• Keine Ergebnisse gefunden

The Conformation of Pentanoates in the Solid and in the Gas Phase Carina Merkens

N/A
N/A
Protected

Academic year: 2022

Aktie "The Conformation of Pentanoates in the Solid and in the Gas Phase Carina Merkens"

Copied!
10
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

The Conformation of Pentanoates in the Solid and in the Gas Phase

Carina Merkensa, Tom Stadtmüllera, Ulli Englerta, Halima Mouhibb, and Wolfgang Stahlb

aRWTH Aachen University, Institute of Inorganic Chemistry, Landoltweg 1, 52056 Aachen, Germany

bRWTH Aachen University, Institute of Physical Chemistry, Landoltweg 2, 52056 Aachen, Germany

Reprint requests to U. E.; E-mail:ullrich.englert@ac.rwth-aachen.de Z. Naturforsch.69a, 303 – 312 (2014) / DOI: 10.5560/ZNA.2014-0013

Received September 24, 2013 / revised January 30, 2014 / published online July 15, 2014 Dedicated to Prof. J. Fleischhauer on the occasion of his 75th birthday

Suitable derivatives of the four isomeric pentanoates have been structurally characterized in the solid and the gas phase. For the latter, the volatile ethyl esters of valeric, isovaleric, methylbu- tyric, and pivalic acid were investigated by a combination of molecular beam Fourier transform mi- crowave (MB-FTMW) spectroscopy and theoretical calculations. Crystalline salts rather than esters were formed by reaction between the carboxylic acids andtrans-1,2-diaminocyclohexane. For both gaseous and crystalline methylbutyrates, an essentially perpendicular arrangement of carboxylate and methyl group was observed; earlier structure determinations documented in the data base agree with this result. Two competing conformers of favourable energy were relevant for the corresponding iso- valerates: They were associated with torsion angles around 20and 50between the carboxylate and the alkyl chain. Good agreements in conformation have also been achieved for our experimentally observed unbranched valerate derivatives and fully branched pivalates in solid and gas phase. De- spite the apparent simplicity of the pentanoates, the identification of their lowest energy conformers represents a challenge for different methods and levels of theory.

Key words:Conformational Studies; X-Ray Diffraction; Microwave Spectroscopy; Quantum Chemical Calculations.

1. Introduction

The conformation of non-rigid molecules may de- pend on their state of aggregation. A comparative anal- ysis of soft degrees of freedom in different phases has to rely on a combination of methods. Single crys- tal X-ray diffraction doubtlessly represents the most common way to determine molecular geometries in the solid, whereas experimental data from microwave spectroscopy, combined with theoretical results, can provide information about the gas phase. The differ- ent requirements for diffraction and microwave spec- troscopy may be challenging: Volatility and elevated vapour pressure are beneficial for the gas phase study but can impede the formation of stable crystalline solids by conventional techniques. In situ crystal- lization [1,2] can allow diffraction studies even on low melting compounds. [3,4] Comparative confor- mational analyses are particularly relevant in the field

of fragrance chemistry: Odorants must be sufficiently volatile to allow for discernible concentrations in the gas phase but also have to interact with their receptors in condensed phase. Investigations concerning possible mechanisms of olfaction should take conformational space rather than individual low-energy conformations into account [5]. In the case of odorants, we have used co-crystallization of these volatile small molecules with cholic acid as an alternative approach [6,7]. In this contribution, we compare the conformation of the alkyl chains in pentanoates, i. e. derivatives of the satu- rated isomeric carboxylic acids with five carbon atoms in a linear or branched chain. These small pentanoates all possess sweet or fruity odors and are thus used in the food and flavour industry. Despite their wide appli- cation, no systematic comparison of the different exist- ing conformers has been performed on such systems.

Figure1compiles the compounds under investigation.

Results for the esters1a–4arely on a combination of

© 2014 Verlag der Zeitschrift für Naturforschung, Tübingen·http://znaturforsch.com

(2)

O O

H

O O

H

O O

H

O O

H

Ethyl esters Salts of

rac-1,2-diaminocyclohexane O

R O

H2N

H2N

1a

2b 1b

4a 3a 2a

4b 3b

Pivalic acid Isovaleric acid 2-Methyl butyric acid

Valeric acid

Fig. 1 (colour online). Summary of the compounds under investigation: Microwave spectroscopy investigations were performed using the corresponding ethyl esters of the pen- tanoates 1a4a. Salt formation of the acids withrac-1,2- diaminocylohexane resulted in crystalline solids1b4bfor X-ray diffraction studies in the solid state.

microwave spectroscopy and quantum chemical cal- culations: Accurate rotational constants were obtained from molecular beam Fourier transform microwave (MB-FTMW) spectrocopy, and the assignment of the observed species was achieved via comparison with ro- tational constants for theoretically predicted minimum structures as potential conformers.

Conformers in the sense of this contribution in the first line address the torsion angle δ, O=C-C-C, be- tween a carboxylate oxygen and an alkyl substituent at theα-carbon atom. For the compounds in Figure1, we can not give a unique definition of the torsion an- gle; the following convention will be used and has been highlighted in red:1and4feature at least one methyl group at theα-carbon; the only methyl group in1and the closest methyl group in4subtend the above men- tioned torsion angle. For2and3, the torsion angle is defined by the only carbon substituent atα-C.

Rac-trans-1,2-diaminiumcyclohexane has been chosen as the counter cation for the salts 1b–4b.

Size matching with the carboxylates and efficient

Table 1. Microwave spectroscopic constants of ethyl 2- methyl butyrate using the xiam code and Watson’sAreduc- tion andIrrepresentation [19,33].

Constant Unit Value

Aa GHz 2.60453135(62)

Ba GHz 0.88540909(11)

Ca GHz 0.79503438(11)

Jb kHz 0.32211(52)

JKb kHz 9.3944(83)

Kb kHz 4.186(86)

δJb kHz 0.03130(20)

δKb kHz 6.377(11)

σc kHz 4.3

κd 0.900

Ne 74

aRotational contants;bCentrifugal distortion contants;cStandard de- viation of the fit;dRay’s asymmetry parameter [18];eNumber of lines included in the fit.

hydrogen bonding were expected, and the cyclohexane backbone reduces the chance to encounter disorder in the cations: Neither in this nor in any earlier work with this constituent, diaminocyclohexane or its derivatives gave rise to disorder [8–11]. As anticipated, 1b–4b are stable crystalline solids, all of them with two carboxylate monoanions per diaminium dication.

1b–4bhave been structurally characterized by X-ray diffraction, and when possible, earlier results from diffraction experiments documented in the Cambridge Structural Database (CSD) [12] have been included in our considerations. Strictly spoken, this approach does not compare molecular conformations for the same molecules in different states of aggregation; rather, we make use of suitable derivatives with the same alkyl residues to map conformational space in a structure correlation approach [13–15].

2. Results and Discussion

We will first focus on the derivatives of 2- methylbutyric acid. For the ester 1a, the torsion an- gle O=C-C-C between the carbonyl CO group and the methyl substituent at the α-carbon atom may adopt values between+and −180; opposite signs can be associated with both enantiomers of this chiral com- pound. In the following discussion, we will refer to the absolute value for this angle. The ethyl ester 1a is a highly volatile compound with a boiling point of

(3)

Table 2. Quantum chemical data obtained for the observed conformer (Conformer 1 at the MP2/6-311++G(d,p) level) of ethyl 2-methyl butyrate at different levels of theory.

Aa dev.b B dev. C dev. µc[Debye] δd[] νe[cm−1]

B3LYP/6-311++G(d,p) 2.537 2.59 0.875 1.22 0.789 0.74 1.84 −79.47 22.27

HF/6-311++G(d,p) 2.680 −2.85 0.875 1.16 0.787 0.98 1.99 −89.72 25.93

MP2/6-311++G(d,p) 2.480 4.77 0.913 −3.16 0.813 −2.28 1.90 −69.96 11.50

MP3/6-311++G(d,p) 2.598 0.23 0.885 0.03 0.795 −0.02 1.89 −83.02 f

MP4/6-31++G(d,p) 2.430 6.71 0.920 −3.91 0.812 −2.16 1.98 −77.85

CCSD/6-311++G(d,p) 2.551 2.06 0.889 −0.41 0.799 −0.50 1.90 −79.22

aRotational constantsA,B, andCin GHz;bRelative deviation in percent with respect to the experimental value;cDipole moment;dDihedral angle as shown in Figure 2;eVibrational frequency of the alkyl group aroundδ;fHarmonic frequency calculations were not performed at theoretical levels higher than the MP3/6-31+G(d,p) level.

408 K [16] and a vapour pressure of approximately 11 hPa [17]. To the best of our knowledge, its confor- mation in the gas phase has not yet been reported. We have assigned the most abundant conformer of ethyl 2- methyl butyrate using MB-FTMW spectroscopy. Us- ing this method, due to the Joule–Thompson effect, rotational temperatures of approximately 1 – 3 K are achieved in the molecular beam. Therefore, we can only observe the most abundant conformer. The exper- imental results can thus be used to validate theoretical results, especially when several alternative conformers with small energy differences are obtained by means of quantum chemical calculations. The microwave spec- troscopic data are shown in Table1. For the inves- tigated ethyl 2-methyl ester, the most abundant con- former in the molecular beam is a near prolate top with an asymmetry parameter of κ =−0.900 [18]. Alto-

Table 3. Crystal data and refinement results forrac-1,2-diaminiumcyclohexane 2-methylbutyrate1band isovalerate2b.

Compound 1b 2b(α) 2b(β)

Crystal System Monoclinic Triclinic Triclinic

Space group (no.) P21/n (14) P¯1 (2) P¯1 (2)

a(Å) 11.8202(15) 8.581(11) 8.5413(10)

b(Å) 16.378(2) 10.222(12) 10.3138(13)

c(Å) 21.009(3) 11.806(14) 12.7046(15)

α() 99.76(2) 67.258(2)

β() 105.167(3) 94.07(2) 78.464(2)

γ() 112.39(2) 67.820(2)

V3) 3925.4(8) 933(2) 953.9(2)

Z 4 2 2

Total/unique reflections 47415/8137 7656/3715 11522/3909

Variables refined 527 (72 restraints) 227 (6 restraints) 257(6 restraints)

Rint 0.0580 0.0673 0.0372

wR2(all reflections) 0.2265 0.1906 0.1068

R1(all/obs) 0.1114/0.0803 0.1083/0.0668 0.0548/0.0440

GOF onF2 1.045 0.980 1.026

Diff. peak/hole ( e Å−3) 0.764/−0.640 0.322/−0.345 0.255/−0.205

CCDC No. 934807 934808 943357

gether, 74 rotational transitions were identified and in- cluded in the fit using the program xiam [19]. The stan- dard deviation of the fit is 4.3 kHz, which is slightly higher than our experimental accuracy of 2 kHz. All strong lines are included in the fit, while some weaker unassigned lines probably belong to other yet unas- signed conformers or to a large number of isotopo- logues. The most abundant conformer in the gas phase is associated with a torsion angle O=C-C-C of 83(see Fig.2).

To determine the geometry of the observed con- former, the experimental results were first compared to a conformational analysis performed at the MP2/6- 311++G(d,p) level of theory. This extended conforma- tional analysis allowed us to identify 23 different en- ergy minima within a range of 10 kJ/mol. The three lowest energy conformers at the MP2/6-311++G(d,p)

(4)

Table 4. Crystal data and refinement results forrac-1,2-diaminiumcyclohexane valerate3band pivalate4b.

Compound 3b 4b

Crystal System Triclinic Triclinic

Space group (no.) P¯1 (2) P¯1 (2)

a(Å) 8.260(2) 9.4700(14)

b(Å) 10.287(3) 9.4957(15)

c(Å) 12.802(4) 11.5278(17)

α() 68.070(5) 99.245(4)

β() 74.981(5) 97.943(4)

γ() 69.499(5) 113.642(4)

V3) 934.8(5) 913.5(2)

Z 2 2

Total/unique reflections 11185/3764 11287/3816 Variables refined 254 (10 restraints) 260 (12 restraints)

Rint 0.0449 0.0660

wR2(all reflections) 0.1130 0.1356

R1(all/obs) 0.0657/0.0461 0.0954/0.0587

GOF onF2 1.044 1.051

Diff. peak/hole ( e Å−3) 0.195/−0.186 0.284/−0.218

CCDC No. 934809 934810

level are found within 1 kJ/mol and possess similar sets of rotational constants and dipole moment com- ponents. This makes the assignment of the observed conformer in the molecular beam ambiguous, but is not a problem for the comparison with the solid phase, as the three lowest energy conformers differ only on the side of the ethyl group. The ethyl group is ei- ther in the same plane as the carboxylic group (Con- former 1, see Table VI (supporting information (SI), can be requested from the authors) or it is bent out of the plane by ±80 (Conformer 2 and 3, see Ta- ble VII and VIII (SI)). However, this is of minor rele- vance here, since we extrapolate structures in the solid phase from derivatives of the carboxylic acid. In the cases of ethyl isovalerate, ethyl valerate, and ethyl pi-

Fig. 2 (colour online). Conformation of1ain the gas phase as obtained at the MP3/6-311++G(d,p) level and observed in the molecular beam; the torsion angle between the carboxylate COO group and the methyl substituent at theα-C has been traced in red.

valate, the ethyl group was always found to be in the COO plane [20–22]. If we assume that this is also the case for ethyl 2-methyl butyrate, it appears rea- sonable that we observe the lowest energy conformer (Conformer 1) at the MP2/6-311++G(d,p) level. It was therefore optimized again using different methods and basis sets. The results of these quantum chemical cal- culations were then validated using the experimental constants (see Tab.2). Figure2displays the best agree- ment obtained at the MP3/6-311++G(d,p) level. The coordinates of this observed conformer are compiled in Table V in the SI.

The crystal structure of1bis surprisingly complex:

The compound crystallizes with two dications and four carboxylate anions in the asymmetric unit. In contrast to the ester1a, these carboxylates feature symmetric COOgroups, and therefore the torsional angleδ can only adopt absolute values between 0 and 90. The space groupP21/nimplies the equimolar occurrence ofRandSconfigured carboxylates; space group sym- metry apart, substitional disorder of both enantiomers occurs. We here will refer to the majority conformer at each site. The four dominant conformers of each inde- pendent carboxylate anion adopt two essentially differ- ent conformations: In three of them, the alkyl chain is very similar to that of the ester1awhereas the fourth methylbutyrate anion is significantly different. 54 crys- tal structures [12] of methylbutyrate esters have been documented in the CSD. The histogram in Figure3 shows the distribution of the torsion angle (as defined

(5)

Fig. 3 (colour online). Histogram of the torsion angle be- tween the carboxylate COO group and the methyl substituent at the α-C in methylbutyrates related to 1. Coloured ar- rows mark experimental results: a) green: values for three of the four symmetrically independent carboxylate anions;

b) green: fourth symmetrically independent carboxylate an- ion; c) red: minimum according to the microwave validated structure shown in Figure2; d) blue: experimental result from the only diffraction experiment documented in the CSD data base (CSD refcode BIJYEY [23]).

in Fig.2) in their alkyl chain and proofs the preference for an almost perpendicular orientation of the methyl group with respect to the carboxylate plane.

Our experimental results as well as the only crys- tal structure determination of a solid which contains methylbutyric acid as a co-crystallized molecule (CSD refcode BIJYEY [23]) are in agreement with the trend in the histogram: Only one of the solid-state conform- ers in 1b features an unusual torsion angle of 37 (marked as ‘b’ in Fig.3). Remarkebly,1aand the three conformationally independent residues of 1bmarked as ‘a’ in Figure3 also agree very well with respect to the orientation of the ethyl group atα-C as shown in Figure2(SI).

As outlined in the introduction, the interpretation of microwave spectroscopic results for such medium- sized molecules necessarily has to rely on a compar- ison with conformers predicted from theory. Inter- estingly, geometry optimizations using different or-

der of perturbation theory disagree with respect to the geometry of the energetically most favourable gas phase conformer of ethyl isovalerate, 2a, despite its apparent simplicity. In a previous microwave spectro- scopic study of ethyl isovalerate, a large amplitude mo- tion was observed around the O=C-C-C bond of the molecule [20]. A similar observation was made in the case of its structural isomer ethyl valerate. However, at this point, the reason for this low torsional potential around the O=C-C-C bond is unknown and rather spec- ulative. To verify this assumption, ideally a conforma- tional study using higher orders of perturbation theory should be performed, but the computational costs for this are too high and surpass the scope of this study. We suggest that the theoretical calculation of this large am- plitude motion depends very much on the method and the basis set used. The good agreement at the MP3/6- 311++G(d,p) level (see Tab.2) is rather coincidental and might be due to error compensations. Altogether, it can be said that we could not determine a specific method and basis set that always yields reliable geome- tries for these ethyl alkyl esters.

In the solid state, 2b is dimorphic and occurs in two different but closely related modifications 2b(α) and2b(β); their relationship and their similarity to3b will be explained below. At the molecular level, each modification contains two symmetrically independent isovalerate anions. In2b(β), one of these residues is disordered and only the majority conformer is taken into account for the following discussion. One iso-

Fig. 4 (colour online). Potential conformations for2. Left:

Conformation adopted by one isovalerate anion in each polymorph of2band conformation of2aoptimized at the B3LYP/6-311++G(d,p) level and observed in the gas phase.

Right: Conformation adopted by the second independent iso- valerate in both crystal forms, conformation of isovaleric acid in the crystal (CSD refcode INUJUW [24]) and con- formation of minimum energy as predicted at the MP2/6- 311++G(d,p) level of theory.

(6)

Fig. 5 (colour online). Displacement ellipsoid plot of the non- disordered valerate anion in crystalline3b. Ellipsoids have been scaled to 50% probability, and the torsion angle be- tween the carboxylate COO group and the alkyl substituent at theα-C has been traced in red.

valerate in each crystal form adopts the same confor- mation as its ethyl ester in the gas phase, obtained at the B3LYP/6-311++G(d,p) level. The conformational similarity between related gas phase and solid state conformers extends to the relative orientation of the isopropyl group as well (Fig. 5 (SI)). In contrast, the second isovalerate residue in each solid corresponds to the geometry of co-crystallized isovaleric acid [24] and to the minimum energy conformation according to cal- culations at the MP2/6-311++G(d,p) level. The occur- rence of both alternative conformations is summarized in Figure4.

Conformations for the unbranched pentanoic acid, its salts and esters, cover a wide range of torsion angles between the carboxylate group and the alkyl chain on theα-carbon. The crystal structure of our salt3bfea-

Fig. 6 (colour online). Pro- jections of the crystal struc- tures of 2b(α),2b(β), and 3b along a. The dominant motif for hydrogen bonding within each double layer has been highlighted in red (dashed ovals), subsequent bilayers are stacked along bvia van-der-Waals (vdW) contacts.

tures two symmetrically independent valerate anions, an ordered and a disordered moiety. The well-ordered carboxylate (Fig.5) is associated with a small torsion angle of ca. 19.

Similar almost eclipsed conformations with tor- sion angles ranging from 2 to 27 have been re- ported for the three crystal structures which contain non-disordered molecules of valeric acid [23,25,26].

Coplanarity is also observed in both experimentally observed gas phase conformations [21]. The disor- dered valerate in 3b corresponds to a majority con- former with a torsion angle of 40and a minority con- former with 59; an earlier report of an inclusion com- pound with disordered valeric acid [27] gives a value of 68.

The solids 2b(α), 2b(β), and 3b are closely re- lated: They do not only show similar lattice parame- ters (Tabs.3 and4) but also agree with respect to the most relevant hydrogen bonds. All three crystal struc- tures are based on hydrogen-bonded double layers of the same composition and very similar geometry in the abplane; the common motif has been highlighted in their projections along the shortest lattice parametera (Fig.6).

The resulting double layers are stacked in slightly different ways alongc. It is important to note that these stacking differences do not affect strong intermolecular interactions such as hydrogen bonds but only van-der- Waals contacts between the alkyl periphery. The spe- cial type of polymorphism encountered for2b(α)and 2b(β)is called polytypism; a compound is polytypic if it occurs in several different structural modifications, each of which may be regarded as built up by stacking

(7)

Fig. 7 (colour online). Histogram of the smallest torsion an- gle between the carboxylate COO group and a methyl sub- stituent at theα-C in pivalate. Arrows mark the experimental results: a) well-ordered and majority conformer in the dis- ordered symmetrically independent carboxylate anions ac- cording to X-ray diffraction of4b; this conformation is also adopted in the gas phase, b) minority conformer in the disor- dered carboxylate anion.

layers of (nearly) identical structure and composition, and if the modifications differ only in their stacking sequence [28].2band3binvolve two different com- pounds and cannot be called polytyps in a strict sense – all three structures form, however, by stacking very similar layers.

An elevated number of derivatives of pivalic acid, the most branched isomeric pentanoate, has already been reported in the CSD data base. The smallest tor- sion angleδ between the carboxylate COO group and one of the three equivalent methyl substituents at the α-C in pivalate covers absolute values between 0 and ca. 30. The histogram in Figure7documents the strong preference for the very small values, i. e. an eclipsed conformation. Both experimentally observed conformers of ethyl pivalate4ain the gas phase agree with respect to this torsion angle between carboxylate group and methyl substituent [22]. The preference for an eclipsed conformation is also reflected in both sym- metrically independent pivalate anions in crystalline 4b: The carboxylate and a methyl group are coplanar both in the well-ordered and the majority conformer of the disordered residue.

3. Conclusions

Soft degrees of conformation in volatile molecules continue to be a challenge which cannot be met in a simple way by replacing potentially demanding ex- periments with theoretical calculations. In the case of the isomeric pentanoates, alternative conformers with competitive energies have been encountered con- comitantly for volatile derivatives in microwave spec- troscopy. In crystalline derivatives, the soft degrees of freedom led to disorder in individual solids and to a polymodal distribution of conformations docu- mented in the database for an ensemble of solids. Al- though the compounds under investigation are small and apparently simple, the ambiguity concerning their conformation of lowest energy cannot be avoided by state-of-the-art theoretical calculations: For the pen- tanoates investigated in this study the results do not only depend on the theoretical approach but also on the basis set chosen, and all caveats concerning the differ- ent experimental approaches must be extended to the quantum chemical calculations as well.

4. Experimental

4.1. Materials and Methods

Chemicals were used without further purifica- tion: (±)-2-methylbutyric acid (98%, Merck), iso- valeric acid (99%, Merck), valeric acid (98%, Merck), pivalic acid (Merck), ethyl 2-methyl butyrate (>90%, Merck). (±)-trans-1,2-diaminocyclohexane (99%, Aldrich) was purified by fractional distillation before use. CHN microanalyses were performed at the Institute of Organic Chemistry, RWTH Aachen Uni- versity, using a HERAEUS CHNO-Rapid.

4.2. Synthesis

Microcrystalline powders of compounds1b,2b,3b, and4b were precipitated from a mixture of diamine and the respective C5acid dissolved in tetrahydrofu- ran (THF) with a 1 : 2 diamine : acid molecular ratio. In each case single crystals were obtained by reactant dif- fusion of a diamine solution and an acid solution using THF as solvent and intermediate layer. All compounds are crystalline colourless solids. Analysis: CHN: Anal.

Calcd. for1b: C32H70N4O9: C 58.69, H 10.77, N 8.55.

Found: C 59.85, H 10.76, N 8.73. 2b: C16H34N2O4:

(8)

C 60.35, H 10.76, N 8.80. Found: C 60.23, H 10.58, N 8.80. 3b: C16H34N2O4: C 60.35, H 10.76, N 8.80.

Found: C 59.95, H 10.35, N 8.79. 4b: C16H34N2O4: C 60.35, H 10.76, N 8.80. Found: C 60.32, H 10.85, N 8.73. Melting points: 1b: 102C.2b: 113C. 3b:

103C.4b: 135C.

4.3. Crystallographic Studies 4.3.1. Single Crystal Studies

Crystal data and refinement results of the salts1b, 2b,3b, and4bhave been compiled in Tables3and4.

Intensity data were collected with a Bruker APEX area detector equipped with an Incoatec microsource (Mo- Kα radiation, λ =0.71073 Å, multilayer optics) at 100 K. Temperature was controlled with an Oxford Cryostream 700 instrument. Intensity data were eval- uated using the program SAINT [29] for integration and SADABS [30] for scaling. Structures were solved by direct methods (SHELXS-97) [31] and were re- fined by full-matrix least squares procedures as imple- mented in SHELXL-97 [31]. Hydrogen atoms of the amino groups were located from the Fourier density map and included with distance restraints. All other hydrogen atoms were included as rigid. In1ball four anions display structural disorder: three of them fea- ture anR,S disorder, the fourth 2-methylbutyrate an- ion is not affected by disorder at the chiral center but by positional disorder of the terminal ethyl group. Non- hydrogen atoms which do not show disorder were as- signed anisotropic displacement parameters.2brepre- sents a fully ordered structure. All non-hydrogen atoms were assigned anisotropic displacement parameters.

3b shows positional disorder of the terminal propyl group in one of the carboxylate anions. In 4b posi- tional disorder of one of the t-butyl group was ob- served. Further details on the crystallographic studies including fractional coordinates, displacement param- eters, and molecular geometry are given in the CIF for- mat in the SI. Crystallographic data for all data collec- tions have been deposited at the Cambridge Crystal- lographic Data Center as supplementary publications;

deposition numbers have been included in Tables 3 and4. Copies of the data can be obtained free of charge (www.ccdc.cam.ac.uk).

4.4. Microwave Spectroscopy

The microwave spectrum of ethyl 2-methyl butyrate was recorded using a MB-FTMW spectrometer operat-

ing in the frequency range from 3.0 to 26.5 GHz [32].

To assign the spectrum, a broad band scan in the fre- quency range from 9.0 to 11.5 GHz was recorded us- ing the low resolution mode of the spectrometer. After recognizing typicala-typeR-branches of the molecule, a first fit of the rotational transitions was performed us- ing the program xiam and Watson’sAreduction andIr representation [19,33]. After assigning theR-branches moreb- andc-type lines were predicted and added to the fit via trial and error. TheR-branches appear in in- tervals ofB+Cin the spectrum. In the particular case of ethyl 2-methyl butyrate the intervals amount to ap- proximately 1.6 GHz. The exact frequencies for each transition were measured again using the high reso- lution mode of the spectrometer and subsequently in- cluded in the fit to accurately determine a set of three rotational and five centrifugal distortional constants.

The experimental rotational constants were then used to validate a large number of quantum chemical calcu- lations.

4.5. Theoretical Calculations

To have an initial guess for the structure of the conformers present in the molecular beam and to start the assignment of the microwave spectrum, quantum chemical calculations were carried out to perform a conformational analysis at the MP2/6-311++G(d,p) level of theory. The MP2/6-311++G(d,p) level was empirically chosen for the geometry optimizations as it usually yields rotational constants that are within 1% deviation of the experimental ones and is therefore widely used in the spectroscopic community [34–36].

The conformational sampling was performed manu- ally using chemical intuition, i. e. by rotatingsp3C–C bonds in steps of 120. For ethyl 2-methyl butyrate we thus obtain 34=81 possible conformers. However, as cis-esters were determined to be 40 kJ/mol higher in energy thantrans-esters [20], we do do not consider them in the conformational search and thus obtain (for three rotatable bonds) a total of 33 =27 start geometries. These geometries were subsequently optimized at the MP2/6-311++G(d,p) level and in our case yielded a total of 23 different energy minima within a range of 10 kJ/mol for ethyl 2-methyl bu- tyrate. Additionally, harmonic frequency calculations were used to determine the nature of the stationary points. All calculations were carried out using the Gaussian03 or Gaussian09 program package [37]. The

(9)

three lowest energy conformers were used to predict theoretical microwave spectra and guide the assign- ment of the recorded microwave spectrum. Additional conformational sampling at higher levels of theory than the MP2/6-311++G(d,p) were not performed as calculations of higher perturbation order, e. g.

MP4, are too expensive to perform a sampling of the conformational space. After assigning the spectrum, the input geometry of the observed conformer in the molecular beam (Conformer 1 as optimized at the MP2/6-311++G(d,p) level of theory) was optimized again using selected methods and different levels of theory such as B3LYP, HF, MP2, MP3, MP4, and CCSD [38–43]. Part of the results are summarized in Table2 (for a complete overview of all methods applied, see SI). We suggest that the theoretical cal- culation of this large amplitude motion depends very much on the method and the basis set used. The good agreement at the MP3/6-311++G(d,p) level is rather

coincidental and probably due to error compensation.

Therefore, we cannot recommend any method or basis set. The Cartesian coordinates of all 23 energy minima optimized at the MP2/6-311++G(d,p) are given as supporting information (Tables VI-XXVIII) in the principal axes system of inertia.

Acknowledgements

We thank RWTH Graduiertenförderung (scholar- ship C. M.) and DFG (International Research Train- ing Group SeleCa) for financial support and the land Nordrhein-Westfalen for funds. Irmgard Kalf is thank- fully acknowledged for experimental support. We thank M. Küppers for his contribution within his stu- dents research project and the center for computing and communication of the RWTH Aachen University for free computer time.

[1] D. Brodalla, D. Mootz, R. Boese, and W. Osswald, J.

Appl. Cryst.18, 316 (1985).

[2] R. Boese and M. Nussbaumer, In Situ Crystallisation Techniques, in: Organic Crystal Chemistry (Ed. D. W.

Jones), Oxford University Press 1994, pp. 20 – 37.

[3] M. U. Kramer, D. Robert, Y. Nakajima, U. Englert, T. P.

Spaniol, and J. Okuda, Eur. J. Inorg. Chem. 5, 665 (2007).

[4] C. Lichtenberg, F. Pan, T. P. Spaniol, U. Englert, and J. Okuda, Angew. Chem. Int. Ed.51, 13011 (2012).

[5] G. Frater, J. A. Bajgrowicz, and P. Kraft, Tetrahedron 54, 7633 (1998).

[6] H. Mouhib, D. Jelisavac, W. Stahl, R. Wang, I. Kalf, and U. Englert, Chem. Phys. Chem.12, 761 (2011).

[7] I. Kalf and U. Englert, Acta Crystallogr. C67, o206 (2011).

[8] I. Kalf, B. Calmuschi, and U. Englert, Cryst. Eng. Com- mun.4, 548 (2002).

[9] I. Kalf, M. Braun, Y. Wang, and U. Englert, Cryst. Eng.

Commun.8, 916 (2006).

[10] Y. Wang and U. Englert, Inorg. Chim. Acta363, 2539 (2010).

[11] C. Wölper, N. Anwar, N. Gulzar, P. Jones, and A.

Blaschette, Acta Crystallogr.C67, o249 (2011).

[12] error-free, 3D coordinates determined, Nov. 2012.

[13] A. Bacchi and G. Pelizzi, J. Comput. Aided Mol. Des.

13, 385 (1999).

[14] H. B. Bürgi and J. D. Dunitz, Acta Crystallogr. B44, 445 (1988).

[15] H. Bürgi and J. Dunitz, Structure Correlation, Vol. 1 and 2, VCH Verlagsgesellschaft mbH, Weinheim 1994.

[16] V. Ferreira, J. Agr. Food Chem.54, 489 (2006).

[17] Scifinder Note: Calculated using Advanced Chemistry Development (ACD/Labs) Software V11.01(1994- 2013 ACD/Labs) (2013).

[18] B. S. Ray, Z. Phys.78, 74 (1932).

[19] H. Hartwig and H. Dreizler, Z. Naturforsch.51a, 923 (1996).

[20] H. Mouhib, D. Jelisavac, L. W. Sutikdja, E. Isaak, and W. Stahl, J. Phys. Chem. A115, 118 (2011).

[21] H. Mouhib and W. Stahl, Chem. Phys. Chem.13, 1297 (2012).

[22] H. Mouhib, Y. Zhao, and W. Stahl, J. Mol. Spectrosc.

261, 59 (2010).

[23] C. Foces-Foces, C. Fernandez-Castano, R. M. Clara- munt, C. Escolastico, and J. Elguero, J. Inclusion Phe- nom. Macrocyclic Chem.33, 169 (1999).

[24] O. Saied, T. Maris, and J. D. Wuest, J. Am. Chem. Soc.

125, 14956 (2003).

[25] M. Gdaniec and H. Czajka, J. Inclusion Phenom. Mol.

Recog. Chem.22, 187 (1995).

[26] A. D. Bond, Chem. Commun. 250 (2003).

[27] X. Wang, M. Simard, and J. D. Wuest, J. Am. Chem.

Soc.116, 12119 (1994).

[28] A. Guinier, G. B. Bokij, K. Boll-Dornberger, J. M.

Cowley, S. Durovic, H. Jagodzinski, P. Krishna, P. M.

de Wolff, B. B. Zvyagin, D. E. Cox, P. Goodman, T.

Hahn, K. Kuchitsu, and S. C. Abrahams, Acta Crystal- logr.A40, 399 (1984).

[29] SAINT: Program for Reduction of Data collected on Bruker CCD Area Detector Diffractometer (2009).

(10)

[30] SADABS: Program for Empirical Absorption Correc- tion of Area Detector Data V 2004/1 (2004).

[31] G. M. Sheldrick, Acta Crystallogr.A64, 112 (2008).

[32] J.-U. Grabow, W. Stahl, and H. Dreizler, Rev. Sci. In- strum.67, 4072 (1996).

[33] K. G. Watson, Vibrational Spectra and Structure, Else- vier, Amsterdam 1977.

[34] B. Velino, A. Maris, S. Melandri, and W. Caminati, J.

Mol. Spectrosc.256, 228 (2009).

[35] E. Hirota, K. Sakieda, and Y. Kawashima, Phys. Chem.

Chem. Phys.12, 8398 (2010).

[36] E. J. Cocinero, A. Lesarri, P. Cija, F. J. Basterretxea, J.- U. Grabow, J. A. Fernandez, and F. Castano, Angew.

Chem. Int. Ed.51, 3119 (2012).

[37] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E.

Scuseria, M. A. Robb, J. R. Cheeseman, G. Scal- mani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnen- berg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Ki-

tao, H. Nakai, T. Vreven, J. A. Montgomery, J. E.

Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Broth- ers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Bu- rant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M.

Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, R. Austin, A. J. Cammi, C. Pomelli, J. W.

Ochterski, R. L. Martin, K. Morokuma, V. G. Za- krzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, and D. J. Fox, Gaussian 09, revision A.02; Gaussian, Inc.: Wallingford, CT (2009).

[38] A. D. Becke, J. Chem. Phys.98, 5648 (1993).

[39] C. Lee, W. Yang, and R. G. Parr, Phys. Rev.B37, 785 (1988).

[40] D. R. Hartree, Proc. Cambr. Phil. Soc.24, 89 (1928).

[41] V. Fock, Z. Phys.61, 126 (1930).

[42] C. Møller and M. S. Plesset, Phys. Rev.46, 618 (1934).

[43] R. J. Bartlett and M. Musial, Rev. Mod. Phys.79, 291 (2007).

Referenzen

ÄHNLICHE DOKUMENTE

Figure 1: The price rises with demand and falls with supply under the condition of a fixed allocation of labor input and a fixed wage rate; demand is here represented by the

We use Erd¨ os’ probabilistic method: if one wants to prove that a structure with certain desired properties exists, one defines an appropriate probability space of structures and

The following theorem (also from Chapter 2 of slides) has an analogous formulation..

Attempts to generate a new framework or new umbrella term (e.g., NPR, 2015; EdWeek, 2015) while well-intentioned, are not designed to address what we see as the

Whereas Ridley ’ s task is to expose Augustus ’ complicated relationship with autobiography (having written the lost de vita sua covering the  rst half of his life and the

states that the Australian languages agree with the Mundä family. I do not think that they are numerous or important

In order to utilise the available and accurate measurements of enthalpy, isobaric heat capacity and speed of sound, an equation of state has been developed for

itself in Eastern Europe, the religious scholars made no attempt to learn. about intellectual movements in Western Europe, doubtless hoping