaining a large number of<br />

renormalized basis for the correlation space.<br />

[1] (a) White, S. R.; Martin, R. L. J. Chem. Phys. 1999, 110, 4127. (b) Mitrushenkov, A. O.; Fano, G.; Ortolani, F.;<br />

Linguerri, R.; Palmieri, P. J. Chem. Phys. 2001, 115, 6815. (c) Chan, G. K.-L.; Head-Gordon, M. J. Chem. Phys.<br />

2002, 116 4462. (d) Legeza, Ö; Röder, J.; Hess, B. A. Phys. Rev. B 2003, 67, 125114. (e) Moritz, G; Reiher, M.<br />

J. Chem. Phys. 2007, 126, 244109. (f) Zgid, D.; Nooijen, M. J. Chem. Phys. 2008, 128, 014107.<br />

[2] (a) Hachmann, J.; Cardoen, W.; Chan, G. K.-L. J. Chem. Phys. 2006, 125, 144101. (b) Hachmann, J.;<br />

Dorando, J. J.; Avilés, M.; Chan, G. K.-L. J. Chem. Phys. 2007, 127, 134309. (c) Ghosh,D.; Hachmann, J.; Yanai,<br />

T.; Chan, G. K.-L. J. Chem. Phys. 2008, 128, 144117.<br />

PP400<br />

Hybrid Functionals and Møller-Plesset Perturbation Theory applied to Extended Systems<br />

Joachim Paier, Andreas Grueneis, Martijn Marsman, Georg Kresse<br />

<strong>University</strong> of Vienna, Computational Materials Physics, Vienna, Austria<br />

The main advantage of density functional theory (DFT) is its computational efficiency, i.e., at the<br />

moment no other ab-initio method provides results at such low computational cost combined with<br />

reasonable accuracy. Therefore, DFT has found its undisputed place for many applications in solid<br />

state physics as well as quantum chemistry.<br />

Of course, DFT in its LDA and GGA flavours still does not provide final and decisive answers to some<br />

of the most delicate problems in computational physics and chemistry: E.g., reaction energies and<br />

barriers of chemical reactions are wrong by up to 100 kJ/mol (= 1 eV/f.u.), some semiconductors are<br />

predicted to be metals, correlated oxides are not well described, equilibrium volumes and elastic<br />

constants are generally wrong by several percent and dispersion interactions are not properly<br />

accounted for. In order to improve and to include the yet lacking physics a possible strategy seems to<br />

be the inclusion of (exact) Hartree-Fock exchange together with a compatible correlation energy, e.g.,<br />

by N th -order Møller-Plesset (MPN) perturbation theory. Note that MP2 is almost routine in quantum<br />

chemistry but due to the computational cost involved hardly ever applied to extended systems.<br />

Recently, the nonlocal Fock operator as well as 2 nd -order Møller-Plesset perturbation theory (MP2)<br />

were implemented within the framework of the full-potential projector augmented plane-wave using<br />

periodic boundary conditions and a plane-wave basis set.<br />

The poster summarizes some of the latest post-DFT results obtained using the projector augmented<br />

plane-wave code VASP (Vienna ab-initio simulation package). We present lattice constants, bulk<br />

moduli, atomization energies and band gaps obtained using the HSE06 Hartree-Fock/DFT hybrid<br />

functional applied to several representative semiconducting and insulating systems. Secondly,<br />

emphasis is put on the performance of hybrid functionals when applied to metallic systems. Finally,<br />

lattice constants, bulk moduli, atomization energies and band gaps of several semiconductors and<br />

insulators obtained using MP2 are presented.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP401<br />

Molecular Docking with Accurate Polarizable Charges: A QM/MM Approach in Discovery Studio<br />

Jiabo Li, Al Maynard, Jurgen Koska, George Fitzgerald, Dipesh Risal, Paul Kung, Jon Sutter, Paul<br />

Flook<br />

Accelrys Inc, San Diego, United States<br />

The importance of using accurate polarizable charges in molecular docking has been recognized in<br />

recent studies. A major defect of using fixed force field charges in docking is that it lacks the flexibility<br />

to reflect the electrostatic polarization of proteins and ligands. In this study, a novel docking method<br />

using accurate electric charges of ligands derived from quantum mechanical calculations has been<br />

reported. This method has been implemented in Discovery Studio 2.1, the most comprehensive<br />

platform for life science applications [1]. Validation studies are also reported.<br />

PP402<br />

Ionization Energy of 1-Hydroxyethyl Radical: The Effects of Hyperconjugation<br />

Boris Karpichev, Hanna Reisler, Anna Krylov, Kadir Diri<br />

<strong>University</strong> of Southern California, Department of Chemistry, Los Angeles, CA, United States<br />

Experimental studies on the hydroxymethyl and 1-hydroxyethyl radicals have revealed an unusually<br />

large difference in their ionization energies (IE). The expected decrease in the IE of the latter radical<br />

due to its larger size does not fully account for the experimentally observed difference of 0.92 eV. In<br />

this work we investigated the problem with the aid of electronic structure calculations. It is found that<br />

the large drop in the IE of 1-hydroxyethyl radical stems from the combined effects of the<br />

destabilization of its highest occupied molecular orbital and the stabilization of the corresponding<br />

cation due to hyperconjugation. This qualitative picture is in agreement with a simple Hückel-like<br />

approach and is also verified with Natural Bond Orbital calculations.<br />

[1] Discovery Studio 2.1, Accelrys Inc, San Diego, 2008


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP403<br />

Oxidase Catalysis Beyond Biology: A Density Functional Theory Investigation of<br />

N-Methyl-Imidazole (and Acetate) Complexes of First-Row Transition Metal Ions<br />

Ivan Taylor, Stephen Colbran, Gary Willett<br />

School of Chemistry, <strong>University</strong> of New South Wales, Sydney NSW, Australia<br />

Imidazolyl (His) and carboxylate (Asp or Glu) ligand environments distinguish the active sites of many<br />

metalloproteins that activate dioxygen for constructive biological reactions or deactivate reactive<br />

reduced oxygen species which are deleterious in living systems. Examples include the characteristic 2<br />

His + 1 carboxylate triad of non-heme iron oxygenases, the iron and copper centres within soluble and<br />

particulate methane monooxygenases, respectively, and the manganese, iron and nickel centres in<br />

the respective superoxide dismutase [1]. The study of metal centres with imidazole and carboxylate<br />

ligands is the key to understanding the physical processes behind catalysis by the aforementioned<br />

metalloproteins. Looking more widely, questions about how analogous complexes of transition metal<br />

ions not utilised in biological systems would behave in dioxygen activation reactions have not before<br />

been addressed.<br />

Using N-methyl imidazole (N-MeIm) and acetic acid (HOAc) as analogues for histidine and for aspartic<br />

and glutamic acids, respectively, the first-row transition metal (Sc through Zn) complexes [M(N-<br />

MeIm)] n+ , [M(N-MeIm) 2 ] n+ , [M(N-MeIm) 3 ] n+ , [M(N-MeIm)(OAc)] n+ and [M(N-MeIm) 2 (OAc)] n+ have been<br />

modelled using Density Functional Theory. These species and their reactions can also be observed<br />

and studied in the gas phase by mass spectrometry. The geometric and electronic structures of these<br />

metal complexes and, in addition, the charge balance between the metal ion and its ligand(s) [2] have<br />

been studied in order to elucidate the chemical reasons for the prolific utilisation of such species in<br />

metalloprotein active sites.<br />

Periodic trends are observed in the charge donation across the first-row transition metal complexes,<br />

possibly correlating with the catalytic activity of these metal centres.<br />

[1] See, for example: Kovacs, J. A. Science, 2003, 299, 1024-1025. Schwartz, J. K.; Rosenzweig, A. C.;<br />

Frederick, C. A.; Lippard, S. J.; Nordlund, P. Nature 1993, 366, 537-543. Solomon E. I. et al. J. Am. Chem. Soc.<br />

2008, 130, 7098-7109. Yoshizawa K.; Shiota Y. J. Am. Chem. Soc., 128, 9873 -9881. Gherman B.F. et al. J. Am.<br />

Chem. Soc. 2001, 123, 3836-3837. Lieberman, R. L.; Rosenzweig, A. C. Nature, 2005, 434, 177-182. Miller, A.-<br />

F., and Sorkin, D. L. Comments Mol. Cell. Biophys. 1997, 9, 1-48.<br />

[2] Solomon E. I.; Gorelsky S. I.; Dey A. J. Comp. Chem. 2006, 27, 1415-1428.<br />

PP404<br />

Development of the Translational and Rotational Free Quantum Monte Carlo Method<br />

Yukiumi Kita 1 , Ryo Maezono 2 , Masanori Tachikawa 1<br />

1 Yokohama city <strong>University</strong>, Yokohama, Kanagawa, Japan, 2 School of Information Science, Japan<br />

Advanced Institute of Science and Technology, Nomi, Ishikawa, Japan<br />

The quantum nature of the light particle (proton and deuteron etc.) has important effects on the many<br />

research fields. Representative topics include the hydrogen/deuterium isotope effect observed in redshift<br />

of vibrational frequencies [1], shift of chemical reaction rates [2], the Ubbelohde effect [3], and<br />

drastic shift in the structural phase transition temperature for hydrogen-bonded dielectric materials [4].<br />

Especially, hydrogen cluster and its isotope (H 2 , HD and H 3 + etc.) are one of the system in which the<br />

nuclear quantum effect appears in dominantly.<br />

For such systems in which we cannot use the Born–Oppenheimer approximation for nuclei, two<br />

distinct theoretical approaches are available. One is the method using the internal coordinate which<br />

allows us to analyze theoretically only physical quantities depending on internal coordinates [5]. The<br />

other approach is the method using the modified Hamiltonian defined by,<br />

where the first two terms are the kinetic and Coulomb interaction energies for N-particle system,<br />

respectively. The last two terms are the kinetic energy of the center of mass of N-particle system and<br />

the rotational energy of the system around the center of mass, respectively. Thus, H TRF means the<br />

internal energy of the N-particle system. This theoretical approach using H TRF , which was first<br />

proposed by Nakai et al. with the ab initio molecular orbital technique [6], has the generalized<br />

frameworks for various systems containing only quantum particles and the ease for generating the trial<br />

wave function used in quantum Monte Carlo method in contrast to the former.<br />

In this study, we developed the translational and rotational free quantum Monte Carlo [TRF-QMC]<br />

method which means the variational and diffusion Monte Carlo method under the Hamiltonian H TRF<br />

using the trial wave function obtained by ab initio Multi-Component Molecular Orbital method [7], and<br />

applied to the system of the positronium and<br />

HD as benchmark models for this method, and<br />

H + -0.18908(4)<br />

3 system.<br />

TF: translational free<br />

VMC<br />

TRF: translational & rotational free<br />

Figure 1 shows the total energy of the<br />

positronium obtained by our methods. From<br />

this figure, we found that the diffusion Monte<br />

-0.249476(9)<br />

Carlo method with the translational and<br />

TF-VMC<br />

rotational free technique gives the exact<br />

-0.249627(7)<br />

internal energy of this system. The results of<br />

TRF-VMC -0.24992(6)<br />

HD and H + -0.25<br />

3 will be shown in the presentation.<br />

TRF-DMC exact<br />

Energy [a.u.]<br />

Fig.1: Total energy of positronium by TRF-QMC method<br />

[1] N. D. Sokolov, M. V. Vener, and V. A. Savel’ev, J. Mol. Struct. 1990, 222, 365.<br />

[2] Reaction Rate of Isotopic Molecules, edited by L. Melander and W. H. Saunders (Wiley, New York, 1980).<br />

[3] A. R. Ubbelohde and K. J. Gallagher, Acta. Cryst. 1995, 8, 71.<br />

[4] N. Dalal, A. Klymachyov, and A. Bussmann-Holder, Phys. Rev. Lett. 1998, 81, 5924.<br />

[5] D. B. Kinghorn and L. Adamowicz, J. Chem. Phys. 2000, 113, 4203.<br />

[6] H. Nakai, M. Hoshino, K. Miyamoto and S. Hyodo, J. Chem. Phys. 2005, 122, 164101.<br />

[7] M.Tachikawa; Chem. Phys. Lett. 2001, 350, 269.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP405<br />

Fragmentation of Peptide Radical Cations: Proton Scissors vs. Proton Patches<br />

Galina Orlova, Matthew MacLennan<br />

St. Francis Xavier <strong>University</strong>, Department of Chemistry, Antigonish, NS, Canada<br />

Collision-induced dissociation (CID) of peptide radical cations is becoming a powerful mean in mass<br />

spectrometry for structural determination and proteomics. The fragmentation behaviour of radical<br />

cations is different from that of protonated species. The fragmentation mechanism of protonated<br />

peptides is based on the “proton scissors” model developed over the last two decades [1]. According<br />

to this model, a proton hops with low activation barriers between the basic centres of a peptide<br />

weakening the skeletal bonds. For example, the cleavage of an amide bond is rationalized by the<br />

protonation at the nitrogen of the peptide bond followed by low-barrier dissociation.<br />

The fragmentation mechanism of peptide radical cations is not yet fully understood. There exist<br />

competitive radical-induced and charge-directed fragmentations observed experimentally, the latter is<br />

rationalized in terms of “proton scissors” model [2]. We performed QM/CPMD study of spontaneous<br />

and low-barrier (ca. 10 kcal/mol) fragmentations of radical-cationic amino acids Thr +• (loss of CH 2 C=O<br />

and CH 2 C=OH + ), Arg +• and Asn +• (loss of CO 2 ), and Trp +• (loss of cationic side chain, 3-methylene<br />

indolenine). QM at the B3LYP/6-311+G(d,p) level predicts that loss of neutral fragments is driven by<br />

proton transfer, in accord with the proton scissors model. However, molecular dynamics simulations<br />

showed that the C-C bonds break first followed by proton transfer as shown below. Moreover, proton<br />

PP406<br />

A Molecular Dynamics study of Protein-Protein Binding between a K + Channel and Peptide<br />

Toxin<br />

Po-Chia Chen, Serdar Kuyucak<br />

School of Physics, The <strong>University</strong> of Sydney, NSW, Australia<br />

Potassium channels have become a popular topic of study due to the complex features of the<br />

selectivity filter essential to its function in vivo, clinical relevance, and complex structural-functional<br />

relationships [1]. Prediction of binding poses and binding energy have become a major focus of some<br />

computational studies [2].<br />

For this latter purpose, umbrella sampling (US) is a straightforward process of calculating the PMF<br />

(and by extension an estimation of energy of binding) of protein-ligand interactions. Our study<br />

represents a novel investigation into its use on protein toxins and K+ channels, a magnitude larger in<br />

size to its previous applications, with the aim of verifying whether the essentials characteristics of<br />

binding can be replicated for a larger system.<br />

We studied the unbinding process between a complex of KcsA and charybdotoxin (ChTX), the<br />

structure of which has been previously elucidated [3]. The toxin was pulled away from its binding site<br />

in the direction along the channel axis, and 1D Umbrella sampling (US) was applied in order to predict<br />

the PMF for this interaction, with the choice of backbone COM as the reaction coordinate.<br />

It was found that the essential features of the underlying PMF have been observed, and key positions<br />

along the reaction coordinate correlated with e.g. the breaking of ionic interactions between charged<br />

side-chains of the participants. We further showed that the final PMF score is unaffected by the<br />

strength of umbrella potential, provided that the windows overlap. Transient differences between the<br />

PMFs deduced for umbrella potential 20 kcal/mol and 40 kcal/mol have been attributed to the stronger<br />

strain produced between the backbone and the side-chains, reducing the entropic gain of the toxin.<br />

CPMD trajectory snapshots of Arg +• fragmenting to lose CO 2 .<br />

[1] See, for example: Mitcheson, J. S. Chem. Res. Toxicol., 2008, 21, 1005-1010.<br />

[2] e.g. Recanatini, M; Cavalli, A; Masetti, M. Chem. Med.Chem., 2008, 3, 523-535.<br />

[3] Yu, L.; Sun, C.; Song, D.; Shen, J.; Xu, N.; Gunasekera, A.; Hajduk, P.J.; Olejniczak, E.T. Biochemistry, 2005,<br />

44, 15834-15841.<br />

transfer affects thermochemistry making fragmentations more exothermic but it does not notably affect<br />

the reaction barriers to fragmentation, which is important since CID is kinetically controlled. When<br />

proton transfer is sterically hindered the C-C bond still cleaves spontaneously or with a low barrier to<br />

produce a radical and a cation (for example, CH 2 C=OH + and [Gly-H] • for Thr +• )<br />

The proton scissors model therefore might be modified to the “proton patches” model when applied to<br />

the radical cations.<br />

[1]. Wysocki, V.H.; Cheng, G.; Zhang, Q.; Herrmann, K.; Beardsley, R. L., Hilderbrand, A. E., in Principles of Mass<br />

Spectrometry Applied to Biochemistry; Laskin J., Lifshitz, C., Eds.; John Wiley and Sons;NJ, 2006, chapter 8.<br />

[2]. see for example, Chu, I.K.; Zhao, J.; Xu, M.; Siu, S.O.; Hopkinson, A.C.; Siu, K.W.M. JACS, 2008, ASAP<br />

Article; DOI: 10.1021/ja801108j


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP407<br />

A Molecule Search System Using Ajax<br />

Lizhu LIU, Shin-ya TAKANE<br />

Department of Information Systems Engineering, Daito, Osaka, Japan<br />

With the spread of the Internet, we have freely accessed to the immense amount of information on the<br />

Internet. However, it is slightly difficult to find out necessary information quickly. Although search<br />

engines such as Google are useful tools when the keywords contain rather general phrases, it is still<br />

difficult to search scientific or technical terms efficiently.<br />

In this study, we present a prototype of the search system for molecular information (structures and<br />

properties) getting from the temporary database which is constructed from the results of the primary<br />

(keyword) search using a spider program. The results are followed by the secondary search, which<br />

refines by filter program. For the usability of the system, we used Ajax (Asynchronous JavaScript +<br />

XML) as an interface between the search page and the database. This is all done asynchronously<br />

without reloading the page and interrupting user activity.<br />

PP408<br />

Ab Initio Calculation of the Coherent 2D Infrared Response Function for Two-Dimensional<br />

Vibrational Spectroscopy<br />

Sangjoon Hahn 1 , Minhaeng Cho 2<br />

1 Korea Science Academy, Busan, Korea, Republic of, 2 Korea <strong>University</strong>, Seoul, Korea, Republic of<br />

The two-dimensional vibrational spectroscopy involving two infrared pulses and a single optical pulse<br />

[1] is studied by using the ab initio calculation method. By obtaining the first- and second-order<br />

derivatives of the molecular dipole moment as well as the polarizability, the coherent 2D IR response<br />

function and its spectrum are calculated with an assumption that the vibrational dynamics can be<br />

described by the Brownian oscillator model.<br />

Internet<br />

Access<br />

Page<br />

Spider<br />

Database<br />

Filter<br />

The origin of each peak in the entire coherent 2D IR spectrum is discussed in detail, and is directly<br />

compared with the coherent 2D Raman scattering spectrum. This comparison demonstrates the<br />

complementary nature between the coherent 2D IR and Raman spectroscopies. A brief discussion on<br />

the coupling patterns is also presented.<br />

User<br />

Keyword<br />

Ajax<br />

Control<br />

Search page<br />

[1] Park, Kisam; Cho, Minhaeng. J. Chem. Phys. 1998, 109, 10559–10569.<br />

Result


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP409<br />

Ab Initio Density Matrix Renormalization Group with Orbital Optimisation and its Application to<br />

β-Carotene<br />

Debashree Ghosh 1 , Johannes Hachmann 1 , Takeshi Yanai 2 , Garnet Chan 1<br />

1 Department of Chemistry and Chemical Biology, Cornell <strong>University</strong>, Ithaca, United States,<br />

2 Department of Theoretical and Computational Molecular Science, Institute for Molecular Science,<br />

Okazaki, Japan<br />

We have implemented orbital optimisation within our local ab-initio Density Matrix Renormalization<br />

Group method. This has allowed us to carry out DMRG-CASSCF with active spaces as big as (24,24).<br />

We demonstrate our algorithm with full-valence space CASSCF studies on β-carotene [1]. These yield<br />

new insight into the nature of the low-lying excited states of carotenoids involved in photosynthetic<br />

light-harvesting.<br />

[1] D. Ghosh, J. Hachmann, T. Yanai, G.K.-L. Chan, J. Chem. Phys. 2008, 128, 144117<br />

PP410<br />

Excited State Dynamics of Molecules by Using Gaussian Wave Packet<br />

Takashi Kuchitsu 1 , Motoyuki Shiga 2 , Masanori Tachikawa 1<br />

1 International Graduate School of Arts and Sciences, Yokohama City <strong>University</strong>, Seto 22-2,<br />

Kangazawa-ku, Yokohama, Kanagawa 236-0027, Japan, 2 Center for Computational Science and E-<br />

Systems, Japan Atomic Energy Agency (JAEA), Higashi-Ueno 6-9-3, Taito-ku, Tokyo 110-0015,<br />

Japan<br />

By recent development of intense laser technology, it is nowadays a realistic challenge to probe the<br />

electron rearrangement within a molecule [1]. For instance, new ideas have been proposed to control<br />

bond breaking/creating processes and charge transfer reactions using ultrafast laser pulses. In<br />

attosecond to femtosecond time region, photo-released electron comes back to the molecule by<br />

oscillating laser field, and this recolliding electronic wavepacket gives many informations of the<br />

molecule. In such femtosecond/attosecond phenomena, electron dynamics is usually important, since<br />

the electronic excited state may not settle in a certain stationary state for an instantaneous nuclear<br />

configuration. In femtosecond to picosecond time region, ultrafast spectroscopies give information of<br />

nuclear motion in electronically excited state. In some experiment, unexpectedly fast decay of excitedstate<br />

spectra are observed, for such as DNA bases. Although it is considered that the molecule is<br />

relaxed by passing through a crossing of adiabatic surface, its dynamical aspect is not yet studied<br />

sufficiently. Especially, it is a difficult challenge to describe electron-nuclear entanglement.<br />

Recently, we have developed a general time-dependent framework to excited state dynamics of<br />

molecular system [2], combining ab initio molecular orbital scheme and Gaussian wave packet<br />

dynamics. Our purpose is to adopt and modify our framework to the dynamics of attosecond,<br />

femtosecond, and picosecond time region given above.<br />

In our approach, similar to Deumens and Öhrn [3], the many-particle wavefunctions are described by<br />

Slater determinants constructed from time-dependent molecular orbitals, that are usually expanded as<br />

the linear combination of atomic orbitals (LCAO). If only LCAO coefficients are chosen as the timedependent<br />

parameters, our approach corresponds to the conventional real-time time-dependent<br />

Hartree-Fock approach. In addition, the center coordinates of Gaussian basis functions can be set to<br />

time-dependent, giving the traveling basis function. Traveling basis function is expected for electronic<br />

basis function which follows nuclear motion, or describing polarization [2] or ionization [4]. The<br />

remaining parameter, the exponent values of Gaussian basis functions can also be time-dependent,<br />

giving the thawed basis function. This thawed basis function makes possible the distribution of the<br />

wavefunction to diffuse or shrink as electron cloud moves [4]. These time-dependent parameters<br />

of the wavefunction are time-developed by the equation of motion<br />

obtained by applying the time-dependent variational principle. We are expecting our approach to be<br />

developed for multiconfigurational wavefunction and electron nuclear dynamics.<br />

[1] Gagnon, E.; Ranitovic, P.; Tong, X.-M.; Cocke, C. L.; Murnane, M. M.; Kapteyn, H. C.; Sandhu, A. S. Science<br />

2007, 317, 1374.<br />

[2] Kuchitsu, T.; Tachikawa, M.; Shiga, M. Chem. Phys. Lett., 2006, 433, 193.<br />

[3] Deumens, E.; Öhrn, Y. J. Phys. Chem. A 2001, 105, 2660.<br />

[4] Kuchitsu, T.; Okuda, J.; Tachikawa, M. Int. J. Quantum Chem., 2008, accepted.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP411<br />

Spin Transition Mechanism and New Necessary Condition of LIESST: DFT Study of<br />

[Fe(2-pic) 3 ] 2+<br />

Hideo Ando, Yoshihide Nakao, Hirofumi Sato, Shigeyoshi Sakaki<br />

Department of Molecular Engineering, Graduate School of Engineering, Kyoto <strong>University</strong>, Kyoto,<br />

Japan<br />

Photo-induced spin transition between low-spin (LS) and high-spin (HS)<br />

states, called Light-induced excited spin state trapping (LIESST), was<br />

discovered by Decurtins et al. in 1984 [1]. Since then, this phenomenon has<br />

drawn considerable attention. We theoretically investigated the electronic<br />

structures and the potential energy curves (PECs) of LS, HS, and<br />

intermediate-spin (IS) states of a well-known LIESST complex, mer-[Fe(2-<br />

pic) 3 ] 2+ (pic = picolylamine, Scheme 1). Our study aims to clarify the spin<br />

transition mechanism in terms of relative positions of PECs and to present<br />

new guidelines for designing LIESST complexes.<br />

Geometry optimization and evaluation of PECs were carried out with the<br />

DFT method, using either the B3LYP or the B3LYP* functional [2]. PECs<br />

were evaluated along the linear internal coordinate [3] which connects the<br />

LS, HS, and IS equilibrium geometries. Excitation energies were calculated by using the TD-DFT<br />

method with B3LYP functional.<br />

The HS state is as stable as the LS state, which is consistent with experimental results [4]. Spin<br />

transition from LS to HS causes elongation of all Fe-N bonds by about 0.2 Å because anti-bonding d<br />

orbitals become occupied in the HS state. In the IS<br />

state, we found three different structures in which<br />

four Fe-N bonds are elongated by Jahn-Teller<br />

distortion. These structures are close in energy to<br />

each other (within 1 kcal/mol). As shown in Figure<br />

1, the IS potential minimum is above the LS and<br />

the HS state at the same geometry, indicating that<br />

both ISLS intersystem crossing (ISC) and<br />

ISHS ISC can occur. This is one of the<br />

important conditions required for LIESST<br />

complexes [5]. The activation barrier is large<br />

enough to suppress thermal spin transition and<br />

tunneling between LS and HS states. The<br />

excitation energies are different between LS and<br />

HS complexes; the d-d transitions appear at 2.05,<br />

2.07, and 2.09 eV in LS state and at 1.46 and 1.64<br />

eV in HS state. Thus, the LSHS and the<br />

HSLS spin transitions can be induced selectively<br />

by different lights. All these results are consistent<br />

with the fact that both LIESST and reverse-LIESST<br />

are observed in mer-[Fe(2-pic) 3 ] 2+ .<br />

Scheme 1.<br />

mer-[Fe(2-pic) 3 ] 2+<br />

0.0<br />

LS (singlet)<br />

-0.2<br />

HS (quintet)<br />

-0.4<br />

LS geometry IS geometry HS geometry<br />

[1] Decurtins, S.; Gütlich, P.; Köhler, C. P.; Spiering, H.; Hauser, A. Chem. Phys. Lett. 1984, 105, 1-4.<br />

[2] Reiher, M.; Salomon, O.; Hess, B. A. Theor. Chem. Acc. 2001, 107, 48-55.<br />

[3] Komornicki, A.; McIver, J. W., Jr. J. Am. Chem. Soc. 1974, 96, 5798-5800.<br />

[4] (a) Goodwin, H. A. Coord. Chem. Rev. 1976, 18, 293-325. (b) Halcrow, M. A. Polyhedron 2007, 26, 3523-<br />

3576.<br />

[5] Ando, H.; Nakao, Y.; Sato, H.; Sakaki, S. J. Phys. Chem. A 2007, 111, 5515-5522.<br />

Energy /eV<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

IS (triplet)<br />

Figure<br />

Li<br />

1.<br />

i<br />

Potential<br />

l<br />

energy<br />

di<br />

curves<br />

obtained by DFT(B3LYP*) method<br />

PP412<br />

Molecular Dynamics Simulation for the Protonation Process in Matrix-Assisted Laser<br />

Desorption Ionization<br />

Makoto Hatakeyama, Masanori Tachikawa<br />

Graduate School of Integrated Science, Yokohama-City <strong>University</strong>, Yokohama, Japan<br />

In the MALDI (Matrix-Assisted Laser Desorption Ionization), protonated target molecules, produced by<br />

proton transfer from matrix molecules of condensed phase to target species, are detected sharply in<br />

the mass spectrum. The ion yields in the MALDI procedure tend to depend on the proton affinity of<br />

matrix molecules. However, the ion yields of molecules such as Lysine et al. are known to be<br />

independent of the proton affinity [1]. We note here that the experimental process of MALDI begins<br />

from the condensed phase, while a proton affinity is a characteristic value in the gas phase. Thus, we<br />

can easily assume that the protonation of Lysine in the condensed phase may be different from that in<br />

the gas phase. In this paper, we will analyze the protonation reaction on the condensed phase in the<br />

MALDI process by using quantum mechanics and molecular mechanics (QM/MM) molecular<br />

dynamics.<br />

First, we focus on the dynamics of only matrix molecules of α-cyano-4-hydroxy-cinnamic acid (CHCA)<br />

as shown in Figure 1. We have employed HF/STO-3G molecular orbital theory and the general Amber<br />

Force Field as QM and MM parts, respectively. Molecular desorption is simulated as the temperature<br />

in the matrix surface is heated up instantly from 300 K to 1300 K.<br />

When quantum molecules desorbed from the matrix surface, proton transfer sometimes occurs in the<br />

intermolecular hydrogen bond at the =O・・・HO- part. Since these protonated molecules are unstable,<br />

these species return to the neutral molecule. Figure 2 shows the potential energy map of =O・・・HO-,<br />

in which the lower left valley corresponds to the optimized structure and the protonation barrier is<br />

greater than 50 kJ/mol. Figure 2 clearly indicates that molecular translational motion is needed for the<br />

protonation, as well as simple HO bonding vibrational motion. However, we have observed that<br />

molecular translational motion is much slower than HO vibrational motion at the early desorption<br />

process in our QM/MM simulation. We will show more detailed results in our poster session.<br />

H<br />

O<br />

N<br />

C<br />

C<br />

×<br />

C<br />

H<br />

O<br />

C H<br />

O<br />

O H<br />

H C C<br />

×<br />

r<br />

O C<br />

C<br />

R<br />

Fig 1. Hydrogen bonding CHCA pair. R and r refer to<br />

the molecular center of mass dis-tance and H-O<br />

bond length respectively.<br />

N<br />

[1] Nishikaze T.; Takayama M. Rapid Commun. Mass Spectrom. 2006, 20, 376-382.<br />

O<br />

H<br />

Fig 2. =O・・・HO- potential energy difference<br />

from optimized structure by UHF/STO-3G.<br />

Contour interval is defined 20 kJ/mol.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP413<br />

Tungsten η 3 -Silaallyl/Vinylsilyl, η 3 -Silapropargyl/Alkynylsilyl, and Silylene Complexes: New<br />

Insight of their Bonding Nature and Electronic Structure<br />

Mausumi Ray, Yoshihide Nakao, Hirofumi Sato, Shigeyoshi Sakaki<br />

Department of Molecular Engineering, Graduate School of Engineering, Kyoto <strong>University</strong>, Nishikyo-ku,<br />

Kyoto 615-8510, Japan<br />

Multiple bonds involving Si are interesting because<br />

Si is considerably different from its carbon<br />

analogue. Syntheses of such chemical species<br />

are interesting and challenging in recent chemistry.<br />

Silaallyl and silapropargyl species are silicon<br />

analogues of allyl and propargyl, respectively<br />

(Scheme 1). They would be reactive, but the<br />

interaction with transition metal (TM) complex is<br />

expected to stabilize them. Also, the interaction<br />

between the TM complex and these Si species is<br />

interesting because it is very much different from<br />

carbon analogues. Hence, bonding nature and<br />

electronic structure of TM silaallyl and silapropargyl<br />

complexes are attractive research subject in<br />

theoretical/computational chemistry.<br />

We theoretically investigated the bonding nature<br />

and stability of tungsten η 3 -silaallyl complex<br />

Cp(CO) 2 W(η 3 -H 2 SiCHCH 2 ) 1, tungsten η 3 -<br />

silapropargyl complex Cp(CO) 2 W(η 3 -H 2 SiCCH) 2,<br />

and tungsten acetylide-silylene complex<br />

Cp(CO) 2 W(CCH)(SiH 2 ) 3 (Scheme 2), using DFT,<br />

MP4, and CCSD(T) methods with triple zeta quality<br />

basis sets.<br />

Our theoretical study presents fundamental<br />

understanding about the electronic structures of 1,<br />

2, and 3 and the interaction between the TM<br />

complex and silicon species, as follows: Though<br />

the non-bonding π orbital (HOMO) of 1 and 2 are<br />

similar to those of TM-allyl and propargyl<br />

complexes, the π orbital (HOMO-1) is much<br />

different from those of the TM-allyl and propargyl<br />

complexes, indicating that the electronic structures<br />

of 1 and 2 are intermediate between those of<br />

completely conjugated 4 and non-conjugated 5<br />

(Scheme 3) [1,2]. 3 exhibits interesting chargetransfer<br />

(CT) interactions; one is the CT from the<br />

lone pair orbital of silylene to the π* orbital of<br />

acetylide and the other is the CT from the π orbital<br />

of acetylide to the empty p orbital of silylene. 1<br />

Detailed discussions will be presented in the<br />

poster.<br />

Allyl Propargyl<br />

Silaallyl Silapropargyl<br />

Scheme 1<br />

1 2 3<br />

H<br />

Si<br />

H<br />

W H<br />

C C<br />

H<br />

H<br />

Scheme 2<br />

H<br />

Si<br />

H<br />

[1] Ray, M.; Nakao, Y.; Sato, H.; Sakaba, H.; Sakaki, S. J. Am. Chem. Soc. 2006, 128, 11927.<br />

[2] Ray, M.; Nakao, Y.; Sato, H.; Sakaki, S. Organometallics 2007, 26, 4413.<br />

H<br />

H<br />

W<br />

C<br />

C<br />

H<br />

H<br />

H<br />

4 5<br />

Scheme 3<br />

W<br />

H<br />

C C H<br />

H<br />

W<br />

C<br />

C<br />

H<br />

PP414<br />

Theoretical Analysis on Positron Halide Complexes by Multi-Component Quantum Monte Carlo<br />

Method<br />

Tomohiro Takeda 1 , Yukiumi Kita 1 , Ryo Maezono 2 , Masanori Tachikawa 1<br />

1 Graduate School of Integrated Science, Yokohama City <strong>University</strong>, Yokohama, Japan, 2 School of<br />

Information Science, Japan Advanced Institute of Science and Technology, Isikawa, Japan<br />

It is well known experimentally that positron injecting into liquid/solid forms the positron-molecular<br />

complex which is the temporary bound state between a positron and atom/molecule. The positronium<br />

halides complexes [X - ;e + ] (X=F, Cl, Br, and I), which are typical positron-molecular complexes, has<br />

been detected experimentally except for [F - ;e + ]. It is found that the order of the positronium (Ps)<br />

binding energies (X BE ) in aqueous solutions is Cl BE < Br BE < I BE [1], while the order in a vacuum is<br />

known theoretically to be Cl BE > Br BE > I BE [2]. In order to make this inconsistency clear, we evaluate<br />

Ps binding energies by using the more quantitatively accurate multi-component quantum Monte Carlo<br />

[3] (MC_QMC) method. MC_QMC method can handle the correlation not only between electrons but<br />

also the electron-positron in more reliable manner, latter of which is known as a difficult quantity to be<br />

evaluated by conventional quantum chemical methods.<br />

In this paper, we analyze the total energy of positronium halides by MC_QMC method as the first<br />

attempt to reveal the inconsistency between the experimental and theoretical results with respect to<br />

the order of X BE .<br />

We used the Slater-Jastrow wave function,<br />

Ψ ( R)<br />

= e<br />

T<br />

J ( R ) ↑<br />

↓<br />

× De<br />

( R<br />

↑<br />

) × De<br />

( R<br />

↓<br />

) × φp(<br />

rp<br />

) ,<br />

e<br />

e<br />

as a trial/guiding wave function for VMC (variational Monte Carlo) and DMC (diffusion Monte Carlo)<br />

methods. D e ↑↓ , φ p , and J(R) denote single Slater determinant of up/down spin electron, a positron<br />

orbital, and Jastrow factor, respectively. Variational parameters in the Jastrow factor are optimized by<br />

the variance minimization [4, 5] under the cusp condition between each charged particles. Tables 1<br />

and 2 show the total energy of [F - ;e + ], F - , and [Cl - ;e + ] systems with various schemes, respectively. Our<br />

results give variationally improved (lower) energy than those by conventional one (MRCISD), as well<br />

as the smaller statistical error than the previous work by Bressanini [6].<br />

Table 1. Total energy (a.u.) of<br />

Table 2. Total energy (a.u.) of<br />

[F - ;e + ] system<br />

[Cl - ;e + ] system Method Energy<br />

Method Energy<br />

VMC -99.9881(9)<br />

MC_MO -459.7085<br />

DMC -100.0713(2)<br />

VMC -460.257(2)<br />

MRCISD [2] -100.0175<br />

MRCISD [2] -460.0378<br />

DMC [6] -100.0719(8)<br />

[1] Mogensen, O. E.; et al. Chem. Phys. 1979, 37, 139.<br />

[2] Saito, S. L.; et al. J. Chem. Phys. 2005, 122, 054302.<br />

[3] Hammond, B. L.; et al. Monte Carlo Methods in Ab Initio Quantum Chemistry; World Scientific: Singapore,<br />

1994<br />

[4] Umrigar, C. J.; et al. J. Chem. Phys. 1988, 60, 1719.<br />

[5] Drummond, N. D.; et al. Phys. Rev. B, 2005, 72, 085124.<br />

[6] Bressanini, D.; et al. J. Chem. Phys. 1998, 108, 12, 22.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP415<br />

New Developments in Spin-Flip Methods<br />

David Casanova 1 , Lyudmila V. Slipchenko 2 , Anna I. Krylov 3 , Martin Head-Gordon 1<br />

1 <strong>University</strong> of California, Berkeley, Berkeley, California, United States, 2 Purdue <strong>University</strong>, West,<br />

Lafayette, Indiana, United States, 3 <strong>University</strong> of Southern California, Los Angeles, California, United<br />

States<br />

One of the main reasons for the introduction of the spin-flip methods was the idea that triplet states<br />

are often easier to model across a potential energy surface than singlets. As a result, these methods<br />

have been shown to be very reliable in the formation/breaking of chemical bounds [1] and in the<br />

description of diradical systems [2, 3] as well. For these reasons, in recent years the spin-flip methods<br />

have become a common tool for molecular electronic structure studies.<br />

This contribution focuses on two recent developments. First, a general extension of the SF method<br />

with multiple spin-flip excitations is introduced [4]. The extension includes all configurations in which<br />

no more than one virtual level of the high spin reference becomes occupied and no more than one<br />

doubly occupied level becomes vacant. As a result, the ground and low-lying states are treated at the<br />

same level. Therefore, these methods appear to be specially suitable in some challenging situations<br />

like bond breaking, multiradiacaloid systems, avoided crossings or conical intersections. This<br />

expansion takes care of the unbalanced treatment of alpha and beta spaces of the standard SF<br />

models, removing the common noticeable presence of spin contamination in some of the computed<br />

states.<br />

PP416<br />

Fullerene Formation from a Benzene Source: Density Functional Tight-Binding Molecular<br />

Dynamics Simulations<br />

Biswajit Saha 1 , Stephan Irle 2 , Keiji Morokuma 1<br />

1 Fukui Institute for Fundamental Chemistry, Kyoto <strong>University</strong>, Kyoto 606-8103, Japan, 2 Institute for<br />

Advanced Research and Department of Chemistry, Nagoya <strong>University</strong>, Nagoya 464-8602, Japan<br />

Molecular dynamics simulations using the density functional tight-binding (DFTB) quantum chemical<br />

method are reported for the fullerene formation mechanism with benzene molecules as carbon<br />

feedstock. We observed that the presence of hydrogen slows down the cluster growth process as well<br />

as fullerene formation. Reactive polycyclic aromatic hydrocarbons (PAHs) without complete saturation<br />

of their edges by hydrogen form in the early stage, and these PAHs aggregate to form giant fullerenes.<br />

The dynamic self-assembly goes through four distinct stages, i.e., ring breaking, nucleation, ring<br />

condensation and cage closure. The effect of temperature on the formation mechanism is discussed,<br />

and the trajectories are analyzed in terms of pentagon/hexagon/heptagon ring count and sp/sp 2 ratio<br />

evolution. As in the case of fullerene formation from C 2 molecules as feedstock, only giant fullerenes<br />

are created, which are expected to shrink following the “Shrinking Hot Giant” road of fullerene<br />

formation, described earlier by our group.<br />

The second part of this work introduces the double spin-flip (2SF) family of methods. The<br />

implementation and application of 2SF within EOM-CC and CI formalism show their capability of<br />

describing several low-lying electronic states. This approach appears to be particularly suitable in the<br />

computation of double bond breaking and tetraradical systems.<br />

[1] Krylov, A. I.; Sherrill, C. D. J. Chem. Phys. 2002, 116, 3194.<br />

[2] Shao, Y.; Head-Gordon, M.; Krylov, A. I. J. Chem. Phys. 2003, 118, 4807.<br />

[3] Levchenko, S. V.; Krylov, A. I. J. Chem. Phys. 2002, 117, 4694.<br />

[4] Casanova, D.; Head-Gordon, M. J. Chem. Phys. 2008, 129, 64104.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP417<br />

The Electronic Structure of Gold Clusters: A Study by Coupled Cluster Calculation with a<br />

Newly Developed Relativistic Model Core Potential<br />

Hirotoshi Mori 1 , Hisaki Nakashima 2 , Eisaku Miyoshi 2<br />

1 Division of Advanced Sciences, Ochadai Academic Production, Ochanomizu <strong>University</strong>, Bunkyo-ku,<br />

Ohtsuka 2-1-1, Tokyo 112-8610, Japan, 2 Interdisciplinary Graduate School of Engineering Sciences,<br />

Kyushu <strong>University</strong>, 6-1 Kasuga Park, Fukuoka 816-8580, Japan<br />

After the discovery of gold cluster’s amazingly high catalytic function by Haruta et al. [1], the clusters<br />

have been considered as a future’s new ecological material and studied by many scientists [2]. For<br />

example, small gold clusters supported on MgO substrates show remarkable catalytic activity for the<br />

oxidation of CO. The partial hydrogenation of acetylene on supported gold clusters is another<br />

example for the catalytic activity of gold nano clusters. Now, gold’s chemistry is considered as one of<br />

the nano science.<br />

In the past decade, many theoretical studies were done to understand the mechanism of catalytic<br />

function of gold clusters. Most of them were done on the basis of density functional theory (DFT) from<br />

the computational costs point of view, and gave many basic and important information of gold’s<br />

chemistry [3]. On the other hand, recently, Olson et al. did ab initio post-HF based study with<br />

CCSD(T) level of theory instead DFT, and they discussed the effect of semi-core electrons’ correlation<br />

effect [3]. Their results gave new insights of gold chemistry, which cannot be discussed with simple<br />

DFT. However, since their CCSD(T) study was done just for Au 6 and Au 8 , there seems to be still<br />

unresolved chemical problems of gold nano clusters.<br />

In this study, the electronic structure of gold nano clusters (Au n ; n=2-10) was studied by using highlevel<br />

ab initio molecular orbital calculations at the MP2, CCSD, CCSD(T) and CR-CCSD(T) L level of<br />

theories with newly developed small core type relativistic model core potentials (spdsMCP) [4]. MCP<br />

is a kind of effective core potential (ECP) methods which is originally proposed by Huzinaga et al. and<br />

unique in that they are naturally capable of producing valence orbitals with nodal structure [5]. Since<br />

MCP reproduces the valence nodal structure, which is important to describe electron-electron<br />

correlation effect, the choice of the combination of MCP and highly correlated post-HF methods shown<br />

above should be reasonable and reliable one. For comparison, corresponding copper and silver<br />

clusters’ electronic structures were also investigated. The electronic structures of gold clusters,<br />

difference from Cu n or Ag n , will be discussed in detail.<br />

PP418<br />

Density Functional Theory Study on the Stacking and Excitation of Metal Ion Containing<br />

Artificial DNA<br />

Hideaki Miyachi 1 , Toru Matsui 1 , Yasuteru Shigeta 2 , Kimihiko Hirao 1<br />

1 Department of Applied Chemitry, School of Engineering, The <strong>University</strong> of Tokyo, Tokyo, Japan,<br />

2 Institute of Picobiology, Graduate School of Lifescience, <strong>University</strong> of Hyogo, Hyogo, Japan<br />

These days, DNA has become attractive again to chemists as molecules having flexible functions.<br />

Recently, Tanaka et al. and many others have succeeded in developing metal ion containing artificial<br />

DNA systems [1]. While having similar structure as natural DNA nucleobases, the artificial DNA<br />

nucleobases posses distinctive shape, size and function. Furthermore, various metal ions in nature<br />

coordinate to DNA in unique ways [2]. Thus, the combination of metal ions and artificial DNA greatly<br />

improves the flexibility of DNA which will lead to progress in nano material science.<br />

After the pioneering work by Tanaka et al., there has been much effort by experimentalists to further<br />

understand the systems. There are also important contributions by theoretical chemists as well. For<br />

example, Voityuk [3] predicted that [T-Hg 2+ -T] paring plays important role in an excess electron<br />

transfer in DNA duplex, which was later experimentally reported. However, there are not many<br />

detailed molecular level analysis by X-ray crystal structures or theoretical calculations. Therefore, our<br />

present research aims to give understandings of these systems.<br />

Among various artificial DNA and metal ion systems, we focused on a system consisted of<br />

hydroxypyridone nucleobase (H) and Cu 2+ ion, [H-Cu 2+ -H] (Fig. 1). The Density Functional Theory<br />

(DFT) and Time-Dependent (TD)-DFT methods was used to calculate the ground state and excited<br />

states. First, we calculated the distance between two [H-Cu 2+ -H] units (Fig. 2).<br />

Fig. 1 The structure of [H-Cu 2+ -H] Fig. 2 Two stacking [H-Cu 2+ -H]<br />

Fig.1 One of the stable geometries of Au 7 and Au 8 predicted by CCSD(T)/spdsMCP.<br />

[1] M. Haruta et al., J. Catal. 1989, 115, 301.<br />

[2] G. C. Bond et al., Catalysis by gold, Imperial College Press (2006).,<br />

[3] R. M. Olson and M. S. Gordon, J. Chem. Phys., 2007, 126, 214310, and references there in.<br />

[4] Y. Osanai et al., Chem. Phys. Lett., 452, 210 (2008)., Y. Osanai et al., Chem. Phys. Lett., submitted (2008).<br />

[5] V. Bonifacic, S. Huzinaga, J. Chem. Phys., 1974, 60, 2779.<br />

Acknowledgement This research was primarily supported by the Ochadai Academic Production project<br />

operated by Japan Science and Technology agency (JST). H.M. and E.M. also should acknowledge JST-CREST<br />

for their great support. Some parts of the calculations were performed on CPU resources on RCCS (research<br />

center for computational science) at Institute for Molecular Science. Authors were grateful to the research center<br />

for giving us enough CPU time for this research.<br />

With the conventional DFT functional, there was no<br />

minimum when the distance was elongated. This result<br />

implies that the van del Waals interaction missing in the<br />

conventional functional is necessary. We confirmed this<br />

by adding the Andersson-Langreth-Lundqvist van del<br />

Waals correlation functional (ALL) [4]. The calculated<br />

result shown in Fig. 2 agrees with 3.7±0.1 Å obtained by<br />

EPR measurement (Fig. 3). In addition, the triplet and<br />

singlet states were nearly degenerate which also suggest<br />

that the stacking interaction between H is more important<br />

than the interaction between Cu 2+ . In the poster<br />

presentation, additional results on ground state properties<br />

and results on excited state of the system [5] will be<br />

presented in detail.<br />

Fig. 3 PES with and without ALL<br />

[1] Tanaka, K.; Clever, G. H.; Takezawa, Y.; Yamada, Y.; Kaul, C.; Shinoya, M.; Carell, T. Nature Nanotech.<br />

2006, 2, 190.<br />

[2] Matsui, T; Shigeta, Y.; Hirao, K. Chem. Phys. Lett. 2006, 423, 331.<br />

[3] Voityuk, A. A. J. Phys. Chem. B 2006, 110, 21010.<br />

[4] Sato, T.; Tsuneda, T.; Hirao, K.; J. Chem. Phys. 2007, 126, 234114.<br />

[5] Matsui,T.; Miyachi, H.; Shigeta, Y.; Hirao, K. submitting.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP419<br />

Diels-Alder Reaction of Crotonyl Phosphonate: Cylopentadiene versus Cyclohexadiene<br />

Nihan Çelebi-Ölçüm 1 , Viktorya Aviyente 1 , K. N. Houk 2<br />

1 Bogaziçi <strong>University</strong>, Department of Chemistry, Istanbul, Turkey, 2 <strong>University</strong> of California, Los Angeles<br />

Department of Chemistry and Biochemistry, CA, United States<br />

Diels-Alder reactions between a diene and a hetero-diene provide a good example for the<br />

periselectivity of the cycloadditions yielding two different products. We have recently found that in the<br />

thermal and SnCl 4 catalyzed reactions of crotonyl phosphonate with cyclopentadiene, the bispericyclic<br />

transition states lead to the formation of both Diels-Alder (DA) and hetero-Diels-Alder (HDA)<br />

products. The Lewis acid catalyst alters the shape of the potential energy surface, ultimately shifting<br />

the product distribution toward the hetero-Diels-Alder product. [1]<br />

H 3 C<br />

O<br />

P(OCH 3 ) 2<br />

O<br />

O<br />

CH 3<br />

O<br />

+<br />

P(OCH 3 ) 2<br />

O P(OCH 3 ) 2<br />

CH 3<br />

DA HDA<br />

In this study, the potential energy surface of the reaction between crotonyl phosphonate and<br />

cyclohexadiene is explored using Density Functional Theory in order to understand the poor<br />

periselectivity of the cycloaddition in contrast to the reaction between crotonyl phosphonate and<br />

cyclopentadiene [2].<br />

[1] Çelebi-Ölçüm, N.; Ess, D. H.; Aviyente, V.; Houk, K. N. J. Am. Chem. Soc. 2007, 129, 4528.<br />

[2] Hanessian, S.; Compain, P. Tetrahedron 2002, 58, 6521.<br />

O<br />

PP420<br />

Solving the Schrödinger Equation of the Hydrogen Molecular Ion in a Magnetic Field by the<br />

Free ICI (Iterative Complement Interaction) Method<br />

Atushi Ishikawa 1 , Hiroyuki Nakashima 2 , Hiroshi Nakatsuji 2<br />

1 Kyoto <strong>University</strong>, Kyoto, Japan, 2 Quantum Chemistry Research Institute, JST-CREST, Kyoto, Japan<br />

The free iterative complement interaction (ICI) method is applied to the hydrogen molecular ion (H 2 + ),<br />

previously in non-magnetic field [1], and here in a strong magnetic field. H 2 + in a strong magnetic field<br />

is interesting from both chemical and astronomical viewpoints, since it is considered to be a interstellar<br />

molecule that maybe in a strong magnetic field. The direction of the applied magnetic field is assumed<br />

to be parallel or perpendicular to the internuclear axis. The ground state total energies are calculated<br />

under various strengths of magnetic fields, and calculated energies are highly accurate for all field<br />

strengths, both parallel and perpendicular cases.<br />

The generated wave function forms depend on the direction of the magnetic field. Under the<br />

perpendicular magnetic field, our ICI wave function is represented by the mixture of various angular<br />

momentum quantum numbers. This clearly indicates that ICI method automatically takes into account<br />

the symmetry-breaking brought about by the applied magnetic field: our ICI procedure automatically<br />

generates the wave functions appropriate for the environment.<br />

The gauge-origin problem is also investigated. We have shown that the gauge-origin dependence of<br />

energy is removed by introducing the gauge-including initial function. This initial function holds its<br />

gauge-including property during the free ICI cycle. Consequently, our highly-accurate method can be<br />

applied to various atoms and<br />

molecules in the magnetic field<br />

without gauge-dependence problem.<br />

This problem is also investigated with<br />

the use of the gauge-dependent type<br />

initial function. The calculated<br />

energy, which is initially gauge-origin<br />

dependent, converges to the gaugeindependent<br />

result when a sufficient<br />

ICI order is applied. This property is<br />

also confirmed by the electron<br />

density plot (see FIGURE). These<br />

results indicate that the wave<br />

functions generated by the ICI<br />

method properly converges<br />

automatically to the exact wave<br />

function, from both energetic and<br />

gauge-dependent point of view.<br />

[1] Ishikawa, A.; Nakashima, H.; Nakatsuji, H.; J. Chem. Phys. 2008, 128, 124103


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP421<br />

Oxygen Atom Transfer Reactions of Iridium and Osmium Complexes: Theoretical Study of<br />

Significantly Large Differences between these Two Complexes<br />

Atsushi Ishikawa, Yoshihide Nakao, Hirofumi Sato, Shigeyoshi Sakaki<br />

Kyoto <strong>University</strong>, Kyoto, Japan<br />

Oxygen atom transfer reaction between O=ML 3 and ML 3 (eq. 1) (L=2,4,6-trimethylphenyl (mes) for<br />

M=Ir and L=tris(2,6-diisopropylphenyl)imide (TIPI) for M=Os) [1] was theoretically investigated by DFT<br />

method. The optimized geometry of the µ-oxo intermediate, (CH 3 ) 3 Ir-O-Ir(CH 3 ) 3 , is considerably<br />

different from the experimental one but that of (mes)Ir-O-Ir(mes) agrees well with the experimental<br />

one.<br />

ML 3 =O + ML 3 →ML 3 + O=ML 3 (1)<br />

Our computational results indicate that the bulky substituent of the ligand plays important rules to<br />

determine the geometry. The calculated activation barriers of the iridium and osmium systems are 2.8<br />

and 33.0 kcal/mol, respectively, which are consistent with the experimental results that the reaction<br />

easily occurs in the iridium system but with difficulty in the osmium system.<br />

The large difference in activation barriers arises from the nuclear and electronic factors.<br />

The nuclear factor is the sum of<br />

re-organization energies of ML 3<br />

and O=ML 3 which are defined as<br />

the energy difference between the<br />

reactant geometry and the<br />

transition state one. This nuclear<br />

factor is much larger in the<br />

osmium system than in the iridium<br />

system because the osmium<br />

reactants take trigonal planar<br />

structure. Therefore, significant<br />

geometrical change is needed for<br />

the osmium reactants to reach the<br />

transition state structure.<br />

Electronic factor arises from<br />

donor-acceptor interaction<br />

between O=ML 3 and ML 3 , and the<br />

energy gap between the donor<br />

orbital of ML 3 and the acceptor orbital of O=ML 3 is much larger in the osmium system than in the<br />

iridium system.<br />

[1] Fortner, K. C.; Laiter D. S.; Muldoon, J; Pu, L.; Braun-Sand, S. B.; Wiest, O.; Brown, S. N. J. Am Chem. Soc.<br />

2007, 129, 588–600.<br />

PP422<br />

Elucidating the Photoelectron Spectrum of CS by Molecular Vibration Calculations of CS +<br />

Nobumitsu Honjou<br />

Oita <strong>University</strong>, Oita, Japan<br />

The objectives of this study are (i) to determine vibrational frequency ω e and anharmonicity constant<br />

ω e x e for the C 2 Σ + state of CS + to investigate the cause of the discrepancy (64 cm -1 ) in the ω e value of<br />

the C state between the photoelectron (PE) spectroscopy value [1] and a recent theoretical value [2]<br />

obtained from polynomial fitting of ab initio configuration interaction (CI) energies, and (ii) to interpret<br />

the vibrational structure on the fourth band in the observed PE spectrum of CS [1] because there has<br />

been no detailed interpretation of the vibrational structure on the fourth band. We calculated potential<br />

energies for low-lying electronic states of CS + , and the 1 1 Σ + state of CS by the ab initio CI method and<br />

also calculated molecular vibrations for the electronic states in which the vibrational Schrödinger<br />

equation was solved using the potential energy functions [3]. The calculated results were used to<br />

deduce spectroscopic constants, including the vibrational constants for the CS + electronic states, and<br />

to compute Franck-Condon factors so as to provide the PE vibrational intensity distributions (VID) for<br />

ionization from the CS 1 1 Σ + state to the CS + electronic states. The VID allows comparison of both the<br />

energy levels and intensities of vibrational components with the experimental vibrational structure. The<br />

ALCHEMY II computer program [4] was used for the CI calculations.<br />

The present study reveals the following information [3].<br />

(1) The calculated potential energies and vibrational term values for the X 1 2 Σ + , A 1 2 Π, and B 2 2 Σ +<br />

states of CS + are accurate because the theoretical spectroscopic constants (electronic term value<br />

T e , equilibrium internuclear distance R e , ω e , and ω e x e ) for the three states are in good agreement<br />

with experiment.<br />

(2) The C 3 2 Σ + state has a double potential well, whose inner well accommodates essentially two<br />

vibrational states with a significant anharmonicity of vibration. The ω e and ω e x e values for the innerwell<br />

are calculated to be 1101 cm -1 and 46.7 cm -1 , respectively. The present ω e value is 46 cm -1<br />

larger than the PE spectroscopy value [1]. No experimental ω e x e value for the C state has been<br />

available. The discrepancy in ω e might be partly due to the experimental determination of ω e from<br />

only the observed PE spectrum because of an insufficient number of inner-well vibrational states to<br />

estimate the anharmonicity constant.<br />

(3) The first and second components of the fourth band in the observed PE spectrum are respectively<br />

assigned to the transitions to the C 3 2 Σ + v’=0 and v’=1 states occupying essentially the inner-well,<br />

where v’ denotes vibrational quantum number. Two transitions to the C 3 2 Σ + v’=4 and v’=5 states<br />

extending across both wells are responsible for the third component. Experiments to resolving the<br />

splitting of the third component into two peaks should provide evidence of the double potential well.<br />

[1] Frost D.C.; Lee S.T.; and McDowell C.A., Chem. Phys. Letters 1972, 17, 153-156.<br />

[2] Honjou N., Chem. Phys. 2006, 324, 413-419.<br />

[3] Honjou N., Chem. Phys. 2008, 344, 128-134.<br />

[4] McLean A.D.; Yoshimine M.; Lengsfield B.H.; Bagus P.S.; and Liu B., in: E. Clementi (Ed.), Modern<br />

Techniques in Computational Chemistry, Escom, Leiden, 1990, 593-638.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP423<br />

DFT Studies into Conjugate Transfer Hydrogentations of Hantzsch Esters<br />

Robert S Paton, Jonathan M Goodman<br />

<strong>University</strong> of Cambridge, Cambridge, United Kingdom<br />

Recent experimental studies have shown that the creation of new C-H stereogenic centres can be<br />

accomplished using the blueprints of biochemical reductions [1]. Nature’s reducing agents can be<br />

mimicked synthetically by small molecule amine catalysts and Hantzsch ester pyridines. The use of<br />

chiral amines, and more recently, chiral amine salts results in highly regio- and stereoselective<br />

hydrogenations of unsaturated enals and enones. The mechanism and competing transition structures<br />

for this process have been studied using DFT calculations.<br />

Calculations show that formation of reactive iminium species leads to a significant reduction in the<br />

energetic barrier to conjugate hydrogenation. The nature of the competing transition structures have<br />

been probed using a variety of spin-restricted and unrestricted density functionals and van der Waals<br />

corrected functionals. The effects of continuum solvation on TS geometries and energies have also<br />

been studied. The stereoinduction observed due to chiral amines is explained in terms of competing<br />

transition structures in which the participation of the amine’s counterion is also critical. For reactions<br />

that employ a chiral phosphate counterion, our calculations show the lowest energy pathway involves<br />

bidentate coordination to protons in the Hantzsch ester and also in the iminium substrate. This “three<br />

point contact model” (which can also be applied to the direct hydrogenation of imines [2]) leads to a<br />

successful explanation for the observed enantioselectivity.<br />

PP424<br />

Theory of Paramagnetic NMR Chemical Shift in the Presence of Zero-Field Splitting<br />

Teemu O. Pennanen, Juha Vaara<br />

Laboratory of Physical Chemistry, Department of Chemistry, <strong>University</strong> of Helsinki, Helsinki, Finland<br />

NMR spectroscopy of open-shell, paramagnetic molecules is of interest when studying, e.g.<br />

,metalloprotein systems in natural solution environment. Computational determination of NMR<br />

shielding tensor is a valuable tool to help structure determination, but until recently, theory only existed<br />

for the special cases of double systems [1] and (as an approximate a posteriori correction) cylindrically<br />

symmetric systems of higher multiplicity [2].<br />

We have developed a general and systematic theory of paramagnetic NMR shielding tensor for<br />

paramagnetic systems in arbitrary spin state and spatial symmetry [3]. As a result of including zerofield<br />

splitting of energy levels, all the contributions [1] to paramagnetic part of shielding tensor have<br />

contributions of more than one tensorial rank, e.g., the contact shift has an anisotropic part, anad there<br />

is an isotropic leading-order dipolar shift.<br />

The theory was implemented into a program that acquires the g-, hyperfine and zero-field splitting<br />

tensors from a calculation with ORCA [4] software and the orbital shielding tensor from calculation with<br />

Gaussian 03 [5]. This program is easily adaptable to use tensor data from any quantum chemistry<br />

code available.<br />

[1] Pennanen, T.O.; Vaara, J. J. Chem. Phys. 2005, 123, 174102.<br />

[2] Hrobarik, P.; Reviakine, R.; Arbuznikov, A. V.; Malkina, O.; Malkin, V. G.; Köhler, F. H.; Kaupp, M. J. Chem.<br />

Phys. 2007, 126, 024107.<br />

[3] Pennanen, T. O.; Vaara, J. Phys. Rev. Lett. 2008, 100, 133002.<br />

[4] Neese, F. Orca, Version 2.6.4; <strong>University</strong> of Bonn, 2007.<br />

[5] Frisch, M. J. et al., Gaussian 03, Revision C.02; Gaussian Inc.; Wallingford CT, 2004.<br />

[1] Ouellet, S. G.; Walji, A. M.; Macmillan, D. W. C. Acc. Chem. Res. 2007, 40, 1327-1339.<br />

[2] Simón, L.; Goodman, J. M. J. Am. Chem. Soc. 2008, ASAP article.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP425<br />

Reversible Interconversion of H 2 and H 2 O by Hydroxo/Sulfido-Bridged Dinuclear Ruthenium-<br />

Germanium Complex. Theoretical Study<br />

Noriaki Ochi, Hirofumi Sato, Yoshihide Nakao, Shigeyoshi Sakaki<br />

Kyoto <strong>University</strong>, Department of Molecular Engineering, Kyoto, Japan<br />

The activation of dihydrogen molecule with transition metal complex is interesting as a model of the<br />

hydrogen metabolism reaction mediated by enzyme such as hydrogenases, in which H 2 is reversibly<br />

converted to two electrons and two protons. Although several activations of H 2 by enzyme models<br />

were reported, the reversible interconversion between H 2 and H 2 O has been limited. Recently, the<br />

reversible interconversion by [(Dmp)(Dep)Ge(µ-S)(µ-OH)Ru(PPh 3 )] + (Dmp=2,6-dimesitylphenyl,<br />

Dep=2,6-diethylphenyl) 1 was experimentally reported by Matsumoto and Tatsumi et at [1].<br />

We theoretically investigated the interconversion with ONIOM(DFT:MM) method, to understand the<br />

reaction mechanism and the reason why 1 efficiently catalyzes this reaction. Because 1 contains the<br />

reactive µ-OH and µ-S groups, there are two possible reaction courses, as shown in Figure 1. In path I,<br />

the reaction occurs on the µ-OH, in which the rate determining step is dissociation of H 2 O (TS OH ), and<br />

its activation barrier is 19.6 kcal/mol. In path II, the reaction occurs on the µ-S, in which the rate<br />

determining step is isomerization (TS S ) of the H atom concomitant with PPh 3 dissociation and its<br />

activation barrier is 32.1 kcal/mol. These results indicate that the interconversion takes place on the<br />

µ-OH.<br />

In TS OH , H 2 O recedes from the Ge atom to afford the product complex (PRD), which indicates the<br />

important role of the Ge atom in the interconversion reaction. To shed some light on the role of the Ge<br />

atom in 1, we investigated the same reaction with the Si analogue [(Dmp)(Dep)Si(µ-S)(µ-OH)Ru<br />

(PPh 3 )] + 2, in which the Ge atom is replaced with Si atom. Except for the transition state (TS Si ) of<br />

dissociation of H 2 O, the energy changes of 2 are similar to those of 1. The activation barrier for TS Si is<br />

30.8 kcal/mol, which is considerably larger than that of 1. This difference is interpreted in the term of<br />

E-OH 2 (E=Ge and Si) bond energy; Ge-OH 2 and Si-OH 2 bond energies are estimated 26.6 and 36.9<br />

kcal/mol, respectively. Because the Ge-OH 2 bond energy is moderate, 1 efficiently catalyzes the<br />

interconversion of H 2 and H 2 O.<br />

PP426<br />

Theoretical Study of the Hydrogen Bonds in Functionalized Molecules using Multi-Component<br />

Molecular Orbital (mc_mo) Method<br />

Masato Kaneko 1 , Taro Udagawa 2 , Masanori Tachikawa 1<br />

1 Graduate School of Integrated Science, Yokohama-City <strong>University</strong>, Yokohama, Japan, 2 Faculty of<br />

Engineering, Gifu <strong>University</strong>, Gifu, Japan<br />

We have recently developed the multi-component molecular<br />

orbital (MC_MO) methods [1], which can directly include nuclear<br />

quantum effect of light particles such as proton. We have clearly<br />

demonstrated that the nuclear quantum effect is important for<br />

adequate description of the various systems in which hydrogen<br />

atom or hydrogen nucleus (proton) takes an important role. In this<br />

research, we applied MC_MO method to hydrogen bonds in HIV-1<br />

Protease (HP) and Naphthalocyanine molecule to theoretically<br />

analyze a role of hydrogen bonds for such systems.<br />

HP is an important enzyme for AIDS therapeutics, because it<br />

dominates the replication of HIV. Porter et al. theoretically found<br />

that one water molecule exists between two Asp residues in HP’s<br />

active site, and three hydrogen bonds stabilize the geometry [2].<br />

Although the water molecule is thought to be important for<br />

electronic state and enzymatic reaction in HP’s active site, the<br />

active site has not been theoretically analyzed sufficiently, and it<br />

has not been revealed which of three bonds is directly involved in<br />

its reaction.<br />

On the other hands, Naphthalocyanine (Figure 1) is expected as a monomeric device owing to double<br />

proton transfer reaction in the center of molecule. Liljeroth et al. [3] found current-induced hydrogen<br />

tautomerization using STM. By this reaction, the shape of molecular orbital and direction of electrical<br />

current in Naphthalocyanine change, and this molecule may serve as single-molecule switch.<br />

Additionally, it is well known that quantum mechanical treatment of proton is important for hydrogen<br />

(or proton) tautomerization. However, there are no theoretical reports on this tautomerization with<br />

using multi-component methods.<br />

r1<br />

r2<br />

Figure 1. Optimized structure of<br />

Naphthalocyanine<br />

In order to reveal a role of nuclear quantum nature of proton for these systems, we have theoretically<br />

analyzed the geometrical isotope effect on these systems with our MC_MO method. In addition, we<br />

have calculated the potential energy surface for Naphthalocyanine molecule for detailed analysis of<br />

hydrogen(proton) tautomerization reaction.<br />

The optimized covalent bond lengths and GIEs of Naphthalocyanine are shown below. As a result of<br />

anharmonicity of the potential, which is included by MC_MO method, N-H covalent bond length in HH<br />

is longer than that in conventional. Detailed results will be presented on the day.<br />

Figure 1. Rate determining step in the interconversion of H 2 and H 2O on µ-OH and µ-S calculated with<br />

ONIOM(DFT:MM). The geometry in square bracket means the transition state of rate determining step.<br />

The energies are in parenthesis (kcal/mol).<br />

TABLE: Optimized Covalent Bond Length and Calculated GIEs a .<br />

HF DFT<br />

MC_HF<br />

(HH) (HH) HH HD DD<br />

r1 0.999 1.019 1.018 1.018 (H→D) 1.012<br />

r2 0.998 1.019 1.015 (H→D) 1.008 1.008<br />

(HH: non-substituted HD: mono-substituted DD: di-substituted)<br />

a Basis sets: 6-31g**(center)/6-31g(outer) for electron, 1sGTF for proton/deuteron<br />

[1] Matsumoto, T.; Nakaya, Y.; Itakura, N.; Tatsumi, K. J. Am. Chem. Soc., 2008, 130, 2458-2459.<br />

[1] Tachikawa, M.; Mori, K.; Nakai, H.; Iguchi, K. Chem. Phys. Lett.1998, 290, 437-442.<br />

[2] Porter, M. A.; Molina, P. A. J. Chem. Theor. Comp. 2006, 2, 1675-1684.<br />

[3] Liljeroth, P.; Repp, J.; Meyer, G. Nature, 2007, 317, 1203-1206.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP427<br />

Role of Electrodes in Characterizing Transport Properties of a Molecular Device<br />

Yeonchoo Cho, Woo Youn Kim, Kwang Soo Kim<br />

Center for Superfunctional Materials, Department of Chemistry, Pohang <strong>University</strong> of Science and<br />

Technology, Pohang, Korea, Republic of<br />

Current-voltage (I-V) characteristics of a molecular device have been extensively studied. To design<br />

novel devices, it is important to manipulate transport properties of the system composed of a<br />

molecular core, electrodes sandwiching the molecule, and linkers connecting them. Although there are<br />

various studies focusing on the role of molecular cores and linkers, few studies have focused on the<br />

role of electrodes in characterizing transport properties.<br />

We have made first principles calculations on systems with different electrodes using the density<br />

functional theory code where the non-equilibrium Green function formalism is implemented [1]. Gold,<br />

ruthenium, and carbon nanotube electrodes are studied, each representing s, d, and p valence metals,<br />

respectively. Using the simplest alkyne molecule, the transmission coefficients and the I-V curves are<br />

calculated. We have found that the transport properties considerably depend on electrode materials.<br />

The gold electrode, with the s valence, shows high conductance due to highly broadened transmission<br />

peaks, whereas the ruthenium electrode having the d valence is less effectively coupled to the<br />

molecule than the gold electrode regardless of linkers. In contrast, the carbon nanotube electrode,<br />

with the p valence, displays discrete bands reflecting molecular features.<br />

PP428<br />

The CASPT2 Method: Current Limitations and Benchmarks.<br />

Valera Veryazov<br />

Lund <strong>University</strong>, Theoretical Chemistry, Lund, Sweden<br />

Multiconfigurational second-order perturbation method CASPT2 is known as a reliable computational<br />

tool for the electronic structure calculations. However, quite often this method is associated with<br />

calculations of very small molecules.<br />

Recent development of CASSCF and CASPT2 codes in Molcas 7 package allows us to extend the<br />

limits for the calculations which can be produced with these methods. The purpose of this study is to<br />

demonstrate the possibility to use CASSCF/CASPT2 approach for relatively large molecules, and<br />

molecular systems. Benchmark tests for CASSCF and CASPT2 implementation in Molcas 7 package<br />

are also presented.<br />

[1] Kim, W. Y.; Kim, K. S. J. Comput. Chem. 2008, 29, 1073-1083.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP429<br />

Cycloaddition Reactions of ICNO: A Theoretical and Experimental Study<br />

Melinda Krebsz 1 , Tibor Pasinszki 1 , Balázs Hajgató 2<br />

1 Department of Inorganic Chemistry, Institute of Chemistry, Eötvös Loránd <strong>University</strong> Budapest,<br />

Budapest, Hungary, 2 Department SBG, Hasselt <strong>University</strong>, Diepenbeek, Belgium<br />

Small nitrile oxides (X―C≡N→O) are important molecules in synthetic organic chemistry, and are<br />

widely utilized for dipolar cycloaddition reactions, in particular for the formation of various substituted<br />

heterocycles. They are unstable in the pure state and solutions, thus they are generated in situ at the<br />

presence of the appropriate dipolarophile. ICNO is a promising precursor for cycloaddition reactions,<br />

however, its generation is not known so far.<br />

This paper discusses the theoretical investigation of cycloaddition reactions of ICNO with various<br />

dipolarophiles (see below) and the dimerisation process, as well as our first experimental results for<br />

the same reactions. The mechanism of the cycloaddition rections have been studied using single- and<br />

multi-reference coupled-cluster methods, as well as density functional theory. Our results indicate a<br />

multi-step dimerisation process for ICNO, but a synchronous cycloaddition with nitriles and ethynyl<br />

derivatives.<br />

PP430<br />

Charge-State Dependent Hydrogen Diffusion on Silicon (001)<br />

Oliver Warschkow<br />

Centre for Quantum Computer Technology, School of Physics, The <strong>University</strong> of Sydney, Sydney,<br />

NSW, Australia<br />

The diffusion of hydrogen atoms is an important reaction in a number of chemical and technological<br />

processes of the silicon (001) surface. This includes the dissociative adsorption of molecules, the<br />

growth of overlayers by chemical vapor deposition (CVD), thermal desorption of various molecules<br />

(including H 2 ) from the surface, and the directed atomic-scale functionalization of the surface by<br />

scanning tunnelling microscopy (STM) hydrogen lithography. The basic inter- and intradimer hydrogen<br />

shift reactions are well studied theoretically (e.g. Refs. 1, 2). STM measurements [2, 3] have yielded<br />

experimental activation energies of ~1.7 eV and between 1.0 and 1.4 eV for inter- and intradimer H-<br />

shift, respectively. With this poster we pose and discuss two pertinent questions: (1) Are single energy<br />

barriers adequate to describe these H-shift reactions, and (2) are STM measurements of H-diffusion<br />

truly representative for hydrogen desorption in the absence an STM tip. These two questions warrant<br />

examination because hydrogen adatoms on Si(001) are known [5-7] to adopt a variety of charge<br />

states depending on factors such as the doping level of the bulk, the defect density on the surface,<br />

and – critical to the STM diffusion measurements – the STM tip to surface bias voltage. Using a highlevel<br />

“cluster compound model” [8] approach which delivers approximate formation energies at the<br />

B3LYP/6-311++(2df,2pd)//B3LYP/6-311++(d,p) level for a large Si 53 H 44 surface cluster, we report<br />

activation energies of diffusion of hydrogen adatoms in positive, neutral, and negative charge states.<br />

We correlate our results with the experimental literature and discuss the effects of charging on<br />

diffusion rates.<br />

[1] Vittadini, A.; Selloni, A.; Casarin, M. Phys. Rev. B 1995, 52, 5885.<br />

[2] Nachtigall, P.; Jordan, K.D. J. Chem. Phys. 1995, 102, 8249.<br />

[3] Owen, J.H.G; Bowler, D.R.; Goringe, C.M.; Miki, K.; Briggs, G.A.D. Phys. Rev. B 1996, 54, 14153.<br />

[4] Hill, E.; Freelon, B.; Ganz, E. Phys. Rev. B 1999, 60, 15896.<br />

[5] Reusch, T.C.G.; Warschkow, O.; Radny, M.W.; Smith, P.V.; Marks, N.A.; Curson, N.J.; McKenzie, D.R.;<br />

Simmons, M.Y. Surf. Sci. 2007, 601, 4036.<br />

[6] Radny, M.W.; Smith, P.V.; Reusch, T.C.G,; Warschkow, O.; Marks, N.A.; Wilson, H.F.; Schofield, S.R.;<br />

Curson, N.J.; McKenzie, D.R.; Simmons, M.Y. Phys. Rev. B 2007, 76, 155302.<br />

[7] See also the gating of a prototype molecular transistor by a change of charge-state of a missing hydrogen<br />

defect on a hydrogenated Si(001) surface reported in Piva, P.G.; DiLabio, G.A; Pitters, J.L.; Zikovsky, J.; Rezeq,<br />

M.; Dogel, S.; Hofer, W.A.; Wolkow, R.A. Nature 2005, 435, 658<br />

[8] Warschkow, O.; McDonnell; T.L.; Marks, N.A. Surf. Sci. 2007, 601, 3020.<br />

Acknowledgement: Hungarian Scientific Research Fund (OTKA T049148)


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP431<br />

Exploiting Next-Gen Computers for Computational Quantum Chemistry<br />

Tirath Ramdas 1 , Gregory Egan 1 , David Abramson 2 , Kim Baldridge 3<br />

1 Monash <strong>University</strong>, Center for Telecommunications and Information Engineering, Melbourne, VIC,<br />

Australia, 2 Monash <strong>University</strong>, Center for Distributed Systems and Software Engineering, Melbourne,<br />

VIC, Australia, 3 <strong>University</strong> of Zürich, Organic Chemistry Institute, Zürich, Switzerland<br />

The merit of the Cell BE (e.g. Playstation 3), graphics processing units (GPUs), and fieldprogrammable<br />

gate arrays (FPGAs) for scientific applications are now sufficiently hyped [1]. That<br />

these devices are capable of impressive performance is widely accepted – the difficulty is in exploiting<br />

this capability. The best performance is usually only obtained with a highly detailed understanding of<br />

the specific architecture of the targeted device. However, there is a common feature to these<br />

architectures (and this a key feature of the Cell BE and GPU, in particular) that may be exploited in<br />

general: data-parallel processing, commonly known as single-instruction multiple-data (SIMD)<br />

processing.<br />

Data-parallelism is a highly efficient form of parallelism; more efficient that the ubiquitous threadparallelism<br />

that is exploited by computational clusters and other conventional parallel processors.<br />

Unfortunately, data-parallelism is not available to many applications. Ab initio computational quantum<br />

chemistry tools (such as GAMESS, Gaussian, etc.) contain computational bottlenecks that belong to a<br />

general class of operations called tensor contractions [2]: these are SIMD-friendly. However, there is<br />

another bottleneck – electron repulsion integrals – which are not SIMD-friendly.<br />

A key thrust of our research has been to overcome this problem through a run-time dynamic workload<br />

reordering approach, where it has been shown that substantial SIMD processing of ERIs is possible<br />

[3]. With that basic issue addressed, we expand our scope of interest to consider other issues, such<br />

as:<br />

1. How to best exploit the large internal bandwidth in next-generation architectures?<br />

PP432<br />

Thermal Isomerization of 11-cis-Retinal. An Unexpected Manifold<br />

Carlos Silva Lopez, Rosana Alvarez Rodriguez, Marta Dominguez Seoane, Olalla Nieto Faza, Angel<br />

R. de Lera<br />

Universidade de Vigo, Galicia, Spain<br />

The native chromophore responsible for vision in vertebrates, 11-cis-retinal, responds very distinctively<br />

to photochemical and thermal conditions. In the vertebrate eye, ligated as a protonated Schiff base to<br />

the opsin apoprotein, and under UV-vis irradiation, 11-cis-retinal undergoes double bond isomerization<br />

to yield efficiently and quantitatively the protonated Schiff base of all-trans-retinal [1]. In solution, under<br />

thermal conditions, 11-cis-retinal is transformed into a mixture of two isomers, namely 13-cis-retinal<br />

and all-trans-retinal in a ~65:35 ratio.<br />

all-trans-retinal<br />

CHO<br />

hν<br />

11-cis-retinal<br />

CHO<br />

80 ºC<br />

+<br />

CHO<br />

13-cis-retinal<br />

all-trans-retinal<br />

13-cis/all-trans ~ 65:35<br />

Control experiments on 13-cis and all-trans-retinal suggest that these two compounds do not<br />

interconvert under thermal conditions, indicating that two independent reaction pathways are operating<br />

to furnish each isomer. Furthermore, deuterium labeling experiments devised to probe if sigmatropic<br />

rearrangements are involved in the product formation provide an intriguing inversion in product<br />

distribution and, to some extent, deuterium scrambling.<br />

CHO<br />

2. How to overcome constraints such as limited memory size in off-load compute engines?<br />

Note that this is a significant reversal in conventional thinking; in the past the implicit assumption has<br />

always been that processors have access to a large memory space but that bandwidth was limited!<br />

We also consider the impact internal interconnection topologies would have on dataflow at a finer<br />

granularity than we have become accustomed to with conventional coarse-grained parallelism. We<br />

pose many of these questions, and consider their implications on the way ab initio computational<br />

quantum chemistry codes should be designed and/or augmented.<br />

CD 3<br />

11-cis-retinal<br />

CHO<br />

80 ºC<br />

CD 3-y (H) y<br />

13-cis-retinal<br />

CD 3<br />

D(H)<br />

+<br />

CHO<br />

all-trans-retinal<br />

13-cis/all-trans ~ 30:70<br />

CHO<br />

[1] See, for example: Williams et al.; Proceedings of ACM Computing Frontiers, 2006, 9-20.<br />

[2] Baumgartner et al; Proceedings of the IEEE, 2005, 93, 276-292.<br />

[3] See, for example, Ramdas, T; Egan, GK; Abramson, D; Baldridge, KK; Comp Phys Commun, 2008, 178, 817-<br />

834.<br />

Experimental kinetic studies and computational modeling are employed to unveil a complex<br />

mechanistic manifold implicated in the thermal isomerization of 11-cis-retinal. The absence or<br />

presence of label scrambling and the inversion of populations upon deuterium substitution can also be<br />

explained via synergistic NMR monitoring experiments and molecular modeling.<br />

[1] This is one of the fastest and most efficient reactions recorded, taking place in only 200 fs. with a quantum<br />

yield of 0.65. Yan, E. C. Y.; Ganim, Z.; Kazmi, M. A.; Chang, B. S. W.; Sakmar, T. P.; Mathies, R. A. Biochemistry<br />

2004, 43, 10867-10876.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP433<br />

Structures and Infrared Spectra of the Topology-Distinct Protonated Water Clusters<br />

H 3 O + (H 2 O) n-1 (n ≤ 6)<br />

Maihemutijiang Jieli, Misako Aida<br />

Center for Quantum Life Sciences & Graduate School of Science, Hiroshima <strong>University</strong>, Higashi-<br />

Hiroshima, Hiroshima, Japan<br />

1. Introduction<br />

Rooted digraphs were used to represent the features of the hydrogen bond (HB) and we enumerated<br />

all possible topology-distinct patterns corresponding to protonated water (PW) clusters containing up<br />

to 8 water molecules by means of the graph theory. From close investigation of the structural patterns<br />

obtained, several restrictions which should be satisfied in the stable structures of PW clusters were<br />

found. The generated hydrogen bond (HB) matrices of the restrictive rooted digraph were used as the<br />

theoretical framework; the local minima on the potential energy surfaces of those PW clusters are<br />

obtained using ab initio MO method and DFT method. For PW pentamers and hexamers we found<br />

some new local minimum structures which had not been obtained previously. The characteristic IR<br />

harmonic vibrational frequencies of stretching modes of different HB types of OH in PW clusters are<br />

generated systematically using the MP2/aug-cc-pVDZ level of theory.<br />

2. Computational Methods and Results<br />

The H-B matrix is used to enumerate all the possible structures, in which the hydrogen bonding<br />

patterns are different. We found several restrictions which should be satisfied in the stable structures<br />

of PW clusters given belove (we call a vertex which corresponds to a protonated water molecule P-<br />

vertex, and a vertex which corresponds to a water molecule W-vertex):<br />

(1) There is no arrow directed toward the P-vertex.<br />

(2) The number of the arrows directed from the P-vertex is 2 or 3.<br />

(3) When two arrows are directed from the P-vertex, a W-vertex which accepts an arrow from the<br />

P-vertex cannot accept any arrow from other vertex.<br />

(4) When three arrows are directed from the P-vertex, all of the three W-vertices, each of which<br />

accepts an arrow from the P-vertex, cannot accept other arrow from other vertex.<br />

The OH stretching modes are divided into eight types by analyzing the characteristic IR frequencies of<br />

local minimum structures of PW clusters H 3 O + (H 2 O) n-1 (n=2~7).<br />

3. Conclusion<br />

We showed here the systematical method to find all possible structures of PW clusters. Combination<br />

of graph theoretical enumerations with ab initio MO calculations allows us to find all topology-distinct<br />

stable structures for PW clusters. We found some new PW pentamer and hexamer structures.<br />

Stretching modes of different OH bonds in local minima of PW clusters are generated systematically at<br />

the MP2/aug-cc-pVDZ level of theory. The vibrational frequencies distribution and their changing order<br />

of different types of OH in PW clusters are classified systematically. Vibrational frequencies of different<br />

OH types agree reasonably well with the recent experimental and theoretical results.<br />

PP434<br />

The Bonding Nature of Dinuclear Cr(II) Complexes. A Theoretical Study with the MRMP2<br />

Method<br />

Yusaku Kurokawa, Yoshihide Nakao, Shigeyoshi Sakaki<br />

Department of Molecular Engineering, Graduate School of Engineering, Kyoto <strong>University</strong>, Kyoto City,<br />

Japan<br />

Metal-Metal multiple bond is one of the interesting and challenging research targets in inorganic,<br />

physical, and theoretical chemistries.[1] Chromium dimer, Cr 2 , is of considerable interest because it is<br />

believed to bear a hextuple Cr-Cr bond, in a formal sense, which is the largest bond order at this<br />

moment. Also, theoretical calculation of similar Cr dinuclear complex is very challenging because of<br />

the presence of very large static correlation effects.<br />

In this study, recently synthesized open-lantern type dinuclear chromium complex, [Cr(R 1 -<br />

NC(R 2 )NR 3 ) 2 ] 2 (R 1 = Et, R 2 = CH 3 , R 3 = t Bu, 1) [2], was theoretically investigated with DFT, CASSCF,<br />

and MRMP2 methods.<br />

R 2<br />

R<br />

R 2<br />

1<br />

N<br />

R 1<br />

R<br />

N<br />

N N R 3<br />

3<br />

Cr Cr<br />

R 1<br />

N<br />

N R<br />

R 3<br />

1<br />

N<br />

R 2<br />

NR R3 2<br />

1<br />

First, we optimized the structure of 1 with the DFT method at various Cr-Cr distances. In the DEToptimized<br />

geometry, the potential energy surface (PES) smoothly decreases as the Cr-Cr distance<br />

decreases, whereas the optimized Cr-Cr distance (1.76 Å) is too short. On the other hand, that of<br />

CASSCF does not present a minimum in the range of the Cr-Cr distance from 1.75 to 2.15 Å. The<br />

MRMP2 method successfully presents a minimum at Cr-Cr distance of 1.851 Å, as shown in Fig. 1,<br />

though it is a little shorter than the experimental value. These suggest that both static and dynamical<br />

correlations are important in this complex.<br />

The occupation numbers of d σ , d π1 , and d π2 orbitals are evaluated to be 1.72, 1.70, and 1.69,<br />

respectively, and that of d δ was 1.30. The bond order between two Cr atoms is 2.40 with the CASSCF<br />

method, which is much smaller than the formal bond order (4) of this complex.<br />

Significantly large differences are observed between this dinuclear chromium complex and Mo<br />

analogue. In the Mo analogue, the DFT, CASSCF, and MRMP2 methods present almost the same<br />

Mo-Mo equilibrium distance, as shown in Fig. 2. The Mo-Mo bond order was evaluated to be 3.41,<br />

which is considerably smaller than the formal value but much larger than the Cr-Cr bond order. The<br />

reason of the differences will be discussed.<br />

Fig. 1 Fig. 2<br />

Jieli M.; Miyake, T.; Aida, M. Bull. Chem. Soc. Jpn. 2007, 80, 2131–2136.<br />

[1] F. Weinhold and C. R. Landis Science, 2007, 316, 61.<br />

[2] A. R. Sadique; M. J. Heeg; C. H. Winter; J. Am. Chem. Soc., 2003, 125, 7774.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP435<br />

Sampling Enhancement for Ab Initio Monte Carlo Calculations of Molecular Clusters<br />

Akira Nakayama, Tetsuya Taketsugu<br />

Division of Chemistry, Graduate School of Science, Hokkaido <strong>University</strong>, Sapporo, Japan<br />

An approach is developed to enhance sampling for ab initio Monte Carlo and ab initio path integral<br />

Monte Carlo calculations of molecular clusters by employing an auxiliary Markov chain to move<br />

configuration space efficiently. In this scheme, an auxiliary Markov chain moves on an approximate<br />

and computationally cheap potential constructed by the interpolation method. The ab initio potential<br />

energies at previous steps are used as reference points for interpolation and these reference points<br />

are updated dynamically in the simulation.<br />

We apply this scheme to protonated water clusters and demonstrate its dramatic enhancement in<br />

sampling efficiency. Also, combination of this technique with other methods, such as umbrella<br />

sampling and parallel tempering, is presented in order to further speed up the convergence.<br />

PP436<br />

Complexity under Control: A Theoretical Mechanistic Study of Gold and Platinum Catalyzed<br />

Rearrangements<br />

Olalla Nieto Faza, Adán Gonzalez Perez, Carlos Silva Lopez, Angel R. de Lera<br />

Universidade de Vigo, Vigo, Spain<br />

The use of gold and platinum as homogeneous catalysts in organic transformations has exploded in<br />

the last years. Their dual activity as electron donors and acceptors, which can be attributed to<br />

relativistic effects [1], provides an excellent scaffold for facile transformations involving the activation of<br />

diverse functional groups. As a result, they have been extensively used to mediate in complex<br />

rearrangements that sometimes coexist with an exquisite stereocontrol.<br />

We have used Density Functional Theory calculations to study the mechanism of two of these<br />

remarkable transformations, namely the Pt-catalyzed pentannulation of propargylic esters containing<br />

oxirane or aziridine moieties reported by Sarpong et al. and the Au-catalyzed<br />

carboalkoxylation/Claisen rearrangement reported by Toste et al. [2]. The theoretical study highlights<br />

the role of the metal in mediating complex, multistep rearrangements of polyfunctionalized substrates,<br />

occurring with chirality transfer.<br />

[1] Gorin, D. J.; Toste, F. D. Nature, 2007, 446, 395-403.<br />

[2] (a) Pujanauski, B. G.; Bhanu Prasad, B. A.; Sarpong, R. J. Am. Chem. Soc., 2006, 128, 6786-6787. (b)<br />

Motamed, M.; Bunnelle, E. M.; Singaram, S. W.; Sarpong, R. Org. Lett., 2007, 9, 2167-2170. (c) Dubé, P.; Toste,<br />

F. D. J. Am. Chem. Soc., 2006, 128, 12062-12063.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP437<br />

Hartree-Fock and Kohn-Sham Response Theory in a Second-Quantization Atomic-Orbital<br />

Formalism Suitable for Linear Scaling<br />

Thomas Kjaergaard 1 , Poul Jørgensen 1 , Jeppe Olsen 1 , Sonia Coriani 2 , Trygve Helgaker 3<br />

1 Aarhus <strong>University</strong> Department of Chemistry, Aarhus C, Denmark, 2 Dipartimento di Scienze Chimiche<br />

Universita degli Studi di Trieste, Trieste, Italy, 3 Department of Chemistry, <strong>University</strong> of Oslo, Oslo,<br />

Norway<br />

We present a second-quantization based atomic-orbital method for the computation of response<br />

functions within Hartree–Fock and Kohn–Sham density-functional theories. The method scales linearly<br />

with the size of the system. Illustrative results are presented for excitation energies, one- and twophoton<br />

transition moments, polarizabilities and hyper-polarizabilities for hexagonal BN sheets with<br />

upto 180 atoms.<br />

PP438<br />

Scrutiny of the Mean-Field Approximation to the Two-Electron Spin-Dependent Terms in the<br />

Two-Component Pseudo-Relativistic Hamiltonian<br />

Mojmir Kyvala<br />

Institute of Organic Chemistry and Biochemistry, Academy of Sciences of the Czech Republic,<br />

Flemingovo namesti 2, 166 10 Praha 6, Czech Republic<br />

A series of ab initio one-electron approximations to the Breit−Pauli two-electron spin-spin Hamiltonian<br />

has been developed under the well established scheme for the derivation of the mean-field spin-orbit<br />

Hamiltonian [1]. Mainly for comparison, similar approximations to the two-electron part of the<br />

nonrelativistic Hamiltonian have also been defined. It is shown that the single-determinant<br />

approximation to the diagonal matrix elements of the spin-spin Hamiltonian (corresponding to the<br />

single-determinant open-shell Hartree−Fock approximation to the expectation values of the<br />

nonrelativistic Hamiltonian) introduced almost fifty years ago [2] and exploited recently in both density<br />

functional theory and correlated ab initio calculations [3] represents the simplest but the least accurate<br />

element in the series.<br />

All the discussed one-electron approximations to the two-electron terms in the two-component<br />

pseudo-relativistic Hamiltonian have been put to the test: several types of matrix elements among<br />

multiconfiguration nonrelativistic states have been evaluated for various small and medium-sized<br />

molecules (biradicals). As a reference, both the matrix elements of the corresponding “exact” twoelectron<br />

operators and the known experimental values of the zero-field splitting parameters D and E<br />

have been used.<br />

[1] Hess, B. A.; Marian, C. M.; Wahlgren, U.; Gropen, O. Chem. Phys. Lett. 1996, 251, 365−371.<br />

[2] McWeeny, R.; Mizuno, Y. Proc. R. Soc. (London) 1961, A259, 554−577.<br />

[3] (a) Petrenko, T. T.; Petrenko, T. L.; Bratus, V. Y. J. Phys: Condens. Matter 2002, 14, 12433−12440. (b) Neese,<br />

F. J. Am. Chem. Soc. 2006, 128, 10213−10222.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP439<br />

Theoretical Study of Multi-Nuclear Complexes, [Pt 2 X 4 (Me 2 pz) 8 ] (X=Ag, Cu, H): Metallophilic<br />

LUMO Participation in Strong Luminescence<br />

Yoshihide Nakao, Shigeyoshi Sakaki<br />

Kyoto <strong>University</strong>, Kyoto, Japan<br />

[Pt 2 Ag 4 (Me 2 pz) 8 ] (1), [Pt 2 Cu 4 (Me 2 pz) 8 ] (2), and [Pt 2 H 4 (Me 2 pz) 8 ] (3) are strongly luminescent<br />

complexes, and luminescent colors are blue(497 nm), orange(625 nm), and yellow(556 nm),<br />

respectively, in the solid state. They have no color in the ground singlet state and their luminescence<br />

exhibits remarkably large Stokes shift. Sasaki et. al. explained that strong luminescence requires the<br />

small geometrical relaxation in the excited state.[1] Recently, [Cu 3 (X 2 pz) 3 ] (X=H, CH 3 , CF 3 )<br />

synthesized by Dias et. al. have large Stokes shifts and metal-metal bonding interaction in the excited<br />

triplet states.[2] Due to creating the metal-metal bonding interaction in excited states, they have very<br />

large geometrical difference between the ground and excited states. We theoretically studied relation<br />

between metal-metal bonding interaction and luminescence of three complexes.<br />

We carried out the geometry optimization of the ground singlet and excited triplet states and evaluated<br />

emission energies from vertical excitation energies by the B3LYP method. Absorption spectrum was<br />

evaluated by the TD-B3LYP method. The Pt-Pt distance of the singlet and triplet states are 5.439 and<br />

5.914 Å, respectively, and the Ag-Ag distance of those are 3.407 and 2.870 Å, respectively. Since<br />

other complexes have large geometrical relaxation in excited states, it considers they should exhibit<br />

large Stokes shifts. Complexes 1 and 2 have very interesting LUMO, which consists of in-phase<br />

bonding interaction among all of 6p orbitals of two Pt atoms and either 5p orbitals of four Ag atoms or<br />

4p orbitals of four Cu atoms, respectively. Luminescence of 1 corresponds with transition from these<br />

LUMO to Pt(d-sigma) and the calculated emission energy, 2.54 eV, agrees with the experimental<br />

value, 2.34 eV, in CH 2 Cl 2 . The calculated and experimental emission energies of 2 are 1.53 and 1.51<br />

eV, respectively. The calculated emission energies are in good agreement with the experimental<br />

values.<br />

Pt II<br />

Pt 2.5<br />

PP440<br />

New Approaches to Large-Scale Multiconfigurational Perturbation Theory Calculations<br />

Francesco Aquilante 1 , Tanya Todorova 1 , Laura Gagliardi 1 , Björn Olof Roos 2<br />

1 Department of Physical Chemistry, Geneva <strong>University</strong>, Geneva, Switzerland, 2 Department of<br />

Theoretical Chemistry, Lund <strong>University</strong>, Lund, Sweden<br />

Two complementary approaches are described which further extend the capabilities of Cholesky<br />

decomposition-based multiconfigurational perturbation theory (CD-CASPT2) calculations [1] to treat<br />

large systems and with high-quality atomic orbital basis sets.<br />

For situations where the active orbitals are localized within a small region of the molecule, a suitable<br />

“active site'” can be identified as the collection of atoms where the active orbitals effectively extend.<br />

Accordingly, the inactive and secondary orbitals can be separately localized and partitioned between<br />

this active site and the remaining atoms (environment). From the basic assumptions of our approach,<br />

these two regions are assumed to be uncoupled, and therefore two separate sets of canonical orbitals<br />

can be deduced for the active site and the environment. It is shown that accurate relative energies can<br />

be computed by performing CD-CASPT2 calculations in which the correlating orbitals are restricted to<br />

only those assigned to the active site. This allows for substantial savings in computational costs – not<br />

uncommonly, of 1-2 orders of magnitude.<br />

For systems with delocalized active orbitals, instead, a method is suggested which allows to truncate<br />

the virtual space in CD-CASPT2 calculations with systematic improvability of the resulting<br />

approximation. The method is based upon a modified version of the Frozen Natural Orbital (FNO)<br />

approach used in coupled cluster theory [2] – similarly, the idea is to exploit approximate linear<br />

dependences among the eigenvectors of the virtual-virtual block of the MP2 density matrix. It is shown<br />

that FNO-CASPT2 recovers more than 95% of the full CD-CASPT2 correlation energy while requiring<br />

only a fraction of the total virtual space, especially when large atomic orbital basis sets are in use.<br />

Tests on various properties commonly investigated with CASPT2 demonstrate the reliability of the<br />

approach and the corresponding dump in computational costs and storage demands of the<br />

calculations.<br />

Ag I Ag I<br />

Ag I Ag I<br />

Ag I<br />

Ag I Ag I<br />

Ag I<br />

[1] Aquilante, F.; Malmqvist, P-Å.; Pedersen, T. B.; Ghosh, A.; Roos, B. O. J. Chem. Theory Comp. 2008, 4, 694-<br />

702<br />

[2] See, for example: Taube, A. G.; Bartlett, R. J. Collect. Czech. Chem. Commun. 2005, 70, 837-850<br />

Pt II<br />

Pt 2.5<br />

singlet<br />

triplet<br />

r(Pt-Pt) = 5.439 Å<br />

r(Pt-Ag) = 3.633 Å<br />

r(Ag-Ag) = 3.407 Å<br />

r(Pt-Pt) = 5.914 Å<br />

r(Pt-Ag) = 3.587 Å<br />

r(Ag-Ag) = 2.870 Å<br />

Figure 1. Geometrical changes between ground singlet<br />

and excited triplet states in 1<br />

HOMO<br />

LUMO<br />

Figure 2. HOMO and LUMO of singlet state in 1<br />

[1] Sasaki, Y. Bull. Jpn. Soc. Coord. Chem. 2006, 48, 50.<br />

[2] Dias, H. V. R.; Diyabalanage, H. V. K.; Eldabaja, M. G. ; Elbjeirami, O.; Rawashdeh-Omary, M.; Omary, M. A.<br />

J. Am. Chem. Soc. 2005, 127, 7489.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP441<br />

Spin-Orbit Coupling Effects in Rhenium Tetra-Hydrides<br />

Shiro Koseki, Toshio Asada, Taka-aki Hisashima<br />

Department of Chemistry, Osaka Prefecture <strong>University</strong>, Sakai, Osaka, Japan<br />

In the series of our research projects, the spin-orbit coupling effects upon the shapes of potential<br />

energy surfaces are analyzed for transition metal compounds. The present poster will report<br />

dissociation potential energy curves of several low-lying electronic states in rhenium tetra-hydride<br />

(ReH4). Since the tetra-hydride at the most stable structure is very close in energy to its corresponding<br />

dissociation limits, the spin-orbit coupling effects must become very important in exploring the<br />

potential energy surfaces of low-lying electronic states and investigating Jahn-Teller and pseudo-<br />

Jahn-Teller deformation paths.<br />

In the present theoretical calculations, the SBKJC basis set was used after augmented by a set of<br />

polarization functions for each atom including hydrogen atoms. The orbitals were optimized by using<br />

full-optimized reaction space (FORS) multi-configuration self-consistent field (MCSCF) method, where<br />

its MCSCF active space includes the orbitals correlating to Re’s 5d and 6sp orbitals and hydrogen’s 1s<br />

orbitals in the dissociation limit. Namely, the active space includes 13 orbitals and 11 electrons. Using<br />

these optimized orbitals, first-order configuration interaction (FOCI) wave functions were constructed<br />

for the purpose of spin-orbit coupling calculations.<br />

The most stable structure of ReH 4 has a planar C 2v structure and its ground state belongs to 4 B 2 . The<br />

ground state in the dissociation limit is sextet, ReH 2 ( 6 Σ + g ) + H 2 ,( 1 Σ + g ), and the energy difference<br />

between these is smaller than 10 kcal/mol. Additionally, the highest symmetric T d<br />

structure has a 2 E ground state and this is distorted into two different D 2d<br />

structures ( 2 A 1 and 2 B 1 ) by the Jahn-Teller effect. The energy differences are very<br />

small between these D 2d structures and the most stable planar C 2v structure and,<br />

as a result, the D 2d structures are very close in energy to the dissociation limit. It is<br />

apparently necessary to carry out more sophisticated calculations in order to<br />

obtain conclusive results. Another highest symmetric D 4h structure has 6 E u ground<br />

state. However, this is considerably higher in energy than the structures described above. This is<br />

distorted into D 2h structures ( 6 B 2u , 6 B 3u ) by the Jahn-Teller effect and into C 2v planar structures ( 6 A 1 ,<br />

6 B 2 ) by the pseudo-Jahn-Teller effect, where the geometry optimization of the C 2v structures derives<br />

the dissociation into ReH 2 + H 2 . As long as the adiabatic approximation is used, this molecular system<br />

has never reached the C 2v structure ( 4 B 2 ) or the T d structure ( 2 E) because of different spin multiplicity.<br />

In order to explain the paths from the D 4h structure to the T d and C 2v structures, the spin-orbit coupling<br />

effects need to be included for calculating the potential energy surfaces of low-lying electronic states<br />

in this molecule. The details of the calculated results will be reported at the symposium. Additionally,<br />

the dissociation of ReH 2 into Re ( 6 S) + H 2 is now being investigated. A large energy barrier exists on<br />

the ground-state sextet potential energy curve along the C 2v path, as previously reported by<br />

Balasubramanian et al. This barrier could be remarkably reduced if a C s path and spin-orbit coupling<br />

effects are considered.<br />

PP442<br />

Chromium-Complexes of Furoxan and Benzofuroxan<br />

Tibor Pasinszki<br />

Department of Chemistry, Institute of Chemistry, Eötvös Loránd <strong>University</strong> of Budapest, Budapest,<br />

Hungary<br />

Furoxans (1,2,5-oxadiazole 2-oxides) are potential organic ligands in organometallic chemistry. They<br />

are aromatic molecules, and may form π-donor and n-donor complexes via the π-system and oxygen<br />

and nitrogen lone electron pairs, respectively. Furoxan-complexes, however, are unknown, so far.<br />

The original aim of the present work was to study structures of potential furoxan and benzofuroxan<br />

complexes of chromium(0), which could be derived by substitution reactions from the stable and easily<br />

accessible Cr(CO) 6 , Cr(CO) 3 (CH 3 CN) 3 , (η 6 -benzene)Cr(CO) 3 , and (η 6 -benzene) 2 Cr complexes.<br />

Another aim was to study the thermodynamics of the formation reactions prior to experimental work.<br />

The studied reactions are as follows:<br />

Cr(CO) 6 + furoxan → Cr(CO) 5 (κ N -furoxan)<br />

Cr(CO) 6 + furoxan → Cr(CO) 5 (κ O -furoxan)<br />

Cr(CO) 5 (CH 3 CN) + furoxan → Cr(CO) 5 (κ N -furoxan)<br />

Cr(CO) 5 (CH 3 CN) + furoxan → Cr(CO) 5 (κ O -furoxan)<br />

Cr(CO) 3 (CH 3 CN) 3 + furoxan → Cr(CO) 3 (η 5 -furoxan)<br />

(η 6 -benzene)Cr(CO) 3 + furoxan → Cr(CO) 3 (η 5 -furoxan)<br />

Cr(η 6 -benzene) 2 + furoxan → Cr(η 5 -furoxan) 2<br />

The structure and formation of target molecules have been studied using density functional theory. As<br />

test systems, the experimentally well known Cr(CO) 6 and (η 6 -benzene)Cr(CO) 3 molecules have been<br />

studied using various DFT functionals, and finally the B3PW91 functional is selected for this study.<br />

Calculations have predicted the formation of n-donor complexes, but unexpectedly indicated the<br />

furoxan-ring opening in hypothetical π-donor mode, leading to novel dinitrosoethylene-derivatives.<br />

Structures and energetics of potential formation reactions have been calculated for all furoxan- and<br />

dinitrosoethylene-complexes.<br />

Acknowledgement: Hungarian Scientific Research Fund (OTKA T049148)<br />

[1] T. Hisashima, T. Matsushita, T. Asada, S. Koseki, and A. Toyota, Theoret. Chem. Acc., 2008, 120, 85.<br />

[2] S. Koseki, Computational Methods in Sciences and Engineering, Theory and Computation: Old Problems and<br />

New Challenges, edited by G. Maroulis and T. Simos, 2008, CP963, Vol. 1, page 257.<br />

[3] D. Dai and K. Balasubramaniana, J. Chem. Phys. 1991, 95, 4284.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP443<br />

Crowding Effects on Protein Folding: A Binary System Simulation<br />

Jian Yin 1 , Jianwen Jiang 2 , Raj Rajagopalan 3<br />

1 Singapore-MIT Alliance, Singapore, Singapore,<br />

2 Department of Chemical & Biomolecular<br />

Engineering, <strong>National</strong> <strong>University</strong> of Singapore, Singapore, Singapore, 3 Singapore-MIT Alliance and<br />

Department of Chemical & Biomolecular Engineering, <strong>National</strong> <strong>University</strong> of Singapore, Singapore,<br />

Singapore<br />

It has been broadly recognized only during this recent decade in biological community that crowding<br />

effect plays an important role in the process of protein refolding and aggregation [1]. The crowding<br />

effect is a result of multi-body interaction. It is created thermodynamically from the difference of<br />

volume exclusion of the macromolecules around the neighbourhoods of a target macromolecule. The<br />

so called depletion force due to the depletion zones as the overlaps of the excluded volumes is one of<br />

these crowding effects.<br />

We are investigating the crowding effect as a function of crowder size and effective volume fraction on<br />

a single peptide with molecular dynamics simulation using coarse-grained model. The attribute feature<br />

of this work is that our simulation system is simple binary system consisting of one coarse-grained<br />

peptide and hundreds of Lennard-Jones crowders. Our major conclusions: 1) Crowding enhances<br />

extraordinarily conformational fluctuation of peptide. It is generally believed that larger conformational<br />

fluctuations are necessary for and favourite to conformational transitions. 2) Crowding raises the van<br />

der Waals energy of the peptide. This will imply an acceleration of the rate of conformational transition<br />

of peptides, which will be beneficial to protein refolding. 3) Crowding enhances the correlation<br />

between its van der Waals energy and its radius of gyration. 4) Crowding enhances the solvent<br />

accessible surface area by enhance its hydrophobic portions. 5) The crowder-excluded space creates<br />

hydrophobic cavities. This will result in the depletion attraction taking over the dominating position of<br />

driving force exerted on the peptide. The major support of this proposal is coming from the works of<br />

Wenner and Bloomfield, etc [2].<br />

PP444<br />

The Free Energy Change of Glycine Tautomerization in Aqueous Solution<br />

Miyamoto Hidenori, Misako Aida<br />

Center for Quantum Life Sciences & Graduate School of Science, Hiroshima <strong>University</strong>, Higashihiroshima,<br />

Japan<br />

Introduction<br />

Glycine, the simplest amino acid, is a compound with an intramolecular hydrogen bond. A glycine<br />

molecule gets converted from neutral form (NF) to zwitterionic form (ZW) in aqueous solution, via<br />

intramolecular hydrogen transfer. In this study, we calculate the relative free energy of this process of<br />

the tautomerization of glycine.<br />

Method<br />

A Monte Carlo simulation with the free energy perturbation method was employed to determine the<br />

relative stability between the glycine neutral form and the zwitterionic form in aqueous solution. The<br />

difference in free energy between solute structures i and j is given by:<br />

∆A ij = – kT ln[〈exp{– (E j – E i ) / kT}〉 (i) ]<br />

The bracket (i) denotes the ensemble average of the difference in total QM/MM energies E i and E j<br />

of the system with solute in the i and j structures respectively. To calculate ensemble average of<br />

energy, MM water molecules were arranged around a glycine molecule, and 20200000 configurations<br />

of MM water molecules were generated using MC method at 298K with the glycine molecule fixed,<br />

and 2020 configurations from those 20200000 configurations were selected randomly. Selected 2020<br />

configurations were calculated by means of QM/MM method, where the QM part was the glycine<br />

molecule and the MM part was the water molecules, and MM water molecules within 2.45 angstrom of<br />

the glycine were treated as QM molecules.<br />

[1] (a) Minton, A. P. Mol. Cell Biochem., 1983, 55: 119-140. (b) Zimmerman, S. B., and Minton, A. P. Annu. Rev.<br />

Biophys. Biomol. Struct., 1993, 22, 27-65. (c) Ellis, R. J. Curr. Opin. Struct. Biol., 2001, 11: 114-119. (d) Hall, D.,<br />

and Minton, A. P. Biochim. Biophys. Acta, 2003, 1649: 127-139.<br />

[2] (a) Wenner, J. R., and Bloomfield, V. A. Biophysical J. 1999, 77, 3234 – 3241. (b) Record, M. T., Courtenay, E.<br />

S., Cayley, D. S., and Guttman, H. J., Trends Biochem. Sci., 1998, 23: 143-148. (c) Dinnbier, U., Limpinsel, E.,<br />

Schmid, R., and Bakker, E. P., Arch. Microbiol., 1988, 150: 348-357; (d) Shepherd, V. A., Curr Top Develop Biol.,<br />

2006, 75:171–223.<br />

The program packages used were HONDO,<br />

GAMESS, Gaussion03 and AIM2000.<br />

Results and discussion<br />

The free energy difference between the NF and<br />

the ZW in water and the activation free energy<br />

barriers are plotted in Fig.1. The solute molecule<br />

was calculated using MP2/6-31G* basis set,<br />

solvent molecules within 2.45 angstrom of the<br />

glycine molecule were treated with HF/6-31G*<br />

basis set, and other solvent molecules were<br />

treated with TIP3P, at each point along the<br />

course of hydrogen atom transfer. The blue line<br />

is the free energy hydration, the black line is the<br />

free energy of the system, and the red line is the<br />

potential energy of the solute (glycine). The sign<br />

of free energy of hydration is inverted: it's<br />

actually with minus sign. The available<br />

experimental values for the NF → ZW<br />

transformation compare with our computational<br />

result. The agreement between the experimental<br />

(7.3 kcal/mol) and our results (7.2kcal/mol) is<br />

fairly good, which may indicate that our method<br />

is effective to calculate solvation effect.<br />

∆A ( kcal/mol )<br />

energy of solute ( glycine)<br />

free energy of hydration<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

Fig. 1 g(MP2) + nw(HF) + (101−n)w(MM)<br />

NF<br />

ZW


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP445<br />

Locating Multiple SCF Solutions: an Approach Inspired by Metadynamics<br />

Alexander Thom, Martin Head-Gordon<br />

<strong>University</strong> of California, Berkeley, Berkeley, California, United States<br />

The SCF equations are non-linear in the electron density and their solutions, which are stationary<br />

points of the energy with respect to changes of the occupied orbitals, are usually located iteratively.<br />

The non-linearity of the equations presents a difficult mathematical challenge, as the number of<br />

solutions is not known, and so there is no guarantee that any solution which has been located is an<br />

appropriate global minimum. Furthermore, in systems with many minima close in energy, the solution<br />

located is extremely dependent on the initial guess used for the iterative procedure.<br />

We propose a method to locate the solutions to the Self-Consistent Field equations, using an<br />

approach based upon metadynamics. Within an SCF calculation, when a solution is found, a biasing<br />

potential is added to the energy to avoid reconvergence to the same solution. Multiple solutions can<br />

therefore be relatively easily found with relatively trivial modifications to existing algorithms. Using this<br />

method we locate all known solutions and one unknown solution of the H4 model [1], one of the few<br />

systems whose many solutions have been investigated, and also apply it to a number of small<br />

molecules.<br />

PP446<br />

Theoretical Study of the Electric Conductivity in Nb-Doped TiO 2<br />

Takahiro Suenaga, Hideyuki Kamisaka, Hisao Nakamura, Koichi Yamashita<br />

<strong>University</strong> of Tokyo, Department of Chemical System Engineering, Tokyo, Japan<br />

Transparent conductive oxides are key components in flat-panel display (FPD), photovoltaic cells and<br />

light emitting diodes (LED). Sn-doped In 2 O 3 (ITO) is the most widely used, but the supply of indium is<br />

expected to be depleted because of its worldwide consumption. So we need to seek the alternative to<br />

ITO. The alternative needs to have wide band gap which leads transparency in visible range and a lot<br />

of carrier which leads high conductivity. T. Hasegawa found that Nb-doped anatase (TNO) has high<br />

conductivity and transparency in visible range. The aim of our research is to clarify the micro<br />

mechanism of TNO by investigating band structure, chemical bond and effective mass.<br />

The structural and electronic properties of TNO have been investigated by a DFT-based first-principle<br />

method using PW91 functional and ultrasoft pseudopotential (USPP). When we first construct a TNO<br />

model, we prepare pure anatase cell containing Ti 16 O 32 and one of Ti atoms is replaced by Nb atom.<br />

Then we optimize it. The lattice constant of the optimized structure corresponds to the experimental<br />

data [1]. From the optimized structure, the chemical properties of the Nb dopant were analyzed.<br />

[1] Kowalski, K.; Jankowski K Phys. Rev. Lett. 1998, 81, 1195-1198<br />

Table 1. Lattice constant of antase and TNO<br />

anatase TNO TNO<br />

(calc.) (calc.) (exp. [1])<br />

a (Å) 7.564 7.647 7.596<br />

c (Å) 9.502 9.560 9.561<br />

We calculate the band structure, density of state, charge density and electron effective mass of the<br />

optimized TNO model and compare with pure anatase. From the band structure, we find that there is<br />

little effect of Nb as a dopant. From the density of state, we confirm that valence band is constructed<br />

by O 2p electron and conduction band is constructed Ti 3d and Nb 4d electron which they are hybrid<br />

and there is no impurity band between valence band and conduction band, so TNO keep transparent.<br />

From the electron effective mass, we find that TNO has anisotropy; z-axis direction’s is ten times as<br />

much as x-axis one. In the theoretical analysis by H. Furubayashi, the effective mass of TNO is 0.4m 0 .<br />

So the result of out calculation accord with the analysis.<br />

Table 2. Effective mass of antase and TNO<br />

anatase TNO Exp. [2]<br />

x-axis direction 0.421 0.419<br />

z-axis direction 4.053 4.053<br />

about 1<br />

[1] Y. Furubayashi et al. Appl. Phys. Lett. 2005, 86, 252101<br />

[2] H. Tang et al. J. Appl. Phys. 1994, 75, 2042


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP447<br />

Understanding Nucleation Phenomena Near the Spinodal<br />

Suman Chakrabarty, Mantu Santra, Prabhakar Bhimalapuram, Biman Bagchi<br />

Solid State and Structural Chemistry Unit, Indian Institute of Science, Bangalore, Karnataka, India<br />

We have revisited the apparently well-studied problem of nucleation and growth of a stable phase<br />

inside a parent metastable phase, particularly near the spinodal curve. We have undertaken extensive<br />

computer simulation studies to probe the molecular mechanism for the onset of instability in a wide<br />

range of systems (both Lennard-Jones fluid and nearest neighbour Ising model in 2- and 3-<br />

dimensions). We have constructed the multidimensional free energy surface of nucleation as a<br />

function of multiple reaction coordinates using a very efficient non-Boltzmann sampling scheme. While<br />

the classical Becker-Döring (BD) picture of homogenous nucleation, that assumes the growth of a<br />

single nucleus by single particle addition, holds good at low to moderate supersaturation, the<br />

formation of the new stable phase becomes more collective and spread over the whole system at<br />

large supersaturation. As the spinodal curve is approached from the coexistence line, the free energy,<br />

as a function of the size of the largest liquid-like cluster, develops a minimum at sub-critical cluster<br />

size. We find the emergence of an alternative free energy pathway (with a barrier less than that in the<br />

BD picture) that involves participation of many sub-critical liquid-like clusters and the growth is<br />

promoted by coalescence with intermediate sized clusters present in the neighbourhood of the largest<br />

cluster. Very close to the spinodal the free energy surface becomes quite flat and the significance of a<br />

‘critical’ nucleus is lost. The scenario seems to be mostly independent of the model chosen.<br />

[1] Bhimalapuram, P.; Chakrabarty, S.; Bagchi, B. Phys. Rev. Lett. 2007, 98, 206104.<br />

[2] Reply to a comment against the Ref. 1: Chakrabarty, S.; Santra, M.; Bagchi, B. Phys. Rev. Lett. 2008 (in<br />

press).<br />

PP448<br />

DFT and Multinuclear NMR ( 17 O, 13 C) Studies of Diazenedicarboxylates and Related<br />

Compounds<br />

Francesca Mocci, Michele Usai, Giovanni Cerion<br />

Università di Cagliari, Dipartimento di Scienze Chimiche, Cagliari, Italy<br />

Few widely employed organic compounds, diazenedicarboxylates (1a) and related maleic (2a) and<br />

fumaric (3a) dimethyl diesters, have been studied by a combined computational and multinuclear NMR<br />

(O-17, C-13) approach. Theoretical calculations were carried out at the density functional theory level<br />

employing the PBE0 functional, which has been shown to give very good results in the prediction of<br />

the chemical shielding, together with the 6-311G(d,p) basis set for geometry optimization, and the 6-<br />

311+G(2d,p) basis set for calculating the NMR shielding. The polarizable continuum model (PCM) was<br />

employed both for geometry optimization and for calculating the NMR shielding constants using the<br />

gauge-including atomic orbitals method.<br />

This combined approach afforded important information about the preferred conformations in<br />

chloroform and their influence on the NMR isotropic shielding, indicating that all of the studied<br />

compounds exist in solution as a mixture of two or more conformations in fast exchange, and that for<br />

1a and 2a most of the conformations have the plane of carboxylate groups deviating from planarity<br />

with respect to the system of delocalized electrons to which they are bound.<br />

The results show how an “exotic” magnetic nucleus such as oxygen-17 can be of great help in the<br />

determination of conformational preferences of molecules in solution.<br />

The lowest energy conformations of 1a, 2a and 3a


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP449<br />

The Double-helical → Ladder Structural Transition in the B-DNA is Induced by a Loss of<br />

Dispersion Energy<br />

Jiri Cerny, Pavel Hobza<br />

Institute of Organic Chemistry and Biochemistry, Academy of Sciences of the Czech Republic,<br />

Prague, Czech Republic<br />

Noncovalent interactions play a unique role in biology since they are responsible for the structure of<br />

biomacromolecules. The double-helical structure of DNA which is responsible for storing and transfer<br />

of genetic information and its structure is governed from the subtle balance of noncovalent interactions<br />

among DNA building blocks. The most prominent role is played by interactions between DNA bases<br />

and two binding motifs can be recognized: planar hydrogen bonding and vertical stacking.<br />

In this work we investigate separately the roles of the dispersion energy and electrostatic energy on<br />

the geometry and stability of B-DNA helix. We performed series of molecular dynamics simulations<br />

with empirical force field and hybrid QM/MM molecular dynamics simulations where the dispersion or<br />

electrostatics term are suppressed or increased. We show that the decrease of the dispersion term<br />

leads to increase of vertical separation of the bases as well as loss of helicity and thus resulting in<br />

ladder–like structure. A decrease of the electrostatic term produces the unwinding of the DNA strands.<br />

We can conclude that the simulations clearly show that the dispersion energy is responsible for the<br />

helical structure of the DNA.<br />

PP450<br />

Enzymatic Reaction Mechanism Revealed by Molecular Docking and QM/MM-MD<br />

Yohsuke Hagiwara 1 , Osamu Nureki 2 , Masaru Tateno 3<br />

1 Graduate School of Pure and Applied Sciences, <strong>University</strong> of Tsukuba, Tsukuba, Ibaraki, Japan,<br />

2 Institute of Medical Science, <strong>University</strong> of Tokyo, Minato-ku, Tokyo, Japan, 3 Center for Computational<br />

Sciences, <strong>University</strong> of Tsukuba, Tsukuba, Ibaraki, Japan<br />

Aminoacyl-tRNA synthetases (aaRS’s) play a critical role in decoding genetic information located on<br />

genome DNA sequence, through catalyzing attachment of their cognate amino acid to 3’-end of the<br />

specific tRNA (aminoacylation). The fidelity of translation is assured by their strict discrimination of the<br />

cognate amino acids from non-cognate ones. However, in the case of valine, isoleucine, and leucine,<br />

all of which are similar to each other in their sizes and hydrophobicity, the cognate enzymes, i.e.<br />

leucyl-, valyl-, and isoleucyl-tRNA synthetases (LeuRS, ValRS, and IleRS, respectively), have<br />

difficulties in the strict discrimination of their specific amino acid if they perform only the single catalytic<br />

reaction, producing mis-aminoacylated tRNA’s, such as Ile-tRNA Leu . To avoid generating such<br />

incorrect products, the enzymes accomplish another reaction, i.e. ‘editing’, through which those<br />

enzymes hydrolyze mis-products by themselves. Although several crystal structures of those aaRS<br />

systems have been determined in complex with the cognate tRNAs, reaction mechanisms of editing<br />

have not yet been clarified. The reasons are as follows: (i) no crystal structures of the enzymes in<br />

complex with the substrate, i.e. the cognate tRNA of which the 3’-terminus is bound to nonspecific<br />

amino acid. (ii) nucleophile to trigger the reaction has not yet been identified in the crystal structures<br />

due to lower quality of X-ray crystallographic data.<br />

Our goal is to elucidate reaction mechanisms of editing by those aaRS systems; in this presentation,<br />

we focus on the LeuRS system. First, in order to carry out molecular docking of the LeuRS•ValtRNA<br />

Leu complex, we adopted our novel molecular docking algorithm. Characteristic features of our<br />

docking scheme are to enable to predict conformational changes of protein induced by interaction with<br />

a substrate and waters (induced fitting), and also, to identify configuration of ordered water molecules<br />

located in the substrate-binding site. In the case of the LeuRS•Val-tRNA Leu complex, structural water<br />

molecules in the active site have not yet been identified experimentally so far, as mentioned above.<br />

This docking simulation is coupled to molecular dynamics (MD) calculations with explicit solvent water<br />

molecules, thus being referred to as the “Fully Solvated Dynamical Docking” (FSDD) scheme [1].<br />

Thereby, we have successfully identified ordered water molecules forming stable hydrogen bond<br />

networks in the active site composed of LeuRS, Val-tRNA Leu . It should be noted here that one of such<br />

waters is located at an appropriate position as nucleophile in our modelled structure.<br />

Thus, for the entire structure of the LeuRS•Val-tRNA Leu complex constructed by the FSDD calculation,<br />

we performed ab initio electronic structure calculations to elucidate the first phase in reaction of<br />

editing. For the purpose, we used our QM/MM program recently developed, which connects QM<br />

(gamess) and MM (amber) engines [2, 3]. For QM regions composed of about 120 atoms, HF/DFT<br />

hybrid all-electron calculations were adopted with use of the B3LYP functional. Thereby, it has been<br />

found that LUMO is located on the reaction point of the substrate, suggesting that the water identified<br />

attacks LUMO as nucleophile. Therefore, we next performed QM/MM MD calculations to simulate<br />

reaction processes. In the poster session, we discuss detailed reaction mechanisms of editing by the<br />

LeuRS system, which is elucidated for the first time.<br />

[1] Hagiwara, Y.; Tateno, M. to be published.<br />

[2] Ohta, T.; Hagiwara, Y.; Kang, J.Y.; Nagao, H.; Tateno, M. submitted.<br />

[3] Hagiwara, Y.; Tateno, M. to be published.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP451<br />

Density Functional Theory Augmented with an Empirical Dispersion Term – DFT-D – What is<br />

the Effect of the Many-Body Terms and Higher Order Contributions to Dispersion?<br />

Petr Jurecka 1 , Mikulas Kocman 1 , Pavel Hobza 2<br />

1 Palacky <strong>University</strong>, Olomouc, Czech Republic, 2 Institute of Organic Chemistry and Biochemistry,<br />

Academy of Sciences of the Czech Republic, Prague, Czech Republic<br />

Recently it was shown that one of the major drawbacks of current density functionals – their inability to<br />

describe the long-range dispersion interaction – can be mitigated by adding an empirical dispersion<br />

correction [1]. With the growing number of the DFT-D applications we are gradually gaining the<br />

necessary validation of the newly developed dispersion parametrization [2]. Fairly good accuracy<br />

comparable with the results of much more demanding wave-function based methods was found for the<br />

interaction energies of various intermolecular complexes as well as for their geometries. DFT-D,<br />

although designed for the intermolecular interactions, succeeded in predicting the lowest energy<br />

conformers of the single peptide molecules, such as phenylalanyl-glycyl-glycine. The dispersion<br />

correction improves also the DFT force constants calculated for small peptides [3].<br />

Here, we investigated whether the DFT-D accuracy can be improved by adding the many-body terms<br />

and C8 contributions to dispersion. Results are judged based on the RMS error for the S22 reference<br />

set of molecules [4] and other small complexes. The effect of the three-body terms is in most cases<br />

rather small. Contribution of the C8 term, which is known to be non-negligible, improves the accuracy<br />

of the original dispersion parametrization only slightly, most likely because in the C6-only scheme it is<br />

in part fitted into the damping function. This explains why the dispersion correction based on the C6<br />

term only is usually fairly successful, in spite of its simplicity.<br />

PP452<br />

The Vibrational Band Origins and Potential Energy Surface of Fluorine Isocyanate and its<br />

Isomers<br />

Frank Pickard, Yukio Yamaguchi, Henry Schaefer III<br />

The <strong>University</strong> of Georgia, Center for Computational Chemistry, Athens, GA, United States<br />

The isomerisation pathways of fluorine isocyanate (FNCO) have been studied using coupled-cluster<br />

theory incorporating all single and double excitations (CCSD), along with the perturbative inclusion of<br />

connected triple excitations [CCSD(T)]. These calculations employed large one particle correlation<br />

consistent basis sets (cc-pVQZ). The final potential energy surface (PES) of this system was<br />

computed using valence focal point extrapolations [1]. Accurate vibrational band origin (VBO)<br />

predictions were made for all minima on the PES. Excellent agreement was found between the<br />

predicted and observed [2] VBOs for FNCO. The VBO predictions for the heretofore unsynthesized<br />

high energy isomers of FNCO should assist in their eventual experimental characterization. The<br />

calculated PES also demonstrates that several high energy isomers should be viable synthetic targets.<br />

[1]. Császár, A. G.; Allen, W. D.; Schaefer, H. F. J. Chem. Phys. 1998, 108, 9751.<br />

[2]. Jacobs, J.; Juelicher, B.; Schatte, G.; Willner, H.; Mack, H. G. Chem. Ber. 1993, 126, 2167.<br />

[1] Grimme, S. J. Comput. Chem. 2004, 25, 1463-1473.<br />

[2] Jurečka, P.; Černý, J.; Hobza, P.; Salahub, D. R. J. Comput. Chem. 2007, 28, 555-569.<br />

[3] Černý, J.; Jurečka, P.; Hobza, P.; Valdés H. J. Phys. Chem. A 2007, 111, 1146-1154.<br />

[4] Jurečka, P.; Šponer, J.; Černý, J.; Hobza, P. Phys. Chem. Chem. Phys. 2006, 8, 1985-1993.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP453<br />

Exploration of Matrix Exponentiation Issues Involved with Non-Reversible Rate Matrices Used<br />

in the Modelling of DNA Sequence Evolution<br />

Harold Schranz<br />

Computational Genomics, John Curtin School of Medical Research, ANU, Canberra, ACT, Australia<br />

Modelling DNA sequence evolution involves the exponentiation of instantaneous rate matrices [1]. The<br />

most commonly employed rate matrices impose the restriction that evolutionary processes are timereversible.<br />

Relaxing the assumption of time-reversibility requires consideration of non-reversible<br />

matrices.<br />

For particular classes of matrices, the most popular eigendecomposition based matrix exponentiation<br />

algorithm is sometimes inaccurate and may even fail completely [2]. Predicting the conditions for<br />

failure is of considerable practical interest as is the application of more robust exponentiation<br />

methods. Eigendecomposition works well for normal matrices (that commute with their conjugate<br />

transposes) but breaks down when the matrix does not have a complete set of linearly independent<br />

eigenvectors or when the condition number of the matrix of eigenvectors is large (near singular) [3].<br />

Here we tested whether such pathological rate matrices exist in nature by constructing them<br />

from concatenated protein coding gene alignments from microbial genomes, primate genomes<br />

and independent intron alignments from primate genomes. The Taylor series expansion and<br />

eigendecomposition matrix exponentiation algorithms were compared to the less widely employed, but<br />

more robust, Padé with scaling and squaring algorithm for nucleotide, dinucleotide and trinucleotide<br />

rate matrices. Our results [4] indicate that development of robust software for computing nonreversible<br />

dinucleotide, codon and higher evolutionary models requires implementation of the Padé<br />

with scaling and squaring algorithm.<br />

[1] Lio P, Goldman N. Genome Res 1998, 8,1233–44.<br />

[2] (a) Moler CB, Van Loan CF. SIAM Review 2003, 45, 3–49. (b) Golub GH, Loan CFV: Matrix computations (3rd<br />

ed.). Baltimore, MD, USA: Johns Hopkins <strong>University</strong> Press, 1996.<br />

[3] (a) Smith R. Numerische Mathematik 1967,10, 232–240. (b) Demmel J: On condition numbers and the<br />

distance to the nearest ill-posed problem. Numerische Mathematik 1987, 51, 251–289. (c) Bai Z, Demmel J,<br />

McKenney A. ACM Trans. Math. Softw. 1993, 19, 202–223.<br />

[4] Harold W. Schranz, Von Bing Yap, Simon Easteal, Rob Knight and Gavin A. Huttley, in preparation.<br />

PP454<br />

The Electronic State of Blue Cu Protein Revealed by the New QM/MM Interface Program<br />

Takehiro Ohta 1 , Yohsuke Hagiwara 2 , Masaru Tateno 3<br />

1 Institute for Materials Chemistry and Engineering, Kyushu <strong>University</strong>, Kasuga, Fukuoka, Japan,<br />

2 Graduate School of Pure and Applied Sciences, <strong>University</strong> of Tsukuba, Tsukuba, Ibaraki, Japan,<br />

3 Center for Computational Sciences, <strong>University</strong> of Tsukuba, Tsukuba, Ibaraki, Japan<br />

Metalloenzymes are involved in various biological functions such as electron transfer, storage of<br />

metals, binding of dioxygen, turnover of substrates, and configuration of protein structure.<br />

Understanding the coordination geometry and electronic structure of metal active sites in the<br />

metalloenzymes is of fundamental biophysical importance to gain insight into the structure/function<br />

correlation of those biological macromolecules. Azurin<br />

is one of the metalloenzymes, for which biological<br />

function is related to electron transfer. The copper ion<br />

bound to its active site is coordinated by five amino acid<br />

residues, cystein (Cys), two histidines (His), methionine<br />

(Met), and a backbone carbonyl oxygen of glycine<br />

(Gly). Here, we present a QM/MM study of azurin, in<br />

which our newly developed program that interfaces the<br />

quantum mechanical (QM) calculation program gamess<br />

with the molecular mechanics (MM) simulation program<br />

amber is used [1, 2].<br />

Our QM/MM energy evaluation method is based on an additive scheme, in which we consider<br />

polarization of the QM region induced by surrounding MM atoms. In this study the spin-unrestricted<br />

Hartree-Fock (UHF) / density functional theory (DFT) hybrid all-electron calculation with the B3LYP<br />

functional was adopted as the QM Hamiltonian. To assess the accuracy of the additive energy<br />

scheme, we employed two computational models, referred to as Model I and Model II. In Model I,<br />

electrostatic interactions between the QM and MM atoms were calculated by incorporating partial point<br />

charges of MM atoms into one-electron integration of QM Hamiltonian. In Model II, the electrostatic<br />

interactions between the QM and MM atoms were calculated at MM level, and thus, the electron<br />

density of the QM region was not allowed to be polarized by the partial point charges: this is referred<br />

to as the subtractive scheme.<br />

In order to obtain equilibrated structure in solvent, we first performed classical molecular dynamics<br />

(MD) simulation for 1ns with explicit solvent water molecules, started from a high resolution crystal<br />

structure of azurin. During the simulation, the protein structure was quite stable; the final snapshot was<br />

subjected to energy minimization at MM level. Then, started from the minimized structure, we<br />

performed QM/MM geometry optimization by using each calculation scheme of Models I or II.<br />

Comparison of the optimized structures obtained through two schemes showed that Model I provided<br />

the electronic property more consistent with spectroscopic experiments than that of the Model II<br />

calculation. This indicates that the explicit inclusion of electrostatic interaction into the QM Hamiltonian<br />

is essential for accurate descriptions of the copper site. With respect to the geometries, different<br />

descriptions of the copper coordination geometry were obtained between Models I and II, particularly<br />

for the coordinated bonds including a large dipole. Thus, it is suggested that modification of the QM<br />

Hamiltonian so as to interact with long-distance partial point charges of the environment is crucial for<br />

an accurate QM/MM description of both the geometrical and electronic properties of metal active sites.<br />

[1] Ohta, T., Hagiwara, Y., Kang, J.Y., Nagao, H., and Tateno, M., to be published.<br />

[2] Hagiwara, Y, Ohta, T., and Tateno, M., to be published.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP455<br />

Investigation of the Invisible Water Effect for Quantitative Docking Analysis of Protein-Ligand<br />

Complexes<br />

Hitoshi Goto 1 , Shigeaki Obata 1 , Taku Sugiyama 1 , Naofumi Nakayama 1 , Kouichi Nishigaki 2<br />

1 Toyohashi <strong>University</strong> of Technology, Toyohashi, Japan, 2 Saitama <strong>University</strong>, Saitama, Japan<br />

X-ray crystal structure analysis of the target protein with specific ligand is very important for<br />

pharmaceutical developments and the drug-discoveries, such as investigation of the protein function,<br />

hypothetical formation for the activation mechanism, purification of the binding compound, the binding<br />

site determination, and etc. Although we can expect the experimental fact provides us the most<br />

appropriate structure for quantitative evaluation of the binding energy with protein-ligand docking<br />

calculations, in many practical cases, we know it is very very difficult.<br />

Part of the problem is, we call, “invisible water effect”, that is, the role of the water molecules and/or<br />

the water cluster not determined in the X-ray crystal structure analysis. In this paper, as the practical<br />

examples of the effect, we introduce our recent observations in the studies on two pharmaceutical<br />

important enzymes with both molecular mechanics and molecular dynamics calculations: 1) the<br />

unfolding transition on HIV-1 protease-inhibitor complexes during crystal structure optimization, and 2)<br />

deformation of A-B domains of cathepsin E crystal structure.<br />

PP456<br />

Density-Functional Study of Acyclic and Cyclic Phosphoryl Transfer Reactions in Solution:<br />

Comparison with Experimental Thio and Isotope Effect Measurements<br />

Francesca Guerra, Jiali Gao, Darrin M. York<br />

Department of Chemistry and Supercomputing Institute, <strong>University</strong> of Minnesota, Minneapolis, MN,<br />

United States<br />

Phosphoryl transfer reactions are ubiquitous in biology, and are important in cell signaling, energy<br />

transfer and the synthesis and breakdown of DNA and RNA. Of particular importance is in the<br />

unraveling of the mechanisms whereby RNA enzymes, or ribozymes, are able to catalyze fairly<br />

intricate reactions with efficiency that rival many protein enzymes. Experimentally, common methods<br />

used to probe mechanism and the nature of the rate-controlling transition state involve the study of<br />

chemically modified and/or isotopically labeled model reactions. Typically, sulfur is used to study socalled<br />

thio effects, and 18 O and 34 S are used to measure isotope effects, both of which provide<br />

information about the reaction. In order to translate experimental measurements into a detailed picture<br />

of mechanism requires the use of a theoretical model. Density-functional theory provides a powerful<br />

tool to aid in the interpretation of experimental thio and isotope effects, and provide detailed insight<br />

into the molecular mechanisms of phosphoryl transfer reactions. In the present work, DFT is used to<br />

study a series of acyclic and cyclic phosphoryl transfer reactions, including sulfur-substitutions, for<br />

which kinetic isotope measurements have been made. A comparison of the performance of modern<br />

DFT functionals for these reactions, including B3LYP and M06-2X, and implicit solvation methods<br />

including COSMO, PCM and SM6, has been made and used to assess the overall reliability and<br />

accuracy of the models and to establish reliable benchmarks in the gas phase and in solution. These<br />

benchmarks serve as key data in the development of new-generation semiempirical quantum models,<br />

such as the AM1/d-PhoT Hamiltonian, that can be used in combined QM/MM simulations in more<br />

complex aqueous and solvated macromolecular environments. Together with experiments, the<br />

present results provide new insight into the mechanisms of phosphoryl transfer reactions in solution,<br />

and catalyzed by enzymes and ribozymes.<br />

Fig. 1 X-ray structure of HIV-1 protease-inhibitor complex and<br />

the optimized structure with appearance of invisible water molecules


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP457<br />

Photodissociation of the Water Dimer: A 12-D Quasiclassical Study<br />

Gustavo Avila, Geert-Jan Kroes, Marc Van Hemert<br />

Leiden <strong>University</strong>, Leiden Institute of Chemistry, Leiden, Netherlands<br />

The ab initio study of the photodissociation of the water dimer in its first absorption band is the first<br />

step in the understanding of the VUV liquid water absorption spectrum that is so remarkably shifted<br />

towards the blue by around 1 eV with respect to the gas phase spectrum.<br />

PP458<br />

Generator Coordinate Gaussian Basis Sets for the First-row Atoms: A New Alternative for<br />

Pople’s and Dunning’s Basis Sets<br />

Moacyr Comar Jr. 1 , Francisco Lima 2 , Albérico Da Silva 3<br />

1 Instituto de Ciências Exatas - UFAM, Manaus, Amazonas, Brazil, 2 Instituto de Química de São Carlos<br />

- USP, Instituto de Química de São Carlos - USP, Brazil, 3 Instituto de Química de São Carlos - USP,<br />

Instituto de Química de São Carlos - USP, Brazil<br />

Since its development in the eighties, the Generator Coordinate Hartree-Fock (GCHF) method [1] has<br />

been used as an important tool to generate universal and adapted Gaussian basis sets to be used in<br />

atomic and molecular calculations. In 2003, we published a new way to discretize the integral equation<br />

of the GCHF method [2] so that we could improve the efficiency of the GCHF method in generating<br />

Gaussian basis sets. This new way of discretizing the integral equation of the GCHF method is done<br />

by a polynomial and the name of the method was then changed to polynomial version of the Generate<br />

Coordinate Hartree-Fock (pGCHF) method. Here, we now present Gaussian basis sets for the firstrow<br />

atoms of the periodic table that are an excellent alternative for the well-known Pople´s and<br />

Dunning´s Gaussian basis sets since with our first-row basis sets one is able to get results comparable<br />

or better than the Pople´s and Dunning´s Gaussian basis sets but with a lower computational cost,<br />

mainly when compared to the Dunning´s correlation consistent Gaussian basis sets.<br />

[1] Da Silva, A.B.F.; Da Costa, H.F.M. and Trsic, M., Mol. Phys. 1989, 68, 433-445.<br />

[2] Barbosa, R.C. and Da Silva, A.B.F., Mol. Phys. 2003, 101, 1073-1077.<br />

water absorption spectrum in the gas phase, 1:50 Ar matrix, liquid and solid (20K)<br />

(MCvH ~1975).<br />

The basic ingredients in the ab initio calculation of absorption spectra are potential energy surfaces for<br />

all degrees of freedom for both ground and excited states and subsequently the complete description<br />

of the nuclear motion on these potential energy surfaces. For the water monomer such a calculation<br />

has been performed successfully and perfect agreement between the theoretical and experimental<br />

spectrum was obtained [1].<br />

For the water dimer the situation is complicated by the high dimensionality. Our strategy in this case<br />

initially has been a reduced dimensionality treatment [2]. That treatment was able to explain the blue<br />

shift but lacked detail. Afterwards a full 12-dimensional study was performed [3] where the essential<br />

parts of the ground and excited state potentials were calculated on the MRCI level using the grow<br />

approach of the Collins group [4]. The necessary first and second order derivatives were calculated at<br />

a lower level of electron correlation, i.e. MCSCF, in order to avoid prohibitive computer times. The<br />

spectra were calculated using the quasiclassical approximation. This appeared to be a successful<br />

approach since now, among others, the predicted ‘red tail’ of the dimer spectrum could be reproduced.<br />

In this contribution we will show how the spectra can still be further improved by calculating the first<br />

and second order derivatives at the same level of electron correlation as the energies.<br />

[1] R. van Harrevelt and M.C. van Hemert, J. Chem. Phys. 2001, 114, 9453.<br />

[2] L. Valenzano, M.C. van Hemert and G.J. Kroes, J. Chem. Phys., 2005, 123, 034303.<br />

[3] G. Avila, G.J. Kroes and M.C. van Hemert, J. Chem. Phys. 2008, 128, 144313.<br />

[4] M. A. Collins, Theor. Chem. Acc., 2002, 108, 313.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP459<br />

Three-Body Corrections to Scaled Opposite Spin Second-Order Møller-Plesset Perturbation<br />

Theory<br />

Robert DiStasio, Alex Thom, David Casanova, Martin Head-Gordon<br />

<strong>University</strong> of California at Berkeley Department of Chemistry, Berkeley, CA, United States<br />

In this work, we present several theoretical frameworks in which three-body electron correlation can be<br />

included within scaled opposite spin second-order Møller-Plesset (SOS-MP2) perturbation theory [1].<br />

We derive these three-body corrections by considering a subset of the triples contribution in fourthorder<br />

Møller-Plesset perturbation theory that only includes excitations involving opposite spin<br />

electrons. These new models include two empirical parameters based on extensive benchmarking<br />

with respect to highly accurate coupled cluster models. As a result, this approach provides a simple<br />

and economical electronic structure method that has the potential of achieving chemical accuracy in<br />

the prediction of thermochemical properties for systems that are currently too large to be treated using<br />

highly-accurate methods like CCSD(T).<br />

PP460<br />

Structures of the Mono-hydrated Guanine-Cytosine Cation<br />

Heather Jaeger, Henry Schaefer III<br />

<strong>University</strong> of Georgia, Center for Computational Chemistry, Athens, GA, United States<br />

A detailed study of the mono-hydrated guanine-cytosine cation potential energy surface, using B3LYP<br />

with a DZP++ basis, reveals seven local minima. Structural changes from the isolated guaninecytosine<br />

cation geometry upon hydration have been observed and are dependent on the location of<br />

the water molecule. The preferred binding site of the water molecule is at the hydrogen atom on<br />

guanine which replaces the deoxyribose in DNA. At this geometry water binds to the base pair with 13<br />

kcal/mol and is the global minimum. Four local minima were found having binding energies 3 kcal/mol<br />

above the global minimum, and two others were found with energies 7 kcal/mol above the global<br />

minimum. The positive charge of the complex is localized on the guanine molecule which agrees with<br />

experimental evidence that guanine has a lower ionization potential than cytosine.<br />

[1] Jung, Y.; Lochan, R. C.; Dutoi, A. D.; Head-Gordon, M. J. Chem. Phys. 2004, 121, 9793-9802.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP461<br />

Application of Ab Initio DMRG to the Physics of Conjugated π-Electron Systems<br />

Johannes Hachmann, Jonathan J. Dorando, Debashree Ghosh, Garnet Kin-Lic Chan<br />

Department of Chemistry and Chemical Biology, Cornell <strong>University</strong>, Ithaca, NY, United States<br />

We present some recent developments in the ab initio density matrix renormalization group (DMRG)<br />

method for quantum chemical problems, in particular our local, quadratic scaling algorithm [1] for lowdimensional<br />

systems. This method is particularly suited for the description of strong nondynamic<br />

correlation, and allows us to compute numerically exact (FCI) correlated energies for large active<br />

spaces, up to one order of magnitude larger then can be obtained by conventional CASCI techniques.<br />

PP462<br />

Molecular Conductance through a Graphene Sheet<br />

Haitao Wang, Garnet Chan<br />

Department of Chemistry and Chemical Biology, Cornell <strong>University</strong>, Ithaca,NY, United States<br />

Graphene, as a single layer of carbon planner sheet, draws great attention since its experimental<br />

realization. Its interesting linear energy dispersion relationship offers a unique platform for the study of<br />

the electronic transport problem. In this poster, we will investigate the conductance through a<br />

graphene sheet when potassium is added on the top to charge the sheet.<br />

We will review the problems, predominantly in the field of organic electronic materials, which we<br />

studied with the ab initio DMRG:<br />

1) metal-insulator transition in hydrogen chains [1]<br />

2) all-trans polyacetylene oligomers [1]<br />

3) acenes [2]<br />

4) polydiacetylene oligomers [3]<br />

5) graphene nanopatches [3].<br />

Organic electronic materials and the correlation in their quasi-1D-conjugated π-backbone pose are<br />

serious challenge to conventional quantum chemical methods. With the ab initio DMRG we can<br />

correlate the complete π-valence space, which determines the qualitative physics of these molecules,<br />

and obtain numerically exact results.<br />

In our studies we stress the analysis of the obtained ground and excited states with respect to<br />

particular physical regimes in order to categorize their inherent nature.<br />

[1] Hachmann, J.; Cardoen, W.; Chan, G. K.-L. J. Chem. Phys. 2006, 125, 144101.<br />

[2] Hachmann, J.; Dorando, J. J.; Aviles, M.; Chan, G. K.-L. J. Chem. Phys. 2007, 127, 134309.<br />

[3] unpublished.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP463<br />

The Role of Ligands in Developing Mn 4 -based Single-Molecule Magnets and Single-chain<br />

Magnets<br />

Nguyen Anh Tuan, Shin-ichi Katayama, Dam Hieu Chi<br />

1 School of Materials Science, Japan Advanced Institute of Science and Technology, 1-1, Asahidai,<br />

Nomi, Ishikawa, 923-1292, Japan, 2 Faculty of Physics, Hanoi <strong>University</strong> of Science, 334 Nguyen Trai,<br />

Thanh Xuan, Hanoi, Viet Nam<br />

Distorted cubane Mn 4 clusters have been known as single-domain nanoscale magnetic particles<br />

(single-molecules magnets, SMMs) [1]. They display slow magnetization relaxation below their<br />

blocking temperature. This behavior results from a significant ground spin state (S T ≈ 9/2) combined<br />

with a large and negative Ising (or easy-axis) type of magnetoanisotropy, as measured by the axial<br />

zero-field splitting parameter, D ≈ −(0.65~0.75) K. This combination leads to a significant barrier (U) to<br />

magnetization reversal, whose maximum value is given by (S T 2 − 1/4)|D| = 20|D|. There are various<br />

distorted cubane [Mn 4 (µ 3 -O 3 )(µ 3 -X)(O 2 CR) 3 (L1,L2) 3 ] (X, R, L1, and L2 = various) have been<br />

synthesized [1]. Each Mn 4 SMM is distinguished from the other by its ligands and also exhibits<br />

different characteristics due to the function by ligands. Besides, the existence of the dimer structure of<br />

distorted cubane Mn 4 O 3 Cl 4 (O 2 CEt) 3 (py) 3 [2] shows that distorted cubane Mn 4 can become building<br />

block for developing new manganese SMMs and single-chain magnets (SCMs). Here the particular<br />

ligand structure in Mn 4 O 3 Cl 4 (O 2 CEt) 3 (py) 3 was observed to be responsible for the dimer formation.<br />

Based on these observations, we realize that ligands must play an important role in determining<br />

magnetic behavior of distorted cubane Mn 4 , as well as in combining Mn 4 building blocks to develop<br />

new manganese SMMs and SCMs. Therefore, we explore systematically the role of ligands in<br />

determining the geometric structure, electronic structure, and magnetic properties of distorted cubane<br />

Mn 4 SMMs by using first-principles calculations based on density-functional theory, in order to support<br />

for designing new Mn 4 SMMs, as well as to look for new Mn 4 building blocks for developing new SMMs<br />

and SCMs. Our results show that the distorted cubane geometry and the ground spin state of Mn 4<br />

SMMs are preserved even if the µ 3 -O atoms are changed by X 2 -type ligand complexes such as NCH 3 .<br />

The results reveal more possibilities of designing new Mn 4 SMMs and combining them for developing<br />

new manganese SMMs. Our results also show that the distorted cubane geometry and the ground<br />

spin state of Mn 4 SMMs are preserved even if the O atoms of O 2 CR are changed by X 2 -type ligand<br />

complexes. The results also give more possibilities of designing new Mn 4 building blocks and<br />

combining them. Our results show that the ground spin states of distorted cubane Mn 4 molecules can<br />

be controlled by substituting L1 and L2 [3]. Besides, the mechanism of exchange couplings between<br />

manganese ions in distorted cubane Mn 4 SMMs is elucidated. The results show the correlation among<br />

the charge and spin states of manganese ions and the magnetic structure of distorted cubane Mn 4<br />

SMMs [3]. These results also reveal the possibility of designing new distorted cubane Mn 4 SMMs with<br />

S T = 15/2. Finally, we observe significant spin polarizations on peripheral ligands L1 and L2 in<br />

distorted cubane Mn 4 SMMs [3]. These spin polarizations can enhance the exchange couplings<br />

between Mn 4 building blocks for developing Mn 4 -based SMMs and SCMs.<br />

PP464<br />

Ring Molecules as Basic Modules in Metamaterials?<br />

Stephan Bernadotte 1,3 , Wim Klopper 1,2,3 , Ferdiand Evers 1,2<br />

1 Institut für Nanotechnologie, Karlsruhe, Germany, 2 Institut für Physikalische Chemie, Universität,<br />

Karlsruhe, Germany, 3 Institut für Theorie der Kondensierten Materie, Universität, Karlsruhe, Germany<br />

An important research direction in material sciences is the design of ''metamaterials''. Metamaterials<br />

exhibit an artificial unit cell, for instance a split ring resonator that is embedded in a crystalline<br />

superstructure consisting of many other such cells. The design of split ring resonators aims at<br />

customizing the electromagnetic properties. Most notably one would like to create a material with an<br />

index of refraction that is negative at optical frequencies. In order to go to such high frequencies the<br />

current resonator size is too large; whether ring molecules may a viable alternative, is a very important<br />

motivation of our research.<br />

The dynamical electromagnetic response of electrons on a one-dimensional ring is studied in the<br />

present work. The applicability of this model reaches from π-systems of organic ring molecules to<br />

clean metal rings. The electrons are treated with a parabolic (''free'') dispersion as well with a tight<br />

binding one. First we calculate the density and current response with the Kubo formula. The<br />

corresponding response kernels are dressed at the RPA level. From this the permittivity, the<br />

permeability and the index of refraction of an array of such rings are calculated.<br />

By introducing a slit on the ring we observe that the resonance of the magnetic response is shifted<br />

near to the electric resonance so that the real part of the index of refraction may become negative. We<br />

study the analytic structure of the index of refraction and its dependency on damping, screening, the<br />

size of the system, the number of electrons on the ring and the fill factor respectively. Finally,<br />

implications for the material design are discussed.<br />

[1] S. M. J. Aubin, M. W. Wemple, D. M. Adams, H.-L. Tsai, G. Christou, and D. N. Hendrickson, J. Am. Chem.<br />

Soc. 1996, 118, 7746-7754.<br />

[2] W. Wernsdorfer, N. Aliaga-Alcalde, D. N. Hendrickson, and G. Christou, Nature (London) 2002, 416, 406.<br />

[3] N. A. Tuan, S. Katayama, D. H. Chi, A Systematic Study of Influence of Ligand Substitutions on the Electronic<br />

Structure and Magnetic Properties of Mn 4 Single-Molecule Magnets, Phys. Chem. Chem. Phys. 2008, B806661B.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP465<br />

Role of Solvent and Dispersion Forces on the Stability of the DNA Double-Helix<br />

Martin Kabelác, Pavel Hobza<br />

Institute of Organic Chemistry and Biochemistry, Academy of Sciences of the Czech Republic,<br />

Prague, Czech Republic<br />

A role of dispersion and presence of the solvent on the stability and geometry of short sequences of DNA (up to<br />

tetramers) was calculated for the first time at ab initio quantum chemical level. The systems were fully relaxed<br />

using the resolution of identity DFT method with/without empirical dispersion term fitted to the CCSD(T) data. The<br />

optimizations were performed in the presence/absence of water which was modelled implicitly using the<br />

conductor-like screening model (COSMO) as well as counterions. The final stabilization energies and geometries<br />

were mutually compared. The absence of dispersion forces leads to significant distortion of the structure of DNA<br />

and separation of stacked bases in comparison with canonical form of DNA. The presence of the solvent leads to<br />

the destabilization of the double-helical structure, however the optimized structure is pretty close to the canonical<br />

form of B-DNA.<br />

PP466<br />

Design of a Synthetic Catalytic Pore for Cation-Olefin Cyclizations; DFT and MD-Simulations<br />

Approach<br />

Daniel Emery, Guillaume Bollot, Jiri Mareda<br />

Department of Organic Chemistry, <strong>University</strong> of Geneva, Geneva, Switzerland<br />

Cation-olefin cyclizations are very useful, albeit complex, carbon-carbon bond forming reactions. For<br />

such cyclizations a high degree of stereocontrol is often achieved by cyclases and antibodies in<br />

biogenesis of natural products. [1] In the laboratory the cation-olefin cyclizations are accomplished by<br />

solvolysis of an appropriate substrate, although with lesser stereocontrol. Here we describe computer<br />

modeling studies of cation-olefin reactions aimed at elucidating the mechanism of the cyclization at the<br />

active site of the catalytic antibody and design of a new synthetic catalyst able to mimic the action of<br />

the antibody.<br />

The electronic structures, energies, and equilibrium geometries of reaction intermediates involved in<br />

the cation-olefin cyclizations were studied first by density functional theory (DFT) methods. Thus,<br />

obtained stationary point geometries were then docked into the active site of the catalytic antibody<br />

4C6. [2] Subsequent MD simulations on the protein part of the complex led to average geometry that<br />

was used to establish the theoretical model of the antibody active site where only the key residues of<br />

the biocatalyst were retained for further investigation by quantum mechanics. Using the ONIOM<br />

method and several density functionals, the theozyme model was used to investigate the key<br />

pathways of antibody catalyzed cation-olefin cyclizations.<br />

Understanding of the cyclization mechanism and mapping of the antibody’s active site provided<br />

necessary data to propose the structure of an artificial chiral catalyst. Based on the concept of rigidrod<br />

β-barrels developed [3], the internal hollow space of the β-barrel was populated with the<br />

appropriate residues in order to mimic the catalytic effect of the antibody. The predicted catalytic<br />

activity of the proposed rigid-rod β-barrel was evaluated by the DFT model calculations.<br />

[1] (a) T. Li; K. D. Janda; R. A. Lerner: Acc. Chem. Res. 30, 115, 1997 ; (b) J. Hasserodt; K. D. Janda:<br />

Tetrahedron 53, 11237, 1997 ; (c) K. Wendt; G. Schulz; E. J. Corey; D. Liu: Angew. Chem. Int. Ed. 2000, 39,<br />

2813.<br />

[2] X. Zhu; A. Heine; F. Monnat; K. N. Houk; K. D. Janda; I. A. Wilson: J. Mol. Biol. 2003, 329, 69.<br />

[3] N. Sakai; J. Mareda; S. Matile : Acc. Chem. Res. 2005, 38, 79.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP467<br />

Projected Entangled Pair States as a Multi-Dimensional Ansatz for Non-Dynamic Correlation in<br />

Huge Active Spaces<br />

Dominika Zgid, Garnet Chan<br />

Department of Chemistry and Chemical Biology, Cornell <strong>University</strong>, Ithaca, NY, 14850, United States<br />

Modern quantum chemistry methods have become very efficient in addressing dynamic correlation.<br />

However, the success of quantum chemical approaches is still limited if one considers the treatment of<br />

non-dynamic correlation. For example, many inorganic systems with many closely degenerate states<br />

still cannot be qualitatively described by current methods. Although the Density Matrix<br />

Renormalization Group (DMRG) has proven to be very successful for the treatment of the<br />

multireference effects in systems with pseudo-one-dimensional connectivity, extended multidimensional<br />

systems with large active spaces are still too demanding for DMRG. Here, we present<br />

generalization of the DMRG ansatz to higher dimensions that is able to effectively recover nondynamic<br />

correlation in two-dimensional lattices. This ansatz is known in physics as the Projected<br />

Entangled Pair State (PEPS) ansatz and has been proven to work successfully in two-dimensional<br />

model systems. We present the theory of PEPS listing possible advantages and problems arising from<br />

such a treatment. Additionally, preliminary applications of our PEPS code are presented, such as to<br />

multi-center molecular magnets which both require very large active spaces and which possess<br />

intrinsic two-dimensional lattice connectivity.<br />

PP468<br />

Self-Cleavage Catalysis of the Hepatitis Delta Virus Ribozyme Investigated by QM/MM<br />

Calculations<br />

Pavel Banáš 1 , Lubomír Rulíšek 2 , Daniel Svozil 2 , Nils Walter 3 , Jiri Šponer 4 , Michal Otyepka 1<br />

1 Palacky <strong>University</strong>, Department of Physical Chemistry, Olomouc, Czech Republic, 2 Institute of<br />

Organic Chemistry and Biochemistry, Academy of Sciences of the Czech Republic, Prague, Czech<br />

Republic, 3 Department of Chemistry, Single Molecule Analysis Group, <strong>University</strong> of Michigan, Ann<br />

Arbor, United States, 4 Institute of Biophysics, Academy of Science of the Czech Republic, Brno,<br />

Czech Republic<br />

The hepatitis delta virus (HDV) ribozyme is a catalytic RNA embedded in human pathogenic HDV<br />

RNA. Published experimental studies have established that the nucleotide C75 is essential for selfcleavage<br />

of the ribozyme. The exact catalytic role of C75 in the self-cleavage reaction remains<br />

debated. Structural data from X-ray indicate that C75 can act as the general base that initiates<br />

catalysis by deprotonating the 2’-OH nucleophile at the cleavage site, while a hydrated magnesium ion<br />

likely protonates the 5’-oxygen leaving group. In contrast, some mechanistic studies support the<br />

hypothesis that C75 may be protonated and acts as the general acid. We report combined quantum<br />

chemical/molecular mechanical calculations (ONIOM: MPW1K/6-31+G(d,p)/parm99) for the C75<br />

general base pathway, utilizing the available structural data for the wt-HDV genomic ribozyme as a<br />

starting point. Several starting configurations differing in magnesium ion position were considered and<br />

both one-dimensional and two-dimensional potential energy surface scans were used to explore<br />

plausible reaction paths. Our calculations show that C75 is readily capable of acting as the general<br />

base, in concert with the hydrated magnesium ion as the general acid. The calculated energy barrier<br />

of the proposed mechanism, ~20 kcal/mol, would lower the reaction barrier by ~15 kcal/mol compared<br />

to the uncatalyzed reaction and is in good agreement with experimental data. Thus, the hypothesis<br />

that C75 serves as the general base has to be considered chemically feasible. Due to the absence of<br />

suitable experimental starting structures that would be consistent with the general acid mechanism we<br />

could not evaluate the alternative pathway in which C75 is protonated and acts as the general acid.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP469<br />

MOLEONLINE: An Interactive Web-Based Tool to Find and Analyze Molecular Channels,<br />

Tunnels and Pores.<br />

Martin Petřek 1 , Jaroslav Koča 1 , Michal Otyepka 2<br />

1 <strong>National</strong> Centre for Biomolecular Research, Masaryk <strong>University</strong>, Brno, Czech Republic, 2 Department<br />

of Physical Chemistry, Palacky <strong>University</strong>, Olomouc, Czech Republic<br />

MOLEOnline provides interactive and easy-to-use web interface to analyze molecular channels by<br />

MOLE, which is an universal toolkit for rapid and fully automated location and characterization of<br />

channels, tunnels and pores in molecular structures [1]. This tool is freely available on the Internet<br />

(http://mole.chemi.muni.cz) and overcomes many of the shortcomings and limitations of the recently<br />

developed CAVER software [2]. The core of MOLE algorithm is a Dijsktra’s path search algorithm,<br />

which is applied to a Voronoi mesh. Robust tests on a variety of molecular structures show that MOLE<br />

is a powerful software for exploring large molecular channels, complex networks of channels and<br />

molecular dynamics trajectories (AMBER) in which analysis of a large number of snapshots is<br />

required.<br />

PP470<br />

A Systematic Study of UVA Chemical Sunscreen Filters.<br />

Jacqueline Cawthray, Mark Buntine, Stephen Lincoln<br />

Department of Chemistry, <strong>University</strong> of Adelaide, Adelaide, South Australia, Australia<br />

Sunscreens are a popular and effective method of protecting against the damaging effects of solar<br />

radiation including skin cancer and immune system suppression. Chemical sunscreen filters achieve<br />

this by absorbing ultraviolet radiation and are classified as UVB (280 – 320 nm) or UVA (320 – 400<br />

nm) sunscreens depending on the wavelengths in which they absorb energy. An efficient sunscreen<br />

must afford protection against both UVB and UVA. The majority of chemical filters approved for use<br />

worldwide are UVB absorbers and the few UVA filters approved provide minimal UVA protection or<br />

show only moderate photostability. For example, the enol form of the β-diketone, BMDBM (I), absorbs<br />

strongly in the UVA region but is prone to photodegradation via the keto form (II).<br />

H<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

I<br />

II<br />

We have been exploring the use of theoretical methods as a tool in the design of potentially new<br />

sunscreens. In particular, we have investigated the ability of the SAC-CI method to represent the<br />

trends and properties important to the photochemistry of a series of known β-diketones. In most<br />

cases, the SAC-CI theoretical spectra accurately describe the experimental spectra. This information<br />

can then be applied to the design of potentially new sunscreen filters having the desired properties.<br />

Our latest results will be presented.<br />

[1]. Petrek, M., Kosinova, P., Koca, J., Otyepka, M. Structure 2007, 15, 1357.<br />

[2]. Petrek, M., Otyepka, M., Banas, P., Kosinova, P., Koca, J., and Damborsky, J. BMC Bioinformatics 2006, 7,<br />

316.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP471<br />

Quantum Chemical Simulations of Acetone Adsorption on SWCNTs<br />

Yoshifumi Nishimura 1 , Dmitry Kazachkin 2 , Eric Borguet 2 , Stephan Irle 1<br />

1 Institute for Advanced Research and Department of Chemistry, Nagoya <strong>University</strong>, Nagoya, Japan,<br />

2 Department of Chemistry, Temple <strong>University</strong>, Philadelphia, United States<br />

Many studies of the interaction between single walled carbon nanotubes (SWCNTs) and small<br />

molecules such as acetone have already been reported both experimentally and theoretically, aiming<br />

at the development of advanced SWCNT-based nanosensors [1,2]. However, many aspects of such<br />

interactions are unknown as of yet. Therefore, using the dispersion-augmented self-consistent-charge<br />

density functional tight binding (SCC-DFTB-D) method, we performed quantum chemical model<br />

studies of acetone molecules physisorbed on SWCNTs.<br />

First, we performed a rigorous benchmark of the SCC-DFTB-D method against ab initio MP2 and<br />

higher levels of theory using large basis sets for a coronene-acetone model system. We found that<br />

physisorption energy differences between SCC-DFTB-D and counter-poise corrected<br />

MP2/QZVPP//MP2/SVP (MP2 with highly polarized quadruple zeta basis set single point energies for<br />

MP2 with polarized split valence basis set geometries) adsorption energies are typically smaller than 2<br />

kJ/mol, which is an excellent agreement.<br />

Second, we applied SCC-DFTB-D to the study of the physisorption of acetone on 10 Å long (6,5) and<br />

(11,9) open-ended SWCNT fragments, terminated by hydrogen atoms. We considered adsorption at<br />

three distinct sites – exohedral, endohedral, and at the groove sites of two tubes. We found a clear<br />

correlation between sidewall curvature and physisorption energy (See Figure 1). In addition,<br />

adsorption at the groove sites was found to be about 1.9 times stronger than exohedral adsorption at a<br />

single tube. The acetone-SWCNT interactions are largely dominated by dispersion. Conformations<br />

where the carbonyl group of acetone is parallel to the six membered rings are energetically most<br />

favorable, since these have the largest molecule-surface contact areas. Three distinct desorption<br />

peaks are observed experimentally in temperature-programmed desorption mass spectroscopy (TPD-<br />

MS) experiments, suggesting three distinct sites [3]. The magnitude of experimentally determined<br />

interaction energies can be compared with the computational studies. We will further quantify our<br />

results by presenting molecular dynamics simulations for the motion of acetone on pristine and<br />

oxidized tubes.<br />

PP472<br />

Acid-Base Properties of Sterically Demanding Borane–Phosphine Pairs and Implications for<br />

Metal-Free Hydrogen Activation<br />

Tibor András Rokob, Andrea Hamza, András Stirling, Imre Pápai<br />

Chemical Research Center of HAS, Budapest, Hungary<br />

The impossibility of Lewis adduct formation due to steric bulk opens promising synthetic pathways.<br />

Among others, metal-free hydrogen activation and catalytic hydrogenation have been achieved using<br />

unquenched donor/acceptor pairs, termed frustrated Lewis pairs (FLPs) [1]. Our recent computational<br />

mechanistic studies on the P(tBu) 3 + B(C 6 F 5 ) 3 + H 2 → [(tBu) 3 PH] + [HB(C 6 F 5 ) 3 ] − reaction have revealed<br />

that a weakly bound (tBu) 3 P···B(C 6 F 5 ) 3 complex enables simultaneous interaction of acidic and basic<br />

functionalities with H 2 , yielding the product in a strongly exothermic reaction [2]. We suggested<br />

reactant-side destabilization owing to frustration to be the key of this remarkable transformation.<br />

To promote understanding of the factors governing this reactivity, we now present systematic studies<br />

[3] on the heterolytic hydrogen splitting by various phosphine/borane pairs that were studied<br />

experimentally [4]. Structures and binding energies of weak R 3 P···BR′ 3 complexes have been<br />

determined and their structural features discussed. Net reaction energies have been calculated; their<br />

decomposition into terms including proton and hydride affinities (see Figure) confirms the importance<br />

of appropriate cumulative strength of the acid and base. Reaction profile and strain analysis of a<br />

frustrated and a classical system of comparable strength demonstrate quantitatively the role of<br />

frustration.<br />

Figure1. SCC-DFTB-D interaction energies for acetone-SWCNT plotted versus sidewall curvature<br />

[1] Stephan, D. W. Org. Biomol. Chem. 2008, 6, 1535–1539.<br />

[2] Rokob, T. A.; Hamza, A.; Stirling, A.; Soós, T.; Pápai, I. Angew. Chem. Int. Ed. 2008, 47, 2435–2438.<br />

[3] Hamza, A.; Stirling, A.; Rokob, T. A.; Pápai, I. manuscript in preparation<br />

[4] Welch, G. C.; Stephan, D. W. J. Am. Chem. Soc. 2007, 129, 1880–1881.<br />

[1] Chakrapani, N.; Zhang, Y. M.; Nayak, S. K.; Moore, J. A.; Carroll, D. L.; Choi, Y. Y.; Ajayan, P. M. J. Phys.<br />

Chem. B, 2003, 107, 9308-9311<br />

[2] Snow, E.S.; Perkins, F.K. Nano Letters, 2005, 5, 2414-2417<br />

[3] Kazachkin, D.; Nishimura, Y.; Irle, S.; Morokuma, K.; Vidic, R.,; Borguet, E. Langmuir, 2008, in press.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP473<br />

Different Localization of Protons in the Neutral and Anionic Complexes of Cytosine with<br />

Guanine and 8-Oxoguanine<br />

Iwona Dabkowska 2 , Monika Kobylecka 1 , Piotr Storoniak 1 , Maciej Gutowski 3 , Janusz Rak 1<br />

1 <strong>University</strong> of Gdansk, Department of Chemistry, Gdansk, Poland, 2 Free <strong>University</strong> of Berlin, Physical<br />

and Theoretical Chemistry, Berlin, Germany, 3 Heriot-Watt Unviersity, Chemistry-School of Engineering<br />

and Physical Sciences, Edinburgh, United Kingdom<br />

In the past we demonstrated that an excess electron can alter localization of protons in hydrogenbonded<br />

complexes of nucleic acid bases with model weak acids. The excess electron typically<br />

localizes on a nucleic acid base and promotes an intermolecular proton transfer from the acid to the<br />

base, with the products being a radical of the hydrogenated nucleic acid base and the deprotonated<br />

acid. We also demonstrated that the hydrogenated nucleic acid base can trigger single strand breaks<br />

in DNA with the barrier as small as ca. 5 kcal/mol.<br />

Here we extend these studies to two very important complexes: Watson-Crick complexes of cytosine<br />

(C) with guanine (G) and 8-oxoguanine (8oG). 8oG is a common lesion that develops in consequence<br />

of the DNA interactions with oxidative agents. The calculations have been performed with the B3LYP<br />

exchange-correlation functional and at the RI-MP2/aug-cc-pVDZ level.<br />

PP474<br />

Theoretical Studies of the CO Adsorption on Pt(111) Surface Using Cluster Model<br />

Yu-Wei Huang, Ting-Yi Chou, Yan-Yu Chen, Shyi-Long Lee<br />

Department of Chemistry and Biochemistry, <strong>National</strong> Chung Cheng <strong>University</strong>, Chia-Yi, Taiwan<br />

Correct theoretical description of the adsorption of CO on Pt(111) has been long a challenge to<br />

computational chemistry community [1-3]. In this work, we would like to report the theoretical studies<br />

of the CO adsorption on Pt(111) surface using cluster model. In our present work, recently developed<br />

density functional theory method, such as BMK [4] and M06 [5], were used to study CO adsorption on<br />

Pt cluster. Ab initio method, such as HF and MP2 were also adopted to calculate the electronic<br />

structure of the adsorbed system for comparison. Model clusters 7-3, 10-6, 19-12 used in this study<br />

were constructed systematically and shown below. CO can then be adsorbed on atop, fcc and hcp<br />

position. Relationship between exchange-correlation functional and site preference is established.<br />

Effect of cluster shape and size on the electronic structure and bonding of CO to Pt(111) surface will<br />

also be discussed. Our results will also be compared to Gil’s results [2].<br />

Our results indicate that the excess electron localizes on cytosine in both the G-C and 8oG-C<br />

complexes and the anionic complexes are adiabatically bound with respect to the corresponding<br />

neutral Watson-Crick pairs. Moreover, the excess electron promotes an intermolecular proton transfer<br />

from G or 8oG to C. There are small barriers for these intermolecular proton transfers on the potential<br />

energy surfaces of electronic energies: 2.6 (B3LYP) and 1.3 (MP2) kcal/mol for GC - and 1.5 (B3LYP)<br />

and 0.2 (MP2) kcal/mol for 8oGC - . The barriers are even smaller for electronic energies corrected for<br />

zero-point vibrations and thermal corrections to Gibbs free energies.<br />

We conclude that the anionic 8oGC complex is particularly susceptible to intermolecular proton<br />

transfer but the barriers are also very low for the GC - complex.<br />

A radical of the hydrogenated cytosine, which is a product of these elementary intermolecular proton<br />

transfer steps, displays an adiabatic electron affinity of ca. 0.5 eV and provides an avenue to a single<br />

strand break in DNA [1].<br />

[1] I. Dąbkowska, J. Rak, M. Gutowski, Eur. Phys. J. D 2005, 35, 429-431.<br />

[1] Feibelman, P.J.; Hammer B.; Nørskov J.K.; Wagner F.; Scheffler M.; Stumpf R.; Watwe R.; Dumesic J. J.<br />

Phys. Chem. B 2001, 105, 4018<br />

[2] Gil, A.; Clotet, A.; Ricart, J.M.; Kresse, G.; Garcia-Hernandez, M.; Rosch, N.; Sautet, P. Surf. Sci 2003, 530,<br />

71<br />

[3] Hu, Q. M.; Reuter, K.; Scheffler, M. Phys. Rev. Lett. 2007, 98, 176103<br />

[4] Boese, A. D.; Martin, J.M.L. J. Chem. Phys. 2004, 121, 3405<br />

[5] Zhao, Y.; Truhlar D. G. J. Chem. Phys. 2006, 125, 194101


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP475<br />

In Silico and In Vitro Mutagenesis of PA-IIL Lectin – Correlation with Structure-Function<br />

Analysis<br />

Jan Adam, Martina Pokorna, Zdenek Kriz, Martin Prokop, Jaroslav Koca, Michaela Wimmerova<br />

Masaryk <strong>University</strong>, <strong>National</strong> Centre for Biomolecular Research, Brno, Czech Republic<br />

Protein-carbohydrate interactions play a crucial role in recognition and adhesion of pathogen to host<br />

tissue cells in the first step of their invasion and infectivity. Pseudomonas aeruginosa is an<br />

opportunistic pathogen, responsible for numerous nosocomial infections in immunocompromised<br />

patients. The bacterium colonises patients with chronic lung diseases and its infections is fatal in<br />

cystic fibrosis patients. P. aeruginosa produces a fucose-binding lectin PA-IIL that plays a role in<br />

virulence and adhesion [1]. Similar lectins have been found in other opportunistic bacteria like<br />

Ralstonia solanacearum [2], Chromobacterium violaceaum [3] and Burkholderia cenocepacia [4], with<br />

different binding preferences despite only slight sequence changes.<br />

Mutations of the PA-IIL lectin were designed according to these differences, cloned/modelled and<br />

investigated both in vitro, and in silico [5,6]. Thermodynamics of binding using isothermal titration<br />

microcalorimetry helped to rationalise importance of particular amino acids in PA-IIL binding site and<br />

to define overall enthalpy and entropy contributions to the interaction. Good correlation between<br />

experimentally constructed mutants and in silico prediction of their sugar specificity allowed to perform<br />

much wider screening of mutated proteins focusing on importance of particular amino acids for<br />

specificity and/or affinity. This approach significantly saves experimental time needed for construction<br />

of mutant protein library.<br />

PP476<br />

Theoretical Study of the minor Channel forming HNO 3 in the HO 2 + NO reaction<br />

Marie-Therese Rayez, Jean-Claude Rayez<br />

Universite Bordeaux1-ISM-groupe THEO, 33405 Talence Cedex, France<br />

The reaction HO 2 + NO → OH + NO 2 (1a) plays a key role in controlling the interconversion between<br />

OH and HO 2 radicals in the troposphere. This reaction is also a major source of tropospheric ozone<br />

through the conversion of NO to NO 2 followed by NO 2 photolysis.<br />

The observation of a minor channel (1b) forming nitric acid in the reaction channel HO 2 + NO → HNO 3<br />

(1b), has been reported by Butkovskaya et al [1]. This minor production of HNO 3 has been shown by<br />

these authors to be favoured by temperature decrease, pressure increase and the presence of water<br />

vapour [2].<br />

The mechanism of the titled reaction has been investigated using electronic quantum chemistry<br />

coupled to statistical RRKM-type calculations. The first step is the formation of the HOONO cis and<br />

trans complexes which can decompose (channel 1a) or isomerise to HNO 3 (channel 1b). We have<br />

demonstrated that the formation of HNO 3 can also occur through a second path along the<br />

decomposition channel thanks to the presence of a loose intermediate.<br />

Calculations also show that the addition of H 2 O molecules makes the transition states governing the<br />

HNO 3 formation more accessible suggesting an explanation for the experimental increase of HNO 3<br />

formation in the presence of water vapour. This study allows us to suggest a mechanism which agrees<br />

with the experimental results described above.<br />

[1]. Mitchell, E.; Houles, C.; Sudakevitz, D.; Wimmerova, M.; Gautier, C.; Perez, S.; Wu, A. M.; Gilboa-Garber, N.;<br />

Imberty, A., Structural basis for oligosaccharide-mediated adhesion of Pseudomonas aeruginosa in the lungs of<br />

cystic fibrosis patients. Nature Structural Biology 2002, 9, (12), 918-921.<br />

[2]. Sudakevitz, D.; Kostlanova, N.; Blatman-Jan, G.; Mitchell, E. P.; Lerrer, B.; Wimmerova, M.; Katcoff, D. J.;<br />

Imberty, A.; Gilboa-Garber, N., A new Ralstonia solanacearum high-affinity mannose-binding lectin RS-IIL<br />

structurally resembling the Pseudomonas aeruginosa fucose-specific lectin PA-IIL. Molecular Microbiology 2004,<br />

52, (3), 691-700.<br />

[3]. Pokorna, M.; Cioci, G.; Perret, S.; Rebuffet, E.; Kostlanova, N.; Adam, J.; Gilboa-Garber, N.; Mitchell, E. P.;<br />

Imberty, A.; Wimmerova, M., Unusual entropy-driven affinity of Chromobacterium violaceum lectin CV-IIL toward<br />

fucose and mannose. Biochemistry 2006, 45, (24), 7501-10.<br />

[4]. Lameignere, E.; Malinovska, L.; Slavikova, M.; Duchaud, E.; Mitchell, E. P.; Varrot, A.; Sedo, O.; Imberty, A.;<br />

Wimmerova, M., Structural basis for mannose recognition by a lectin from opportunistic bacteria Burkholderia<br />

cenocepacia. Biochem J 2008, 411, (2), 307-18.<br />

[5]. Adam, J.; Pokorná, M.; Sabin, C.; Mitchell, E.; Imberty, A.; Wimmerová, M., Engineering of PA-IIL lectin from<br />

Pseudomonas aeruginosa – Unravelling the role of the specificity loop for sugar preference. BMC structural<br />

biology 2007, 7, (1), 36.<br />

[6]. Adam, J.; Kriz, Z.;Prokop, M.; Wimmerova, M.; Koca, J., Mutagenesis and docking studies of Pseudomonas<br />

aeruginosa PA-IIL lectin – predicting binding modes and energies in silico, manuscript submitted<br />

[1] Butkovskaya, N.; Kukui, A.; Pouvesle, N.; Le Bras, G. J. Phys. Chem. A 2005, 109, 6509<br />

[2] Butkovskaya, N.; Kukui, A.; Le Bras, G. J. Phys. Chem. A 2007, 111, 9047


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP477<br />

Solvent Effect for the Interaction of Adenine Adducts with Thymine<br />

Prabhat K. Sahu, Yu-Wei Huang, Shyi-Long Lee<br />

Department of Chemistry and Biochemistry, <strong>National</strong> Chung Cheng <strong>University</strong>, Chia –Yi, Taiwan<br />

The existence of DNA adducts bring the danger of carcinogenesis because of mispairing with normal<br />

DNA bases. Solvation free energies (∆G solv ) calculations for the 1,N 6 -ethenoadenine adducts (εA) and<br />

1,N 6 -ethanoadenine adducts (EA) with thymine have been achieved by using COSMO model (C-PCM)<br />

at B3LYP/6-31+G*, so as to predict the association energies for the adenine adducts with thymine.<br />

The energetic computed in solvent phase has also been compared with those reported earlier in gas<br />

phase 1 . The calculated Gibbs free energy corrected binding energy (aq) value for εA(2)-T(I) is 3.51<br />

Kcal/mol, which is highest among the etheno adduct-thymine complexes and about 0.03 Kcal/mol less<br />

than those obtained for Watson-Crick A-T base pair and similarly, the Gibbs free energy corrected<br />

binding energy (aq) value for EA(2)-T(I) is 3.94 Kcal/mol, which is highest among the ethano adductthymine<br />

complexes and about 0.4 Kcal/mol more than those obtained for Watson-Crick A-T base pair.<br />

It has also been observed that EA(2)-T(I) and εA(2)-T(I) have been identified to be best possible<br />

mispair out of several different stable conformers for each type of adenine adduct with thymine in<br />

terms of current calculated results (binding energies) and the reported gas phase results (binding<br />

energies and reaction entalpies) with regard to Watson-Crick A-T base pair.<br />

PP478<br />

Ab Initio Molecular Energies by Fragmentation: A Feasible Strategy for Non-Bonded<br />

Interactions<br />

M. A. Addicoat, M. A. Collins<br />

Research School of Chemistry, <strong>Australian</strong> <strong>National</strong> <strong>University</strong>, Canberra, ACT 0200, Australia<br />

Ab initio quantum chemistry can calculate the total electronic energy for small to moderate sized<br />

molecules with reliable accuracy. However, the computational cost of these methods becomes<br />

prohibitive for many larger molecules of interest to chemists. Theoretical studies of the dynamical<br />

mechanisms of chemical reactions have been even more severely restricted to the smallest<br />

molecules, partly because the electronic energy must be calculated for a large number of molecular<br />

geometries. We have shown how the bonding energy of large organic molecules can be estimated<br />

from the energies of relatively small molecular fragments. We report current progress on the<br />

incorporation of weak non-bonded interactions to achieve chemical accuracy for the total molecular<br />

energy.<br />

Sahu, P. K.; Kuo, C. W.; Lee, S. L. J. Phys. Chem. B 2007, 111, 2991


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP479<br />

Agonist Binding at the Calcium Sensing Receptor: A Computational Approach<br />

Susan Corley, Hoi-Ming Chan, Meredith Jordan<br />

School of Chemistry, The Univeristy of Sydney, Sydney, Australia<br />

The divalent metal cations Ca 2+ , Mg 2+ and Sr 2+ all activate the calcium sensing receptor (CaSR) with<br />

differing potencies. [1] Amino acid residues likely to be important to be important in the CaSR binding<br />

site are identified using DFT, and their interactions with the divalent cations examined. Calculations<br />

are performed for both gas phase interactions and using the polarized continuum model (PCM),<br />

whereby a dielectric constant D=4 is used to represent a partially open protein environment. The<br />

thermodynamics of binding indicate that charged residues glutamate and aspartate, along with the<br />

polar residues asparagine and glutamine, together with the protein backbone group represent the<br />

most likely residues for metal binding.<br />

Binding energies alone, however, do not account for the different observed potencies of the three<br />

metal cations. Ca 2+ and Sr 2+ support a coordination number of seven, whereas Mg 2+ has a maximum<br />

coordination number of six. Thus there is more facile migration of amino acid residues from the second<br />

to the first coordination shell of Ca 2+ and Sr 2+ as compared to Mg 2+ , which requires substantial<br />

rearrangement of water molecules in the first coordination shell and is thus kinetically unfavorable due<br />

to a large activation barrier. This is reflected in the significantly higher activation energies for the<br />

transition state between second coordinate and first coordinate shell binding for model amino acid<br />

residue binding to Mg 2+ compared to Ca 2+ and Sr 2+ (figure 1).<br />

PP480<br />

Performance of Density Functionals in a QM and QM/MM study of the Catalytic Mechanism of<br />

β-Galactosidase<br />

Pedro Alexandrino Fernandes, Natércia Brás, Sara Moura-Tamames, Ramos Maria<br />

Requimte/Faculty of Sciences, <strong>University</strong> of Porto, Porto, Portugal<br />

Galactooligosaccharides (GOS) are special glycosides produced by the enzyme β-galactosidase.<br />

They contain galactose and glucose molecules. These sugars are considered to be functional food<br />

ingredients because they stimulate the growth of bifidobacteria and lactobacilli, combining prebiotic<br />

properties beneficial to the human health with physico-chemical properties beneficial in food<br />

processing [1].<br />

The catalytic mechanism on retaining glycosides has been largely studied experimentally [2]. These<br />

studies propose that occurs via a double displacement mechanism involving a glycosylation and a<br />

deglycosylation step, which is shown in Figure 1.<br />

Our goal is to perform a theoretical model, which elucidates at atomic-level detail the catalytic<br />

mechanism of the β-galactosidase from E.coli. A minimal gas-phase model has been used previously,<br />

to test the performance of a large number of density functionals [3], namely B3LYP, BHandHLYP,<br />

B97-2, B98, PBE1PBE, B1B95, BB1K, MPW1K and MPWB1K. A significant dispersion of results has<br />

been found (up to 7.6 kcal.mol -1 ) in the activation energies. Conversely, the catalytic effect of a<br />

fundamental hydrogen bond was well reproduced by all functionals (with a dispersion smaller than 1.1<br />

kcal.mol -1 ). The dispersion was also small in reaction energies. The results of selected functionals<br />

were recalculated with a smaller model, using DFT and higher levels of theory, namely MP2, MP4 and<br />

QCISD(T), for comparison.<br />

Once set on the adequate density functional a large model, including a radius with 15 Å around the<br />

lactose substrate, was used to follow the catalytic mechanism. The ONIOM method was employed to<br />

deal with such a large system, at the BB1K:AMBER theoretical level [4]. The results of the larger<br />

model captured the catalytic effect of the enzyme, lowering the barrier by more than 12 kcal.mol -1 . The<br />

origin of the catalytic power of the enzyme was also discussed.<br />

Figure 1) Energy profile diagrams for the reaction M[(H 2O) 6] 2+ .L M[(H 2O) 6L] 2+ ‡ M[(H 2O) 5L] 2+ .H 2O where L =<br />

formamide (left) and methanol (right).<br />

[1] Holzapfel, W.H.; Schillinger, U. Food Res. Int., 2002, 35, 109.<br />

[2] Juers, D.H.; Heightman, T.D.; Vasella, A.; MacCarter, J.D.; Mackenzie, L.; Withers, S.G.; Matthews, B.W.<br />

Biochemistry, 2001,40, 14781–14794.<br />

[3] Brás, N. F.; Moura-Tamames, S.; Fernandes, P. A.; Ramos, M. J. J. Comput. Chem., ASAP, DOI:<br />

10.1002/jcc.20977<br />

[4] Brás, N. F.; Fernandes, P. A.; Ramos, M. J. submitted.<br />

Similarly, the interconversion between monodentate and bidentate binding for residues modeled with<br />

formate was observed to have a significantly higher reaction barrier for Mg 2+ (figure not shown).<br />

Therefore, we suggest that the selectivity of calcium over magnesium binding at the calcium sensing<br />

receptor and similar proteins is due to kinetic rather than thermodynamic considerations.<br />

[1] Brown, E. M.; Macleod, R. J. Physiological Reviews 2001, 81, 239-297.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP481<br />

Molecular Modelling of Hydrotalcite Layered Double Hydroxide Structures Intercalcated with<br />

Transition Metal Oxide Anions<br />

Vinutha Murthy 1 , Howard D Smith 2<br />

1 Charles Darwin <strong>University</strong>, School of Environmental & Life Sciences, Darwin, NT, Australia, 2 Curtin<br />

<strong>University</strong> of Technology, Nanochemistry Research Institute, Bentley, WA, Australia<br />

Hydrotalcite is a mineral that belongs to a class of chemicals often referred to as layered double<br />

hydroxides and which can be produced synthetically by reacting magnesium II and aluminium III<br />

solutions in the presence of a strong base. In-situ absorption of transition metal anions may occur if<br />

they are present at the time the hydrotalcite is prepared. Our previous experimental studies have<br />

shown that while some transition metal oxide anions such as VO 4 3- and CrO 4 2- may be easily absorbed<br />

into the mineral structure, other anions such as MoO 4 2- are not.<br />

Intercalation, the process where anions become contained within the interlayer spaces between the<br />

octahedral magnesium-aluminium hydroxy layers of hydrotalcite is one possible mechanism of in-situ<br />

absorption. Initial studies using Molecular modelling with empirical forcefield is being carried out in<br />

conjunction with X-ray Diffraction and Electron-microscope data to determine the size or orientation of<br />

transition metal oxide anions and their impact upon hydrotalcite crystal structure.<br />

PP482<br />

Computational Studies on the Thermal Fragmentation Reactions of 1,4,2-Oxathiazole<br />

Derivatives – A Convenient Synthesis of Isothiocyanates<br />

Robert O'Reilly, Leo Radom<br />

School of Chemistry, <strong>University</strong> of Sydney, Sydney, NSW, Australia<br />

Thermolysis of 1,4,2-oxathiazole derivatives has been shown to result in the formation of<br />

isothiocyanates (ITCs) and the corresponding carbonyl derivatives [1].<br />

R 2<br />

R 1 S<br />

R 3 ∆<br />

O<br />

O N R 1<br />

R 3<br />

R 2 + N C S<br />

ITCs are gaining increasing notoriety due to their promising anticancer properties [2].<br />

Whilst relatively few reports on ITC synthesis via this fragmentation reaction exist in the literature, a<br />

polymer-supported synthesis of ITCs has recently been published [3]. Given the potentially useful<br />

applications of this approach to ITC synthesis, and the ability to generate libraries for biological testing<br />

via the polymer-supported approach, a knowledge of the mechanism(s) by which the fragmentation<br />

proceeds is valuable. This poster outlines a mechanistic study carried out using quantum chemical<br />

procedures.<br />

[1] Huisgen, R.; Mack, W.; Anneser, E. Angew. Chem. 1961, 73, 656-657.<br />

[2] Conaway, C. C.; Yang, Y.; Chung, F. –L. Curr. Drug Metab. 2002, 3, 233-255.<br />

[3] Burkett, B. A.; Kane-Barber, J. M.; O’Reilly, R. J.; Shi, L. Tet. Lett. 2007, 48, 5355-5358.<br />

[1]. Smith HD, Parkinson GM and Hart RD, In situ absorption of molybdate and vanadate during precipitation of<br />

hydrotalcite from sodium aluminate solutions. Journal of Crystal Growth, 2005, 275 (1-2): e1665-e1671.<br />

[2]. P. Kovar, M. Pospisil, M. Nocchetti, P. Capkova, K. Melanova. Molecular modeling of layerd double<br />

hydroxides intercalcated with benzoate, modeling and experiment. J Mol Model, 2007; 13: 937-942


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP483<br />

Potential Energy Hypersurface for the Methane Radical Cation<br />

Lucas D. Speakman, Justin M. Turney, Henry F. Schaefer III<br />

<strong>University</strong> of Georgia, Center for Computational Chemistry, Athens, GA, United States<br />

The methane cation has been studied extensively because of its fundamental importance in chemistry.<br />

This cation is<br />

1) prevalent in interstellar clouds and planetary atmospheres [1].<br />

2) likely involved in the chemical evolution that must have preceded the origin of life [2].<br />

3) highly reactive and leads to other fundamental molecular ions such as CH 5 + and C 2 H 5 + .<br />

4) a major radiation product of methane, a species involved in many chemical and biochemical<br />

processes.<br />

5) one of the simplest systems subject to Jahn-Teller distortion.<br />

Because of its significance, an accurate potential energy surface is required to elucidate spectroscopic<br />

properties. The tetrahedral configuration has a triply degenerate 2 T 2 which could lead to C 2v , D 2d , C 3v ,<br />

or lower symmetry distortions because of Jahn-Teller splitting. Due to this degeneracy, several<br />

theoretical methods fail to sufficiently describe these states in the tetrahedral configuration. The<br />

research method utilized is an equation-of-motion ionization potential coupled cluster (EOMIP-CC)<br />

theory which adequately represents the three degenerate states and has not yet been applied to this<br />

cation. Several low lying conformers of CH 4 + will be investigated for a thorough potential energy<br />

hypersurface. Vibrational perturbation theory will also be computed at each stationary point to yield<br />

accurate fundamental frequencies. Finally, dissociation paths of the methane cation into CH 3 + + H and<br />

CH 2 + and H 2 will be explored.<br />

[1] Miller, S. L.; Urey, H. C.; Science, 1959, 130, 245.<br />

[2] Arents, J.; Allen, L. C.; J. Chem. Phys. 1970, 53, 73.<br />

PP484<br />

A Computational Study of Substrate Mechanism of Pyruvate-Formate Lyase<br />

Karmen Condic-Jurkic 1 , V. Tamara Perchyonok 2 , Hendrik Zipse 3 , David M. Smith 1<br />

1 Rudjer Boskovic Institute, Centre for Computational Solutions in the Life Sciences, Zagreb, Croatia,<br />

2 Monash <strong>University</strong>, School of Chemistry, Melbourne, Australia, 3 Ludwig-Maximilians-Universität,<br />

Department Chemie und Biochemie, München, Germany<br />

DFT and high-level quantum mechanical calculations have been performed on small model systems<br />

relevant to the substrate mechanism of Pyruvate Formate-Lyase (PFL), comparing the reactivity of the<br />

natural substrate pyruvate with that of the known inhibitor oxamate. For both substrates the presently<br />

accepted (consensus) mechanism, involving the addition of a cysteine-based thiyl radical to the<br />

carbonyl carbon of the substrate, is compared to an alternative mechanism involving hydrogen<br />

abstraction as the primary step. This mechanism, relevant because of the known structural homology<br />

between PFL and the ribonucleotide reductase family of enzymes,<br />

is found to display similar reaction barriers to the consensus mechanism, but is much less favorable<br />

with respect to the reaction energetics. The inhibitory effect of oxamate can be traced back to an<br />

increase in reaction barrier and reaction energy along the consensus mechanism pathway. The highlevel<br />

quantum results have been subsequently used as benchmarks to examine the suitability of<br />

applying a coupled quantum-mechanical/molecular-mechanical (QM/MM) approach to this system.<br />

This comparison shows that while the QM/MM technique gives reaction barriers and enthalpies in<br />

good agreement with the complete quantum treatment, some caution must be exercised for properties<br />

such as complexation energies. It also seems that DFT tends to underestimate complexation<br />

energies, compared to high-level results.<br />

[1] Condic-Jurkic, K.; Perchyonok, T.V.; Zipse, H.; Smith, D. M., On the modeling of arginine-bound carboxylates:<br />

A case study with Pyruvate Formate-Lyase, J. of Comp. Chem., 2008, in print<br />

[2] Knappe, J.; Elbert, S.; Frey., M.; Wagner, A.F.V. Biochem. Soc. Trans. 1993, 21, 731-734<br />

[3] Becker, A.; Kabsch W J. Biol. Chem. 2002, 277, 400036-40042<br />

[4] Becker, A.; Fritz,-Wolf, K.; Kabsch, W.; Schultz, S.; Wagner, A.F.V.; Knapper J. Nat. Struc. Bio. 1999, 6, 969-<br />

975


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP485<br />

The Enthalpy of Formation and Anharmonic Force Field for Diacetylene<br />

Andrew Simmonett, Wesley Allen, Henry F. Schaefer III<br />

Center for Computational Chemistry, <strong>University</strong> of Georgia, Athens, GA, United States<br />

The enthalpy of formation for the diacetylene molecule (C 4 H 2 ) is pinpointed using state-of-the-art<br />

computational methodology, accounting for high-order correlation effects, relativity, non-Born<br />

Oppenheimer effects and explicit consideration of vibrational anharmonicity. Molecular energies are<br />

obtained through the coupled cluster hierarchy of methods with singles and doubles (CCSD),<br />

quasiperturbative triples [CCSD(T)], full triple excitations (CCSDT) and perturbative quadruples<br />

[CCSDT(Q)] in concert with Dunning’s correlation consistent family of basis sets (cc-pVXZ,<br />

X=D,T,Q,5,6). Our fundamental vibrational frequencies are derived from an all electron CCSD(T) full<br />

quartic force field, computed using Dunning’s quadruple-ζ cc-pCVQZ basis, which includes tight<br />

functions to provide a satisfactory description of core correlation. Second order vibrational<br />

perturbation theory (VPT2) is used to extract the fundamental vibrational frequencies from the<br />

anharmonic force field; we stress that no empirical scale factors are used.<br />

PP486<br />

Perturbative Triples Correction for State-Specific Multireference Coupled Cluster<br />

Francesco A. Evangelista, Wesley D. Allen, Henry F. Schaefer III<br />

Center for Computational Chemistry, <strong>University</strong> of Georgia, Athens, Georgia, United States<br />

In order to generalize the single reference CCSD(T) method to the multireference case, we have<br />

derived a new perturbative correction scheme for triples excitations within the framework of statespecific<br />

multireference coupled cluster theory based on the Jeziorski-Monkhorst ansatz. Our<br />

derivation is based on the Lagrangian formalism and is applicable to Mukherjee (Mk-MRCC) as well as<br />

Brilluoin Wigner (BW-MRCC) multireference coupled cluster. When the model space consists of one<br />

determinant, and Hartree Fock orbitals are used, we recover the conventional CCSD(T) energy. In the<br />

case of Mk-MRCC, we can achieve a non-iterative implementation in two different ways: 1) neglecting<br />

the off diagonal elements of the Fock matrix for each reference, or 2) neglecting the coupling terms in<br />

the triples equations. We compare our formulations to others previously advanced and also comment<br />

on the possibility of treating quadruple excitations within the same scheme.<br />

Using this ab initio methodology, we compute the enthalpy for the isogyric reaction<br />

2 H–C≡C–H → H–C≡C–C≡C–H + H 2<br />

to be (0.03, 0.81) kcal mol -1 at (0, 298.15) K which, combined with acetylene’s enthalpy of formation<br />

from the literature yields ∆ f<br />

H o 0<br />

(diacetylene) = 109.4 ± 0.3 kcal mol -1 o<br />

and ∆ f<br />

H 298<br />

(diacetylene) =<br />

109.7 ± 0.3 kcal mol -1 .


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP487<br />

A Theoretical Study on the S 1 /S 2 Conical Intersection Along the Amino Inversion Coordinate of<br />

Aminonaphthalene<br />

Masayuki Nakagaki, Haruyuki Nakano<br />

Department of Molecular Chemistry, Graduate School of Sciences, Kyushu <strong>University</strong>, Fukuoka,<br />

Japan<br />

Aniline is the most basic aromatic molecule with an amino group. Many studies therefore have been<br />

done both theoretically and experimentally, and its electronic structure has been well elucidated so far<br />

in this molecule, the πσ* Rydberg-type excited state corresponds to the S 2 state and lies near the<br />

valence ππ* S 1 state [1, 2]. The potential energy surface of the amino group depends on the electronic<br />

state; that is the amino group in S 0 state has pyramid-like non-planar structure while it is almost planar<br />

in S 1 state [3].<br />

Similar to aniline, 1-aminonaphthalene (1AN) and 2-aminonaphthalene (2AN) (Fig.1), which are<br />

monosubstituted acenes including amino group, are also fundamental molecules. In AN, when the<br />

amino group has non-planar structure, the ππ* and πσ* states belong to the same irreducible<br />

representation due to the lower molecular symmetry. This fact implies to us that the inversion mode of<br />

the amino group plays an important role in the interaction between ππ* and πσ* states. In order to<br />

clarify the interaction in details, we theoretically examined the potential energy surfaces of the ππ* and<br />

πσ* states along the stretching and inversion motions of amino group coordinates. In particular, the<br />

focus was on the seeking of the conical intersection and evaluation of the interaction between the two<br />

states.<br />

We first performed geometry optimization for 1AN and 2AN in ππ* and πσ* states using the complete<br />

active space self-consistent field (CASSCF) method with the 6-31++G(d,p) basis set. The active space<br />

was constructed from 12 electrons in 12 active orbitals including 6 π orbitals and 5 π* orbitals and 1 σ*<br />

orbital. Then we calculated the two-dimensional potential surfaces V(θ,r) for two lowest singlet excited<br />

states in both diabatic and adiabatic representations, using the state averaged (SA-) CASSCF<br />

method, where two parameters θ and r are defined in Figure 2. We evaluated also the non-adiabatic<br />

coupling matrix elements < Ψi | ∂ / ∂r | Ψ j > and < Ψi | ∂ / ∂θ | Ψ j > using the finite difference method.<br />

PP488<br />

Non Specific Binding of Positron Emission Tomography (PET) Ligands as Seen from an Ab<br />

Initio Computational Study<br />

Lula Rosso 1 , Antony Gee 2 , Ian Gould 1<br />

1 Department of Chemistry, Imperial College London, London, United Kingdom, 2 Clinical Imaging<br />

Centre, GlaxoSmithKline, United Kingdom<br />

In developing new PET radiotracers, a common reason for candidate failure is that high non-specific<br />

and sub cellular binding obscures the specific binding to the molecular target and, thus, reduces the<br />

quality of the PET scan data. Non-specific binding is thought to be correlated in part to a molecule’s<br />

lipophilicity, typically estimated by measuring (or calculating) octanol-water partition coefficient. This is,<br />

however, a gross simplification of a complex phenomenon. The purpose of this ab-initio computational<br />

study is to increase our understanding of non-specific binding by investigating the molecular basis of<br />

ligand-membrane interaction.<br />

Ten well-characterized central nervous system PET radiotracers acting on a variety of molecular<br />

targets were used as primary set. Quantum mechanical methods were used to estimate accurately the<br />

strength of the electronic interaction between individual drug molecules and a single phospholipid<br />

molecule commonly present in mammalian membranes. This was achieved by finding the lowest<br />

optimized ground state energy of several drug-lipid complexes at Hartree-Fock and B3LYP level of<br />

theory, without using experimental data.<br />

The computed energies showed a non-trivial correlation with the in vivo non-specific binding<br />

distribution volumes relative to the free tracer plasma concentration, Fig. 1. Significantly, the drugs<br />

with stronger lipid interaction possessed a higher non-specific binding value. While, for the drugs<br />

taken in consideration in this study, the water-octanol partition coefficient, logP, did not show good<br />

predictive power of the in vivo non-specific binding, Fig.2.<br />

The computational method used provided an interesting insight into non specific binding behaviour in<br />

terms of membrane interaction. This method was further validated on a blinded set of 9 candidate<br />

radiotracers and correctly identified all 3 successful drugs in the set as the molecules forming the<br />

weakest lipid complexes.<br />

We obtained the global minima of the ππ* and πσ* states. For the optimized structure of the ππ* states,<br />

ππ* state was the lowest excited state, while for the optimized structure of the πσ* states, πσ* state<br />

was the lowest excited state. The major difference of structures in two states was the C-N bond length<br />

r. The bond length in πσ* state is shorter by about 0.1 Ǻ than that in the ππ* state. These results<br />

indicate that potential energy curves along r cross each other at the CASSCF level. We also found<br />

that the interaction between the two states in diabatic representation is proportional to θ, whereas it is<br />

almost independent of r.<br />

Fig.1 Structures of aminonaphthalene Fig. 2 Definition of parameters θ and r<br />

Fig. 1 Fig. 2<br />

[1] Y. Honda et al. J. Chem. Phys. 2002, 117, 2045-2052.<br />

[2] T. Ebata et al. J. Phys. Chem. A, 2002, 106, 11070-11074.<br />

[3] J. C. Jiang, C. E. Lin, J. Mol. Struct. (Theochem), 1997, 392, 181-191.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP489<br />

Laser Control of Photoassociation in Diatomic Molecule<br />

Kiyoshi Nishikawa 1 , Masatoshi Sano 1 , Atsuhito Tawada 1 , Shunichi Taniguchi 1 , Kimikazu Sugimori 2 ,<br />

Hidemi Nagao 1<br />

1 Kanazawa <strong>University</strong>, Division of Mathematical and Physical Science, Kanazawa, Japan, 2 Kinjo<br />

<strong>University</strong>, Hakusan, Japan<br />

In this work, we simulate the laser-induced photoassociation of the collision reaction O + H to generate<br />

the OH molecule. The photoassociation process involves the continuum-discrete transition, and the<br />

initial wavepacket corresponding to the collision pair is generally given by a superposition of<br />

continuum state. By numerically solving the time-dependent Schrödinger equation (TDSE) with split<br />

operator method (SOM), we can take into account implicitly but completely for the continuum state in<br />

the photoassociation process. In our simulation, we find interestingly the above threshold dissociation<br />

(ATD) spectrum due to the continuum-continuum transition as well as the target bound states by the<br />

photoassociation. In order to analyze the continuum state involved, we show the effective method by<br />

means of the quasicontinuum state on the Morse potential obtained by numerically diagonalizing the<br />

Fourier grid Hamiltonian (FGH) of the total system. We finally find out the effective laser pulse, which<br />

induces the complete generation of the target bound state from the initial wavepacket.<br />

PP490<br />

Photochemistry of Fischer Carbene Complexes<br />

Israel Fernandez, Miguel A. Sierra<br />

Dpto. de Química Orgánica, Universidad Complutenses de Madrid, Madrid, Spain<br />

Fischer carbene complexes are irreplaceable building blocks. While the thermal reaction chemistry of<br />

these organometallic complexes that has been widely studied, their photochemical reactivity and the<br />

mechanisms of the photochemical reactions mechanisms were at the beginning of our work barely<br />

explored.[1]<br />

By means of experimental and computational tools, we shall discuss the intimacies of the photocarbonylation<br />

process and the reaction of the produced metalla-ketenes with imines (Figure 1).[2]<br />

Furthermore, we will present that other photochemical reaction pathways, which are effective also for<br />

tungsten carbene complexes, are possible after proper selection of the ligands surrounding the metal<br />

centre.[2d,e]<br />

CO<br />

CO XR 1<br />

OC Cr<br />

OC R<br />

CO<br />

hν<br />

∆<br />

(CO) 4 Cr<br />

XR 1<br />

R<br />

O<br />

?<br />

R XR 1<br />

(CO) 4 Cr<br />

O<br />

R 2 R 4<br />

R 3<br />

N<br />

R<br />

R 1 X<br />

R 2<br />

R 3<br />

N<br />

O R 4<br />

Figure 1. Photocarbonylation reaction of Fischer Carbene Complexes<br />

( Initial wavepacket) ( target bound state and outgoing wave)<br />

[1] Hegedus, L. S., Tetrahedron 1997, 53, 4105.<br />

[2] (a) Fernández, I.; Sierra, M. A.; Mancheño, M. J.; Gómez-Gallego, M.; Cossío, F. P. Chem. Eur. J. 2005, 11,<br />

5988. (b) Lage, M. L.; Fernández, I.; Mancheño, M. J.; Sierra, M. A. Inorg. Chem. 2008, 47, 5253. (c) Arrieta, A.;<br />

Cossío, F. P.; Fernández, I.; Gómez-Gallego, M.; Lecea, B.; Mancheño, M. J.; Sierra, M. A. J. Am. Chem. Soc.<br />

2000, 122, 11509. (d) Sierra, M. A.; Fernández, I.; Mancheño, M. J.; Gómez-Gallego, M.; Torres, M. R.; Cossío,<br />

F. P.; Arrieta, A.; Lecea, B.; Poveda, A.; Jiménez-Barbero, J. J. Am. Chem. Soc. 2003, 125, 9572. (e) Fernández,<br />

I.; Sierra, M. A.; Gómez-Gallego, M.; Mancheño, M. J.; Cossío, F. P. Angew. Chem. Int. Ed. 2006, 45, 125.<br />

[1] Sugimori, K,; Ito, T.; Takata, Y.; Ichitani, K.; Nagao, H.; Nishikawa, K. Int. J. Quantum Chem. 2006, 106, 3079-<br />

3086.<br />

[2] Sugimori, K,; Ito, T.; Takata, Y.; Ichitani, K.; Nagao, H.; Nishikawa, K. J. Phys. Chem. A, 2007, 111, 9417-<br />

9423.<br />

[3] Nishikawa, K.; Sugimori, K.; Nagao, H. AIP Conference Proceedings, 2007, 963, 843-846.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP491<br />

Interaction of Roxithromycin with 50s Large Ribosomal Subunit: Molecular Dynamics<br />

Simulation.<br />

Wai Keat Yam, Habibah A Wahab<br />

Pharmaceutical Design and Simulation Laboratory, School of Pharmaceutical Sciences, Universiti<br />

Sains Malaysia, Penang, Malaysia<br />

Roxithromycin (ROX) is a second generation, semi-synthetic macrolide antibiotic derived from<br />

erythromycin. It has a broader antibacterial spectrum compared to earlier generation of macrolides<br />

and is clinically effective towards treatment of respiratory tract infection, bacterial infection associated<br />

to stomach and intestinal ulcers and etc. The structure has a 14-membered lactone ring, with a<br />

desosamine sugar attached to the carbon at position 5 and a cladinose sugar branched from the<br />

carbon 3 of the lactone ring, with the addition of etheroxime chain at position 9 of the lactone ring.<br />

Similarly to other 14-membered macrolide, ROX acts by selectively bind to the Domain V at 23S<br />

ribosomal RNA of the 50S large ribosomal subunit and blocks tunnel that channels nascent peptide<br />

away from peptidyl transferase center (PTC). Blockage of exit tunnel induces premature dissociation<br />

of peptidyl-tRNAs from the ribosome, thus inhibiting bacteria’s protein elongation process. In this<br />

study, we used molecular dynamics (MD) simulation to elucidate the structure and dynamics of the<br />

complex between ROX and ribosomal 50S large subunit. MD results showed the importance of water<br />

molecules in hydrating the binding pocket and their participation in forming and mediating hydrogen<br />

bond with ROX in the binding pocket. Besides that, explicit analysis of individual direct hydrogen bond<br />

between ROX and the binding pocket also enabled us to study their occupancies and assignments to<br />

the overall drug binding mechanism in this system. Our MD simulation also showed direct interaction<br />

of ROX to Domains II, IV and V through the above analyses. Using insights obtained in this study, we<br />

believe it could contribute in understanding the detailed mechanism of action in the binding of ROX to<br />

the 50S large ribosomal subunit, and thus assist in the development of safer macrolide antibiotics.<br />

PP492<br />

Computational Studies on Bimetallic Clusters<br />

Ying-Chan Lin 1 , Li-Feng Cui 2 , Lai-Sheng Wang 3 , Dage Sundholm 1<br />

1 Department of Chemistry, <strong>University</strong> of Helsinki, Helsinki, Finland,<br />

2 Department of Physics,<br />

Washington State <strong>University</strong>, Pullman, United States, 3 Sciences Laboratory and Chemical Sciences<br />

Division, Pacific Northwest <strong>National</strong> Laboratory, Richland, United States<br />

Aromaticity is a concept introduced to account for the unusual stability of an important class of organic<br />

aromatic compounds. In recent times the concept of aromaticity has been extended to all metal<br />

molecules in order to explain the stability of some small clusters observed in laser vaporization<br />

experiments. Although much studied, the exact nature of their aromaticity has still not been completely<br />

determined.<br />

In the present work, we first report the generation of Cu 4 Na - and Au 4 Na - clusters in the gas phase and<br />

their characterizations experimentally using PES and computationally at correlated ab initio levels [1].<br />

The molecular structures of Cu 4 Na - and Au 4 Na - are found to be pyramidal of D 4h symmetry and planar<br />

cluster of D 2v symmetry, respectively. It shows the importance of both PES experiments and<br />

computational spectrum in order to identify the molecular structures of observed clusters.<br />

Moreover, we have applied the GIMIC method [2] to three different metallic systems for understanding<br />

their nature of aromaticity: Al 4 2- , Al 4 4- , and Cu 4 2- [1,3]. The magnetically induced current is invariant to<br />

transformations of the magnetic gauge origin. In practice, however, gauge invariance is hard to<br />

achieve. GIMIC is a method for calculating the components of the magnetically induced current tensor<br />

using gauge-including atomic orbitals (GIAO). The use of GIAOs represents an elegant solution to<br />

remedy the gauge-origin problem. The ring-current strengths are obtained directly via explicit<br />

numerical integration over the net current flow through one of the bond cross -section within the<br />

considered molecular ring.<br />

[1] Lin, Y.-C.; Jusélius, J.; Sundholm, D.; Gauss, J. J. Chem. Phys. 2004, 122, 214308.<br />

[2] Jusélius, J.; Sundholm, D.; Gauss, J. J. Chem. Phys. 2004, 121, 3952.<br />

[3] Lin, Y.-C.; Sundholm, D.; Jusélius, J.; Cui, Li-Feng; Li, Xi; Zhai, Hua-Jin; Wang, Lai-Sheng J. Phys. Chem. A<br />

2005, 110, 4244.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP493<br />

Are Isomers of the Vinyl Cyanide Ion Missing Links for Interstellar Pyrimidine Formation?<br />

Partha P. Bera 1 , Timthy J. Lee 1 , Henry F. Schaefer III 2<br />

1 NASA Ames Research Center, Space Science and Astrobiology, Moffett Field, CA, United States,<br />

2 <strong>University</strong> of Georgia, Center for Computational Chemistry, Athens, GA, United States<br />

The low energy isomers of the acrylonitrile ion (C 3 H 3 N + ) have been investigated using high level ab<br />

initio quantum mechanical methods, coupled cluster single and double [CCSD], and CCSD with<br />

perturbative triples [CCSD(T)], in conjunction with correlation consistent valence triple zeta and<br />

quadruple zeta cc-pVXZ [X=T,Q] basis sets, as well as simple density functional methods. An<br />

automated stochastic search procedure, Kick, has been employed to find isomers on the ground state<br />

doublet potential energy surface of the C 3 H 3 N + ion. Several new structures, along with all the<br />

previously reported minima, have been found. The global minimum H 2 CCCNH + is energetically much<br />

lower than either H 2 CC(H)CN + the acrylonitrile ion, or HCC(H)NCH + , the most likely intermediate of the<br />

reaction between HCCH + and HCN. These isomers are connected to the global minimum via several<br />

transition states and intermediates. The results indicate that not only the global minimum, but also<br />

several higher energy isomers of C 3 H 3 N + ion, could be important in interstellar pyrimidine formation.<br />

The isomeric molecules have the necessary CCNC backbone needed for the reaction with HCN to<br />

form the cyclic pyrimidine framework. The structural and rotational parameters of the isomers<br />

previously investigated, as well as newly identified by this work, have been predicted at the ccpVQZ/CCSD(T)<br />

level of theory with the hope that it will expedite their laboratory as well as<br />

astronomical identification.<br />

PP494<br />

Molecular Shape Analysis Using Ion Probes and Its Applications<br />

Manabu Sugimoto<br />

Kumamoto <strong>University</strong>, Kumamoto, Japan<br />

Molecular recognition is one of the key phenomena in biochemistry, supramolecular chemistry, gas<br />

sorption/separation sciences. The concepts of molecular surface and volume are important in<br />

discussing the host-guest chemistry. In order to evaluate these parameters quantitatively, it is required<br />

to define the outer surface of a molecule, i.e. the shape of a molecule. Here we report a computational<br />

method for quantitatively evaluating the molecular surface where the interaction energy between the<br />

target and a probe ion is used for mapping the topology. Applications to various molecules indicate<br />

that the evaluated topologies reflect chemically important characteristics of the targets and provide<br />

useful information on the molecular interaction. The method is also applied to investigate some<br />

characteristics of the hydrogen molecule, leading to the design of molecules for the hydrogen sorption.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP495<br />

The Hydrogen Bond in Transmembrane Proteins<br />

Dah-Yen Yang<br />

Institute of Atomic and Molecular Sciences, Academia Sinica, Taipei, Taiwan<br />

In our previous work, using molecular dynamics, we showed that the strength of a given hydrogen<br />

bond is much stronger in the isolated molecule than it is in water hence making the latter molecularly a<br />

much more flexible environment, as is desirable for many biological processes. Furthermore we<br />

measured very large entropic changes in the rupture of hydrogen bonds in water, whereas no such<br />

effects were seen for the isolated molecule. Hence for the H-bond in water the low energy of rupture is<br />

facilitated by the presence of water, whereas on the other hand the large entropy change is seen<br />

to reduce the rate by some two orders of magnitude. Recent MD experiments in D 2 O substantiate this<br />

model and show a large solvent isotope effect.<br />

Here we use lipids as an environment for the hydrogen bond and discover that the energy is also<br />

reduced from the isolated molecule, but not as far as in water. On the other hand there is now no<br />

entropy penalty for breaking the hydrogen bond in lipids, as we had seen for water. The result is that<br />

the energy is some two times larger, but nevertheless the rate is faster than in water. This is a very<br />

important result for understanding the reactivity and the dynamics for proteins in lipids.<br />

PP496<br />

Molecular Dynamics of Hydrogen Bonds in Protein-D 2 O<br />

Sheu-Yi Sheu<br />

Faculty of Life Sciences and Institute of Genome Sciences, <strong>National</strong> Yang-Ming <strong>University</strong>, Taipei,<br />

Taiwan<br />

We suggest that the H-bond in proteins is not atomistically related to the motion of hydrogen in its<br />

bond alone and thus does not relate directly to the strength of a hydrogen bonded to its neighbors. It<br />

rather has its origin in the collective motion of the hydrogen bond in a widespread surrounding lattice<br />

of H 2 O molecules which is being perturbed in its optimal entropic configuration by motion of the H-<br />

bond. It is for this reason that the H-bond energy drops from some 5 kcal/mol for the pure system in<br />

the absence of H 2 O, to some 1.4 kcal/mol in the presence of the H 2 O medium. This low value here is<br />

determined by MD calculations and corresponds to the generally accepted value for biological<br />

systems and the calculations are tested by changing to D 2 O as the surrounding medium. We observe<br />

in accordance with this model that under ambient conditions the H-bond kinetics is seriously<br />

depressed, hence confirming the subtle effect of the H 2 O medium as a whole interacting with the H<br />

bond. We observe that there is an entire ensemble of H-bond structures – rather than a single<br />

transition state – all of which contribute to this H-bond.<br />

We distinguish three different cases of the behavior of hydrogen bond/complexes, strongly depending<br />

on the environment.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP497<br />

Atomistic Modelling of Soot Particles and of their Interaction with Surrounding Molecules<br />

J C Rayez 1 , M T Rayez 1 , B Collignon 2 , S Picaud 2 , P N M Hoang 2<br />

1 Universite Bordeaux1, Institut des Sciences Moleculaires (ISM), CNRS UMR5255, 33405, Talence<br />

Cedex, France, 2 Université de Franche-Comté, Institut UTINAM, CNRS UMR 6213, 25030, Besançon<br />

Cedex, France<br />

Nowadays, understanding the aviation’s impact on radiative forcing, climate change, air quality and<br />

human health is a challenging task of great importance. In spite of the efforts undertaken to date by<br />

the scientific community, there is a lack of knowledge about the structure, the morphology, the<br />

composition, and the physico-chemical properties of aircraft engine soot that are released in the<br />

atmosphere.<br />

We present here a modelling study of the soot particles and of their interaction with water and<br />

Polycyclic Aromatic Hydrocarbons (PAHs). This work may lead to a better understanding, at the<br />

molecular level, of the role of the soot structural and chemical characteristics on water condensation<br />

and on the ability of these particles to act as condensation clouds nuclei. It may also lead to a better<br />

understanding of the chemical reactivity at the surface of these soot particles and thus on how soot<br />

may influence the atmospheric chemistry.<br />

Molecular dynamic simulations are used to study the adsorption of water molecules on partially<br />

oxidized graphite containing COOH and OH sites. More specifically, the competition between the OH<br />

and COOH sites with respect to water adsorption is characterized at three different temperatures (200,<br />

250 and 300K). The simulations show a strong preferential clustering of water molecules around the<br />

COOH sites irrespective of the temperature. In fact water molecules can be trapped by the OH sites<br />

only when the temperature is sufficiently low, or when the local density of OH sites is sufficiently high.<br />

These results give insights into the water adsorption mechanisms on oxidized graphite surfaces<br />

constituting black carbons or soot particles by aircraft.<br />

PP498<br />

Phosphotriesterase-Catalyzed Hydrolysis of Toxic Organophosphorus Compounds: Modelling<br />

of the Enzyme-Substrate Complex<br />

Edyta Dyguda-Kazimierowicz 1 , Jolanta Zurek 2 , Adrian Mulholland 2 , W. Andrzej Sokalski 1 , Jerzy<br />

Leszczynski 3<br />

1 Institute of Physical and Theoretical Chemistry, Wroclaw <strong>University</strong> of Technology, Wroclaw, Poland,<br />

2 Centre for Computational Chemistry, School of Chemistry, <strong>University</strong> of Bristol, Bristol, United<br />

Kingdom, 3 Computational Center for Molecular Structure and Interactions, Jackson State <strong>University</strong>,<br />

Jackson, MS, United States<br />

Phosphotriesterase (PTE) is a bacterial enzyme exhibiting the most promising capability of<br />

biodegradation of organophosphorus pesticides and related chemical warfare agents in terms of both<br />

wide substrate acceptance and high turnover number. Available experimental data regarding PTE<br />

properties have not yet allowed us to establish the molecular basis of its action. Since the suggested<br />

PTE catalytic mechanism has not been unequivocally verified by means of theoretical simulation, this<br />

research has been aimed at computational evaluation of the way PTE achieves its enormous rate<br />

enhancement. The knowledge of factors affecting substrate specificity and/or catalytic efficiency would<br />

allow for rational engineering of PTE mutants designed specifically to suit broad range of purposes.<br />

Considering the essential prerequisites necessary for the reliable modeling of an enzyme-catalyzed<br />

reaction, a well-founded model of enzyme-substrate complex is required as a starting point. Since no<br />

experimental structure of the latter is available, the consecutive steps employed in building and<br />

validation of PTE-sarin and PTE-paraoxon systems will be the subject of this contribution. Apart from<br />

classical molecular dynamics simulation, hybrid quantum mechanical and molecular mechanical<br />

(QM/MM) methodology will be employed to reveal molecular details of enzyme-substrate recognition<br />

along with its implication for PTE catalytic properties.<br />

Soot particles are also suspected to modify the atmospheric chemistry by providing surfaces for<br />

heterogeneous reactions. However, because ab initio calculations on such large systems are not<br />

realistic, we have developed mixed classical/semi-empirical calculations (hereafter called the SE-D<br />

method) to characterize the oxidation by OH of PAHs on small graphite cluster modelling soot<br />

surfaces.<br />

[1] Hamad, S.; Mejias, J.A.; Lago, S.; Picaud, S.; Hoang, P.N.M. J. Phys. Chem. 2004, B 108, 5405<br />

[2] Picaud, S.; Hoang, P.N.M.; Hamad, S. ; Mejias, J.A. J. Phys. Chem. 2004, B 108, 5410<br />

[3] Collignon, B.; Hoang, P.N.M.; Picaud, S.; Rayez, J.C. Chem. Phys. Lett. 2005, 406, 431.<br />

[4] Collignon, B.; Hoang, P.N.M., Picaud, S.; Rayez, J.C. Comp. Lett. 2005, 1, 277<br />

[5] Picaud, S.; Collignon, B.; Hoang, P.N.M.; Rayez, J.C. J. Phys. Chem.2006, B 110, 8398<br />

[6] Collignon, B.; Hoang, P.N.M.; Picaud, S.; Liotard, D.; Rayez, M.T.; Rayez, J.C. J. Mol. Struct. Theochem<br />

2006, 772, 1.<br />

[7] Picaud, S.; Collignon, B.; Hoang, P.N.M.; Rayez J.C., Phys. Chem. Chem. Phys. 2008 submitted.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP499<br />

Combining the Time Reversal Symmetry with Point Group Symmetry in Relativistic Molecular<br />

Calculations<br />

Daoling Peng, Kimihiko Hirao<br />

Department of Applied Chemistry, The <strong>University</strong> of Tokyo, Tokyo 113-8656, Japan<br />

Relativistic four components and quasi-relativistic two components quantum chemistry methods for<br />

molecular calculation usually employ double point group symmetry, like double point group symmetry<br />

adapted linear combination, to improve the computational efficiency. Time reversal symmetry reduces<br />

the computational cost by means of properly chose the symmetry adapted linear combination of<br />

electron orbitals to be time reversal pairs. A systematical method has developed for arbitrary point<br />

groups. It fully combine the time reversal symmetry with double point group symmetry in both SCF<br />

calculation (using one electron functions as basis) and post-SCF calculation (using electron pair<br />

functions as basis). Latter is achieved by fixed the phase factor of vector coupling coefficients<br />

(Clebsch-Gordan coefficients) via a simple step by step procedure.<br />

Visscher [1] has tried to introduce time reversal symmetry for point group D 2h and its subgroup. But he<br />

did not solve the whole problem of real representations which occurs in the double point group of C 2v<br />

and D 2 . Saue [2] successfully solved this problem within D 2h and its subgroups via quaternion algebra.<br />

Meyer [3] developed a systemic method for all point groups, but he only got an intermediate form. Our<br />

approach is valid for all point groups and can extend to the case of vector coupling.<br />

Standard z-axis orientation non-polyhedral point groups contain only one time reversal symmetrized<br />

irreducible Clebsch-Gordan coefficients for the direct product of fermion irreducible representations,<br />

like the properties of boson irreducible representations in Abelian point groups. This may be useful for<br />

the simplification of relativistic two electron repulsion integrals calculations in post-SCF methods like<br />

relativistic configuration interaction and couple cluster. An independent program for all point groups<br />

based on these methods is being developed. It will provide the symmetry adapted linear combination<br />

functions of a molecule with given structure, and output the time reversal symmetrized CG coefficients.<br />

PP500<br />

The FMO/EFP Energy Gradient and its Applications to Solvated Peptide Molecules<br />

Takeshi Nagata 1 , Dmitri Fedorov 1 , Mark Gordon 2 , Kazuo Kitaura 3<br />

1 AIST, Tsukuba, Japan, 2 Iowa State <strong>University</strong>, Ames, United States, 3 Kyoto <strong>University</strong>, Kyoto, Japan<br />

We have interfaced Fragment Molecular Orbital method (FMO) [1] and Effective Fragment Potential<br />

method (EFP) [2]. In this study, we developed the FMO/EFP energy gradient and then applied it to<br />

water-solvated glycine tetramers. It is confirmed computationally that zwitterionic glycine stabilizes in<br />

solution [3]. The structures and energies of the electrically neutral and the zwitterionic tetraglycine<br />

molecules immersed in EFP water molecules are investigated using the FMO/EFP geometry<br />

optimization technique. The geometries optimized at RHF/cc-pVDZ level of FMO/EFP are compared<br />

with the conventional MO/EFP optimized geometries by taking the superposition of their geometries<br />

and estimating the root mean square (RMS) deviations. We found that the number of the geometry<br />

optimization cycles was significantly reduced when FMO external electrostatic derivatives were<br />

introduced [4], and the RMS deviations of superposition were small. Using the optimized geometries,<br />

we discuss the stability of the zwitterionic and neutral structures with the systematic increment of water<br />

molecules graphically and numerically.<br />

[1]. Kitaura, K.; Ikeo, E.; Asada, T.; Nakano, T.; Uebayasi, M. Chem. Phys. Lett. 1999, 313, 701.<br />

[2]. Day, P. N.; Jensen, K. H.; Gordon, M. S.; Webb, S. P.; Stevens, W. J.; Krauss, M.; Garmer, D.; Basch, H.;<br />

Drora, C. J. Chem. Phys. 1996, 105, 1968.<br />

[3]. Aikens, C. M.; Gordon, M. S.; J. Am. Chem. Soc. 2006, 128, 12835.<br />

[4]. Nakano, T.; Kaminuma, T.; Sato, T.; Fukuzawa, K.; Akiyama, Y.; Uebayasi, M.; Kitaura, K. Chem. Phys. Lett.<br />

2002, 351, 475.<br />

[1] Visscher, L. Chem Phys Lett 1996, 253, 20.<br />

[2] Saue, T.; Jensen, H. J. A. J Chem Phys 1999, 111, 6211.<br />

[3] (a) Meyer, J. Int J Quantum Chem 1988, 33, 445. (b) Meyer, J. Int J Quantum Chem 1994, 52, 1369. (c)<br />

Meyer, J.; Sepp, W.-D.; Fricke, B.; Rosen, A. Comput Phys Commun 1996, 96, 263.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP501<br />

Dense Packing of Binary Spheres<br />

Gavin Marshall, Jonathan Kummerfeld, Toby Hudson, Peter Harrowell<br />

School of Chemistry, The <strong>University</strong> of Sydney, NSW, Australia<br />

The structure of close packed uniform spheres is an essential reference structure in numerous<br />

materials problems where high density is desirable. This work presents analogous dense packing<br />

reference structures for systems of binary spheres. They are obtained by data mining all known<br />

inorganic crystal structures, simulated annealing, and geometrical methods. Solutions are given for<br />

two subproblems: the case of homogeneous equimolar compound structures; and the case of particle<br />

filling of uniform honeycombs. The structures found include previously unknown structures, as well as<br />

known structures that were not identified as special by previous studies of packing.<br />

PP502<br />

Spectroscopic and Electric Properties of Tungsten Carbide<br />

Ivan Černušák, Miroslav Urban, Vladimír Kellö, Martina Čurkovičová<br />

Department of Physical and Theoretical Chemistry, Comenius <strong>University</strong>, Bratislava, Slovakia<br />

We have investigated the ground X 3 ∆ electronic state and the first excited 5 Σ electronic state of WC<br />

using Coupled Cluster (CC) approximation with single and double excitations corrected for the triple<br />

excitations (CCSD(T)) with the inclusion of scalar relativistic effects within the Douglas-Kroll-Hess<br />

formalism (DK-CCSD(T) in spin-free approximation). Molecular orbitals for tungsten carbide were<br />

expanded in terms of relativistic ANO-RCC atomic sets (triple-zeta, quadruple zeta and large<br />

contractions). The spectroscopic properties are calculated from the DK-CCSD(T) energies using<br />

Dunham sixth degree polynomial. In addition, fir the large contraction the dipole moments and dipole<br />

polarizabilities were evaluated within finite-field scheme. Comparison with the experimental [1] and<br />

previous theoretical [2] data is given.<br />

[1] S. M. Sickafoose, A. W. Smith, M. D. Morse, J. Chem. Phys. 2002, 116, 993.<br />

[2] K. Balasubramanian, J. Chem. Phys. 2000, 112, 7425.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP503<br />

Acyl Radical Addition to Pyridine: Multiorbital Interactions.<br />

Ruth Amos 1,2 , Carl Schiesser 2,3,4 , Jason Smith 1,2 , Brian Yates 1,2<br />

1 School of Chemistry, <strong>University</strong> of Tasmania, Hobart, Tasmania, Australia, 2 ARC Centre of<br />

Excellence for Free-Radical Chemistry and Biotechnology, Country wide, Australia, 3 School of<br />

Chemistry, The <strong>University</strong> of Melbourne, Melbourne, Victoria, Australia, 4 Bio21 Molecular Science and<br />

Biotechnology Institute , The <strong>University</strong> of Melbourne, Melbourne, Victoria, Australia<br />

O<br />

N1<br />

H<br />

N<br />

NH2<br />

N<br />

NH 2<br />

N<br />

N<br />

N<br />

OH HO<br />

O<br />

O<br />

O<br />

P<br />

OHO<br />

O<br />

O<br />

P<br />

Isoniazid, the hydrazide of isonicotinic acid (1) is an important antibiotic used to combat tuberculosis<br />

(TB). Isoniazid is a prodrug which is activated to form a reactive intermediate. Recently we reported<br />

studies supporting the hypothesis that the reactive intermediate formed is an acyl radical produced by<br />

oxidation of the hydrazide. An important step in the chemistry of isoniazid is the addition of this acyl<br />

radical to the pyridinium ion in NAD + to form the true drug (2). We present results from a<br />

computational study into the addition of acyl radicals to both pyridine and the pyridinium ion. Natural<br />

bond orbital analysis predicts simultaneous SOMO→π*, Lone pair→SOMO and Lone pair → π* c=o<br />

interactions for the addition of the acetyl radical to the nitrogen in pyridine. These multiorbital<br />

interactions are responsible for the unusual motion vector associated with the transition state in this<br />

reaction and for the lowering in energy barrier.<br />

2<br />

OH<br />

OH<br />

O<br />

O<br />

OH<br />

N<br />

H<br />

N<br />

O<br />

NH 2<br />

PP504<br />

Selenium in Pyrite - Why is the XANES Spectrum of Iron Selenide not Observed?<br />

Nicholas Lambropoulos, Ken Riley, David French<br />

CSIRO Energy Technology, Lucas Heights, NSW, Australia<br />

There is much evidence for many forms of selenium in coal and the associated mineral matter.<br />

Examination of a number of <strong>Australian</strong> coals using the beam line at the Photon Factory did not<br />

produce x-ray absorption near edge structure (XANES) spectra for which there were unique fits of<br />

spectra from reference materials. Selenium is known to be present as organically bound forms and<br />

also as inorganic species including those associated with pyrite. Thus one of the likely forms is iron<br />

selenide, most likely FeSe 2 . The spectra of reference materials including FeSe and FeSe 2 did not fit<br />

the spectra of coals containing selenium associated with the sulfide fraction.<br />

The conundrum is not readily explained without consideration of the mineralogy of the likely sulfide<br />

associated selenium. With hindsight, it is obvious that the orthorhombic FeSe 2 is unlikely to present in<br />

the cubic FeS 2 mineral phase. Rather the Se is likely present at low concentrations as FeS 2-x Se x in the<br />

pyrite.<br />

The theoretical XANES spectrum of such a species has been calculated drawing on ab initio selfconsistent<br />

real space full multiple-scattering methods [1]. Outcomes from such a theoretical approach<br />

results in determination of optimised electronic structures and XANES spectra using built in core-hole<br />

effects. A comparison with the spectra obtained from coals containing selenium associated with the<br />

sulfide fraction is made [2].<br />

[1] (a) Kresse, G.; Furthmüller, J. Phys. Rev. B. 1996, 54:11169. (b) Ankudinov, A.L.; Ravel, B.; Rehr, J.J.;<br />

Conradson, S.D. Phys. Rev. 1998, B58, 7565.<br />

[2] Riley, K. W.; French, D. H.; Lambropoulos, N. A.; Farrell, O. P.; Wood, R. A.; Huggins, F. E. Int. J. Coal<br />

Geology, 2007, 72(2), 72-80.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP505<br />

Density Matrix Renormalization Group Response Theory<br />

Jon Dorando, Johannes Hachmann, Garnet Chan<br />

Cornell Chemistry and Chemical Biology, Cornell <strong>University</strong>, Ithaca, NY, United States<br />

Response theory in Density Matrix Renormalization Group (DMRG) has been formulated using two<br />

methods, Lanczos DMRG and the correction vector method [1]. Lanczos DMRG is considered to be<br />

accurate for obtaining the static response, but it gives less accurate results for the dynamic response.<br />

On the other hand, the correction vector method is accurate in obtaining both the static and dynamic<br />

response, but this method is rather expensive [2].<br />

In our current work, we have formulated a new method to solve for the response analytically. This<br />

approach gives accurate results for both static and dynamic response, while also utilizing less<br />

computer resources. We compare the three dynamic response methods: the correction vector method<br />

without adapting the DMRG basis, the correction vector method with adapting the DMRG basis, and<br />

our newly formulated analytic response method.<br />

[1] Kuhner, T.D.; White, S.R. Phys. Rev. B 1999, 60<br />

[2] Jeckelmann, E. Phys. Rev. B 2002, 66<br />

PP506<br />

Things You Take For Granted in Closed-Shell MP2: The Subtleties of ROHF References in F12<br />

Methods<br />

Jeremiah Wilke, Henry F. Schaefer III, Wesley Allen<br />

Center of Computational Chemistry, <strong>University</strong> of Georgia, Athens, GA, United States<br />

The slow convergence of the correlation energy with respect to the size of the orbital basis has long<br />

been known. The error only decays as t -1/4 in the computational time! The slow convergence occurs<br />

since the orbital expansion is unable to describe the shape of the coulomb hole around the electron<br />

coalescence point. If two-particle basis functions that depend explicitly on the interelectronic distance<br />

supplement the orbital basis, the shape improves and convergence is greatly accelerated. In the last<br />

ten years, the F12 methods originally introduced by Kutzelnigg and Klopper have been the most<br />

successful in this regard. Efficient F12 codes for closed-shell molecules are now widely available. In<br />

contrast, F12 methods based on ROHF reference functions have only recently been investigated. In<br />

the context of perturbation theory, ROHF reference functions are not an eigenfunction of the Fock<br />

operator. The Hamiltonian partition must therefore be carefully redefined before perturbation theory<br />

can be applied. Properties such as orbital invariance, same orbitals for different spins, the Brillouin<br />

condition, and spin-restriction which are automatically true for closed shell MP2 all change in subtle<br />

ways for ROHF references. In the current work, we weigh the advantages of four different ROHF<br />

perturbation theories in the context of F12 methods - RMP, OPT1, OPT2, and ZAPT. ZAPT seems to<br />

provide the best balance of accuracy and efficiency, and we therefore derive a complete ZAPT2-F12<br />

theory, also considering future extensions to coupled-cluster methods.<br />

[1] Crawford, T.D. and Schaefer, H.F. Journal of Chemical Physics. 1996. 105. 1060-1069<br />

[2] Klopper, W., Valeev, E., Manby, F., and Ten-no, S. International Reviews in Physical Chemistry. 2006. 25.<br />

427-468.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP507<br />

How Strong Is It? The Interpretation of Force and Compliance Constants as Bond Strength<br />

Descriptors<br />

Kai Brandhorst, Jörg Grunenberg<br />

Technische Universität Braunschweig, Institut für Organische Chemie, Braunschweig, Germany<br />

Recent developments of new descriptors for covalent and noncovalent interaction strengths based on<br />

potential constants are summarized. Several publications in favour and against the use of compliance<br />

matrices (inverse force constants matrix) appeared in the literature during the last few years [1] while<br />

the mathematical basis for the understanding and therefore interpretation of compliance constants is<br />

still not well developed. This contribution therefore summarizes the theoretical foundations and points<br />

to the advantages and disadvantages of force constants versus compliance constants for the<br />

description of both, non-covalent and covalent interactions.<br />

We will use simple matrix algebra to illustrate the properties of compliance constants and their<br />

uniqueness and will present some recent applications concerning individual covalent and non-covalent<br />

bond strenghts in molecules.<br />

A clear correlation between compliance constants and interaction energies is demonstrated based on<br />

a compilation of 40 base-pairs from the JSCH2005 dataset [2].<br />

We find, that compliance constants or relaxed force constants offer not only a possibility to compare<br />

bond strengths in related molecules in a unique manner, but also allow the estimation of the overall<br />

interaction energies of weakly bound complexes in cases where several hydrogen-bonds are involved.<br />

PP508<br />

Efficient Computation of Inverse Hessians and Compliance Matrices in Redundant Internal<br />

Coordinates from Cartesian Hessians<br />

Kai Brandhorst, Jörg Grunenberg<br />

Technische Universität Braunschweig, Institut für Organische Chemie, Braunschweig, Germany<br />

We present a new algorithm for the computation of either the hessian or the compliance matrix [1] in<br />

redundant internal coordinates. We will show, that the complete redundant compliance matrix of a<br />

coordinate system of all N(N-1)/2 possible interatomic distances in a molecule can be computed with a<br />

complexity of only O(N 4 ), while the computation of the hessian itself in the same set of coordinates is<br />

of O(N 5 ) complexity, which nevertheless is one order of magnitude lower than the conventional<br />

algorithm that scales like O(N 6 ) [2].<br />

The new algorithm has been implemented in our standalone computer code Compliance 2.0 which is<br />

able to compute the complete and redundant compliance matrix of all interatomic distances, valence<br />

angle bendings and dihedral torsions from a Cartesian hessian. Using our new algorithm avoids the<br />

time consuming and error prone definition of a non redundant set of coordinates and allows the<br />

straightforward computation of the complete compliance matrix even for complex molecular<br />

topologies.<br />

We will make available the code at no cost to anyone who is interested in using it.<br />

[1] (a) Grunenberg, J. J. Am. Chem. Soc. 2004, 126, 16310-16311. (b) Brandhorst, K.; Grunenberg, J.<br />

ChemPhysChem. 2007, 8, 1151-1156. (c) Baker, J. J. Chem. Phys. 2006, 125, 014103. (d) Baker, J.; Pulay, P.<br />

J. Am. Chem. Soc. 2006, 128, 11324-11325.<br />

[2] Jurecka, P.; Sponer, J.; Cerny, J.; Hobza, P. Phys. Chem. Chem. Phys., 2006, 8, 1985-1993.<br />

[1] (a) Decius, J. C. J. Chem. Phys. 1963, 38, 241-248. (b) Jones, L. H.; Swanson, B. I. Acc. Chem. Res. 1976, 9,<br />

128-134. (c) Brandhorst, K.; Grunenberg, J. ChemPhysChem, 2007, 8, 1151-1156.<br />

[2] (a) Pulay, P.; Fogarasi, G. J. Chem. Phys. 1992, 96, 2856-2860. (b) Peng, C.; Ayala, P. Y.; Schlegel, H. B.;<br />

Frisch, M. J. J. Comput. Chem. 1996, 17, 49-56.


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP509<br />

Quantized-Liquid Density Functional Theory of Hydrogen Adsorption in Nano-Porous<br />

Substrates.<br />

Serguei Patchkovskii 1 , Thomas Heine 2<br />

1 NRC Canada, Ottawa, Ontario, Canada, 2 Jacobs <strong>University</strong>, Bremen, Germany<br />

Compact storage of molecular hydrogen is the major obstacle to the development of practical<br />

hydrogen-powered vehicles. Physisorption of hydrogen in nano-porous materials (e.g. clathrates,<br />

metal-organic frameworks, and exfoliated graphite) is among the most practically successful<br />

approaches, but the fundamental limits on the storage capacities of these systems are not fully<br />

understood.<br />

Building on a recently-introduced [1] quantized ideal-gas treatment, we develop a simple method for<br />

estimation of the adsorption free energy of lightweight molecules in nanoporous structures: the<br />

quantized-liquid density functional theory (QLDFT). The approach allows direct calculation of the free<br />

adsorption energy in nano-scale systems, taking into account quantization of the nuclear motion of the<br />

guest molecules. The main ingredients of the technique are the kinetic energy of the quantized ideal<br />

gas, the mean-field interaction potential, and the excess free energy functional. The free-energy<br />

functional is constructed to reproduce the experimental equation of state for the uniform hydrogen<br />

fluid.<br />

PP510<br />

XUV Probing of a Recollision Electron Wavepacket: An Attosecond Electron Microscope.<br />

Olga Smirnova, Serguei Patchkovskii, Michael Spanner<br />

NRC Canada, Ottawa, Ontario, Canada<br />

Ionization of atoms and molecules, followed by recollision with the parent ion, is fundamental to many<br />

strong field phenomena. We propose a technique for direct imaging of the recolliding electron wave<br />

packet with sub-Å spatial and attosecond temporal resolution [1]. The approach relies on adsorption of<br />

an XUV photon at the time of recollision, with subsequent detection of the high-energy photoelectrons.<br />

Delicate numerical simulations show that the proposal is experimentally realizable, despite the very<br />

small cross-section of the desired process.<br />

[1] Smirnova, O., Patchkovskii, S., Spanner, M. Direct XUV Probing of Attosecond Electron Recollision, Phys Rev<br />

Lett 2007, 98, 123001<br />

We consider three increasingly sophisticated forms of the free-energy functionals. In the localinteraction<br />

expression (LIE) approximation, the exchange-correlation is in the local-densityapproximation<br />

form, while the mean-field interaction potential is set to zero. The LIE functional<br />

successfully reproduces adsorption thermodynamics in open pores, but fails to describe the<br />

microscopic fluid structure. Due to the large self-interaction error, it also fails for isolated adsorption<br />

sites. Incorporation of the mean-field interaction potential, together with a scaled-density<br />

approximation (SDA) non-local free energy functional, allows description of the hydrogen fluid<br />

structure, as well as phase transitions. However, the SDA-v12 form of QLDFT still fails for isolated<br />

adsorption sites. This remaining defect is corrected in the self-interaction corrected neighbour-density<br />

approximation form of QLDFT.<br />

[1] Patchkovskii, S., Heine, T. Evaluation of the adsorption free energy of light guest molecules in nanoporous<br />

host structures, Phys. Chem. Chem. Phys. 2007, 9, 2697


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP511<br />

QSPR Studies on the Prediction of Physicochemical Properties of High Energetic Materials:<br />

Melting Point<br />

Chan Kyung Kim 1 , Soo Gyeong Cho 2 , In Suk Han 1 , Junxian Chen 1 , Hyok Yoon 1 , Hai Whang Lee 1 ,<br />

Bon-Su Lee 1<br />

1 Inha <strong>University</strong>, Department of Chemistry, Incheon, 402-751, Korea, Republic of, 2 Agency for<br />

Defense Development, P. O. Box 35-42, Yuseong, Deajeon, 305-600, Korea, Republic of<br />

Prediction of physicochemical properties of high energetic materials (HEM) is an important study in<br />

developing new energetic materials. To predict melting points of HEMs with at least one –NO 2 group,<br />

several methodologies based on group additivity, molecular topology and atom/specific functional<br />

group additivity are examined. As a continuing work on the QSPR works on HEMs, all the HEMs are<br />

fully optimized at B3LYP/6-31G(d) level of theory and confirmed to be genuine minima using Gaussian<br />

03. From the optimized structures, van der Waals surface electrostatic potentials are calculated using<br />

the method developed earlier [1]. QSPR studies are performed to find a good linear correlation using<br />

the descriptors obtained from GIPF variables [2]. Melting points predicted from this work are compared<br />

with the experimental values as well as other methods.<br />

PP512<br />

Theoretical Studies on Nucleophilic Substitution Reactions of Acetyl and Thioacetyl Halides<br />

with NH 3 in the Gas Phase and in Aqueous Solution<br />

In Suk Han, Chang Kon Kim, Hyok Yoon, Hai Whang Lee, Chan Kyung Kim<br />

Inha <strong>University</strong>, Department of Chemistry, Incheon, 402-751, Korea, Republic of<br />

The reactions of acetyl halides, CH 3 C(=O)X, and corresponding sulfur analogues, thioacetyl,<br />

CH 3 C(=S)X, where X = F and Cl with NH 3 have been studied theoretically at MP2 level of theory in the<br />

gas phase and in aqueous solution. All the reactions occurred via the T ± -typed species and the<br />

reactions through neutral intermediates might be ruled out in both phases except for the gas-phase<br />

reaction of acetyl fluoride. In the reaction of acetyl chloride, the tetrahedral species was a transition<br />

structure (TS) but a stable intermediate in the reactions of thioacetyl halides. This could be caused by<br />

different π-bond strengths of C=O and C=S. For acetyl fluoride, the tetrahedral species was neither a<br />

saddle point species nor an energy minimum in the gas phase, but existed as a stable intermediate in<br />

aqueous solution due to solvation effects. Moreover, in the reactions of thioacetyl halides the ratelimiting<br />

step was changed from the first step in the gas phase to the second step in aqueous solution,<br />

since the T ± -typed intermediates became much more stabilized in aqueous solution. However,<br />

lowering in the activation barriers ( ∆ E ≠<br />

ZPE<br />

) in aqueous solution compared to those in the gas phase<br />

was not caused by the solvent effects but smaller deformation energies on going from reactants from<br />

TS.<br />

H 3 C<br />

Y<br />

C<br />

(Y = O or S and X = F or Cl)<br />

Y δ-<br />

H 3 C C<br />

δ+<br />

Y-<br />

X + NH 3<br />

H 3 C<br />

C<br />

+<br />

NH 3<br />

(T + )<br />

δ-<br />

X<br />

X<br />

=<br />

Products<br />

Products<br />

(1a)<br />

(1b)<br />

NH 3<br />

Products (1c)<br />

YH<br />

H 3 C<br />

C<br />

NH 2<br />

[1] Kim, C. K.; Lee, K. A.; Hyun, K. H.; Park, H. J.; Kwack, I. Y.; Kim, C. K.; Lee, H. W.; Lee, B.-S. J. Comput.<br />

Chem. 2004, 25, 2073.<br />

[2] Politzer, P.; Murray, J. S. Quantitative Treatments of Solute/Solute Interac-tions; Elsevier: Amsterdam, 1994;<br />

p. 243.<br />

X<br />

(T)


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP513<br />

Effects of Basis Set Superposition Error on Optimized Geometries and Energies of Organo-<br />

Alkali Metal Cation and its Halide Complexes<br />

Chang Kon Kim 1 , In Suk Han 1 , Hyok Yoon 1 , Jongok Won 2 , Chan Kyung Kim 1<br />

1 Inha <strong>University</strong>, Department of Chemistry, Incheon, 402-751, Korea, Republic of, 2 Sejong <strong>University</strong>,<br />

Department of Applied Chemistry, Seoul, 134-747, Korea, Republic of<br />

Theoretical studies were performed to study the binding of the alkali metal cation, X + (X = Li, Na, K), to<br />

poly(ethylene oxide) (PEO, I), poly(ethylene amine, II) (PEA) and poly(ethylene N-methylamine)<br />

(PEMA, III) and dissociation processes in the same polymer electrolytes with alkali-metal iodides via<br />

the Hartree-Fock (HF) and B3LYP methods using the 6-31G(d) and 6-311+G(d,p) basis sets. Two<br />

types of complex were considered in this study: singly-coordinated system (SCS) and doublycoordinated<br />

system (DCS). Complexation and dissociation energies were calculated both without and<br />

with basis set superposition error (BSSE). Three possible counterpoise (CP) approaches were<br />

examined in detail. In the case of the function CP (fCP) correction, the complexation energies<br />

exhibited an unusual trend due to deformation of the subunits. This problem was solved by including<br />

geometry relaxation in the CP-corrected (GCP) interaction energy. The effects of structures and<br />

vibrational frequencies were small when the complexes were re-optimized on the CP-corrected,<br />

potential energy surface (PES).<br />

H 3 C<br />

H 2<br />

X C CH 3<br />

+ M +<br />

C X<br />

H 2<br />

H 3 C<br />

Singly-Coordinated System(SCS)<br />

H 2<br />

X C CH 3<br />

C X<br />

H 2<br />

M +<br />

PP514<br />

Theoretical Study of the Acid-Base Properties of Piroxicam and Isoxicam<br />

Marco Antonio Franco-Pérez, Rosario Moya-Hernández, Rodolfo Gómez-Balderas<br />

FESC-Cuautitlán, UNAM, Departamento de Química, México<br />

Oxicams are widely used for treatment of chronic inflammatory conditions. In this contribution, we<br />

report results of a theoretical study of Piroxicam (P) and Isoxicam (I). It is well known that acid-base<br />

properties of drugs affect their bio-availability and pharmacological action. In order to consider the<br />

different acid-base species we included the HP + , P, P – , I and I – species. Geometry optimizations and<br />

conformational scans have been done at the B3LYP/6-31G (d, p) level of theory. The gas phase<br />

minima were then reoptimized using the PCM scheme to include the effect of the solvent (H 2 O) at the<br />

B3LYP/6-311++G(d,p) level. Estimation of the H-NMR chemical shifts, including the main contributions<br />

pondered by energy of several conformers agrees reasonably well with the experimental observations.<br />

Additionally, we use Fukui functions, frontier orbital´s and ionization potentials to characterize these<br />

oxicams.<br />

O<br />

S<br />

OH<br />

O<br />

N<br />

CH 3<br />

O<br />

H<br />

N<br />

N<br />

O<br />

S<br />

OH<br />

O<br />

N<br />

CH 3<br />

O<br />

H<br />

N<br />

N<br />

O<br />

H 3 C<br />

X<br />

H 2<br />

C CH 3 + M +<br />

C X<br />

H 2<br />

M+<br />

H 3 C X X<br />

H 2 C CH 2<br />

CH 3<br />

Piroxicam Isoxicam<br />

CH 3<br />

Doubly-Coordinated System(DCS)<br />

H 3 C<br />

H 2<br />

X C CH 3<br />

C X<br />

H 2<br />

M +<br />

Singly-Coordinated System(SCS)<br />

H 3 C<br />

H 2 C<br />

X<br />

+<br />

M<br />

I - I -<br />

CH 2<br />

X<br />

CH 3<br />

Doubly-Coordinated System(DCS)<br />

where X = O (I), NH (II), and N(CH 3 ) (III) and M = Li, Na, and K


<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

<strong>Poster</strong> <strong>Session</strong> 3 – Tuesday September 16 – 5.30-7.30pm<br />

PP515<br />

Accurate Hartree-Fock Calculation Using Perturbation Theory<br />

Jia Deng, Peter Gill<br />

Research School of Chemistry, <strong>Australian</strong> <strong>National</strong> <strong>University</strong>, Australia<br />

The slow convergence of Quantum chemical methods with respect to the size of the one-electron<br />

Gaussian basis set is a major obstacle in computational chemistry, therefore, exploring the basis set<br />

limit at a relatively inexpensive cost is very desirable. Recently, we have discovered a mathematically<br />

rigorous approach, based on perturbation theory, to explore the basis set limit at Hartree-Fock level of<br />

theory. In this poster, we outline the Hartree-Fock Pertubation Theory (HFPT), and show that our<br />

method, at first order, exhibits quadratic convergence. We apply HFPT to several small atomic and<br />

molecular systems. Our numerical investigation shows that HFPT yields very accurate Hartree-Fock<br />

energies, and the convergence behaviour confirms our theoretical result.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!