Вы находитесь на странице: 1из 14

Journal of Molecular Structure 749 (2005) 155–168

www.elsevier.com/locate/molstruc

Steric compression and twist in o-hydroxy acyl aromatics


with intramolecular hydrogen bonding
Poul Erik Hansena,*, Simon Bolvigb, Krzysztof Wozniakc
a
Department of Life Sciences and Chemistry, Roskilde University, P.O. Box 260, DK-4000 Roskilde, Denmark
b
Department of Spectroscopy, Leo Pharmaceutical Products, DK-2750 Ballerup, Denmark
c
Department of Chemistry, Warsaw University, ul. Pasteura 1, 02 093 Warsaw, Poland
Received 2 July 2004; revised 18 March 2005; accepted 22 March 2005
Available online 31 May 2005

Abstract
A series of o-hydroxy acyl aromatics of the type 1,3-diacetyl-2,4,6-trihydroxybenzene (1), 6-methoxy-1,3-diacetyl-2,4-dihydroxy- (2),
2,4,6-trihydroxy-1,3,5-triacetylbenzene (3) and 1-acetyl-2-naphthol (4) have been investigated by means of single crystal X-ray diffraction,
solution and solid state NMR spectroscopy and theoretical calculations. The structures of 1 and 2 exhibit interesting hydrogen bonds, planar
structures and have, as a consequence, unexpected geometrical parameters (interatomic distances, bond lengths and valence angles) and
depletion of electron density of the aromatic rings. Also significant packing effects are present. For 4, a twist of the carbonyl group is
observed together with an out of plane bending of the C–C]O bond leading to the formation of a C]O/H–O–C hydrogen bond which is
almost coplanar with the naphthalene rings. Solid state NMR spectra show lack of C3 symmetry for 3. Solution NMR spectra show-in the case
of 1 and 2-quite different behaviour. A complex averaging-observed for 1 in solution-is unravelled at low temperature. Compounds 1–3 show
large two-bond deuterium isotope effects, 2DC-2(OD), on 13C chemical shifts. This indicates strong hydrogen bonds. These can be
understood in terms of an electronic effect caused by bond localisation of the benzene ring and a steric effect caused by either neighbouring
CH3CO, OH or OCH3 groups leading to shorter OH/O and O/O distances and, consequently, stronger hydrogen bonds. A general scheme
for distinguishing between steric twist (as seen in 4) and steric compression as seen in 1–3 is suggested. An experimental method based on
isotope effects at the chelate proton of compounds deuteriated at the CD3CO groups is demonstrated.
A Bader atom in molecules is done to investigate hydrogen bonding.
q 2005 Elsevier B.V. All rights reserved.

Keywords: Isotope effects; Hydrogen bonding; NMR; X-ray diffraction; Structure calculations; Aim analysis

1. Introduction described by the oxygen–oxygen distance (RO/O). This


parameter is also a reasonable measure of the strength of the
Strong intramolecular hydrogen bonds have been found hydrogen bond [9]. The RO/O distance may be altered by
in mono anions of dicarboxylic acids [1] but may also be steric compression as seen in, e.g. the enolic form of 2,2,6,6-
present in o-hydroxyacyl aromatics [2]. A prerequisite for tetramethyl-3,5-heptanedione compared to that of acetyla-
the latter type is the possibility of resonance assisted cetone [1]. Deuterium isotope effects on chemical shifts
hydrogen bonding (RAHB) [3–5]. This type of hydrogen have been used extensively to describe hydrogen
bonding is also found in tautomeric compounds like enolic bond strength in intramoleculary hydrogen bonded systems
forms of b-diketones [1,6]. Much effort is concentrated on [2,5,11–29] and are shown to correlate with the oxygen–
characterisation of hydrogen bonded systems, especially oxygen distance (R O/O ) [2]. Recently, theoretical
to provide parameters to describe hydrogen bond strength calculations of a series of o-hydroxy acyl aromatics
[1–4,7–10]. The hydrogen bond can advantageously be (2-hydroxyacetophenone, 1,3-diacetyl-2,4-dihydroxyben-
zene and 3 have shown that steric effects are also important
* Corresponding author. Tel.: C45 46742432; fax: C45 46743011. determinants of the RO/O distances and for the hydrogen
E-mail address: poulerik@ruc.dk (P.E. Hansen). bond strength in these systems [2].
Both types of molecules investigated have important
0022-2860/$ - see front matter q 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.molstruc.2005.03.034 non-bonded interactions in addition to the hydrogen bonds.
156 P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168

An analysis based on the Bader principles [30] seemed very 2.2.2. Solid state
appropriate. Solid-state 13C NMR with cross-polarization (CP) and
Steric interactions may for groups like CH3C]O, magic-angle spinning (MAS) was done at 50.3 MHz on a
HC]O, NO2 etc. involved in hydrogen bonding not only Varian Unity-200 spectrometer using a high-speed double
lead to a reduction of the heavy atom distance RO/O but can bearing probe and silicon nitride rotors spun in dry air. The
of course also lead to a twist of the group (non-planarity) single-contact 13C CP/MAS spectra were recorded at 293 K
[29]. It is the aim of the present paper to unravel the factors with the optimised-contact time of 4 ms (OCT) and with the
leading to compression vs. twist and to characterise the short-contact time of 50 ms (SCT). We used the p/2 pulses
angles involved in the non-planar structures. of 5 ms, the recycle delays of 3 s and the MAS rate of 7 kHz.
Steric and electronic factors have previously been The 13C dipolar-dephased spectra (DD) were recorded with
discussed in push–pull benzenes [31] and in Schiff bases the contact time of 4 ms and the 50-ms period without
[32]. Steric compression effects have been observed in decoupling prior to acquisition. The DD experiments expose
ketoenamines with substituents at the C-a carbon as quaternary carbon peaks and the SCT experiments highlight
revealed by on 2DC(ND) isotope effects [10]. The present the peaks from carbons with adjacent protons. Ordinary
findings are used to classify and explain existing data [23]. OCT spectra contain all these peaks.
A prerequisite for structural comparisons is an access to
structural data. X-ray structures and ab initio calculations 2.3. X-ray diffraction
are performed to provide the data whereas the solid state
NMR provides complementary information about the solid Single crystal X-ray diffraction data were collected on a
state. KM4 (1) and KUMA4CCD (2–4) diffractometers with use
In the present study, deuterium isotope effects on of graphite monochromated Cu Ka (1) and Mo Ka (2 and 4)
chemical shifts in compounds with motifs leading to steric radiations. The data were collected at room temperature
compression (Fig. 1(A)) are described and discussed in using uK2q (1) and u (2 and 4) scan techniques. Data
relationship to isotope effects caused by steric twist of the reduction was carried out with the KUMA (1) and
carbonyl groups involved in intramolecular hydrogen KUMARED (2 and 4) software. Because of very small
bonding. values of the absorption coefficients, no absorption correc-
tion was applied, but the data were corrected for the Lorentz
and polarisation effects. The intensities of three standard
2. Experimental reflections (1) per every 100 collected were used for scaling
the data and they varied by less than 3%. In the case of the
other crystals a reference frame was repeated every 25
2.1. Compounds frames collected. The program SHELXS [35] was used for
structure solution and the SHELXL [36] for refinement. The
2,4,6-Trihydroxy-1,3-diacetylbenzene (1) and 2,4-dihy- refinement was based on F2 for all reflections except those
droxy-6-methoxy-1,3-dicacetylbenzene (2) were purchased with very negative F2. Weighted R factors wR and all
from Maybridge Chemical Company, Tintagel, UK. goodness-of-fit S values are based on F2. Conventional R
2,4,6-trihydroxy-1,3,5-triacetylbenzene (3) was prepared factors are based on F with F set to zero for negative F2. The
as described in Ref. [2] and 1-acetyl-2-naphthol (4) was F02 O 2sðF02 Þ criterion was used only for calculating R
purchased from Aldrich, Weinheim, Germany. factors and is not relevant to the choice of reflections for the
refinement. The R factors based on F2 are about twice as
2.2. NMR large as those based on F. Anisotropic temperature factors
were used to describe the thermal motions of nonhydrogen
2.2.1. Solution atoms. Almost all hydrogen atoms were located from
The 1H and 13C NMR spectra were recorded on a Bruker differential maps and refined isotropically. Atomic scatter-
AC250 MHz instrument using CDCl3 as solvent except for ing factors for C, H and O were derived from wave functions
low temperature spectra for which CD2Cl2 was used. TMS tabulated in Tables 6.1.1.4 and 4.2.4.2 from the Ref. [37].
was used as internal reference. For 1 a mixture of CD2Cl2 Other data are given in Table 1.
and THF-d8 was used. 13C NMR spectra were recorded with Labelling of atoms and an illustration of thermal motions
a spectral resolution of 0.55 Hz/point. The COLOC [33] of atoms are shown in Fig. 1(B).
spectra were recorded as described earlier [24]. Spectra of Crystallographic data (excluding structural factors) for
isotopologs were repeated with different deuterium the structures reported in this paper have been deposited
contents. with the Cambridge Crystallographic Data Centre and
The assignments of 1 and 2 were achieved by COLOC allocated the deposition numbers: CCDC 151493, CCDC
[33] experiments. That of compound 2 is further confirmed 151494 and CCDC 151496 for 1, 2 and 4, respectively.
by a SELJRES experiment irradiating the two OH Copies of the data can be obtained free of charge on
resonances, respectively [34]. application to CCDC, 12 Union Road, Cambridge CB2
P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168 157

A O O
H H
HO O O O

O O O O
H H
1 2
O
H
O O O
H H
O O

O O
H
3 4

1 2

Fig. 1. (A) Compounds investigated. (B) Labelling of atoms and an illustration of thermal motions of atoms in 1 (a), 2 (b) and 4 (c). The numbering schemes for
1 and 2 are based on the one for 3. The most interesting geometrical parameters for 4 are the following: C1C2Z1.390(3) Å, C1C11Z1.463(3) Å, C11O1Z
1.249(2) Å, C2O2Z1.341(3) Å, H8H122Z2.22 Å, H8H121Z2.08 Å, O1C11C1Z119.6(2)8, C11C1C2Z117.2(2)8, C1C2O2Z122.8(2)8, C2O2H2oZ
107(2)8, C3C2C1Z121.6(2)8, C9C1C2Z118.0(2)8, C11C1C2O2ZK10.0(3)8, C11C1C9C10ZK167.3(3)8, O1C11C1C2Z18.8(3)8,
O1C11C1C9ZK167.7(3)8, O1C11C2O2Z7.8(3)8.
158 P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168

Table 1
Crystal data for compounds 1, 2 and 4

Identification code 1 2 4
Empirical formula (C10H11O5)2!H2O (C11H12O5)2 C12H12O2
Formula weight 438.38 448.4 186.2
Temperature (K) 293 293 293
Wavelength (Å) 1.54178 0.71073 0.71073
Crystal system Triclinic Triclinic monoclinic
Space group P-1 P-1 P21/n
Unit cell dimensions (Å and 8) aZ7.539(2) aZ3.9310(8) aZ9.386(2)
bZ9.972(2) bZ9.745(2) bZ10.569(2)
cZ14.331(3) cZ27.210(5) cZ9.727(2)
aZ86.11(3) aZ89.94(3) aZ90
bZ82.62(3) bZ89.52(3) bZ109.62(3)
gZ68.40(3) gZ87.80(3) gZ90
Volume (Å3) 993.2(4) 1041.5(4) 908.9(3)
Z 2 2 4
Density (calculated) (mg/m3) 1.466 1.430 1.361
Absorption coefficient (mmK1) 1.038 0.114 0.092
F(000) 460 472 392
Crystal size (mm3) 0.30!0.30!0.30 0.40!0.08!0.02 0.30!0.30!0.25
Theta range for data collection (8) 3.1–75 3.65–20 3.6–20
Index ranges 9%h%9 K3%h%3 K9%h%9
K12%k%12 K9%k%9 K10%k%10
K1%l%17 K26%l%26 K9%l%9
Reflections collected 4378 9596 8397
Independent reflections 3909 [R(int)Z0.027] 1934 [R(int)Z0.1010] 843 [R(int)Z0.065]
Refinement method Full-matrix least-squares on F2
Data/restraints/parameters 3909/0/369 1934/0/383 843/0/168
Goodness-of-fit on F2 1.047 1.200 1.139
Final R indices [IO2s(I)] R1Z0.0530 R1Z0.0872 R1Z0.0387
wR2Z0.1407 wR2Z0.1583 wR2Z0.093
R indices (all data) R1Z0.0685 R1Z0.1115 R1Z0.0402
wR2Z0.1566 wR2Z0.1718 wR2Z0.0952
Extinction coefficient 0.013(2) 0.004(2) 0.025(5)
Largest diff. peak and hole (e AK3) 0.36; K0.22 0.23; K0.21 0.11; K0.13

1EW, UK (fax: Int code C(1223) 336 033; E-mail: by a SELJRES experiment irradiating the two OH
deposit@ccdc.cam.ac.uk). resonances, respectively [34].
A comparison of the low temperature OH resonance
2.4. Theoretical calculations positions of 1 and 2 revealed that the high frequency ones
are almost identical (the isotope effects at these protons are
The molecular geometries were optimised using the also identical (Scheme 1). This strongly suggested that the
Gaussian98 suite of programs [38] and B3PW91 Density low frequency (11.10 ppm) resonance of 1 belongs to OH-6
Functional Theory (DFT) (Beckes exchange [39] and of 1 as this OH group does not occur in 2.
Perdew–Wang correlation term [40]) and the 6-31G(d,p) The assignment of the isotope effects of compounds with
basis set. more than one exchangeable OH proton was done in the
Wave functions for atoms in molecules analysis [30] following fashion. The isotope effects in o-hydroxy
were of RHF 6-31G* type. The bond path critical points, aromatics have been grouped into categories as described
charges etc. were analysed using the MORPHY program in Refs. [15,18]. As an example, at the methyl group of the
[41]. o-acetyl group, a negative isotope effect proportional to
dOH was observed. A very large and unusual four bond
isotope effect, 4DC(OD) is observed in compounds 1 and 2.
3. Results This is assigned to 4DC-4(OD-2) (Scheme 1).

3.1. NMR assignments 3.1.2. Solid state


The assignment of 13C NMR solid state spectra is, in
3.1.1. Solution general, based on the solution spectra, so the assignment
The assignments of 1 and 2 were achieved by COLOC of 2 is straightforward. The distinction of the carbonyl
[33] experiments. That of compound 2 is further confirmed carbons is based on the fact that the carbonyl groups
P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168 159

-0.141d,-0.123e,-0.141f ~0
-0.113e H3C O
H (D)
10.4 (br)c,
11.01d HO O 0.664c,0.652d,0.649e,0.654f
0.045f 0.096e,0.094f
CH3

0.226d,0.225e,0.215f O O
H
14.5 b(br)c,
14.69 (0.044)b,d
1-2OD

16.24c,
0.04e 16.37 (0.029)d
H3C O
H
HO O 0.099d,0.086e,0.079f
0.050f 0.044e
CH3

0.442d,0.440e,0.440f O O -0.207d,-0.209e,-0.222f
H
(D) 0.117d,0.112e,0.108f

1-4OD
0.076e H3C O
H
(D) HO O 0.017e
0.183d,0.191e,0.197f 0.017e
CH3
-0.036h
-0.128h O O
H
H3C O
-0.045h H (D) 1-6OD
O O
H3C 0.686h/0.686g
0.051g CH3 -0.016
-0.094g/-0.088h
0.217g O O
H
-0.035h
15.0 (0.047)g
2-2OD H3C O
H 16.4 (0.030)g
0.051g
O O
H3C 0.083g
0.097g
0.095g/0.100h CH3 -0.203h

0.467g O O
H 0.121h
(D)
2-4OD

Scheme 1. Deuterium isotope effects on 13C chemical shifts for 1 and 2. aData for 3 and 4 see Refs. [2,18,29]. b Numbers in brackets are nDOH(OD) isotope
effects. Numbers in italics are OH chemical shifts. c300 K. Solvent CD2Cl2/THF-d8. d230 K. Solvent as for c. e190 K. Solvent as for c. f170 K. Solvent as for c.
g
250 K. Solvent CD2Cl2. h190 K. Solvent as for g.

having the weaker hydrogen bonds are at highest 3.2. Solution NMR
frequency. For C-2, C-4 and C-6 of 1 the two resonances
at lowest frequency are the ones not involved in the 3.2.1. Temperature effects
intramolecular hydrogen bonding. The one at the higher The deuterium isotope effects on 13C chemical shifts are
frequency of the two is assumed to be the OH group of measured in one tube experiments with both the protio
the A-molecule, which is hydrogen bonded to a water and the deuterio species present and in varying ratios. A
molecule in the crystal. prerequisite for this type of measurement is slow
160 P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168

intermolecular exchange of the OH proton at the NMR time Another significant observation is the rather small
4
scale. DCO(OD-2) observed in 1 and 2 and a negative
5
The deuteriated form of 2 showed an extensive set of DCH3(OD-2).
isotope effects as seen in Scheme 1. Upon cooling to 250 K A methoxy or a hydroxy group in position 6 leads to an
the deuterium isotope effects did not vary with temperature. increase of 2DC-2(OD), [18] thus explaining the larger
At low temperature, a deuterium isotope effect is seen at values found for 1 and 2 compared to 1,3-diacetyl-2,4-
H-5. Such effects are usually not seen at the hydrogen ortho dihydroxybenzene [2]. Cooling had only little effect on the
position to a hydrogen bonded OH group. Large positive magnitude of the isotope effects.
isotope effects are found at both OH-2 and OH-4 chemical Relatively large deuterium isotope effects on OH
shifts upon deuteriation at the other OH group. This is chemical shifts are observed over six bonds, 6DOH(OD) in
probably not a simple long-range effect, as such an effect 1 and 2 and 6DOH-2(OD-4)!6DOH-4(OD-2) (Scheme 1).
is not seen at H-6 of the ‘frozen’ form of deuteriated Similar effects are found in 1,3-diacetyl-2,4-dihydroxyben-
4-methoxy-2,6-dihydroxyacetophenone [24]. zene and 3.[2] The OH chemical shifts are dOH-2OdOH-4.
The 1H NMR spectrum of 1 at ambient temperature Deuterium exchange of the methyl group protons leads to
showed a sharp resonance at 16.24 ppm in CDCl3. In a isotope effects at the chelate proton, DOH(CD3) [29]. In the
mixture of CD2Cl2 and THF-d8 a sharp resonance was seen case of 4 this was seen to be large, [18,29] whereas for 2
at 16.32 ppm together with two broad resonances at w14.5 they were zero for both OH groups and for 3 it was
and w10.5 ppm. The 13C spectrum in CDCl3 at ambient negligible, 0.004 ppm [2].
temperature showed a broad and merged resonance for C-4,
C-6. Nevertheless, isotope effects are observed at C-1, C-2, 3.3. Solid state NMR
C-3, C-5 and OH-6. Upon cooling to 230 K all 13C
resonances sharpened and the OH resonances showed up 13
C NMR spectra of 1, 2 and 3 have been recorded. 2
as three resonances at 16.40, 14.73 and 11.01 ppm. showed a straightforward spectrum rather similar to the
solution spectrum (Table 2). For 1 two sets of resonances in
3.2.2. Isotope effects on chemical shifts the ratio 1:1 are observed. For the C]O part two carbonyl
The two bond deuterium isotope effects (nDC-2(OD)) on resonances per molecule are seen, whereas for the C–OH
13
C chemical shifts of 1 and 2 are unusually large for carbons three resonances per molecule are observed,
benzene derivatives and only surpassed by those of 3 [2] and showing that the molecule is not symmetrical. The
so are a number of long range isotope effects. A comparison differences between resonances belonging to the two
of the long-range isotope effects showed a very large four- molecules are quite distinct (see Table 2). For 3, the
bond effect, 4DC-2(OD-4) in 1 and 2. An unusual isotope resonances corresponding to the carbonyl carbon show two
effect is observed at C-7 of 2. This type of isotope effect is resonances in the ratio 2:1, the same is true for C-2, C-4 and
normally positive [18]. For 1,3-diacetyl-2,4-dihydroxyben- C-6, whereas those corresponding to CH3 and C-1, C-3 and
zene the effect is small for C-7, 0.090 ppm [2]. This is C-5 show no resolved splittings.
a general trend for hydroxy and methoxy substituted For compound 2 two structures are seen in the
compounds in general as judged from data of Ref. [18]. asymmetric unit. However, these are very similar (see
For 1 the effects are only assigned tentatively due to overlap. Table 3) and so similar that they only give rise to one set
Table 2
13
C chemical shifts of 1–3 in ppm either in solution or in the solid state

C-1 C-2 C-3 C-4 C-5 C-6 C-7 C-8 C-9 C-10 C-11
a
1 (l) 104.4 172.2 104.4 169.0 95.0 169.0 204.2 32.8 204.2 32.7 –
1 (l)b 103.8 172.4 104.3 171.7 94.9 167.35 204.4 33.7 204.4 33.5 –
l (s)c 104.4 173.8d 104.4 172.0d 97.2 169.1 206.2e 32.6 205.1e 32.6 –
173.0d 172.0d 95.1 166.2 204.3e 203.2e
2 (l) 104.3 171.7 105.1 172.1 91.4 167.1 203.3 33.0 204.4 32.8 56.0
2 (s) 103.5 172.4 104.1 172.7 92.5 167.7 201.9 34.0 205.8 34.0 58.9
3 (l)f 103.2 175.7 205.0 32.9
3 (s) 101.2(1)g 175.5 (br)h 204.7(2)g 33.0
101.6(2) 203.7(1)
a
l means liquid. Temperature 300 K. Solvent CD2Cl2/THF-d8 (1:1).
b
Temperature 230 K. Solvent CD2Cl2/THF-d8 (1:1).
c
s means solid.
d
May be interchanged.
e
For assignments see Section 3.
f
Taken from Ref. [30].
g
Numbers in brackets give relative intensities of the resonances.
h
br means broad.
P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168 161

Table 3
Geometry of hydrogen bonds formed in 1, 2 and 4

Molecule D–H/A symmetry D–H H/A D/A D–H/A


1 O2a–H2oa/O1a O2a–H2oa H2oa/O1a O2a/O1a O2a–H2oa/O1a
0.95 1.54 2.449 160
O4a–H4oa/O3a O4a–H4oa H4oa/O3a O4a/O3a O4a–H4oa/O3a
0.93 1.60 2.480 157
O2b–H2ob/O1b O2b–H2ob H2ob/O1b O2b/O1b O2b–H2ob/O1b
0.91 1.63 2.483 154
O4b–H4ob/O3b O4b–H4ob H4ob/O3b O4b/O3b O4b–H4ob/O3b
0.88 1.65 2.480 155
O1w–H1w/O1a O1w–H1w H1w/O1a O1w/O1a O1w–H1w/O1a
0.79 2.04 2.830 177
O1w–H2w/O3b O1w–H2w H2w/O3b O1w/O3b O1w–H2w/O3b
XC1,Y,ZK1 0.87 1.97 2.832 170
O6a–H6a/O1b O6a–H6a H6a/O1b O6a/O1b O6a–H6a/O1b
X,YC1,Z 0.95 1.80 2.729 165
O6b–H6ob/O1w O6b–H6ob H6ob/O1w O6b/O1w O6b–H6ob/O1w
X,YK1,ZC1 0.93 1.71 2.606 163
2 O2a–H2oa/O1a O2a–H2oa H2oa/O1a O2a/O1a O2a–H2oa/O1a
1.01 1.55 2.424 142
O4a–H4oa/O3a O4a–H4oa H4oa/O3a O4a/O3a O4a–H4oa/O3a
0.80 1.69 2.457 160
O2b–H2ob/O1b O2b–H2ob H2ob/O1b O2b/O1b O2b–H2ob/O1b
0.85 1.66 2.408 146
O4b–H4ob/O3b O4b–H4ob H4ob/O3b O4b/O3b O4b–H4ob/O3b
1.07 1.44 2.469 159
4 O2–H2o/O1 O2–H2o H2o/O1 O2/O1 O2–H2o/O1
1.01 1.56 2.474 147

Bond lengths and nonbonding distances in Å and valence angles in degrees.

of NMR resonances. The observation of two sets of 3.4. X-ray diffraction


resonances for 1 is well in line with the finding of
distinctly different structures as observed by X-ray. For Two o-hydroxy acyl aromatic compounds (1 and 2)
the set of C-2, C-4 and C-6 resonances one is seen not to crystallise in P-1 triclinic space group with two molecules in
be hydrogen bonded whereas the two others are hydrogen the independent part of the unit cell. Compound 3 crystal-
bonded. This is true for both molecules. The two non- lises in the monoclinic space group P21/c and 4 crystallises
hydrogenbonded carbons show quite different chemical in P21/n, both with one independent molecule in the unit
shifts (Table 2) probably due to the fact that in molecule- cell. There are only minor differences in geometrical
A the OH group is hydrogen-bonded to a water in the parameters of the two independent moieties in 1 and 2.
crystal (RO/OZ2.61 Å). The most significant of them are equal to ca. 0.01 Å for
For 3 two molecule are found the asymmetric unit [42]. bonds and 0.58 for valence angles of 1 and 0.03 Å for bonds
The observation of C]O resonances in the ratio 2:1 and 28 for valence angles of 2. However, none of these
indicates that the molecule has no C3 axis. The finding that differences is significant. The two independent moieties in 1
the most intense resonance is at higher frequency shows and 2 are located in the crystal lattices of these compounds
that the carbonyl groups are less strongly hydrogen bonded in such a manner that they form columns of stacked
than the third. The solid state spectrum supports a low molecules arranged in head-to-head way. An angle between
resolution neutron diffraction structure, which indicates that the best planes of molecules forming independent columns
intramolecular hydrogen bonds are rather similar, whereas is equal to 558 and 448 for 1 and 2, respectively, whereas the
the third is different [42]. A comparison to the solution distances between the planes of the closest molecules in
structure shows remarkable differences as the solution stacks is equal to 3.49 Å for 1 and 3.83 Å for 2 for the
spectrum indicates three equally strong hydrogen bonds on columns along X and 3.56 Å along Z. Characteristic
the NMR time scale [2]. examples of the 3D packing of o-hydroxy acyl aromatics
The infra red spectrum in CDCl3 showed two absorptions are shown for molecule 2 in Fig. 2.
of equal intensity at 1614 and 1590 cmK1 due to carbonyl The independent columns of molecules A and B in 2 go
stretching in line with C3h symmetry of the molecular in along the crystallographic x axis characterised by the
solution. The low frequencies likewise support the finding shortest unit cell parameter. Crystals of 1 contain also one
of weak C]O double bonds. molecule of water in the independent part of the unit cell.
162 P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168

Table 4
The most important structural parameters of 1 and 2

Parameter 1 2
A B A B
C1C2 1.425(2) 1.427(2) 1.433(9) 1.420(9)
C2C3 1.410(2) 1.419(2) 1.40(1) 1.381(9)
C3C4 1.421(2) 1.421(2) 1.40(1) 1.40(1)
C4C5 1.391(3) 1.381(3) 1.37(1) 1.36(1)
C5C6 1.374(2) 1.383(2) 1.37(1) 1.35(1)
C6C1 1.431(2) 1.420(2) 1.440(9) 1.408(9)
C1C7 1.457(2) 1.457(2) 1.47(1) 1.46(1)
C7O1 1.248(2) 1.249(2) 1.227(9) 1.233(8)
C2O2 1.330(2) 1.323(2) 1.298(8) 1.324(9)
C9O3 1.237(2) 1.245(2) 1.268(8) 1.242(9)
C4O4 1.332(2) 1.338(2) 1.340(8) 1.337(8)
C6O6 1.348(2) 1.340(2) 1.347(8) 1.354(8)
C2C1C6 116.7(1) 116.9(1) 116.0(7) 115.6(7)
C1C2C3 122.7(1) 122.6(2) 122.3(7) 124.0(7)
C2C3C4 116.9(2) 116.6(1) 117.0(7) 115.3(7)
C3C4C5 121.8(2) 122.0(1) 123.5(8) 123.4(7)
C4C5C6 120.3(2 120.4(2) 119.5(9) 119.7(8)
C5C6C1 121.5(2) 121.5(1) 121.7(7) 122.1(7)
C7C1C2 119.0(1) 119.5(1) 118.4(7) 119.9(7)
C1C2O2 119.5(2) 120.0(1) 118.7(7) 118.1(7)
O2C2C3 117.8(2) 117.4(1) 119.0(7) 117.8(7)
C2C3C9 124.2(2) 124.3(2) 123.1(7) 125.1(7)
C9C3C4 118.9(2) 119.1(1) 119.9(7) 119.7(7)
C3C4O4 121.6(2) 121.6(2) 121.2(7) 120.3(7)
O4C4C5 116.7(2) 116.4(2) 115.3(8) 116.3(8)
C5C6O6 119.7(1) 119.9(2) 123.6(7) 122.2(7)
O6C6C1 118.8(2) 118.6(1) 114.7(6) 115.7(7)
C6C1C7 124.3(2) 123.7(1) 125.6(7) 124.5(7)
C1C7C8 123.7(2) 123.0(2) 123.6(8) 124.8(8)
C1C7O1 119.7(2) 119.9(2) 119.5(7) 118.3(7)
O1C7C8 116.6(2) 117.1(2) 116.8(8) 116.8(7)
C3C9C10 124.1(2) 123.8(2) 124.6(8) 122.7(9)
C3C9O3 118.9(2) 119.0(2) 117.9(7) 119.8(8)
C10C9O3 117.0(2) 117.1(2) 117.5(8) 117.5(8)
C5C11C12
C5C11O5
O5C11C12

Bond lengths in Å and valence angles in degrees.

It plays the role of a glue joining-via intermolecular


hydrogen bonds—the two independent A and B moieties
present in this structure. The hydrogen bonds are formed to
the oxygen atoms from carbonyl groups. The other lone
pairs of those oxygens are involved in the intramolecular
hydrogen bonding with donating O–H groups.
The crystal structure of compound 4 consists of pairs of
molecules arranged tail-to-head due to apparent dipole–
dipole interactions. The distance between the molecules in a
pair is equal to ca. 3.67 Å. Such pairs of molecules are
located in a T-shape manner in the crystal lattice utilising a
quadrupolar character of the naphthalene rings. All four
molecules also form intramolecular hydrogen bonds. The
geometry of all hydrogen bonds is shown in Table 3.
The most important geometrical parameters of 1 and 2
Fig. 2. 3D-packing of molecules of 2: (a) projection along X-axis,
are collected in Table 4, whereas for 4 they are given in the
(b) projection showing difference between columns formed by independent
molecules. caption of Fig. 1.
P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168 163

3.5. AIM analysis hydroxy groups influences the aromatic ring angles
decreasing the ipso acyl angles up to 115–1178 and
For compounds 1, 2 and 4 ring critical points are increasing the ipso OH angles to 123–1248. This is also
identified corresponding to the rings C1, C7, O1, H2O, O2 associated with the changes of the aromatic bond lengths-
and C2; C3, C9, O3, H4O, O4 and C4; C2, O2, C10, C9 and the ones involved in the RAHB are becoming longer (above
C3 and C6, O6, C8, C7 and C1 in addition to that of C1, C2, 1.4 Å). A comparison with acetophenone [43] shows that
C3, C4, C5 and C6 for 1 and 2. the bond lengths in general are longer in 1 although
For 4 only two are identified, corresponding to C2, O2, acetophenone also have one electron withdrawing group.
H2O, O1, C11 and C1; C1, C11, C12, H8, C8 and C9. The data for 2, which are clearly less well determined due to
Data for point critical points are given in Table 1S. difficulties in growing good crystals (see crystal packing
For C–O bonds the electron density at the bond critical pattern) show similar features.
points depends on the amount of intramolecular hydrogen The most striking feature of 4, except hydrogen bonding,
bonding. The interactions C/O of 1 and 2 and H/C of 3 is the out-of-plane deformation of the acyl group which is
are characterised by high ellipticities (Table 1S). arranged in such a manner that the O2–H2O/O1 hydrogen
bonding form in this structure is almost co-planar with the
best plane of the aromatic fragment. This is well illustrated
4. Discussion in Fig. 3. The angle between the best planes of both (acyl
and aromatic) fragments is equal to 24.58. There seem to be
two sources of this deviation. One intramolecular-repulsive
4.1. Structural data nonbonding interactions between H8 and the methyl
hydrogen atoms (see caption of Fig. 1). Whereas the other
The structures of 1 and 2 are discussed together, whereas one-intermolecular-are the Me group interactions with pi-
that of 4 is discussed separately. In the discussion of strong electron electron density of the neighbouring molecule in
hydrogen bonds it is of great importance to find that in 1 and the crystal lattice of 4 (Fig. 3(b,c)). Deformation of the
2 the oxygen–oxygen distances have become short. The molecule is also confirmed by values of torsion angles
shortest O/O distance is in the moiety 2B (2.408 Å). Such formed by atoms of the acyl group which are equal to
relatively strong hydrogen bonds in 1 are almost coplanar ca. 10–208. Such a deformation enables better co-planarity
with the aromatic rings with the largest deviation from of the O1/H2o–O2 hydrogen bond.
planarity-measured by an angle between the least-square
planes of the aromatic and hydrogen bonded molecular
fragments of 1-equal just 15.48 for O4b–H4bo/O3b H- 4.2. Theoretical calculations
bond. Due to involvement in hydrogen bonding also the
carbonyl C]O bonds are longer-in the range from 1.237 to 4.2.1. Structures
1.249 Å as compared to acetophenone (1.216 Å) [43] and The DFT calculated structures of 1–2 confirm the planar
1,3-diacetyl-2,4-dihydroxybenzene (1.238 Å) [44]. Electron structures and the short O/O distances in general. The data
withdrawing and electron donating character of the acyl and are presented in Table 5. A comparison with X-ray data for 1

Fig. 3. (a) Out-of-plane deformation of the acyl group in 4, and (b) and (c) Me/pi-electron interactions in the crystal lattice of 4.
164 P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168

Table 5 saturation of the chelate protons. However, the trends in,


Calculated distances and angles of compounds 1–3a e.g. the OH bond lengths are similar although the numbers
Com- 1 2 3 3 (non- 2-hydro- are different. For 3 the structure with the lowest energy is
pound planar)b xyaceto- planar and have in practise C3h symmetry. This is similar to
phenone what is found in solution as shown by liquid state NMR and
Bond lengths IR spectroscopy, but clearly different from the findings by
C1C2 1.435 1.434 1.431 1.439 1.420 solid state NMR spectroscopy. An interesting feature of 3 is
C2C3 1.420 1.421 1.421 1.410 1.400
a calculation of a non-planar, higher energy structure in
C3C4 1.430 1.427 1.421 1.430 1.380
C4C5 1.400 1.403 1.421 1.400 1.400 which the acetyl group is twisted 568 out of the ring plane
C5C6 1.382 1.382 1.431 1.390 1.380 (see Table 5). The structure of this compounds is very
C6C1 1.423 1.430 1.421 1.432 1.410 similar to those of 1 and 2 emphasising that as the steric
C1C7 1.450 1.456 1.458 1.450 1.460 strain and the cooperativity is partly lost in the non-planar
C7O1 1.253 1.254 1.254 1.250 1.240
structure so that the hydrogen bonds become less strong.
C2O2 1.317 1.317 1.310 1.310 1.330
C3C9 1.462 1.462 1.458 1.460 –
C9O3 1.249 1.250 1.254 1.240 – 4.2.2. Bader’s atoms-in-molecules analysis
C4O4 1.318 1.318 1.310 1.310 – Once theoretical electron density has been established,
C5C11 – – 1.456 1.515 – the ‘Bader’s Atoms-in-Molecules approach’ [30,45] pro-
C11O5 – – 1.254 1.210 –
vides an excellent tool for the interpretation of topological
C6O6 1.351 1.347 1.310 1.340 –
O2H 1.032 1.034 1.046 1.020 0.990 properties of ab initio calculated charge density r(r).
O4H 1.020 1.020 1.044 1.020 – Any bonded pair of atoms has a bond path, i.e. a line of
O6H – – 1.044(5) 0.096 – the highest electron density linking them. The point on this
O1O2 2.419 2.411 2.401 2.420 2.536 line where the gradient of r, V(r), is equal to zero, is termed
O3O4 2.452 2.451 2.404 2.430 –
the bond critical point (BCP) and the properties of the
O5O6 – – 2.403 2.850 –
Bond angles
density at this point, rb, give quantitative information on
C2C1C6 117.06 117.30 118.28 117.51 118.70 that bond’s character. A bond path between a pair of non-
C1C2C3 122.03 121.99 121.72 121.70 119.50 covalently bonded atoms is called an interaction line. Its
C2C3C4 117.66 117.45 118.26 117.73 120.16 length can be different from the length of the internuclear
C3C4C5 120.98 121.41 121.76 121.97 121.05 vector. It could represent the influence of the local
C4C5C6 121.42 120.28 118.25 119.06 119.18
C5C6C1 121.93 121.59 121.73 122.13 121.39
electronic environment on a given interaction.
C7C1C2 118.17 117.84 117.96 117.33 122.30 The Laplacian of the density (V2r(r))-being the second
C1C2O2 119.06 119.29 119.14 119.39 122.09 derivative of the electron density-indicates where the
C9C3C4 117.95 118.03 118.01 118.13 – density is locally concentrated (V2r(r)!0) and depleted
C3C4O 121.26 121.21 119.14 120.09 – (V2r(r)O0), and hence it can show features such as bonds
C5C6O6 120.00 121.97 119.15 114.94 –
C1C7O8 123.32 123.78 123.50 123.68 –
and lone pairs, which are not observable in r(r) itself.
C1C7O1 119.98 120.01 119.56 121.08 – Ellipticities at BCPs are defined in terms of the ratio of the
C3C9C10 123.05 123.08 123.46 123.24 – curvatures of rb in directions normal to the bond, and
C3C9O3 119.54 119.54 119.60 119.52 – represent the deviations of the bonding density from
C11C5C6 – – 118.00 120.02 – cylindrical symmetry.
C5C11- – – 123.49 117.32 –
C12
Koch and Popelier [46] have utilised Bader’s AIM theory
C5C11- – – 119.51 120.27 – to produce specific criteria to characterise weak inter-
O5 actions, and thus classify hydrogen bonds in particular. The
C2O2H 105.14 105.03 104.84 105.70 – most important of them say that bond critical points should
C4O4H 104.82 104.80 104.92 105.11 – be found between the donor hydrogen atom and the acceptor
C6O6H 108.67 – 104.90 111.45 –
atom of the hydrogen bond, and these two atoms should be
a
Ab initio calculations of the DFT type (see Section 2). linked by a bond path. Additionally, a relationship should
b
Dihedral angles for 3 non-planar C6–C5–C11–O5Z56.378 and C5–C6– exist between rb and overall hydrogen bond energy. A
O6–HZ175.118.
positive Laplacian of the charge density at BCP is typical of
the ionic type of interactions and a negative Laplacian
characterises bonds with covalent character. The calculated
also reveals good general agreement and confirm specific values for V2rb should correlate with the interaction energy.
detail such as RO1–O2ORO3–O4 as seen from a comparison of The other criteria regards mutual penetration of the
data of Tables 2, 3 and 5. hydrogen and acceptor atoms. In a hydrogen bond, the
Some of the details of the calculated structure of hydrogen atom should be more penetrated than the acceptor
compound 3 is slightly different from that reported earlier atom. So when the position of CP is established and
[2]. This can be ascribed to the different schemes used compared with the van der Waals spheres of the donor and
B3PW91 6-31G(d,p) vs. BPW91 6-31G(p) with local acceptors atoms, it appears to be inside of both the spheres.
P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168 165

There are also four additional necessary criteria which are fact that only part of electron correlation is taken into
much more expensive to calculate since they require account by DFT methods. It appears, that there is much less
integration over atomic basins: loss of charge of the of the electron density at the critical points of the C–O bonds
hydrogen atom, destabilisation of the hydrogen atom, than reported by Schiøtt et al. [56] for their transition state
decrease of the dipolar depolarisation of the hydrogen structure of benzoylacetone. A larger penetration by CP of
atom, and decrease in hydrogen atom volume. Properties of the van der Waals sphere of the hydrogen comparing with
different hydrogen bonds were already studied by Espinosa the acceptor oxygen’s sphere is present. Also a small
et al. [47–52] and Mallinson et al. [53–55] who found a accumulation of charge in the donating O–H dipole
number of interesting relations between charge, Laplacian, resulting from the depletion of charge from the space
local kinetic and potential energy densities and the length of between the hydrogen and the acceptor takes place. A
interactions lines. comparison with the C6–O6 bond of 1 not involved in
Our analysis is concentrated around the key parameters, hydrogen bonding shows a larger ellipticity in this case than
those involved in the formation of the pseudo-aromatic ring in the hydrogen bonded ones.
systems. These are the bonds C–O, O–H, H/O C]O, C–C A comparison of two of the other key bonds of the RAHB
and C]C of the RAHB [3,4] intramolecular hydrogen system the OH and the C]O bonds show rather regular
bonds and those involving O/.C interactions in 1 and 2 and features. For the OH bond we find that both the charge
the O/H interaction in 4. A comparison of bond critical density, the Laplacian behave in a parallel fashion (see
Table 1S in Supplementary information) to the hydrogen
point data show some distinct features. It appears that the
bond strength as defined by, e.g. the 2D(OD) isotope effects
calculated data are very consistent with small positive
or the OH chemical shifts. A similar feature is seen for the
Laplacians and ca. 0.5–0.6 e ÅK3 at the critical points of O–
Laplacian of the C]O bond.
H/O hydrogen bonds. This suggests that the O–H/O
hydrogen bonds have an ionic nature with electron
4.3. Separation of twist and compression effects
depletion between the hydrogens and the acceptor oxygen
atoms. Such relatively significant charges at the hydrogen
In Fig. 4(A) the motifs for steric compression are given.
BCPs seem to be overestimated when compared with They are all characterised by six atoms in the interaction
experimental results of charge density studies of the region. As an example the six atoms of 1-formyl-2-naphtol
continuum of weak interactions [54,55] although are C-1, C]O, Hald, H-8, C-8 and C-8a. Contrary to
the hydrogen bonds present in our compounds are sterically twisted compounds, which have seven as seen in
additionally supported by resonance (they are of RAHB 1-acetyl-2-napththone (C-1, C]O, CH3, CH3, H-8; C-8 and
type). Additionally the carbonyl oxygen atoms have close to C-8a (Fig. 4(B)). It should of course be noted that no such
zero Laplacian values which are even positive for those thing as a pure steric twist exist as the effect typically will
oxygens involved in hydrogen bonding and slightly negative include bond bending (see X-ray results for 4).
for those which are not involved. This may result from the An important feature of 1–3 is that the interactions are
mutual and between two intramolecularly hydrogen bonded
A H O
H systems so that the sterically perturbing group is kept in the
H O ring plane. The 1HO chemical shifts and 2DC(OD) isotope
H O
H effects obtained for a series of sterically hindered o-hydroxy
O O H H
R acetyl aromatics in which the carbonyl group is twisted out
H
of the ring plane can roughly be summarised as follows: a
O O
H comparison of 1HO chemical shifts 2- Hydroxyacetophe-
none (12.76 ppm) [18] with that of 6-methyl-2-
H O
H H O H
Two-bond isotope effects

O 1.2
O
1
H 0.8
O H 0.6
H
0.4
B H 0.2
H
O 0
H
HH H O 10 12 14 16 18
H H H
O H OH Chemical Shifts
H O
H Fig. 5. Plot of 2DC(OD) vs. dOH. Data for non-sterically hindered
compounds, :, from Refs. [15,18]. Data for twisted compounds from
Ref. [39] and data for sterically compressed compounds from this paper and
Fig. 4. Motifs leading to steric compression (A) and steric twist (B). Ref. [2].
166 P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168

Hydroxyacetophenone (12.63 ppm) [29] shows that d1H is two-bond isotope effects in model o-hydroxy acyl aromatics
decreasing slightly whereas 2DC(OD) is increasing (0.276 [2]. Steric compression effects on deuterium isotope effects
vs. 0.406 ppm) [29]. A comparison of data (2DC(OD), dOH) have been inferred to explain the larger isotope effects in a-
for 2-hydroxyacetophenone (0.276, 12.76) [18] and 2- substituted vs. non-substituted ketonenamines [10].
hydroxypivalophenone (0.312, 12.68) [29] shows a similar Looking at some unusual results published previously,
trend as also seen comparing data for 2-acetyl-1-hydro- we find that some of these can be explained by steric
xynaphthalene (0.470, 13.99) [18] with those of 4 (0.656, compression effects. The finding that 2DC(OD) isotope
13.44) [29] and 1-propionyl-2-hydroxynaphthalenes (0.648, effect and the d1H chemical shift are both larger for
15.13) [29]. A comparison of 1H and 2DC(OD) data for 1, 2 2-hydroxy-1-naphtaldehyde than for 1-hydroxy-2-
(see Scheme 1), 1,3-diacetyl-2,4-dihydroxybenzene (0.506, napththaldehyde can be assigned to a steric compression
14.79; 0.426,14.26)2 and 3 (0.722,17.09)2 shows that a large effect in the former (which is known to be planar) (Fig. 4).
dOH is followed by a large 2DC(D). A plot of dOH vs. Another unusual example is the symmetrical structure of
2 1,3-diacetyltetrahydropyrane-2,4,6-trione [23]. In this com-
DC(D) is shown in Fig. 5. This includes non-sterically
hindered compounds as well as twisted and sterically pound, despite the double hydrogen bond to a common
compressed ones. It is seen that for the twisted compounds carbonyl oxygen, the 2DC(OD) is unusually large [23]. In
dOH is much smaller than for the sterically compressed this structure steric compression is clearly present.
compounds and the slope is larger.
A second distinctive difference between sterically 4.5. Conformational analysis
twisted and strained compounds lie in large 4DC]O(OD)
isotope effects. This is positive for non-compressed The rotamers of 1 are shown in Fig. 6. The 13C spectrum
compounds as well as for sterically compressed compounds, at ambient temperature excludes the symmetrical C
but clearly negative for twisted compounds [29]. A third structure as the only one, as this would show fewer
way of distinguishing steric strain and steric twist is to resonances. As the 13C chemical shifts of C-2 and C-5
measure DO1H(CD3) isotope effects. In case of steric twist (Table 2) do not change very much with temperature an
the isotope effects are large [29] as found for 4, whereas in equilibrium between A and B tautomers is more likely than
case of steric compression they are negligible as seen for 2 one between A and C. The equilibrium has to be slow
and 3. These indicators should be used together, as an enough on the NMR time scale to cause a broadening of C-4
isolated example could be misleading, e.g. as 4DC-7(OD) in and C-6 resonances, but also fast enough to broaden the OH-
2 which is negative probably due to substituent effects. 4 and OH-6 resonances. A possible explanation for C being
disfavoured is that this structure requires an electron
arrangement with non-alternant double bonds in order to
4.4. Steric effects form two simultaneously strong hydrogen bonds. This form
is similar to that found for 1,3-diacetyl-4,6-dihydroxyben-
Steric compression effects on the O/O distance are seen zene in which no strong hydrogen bonds were formed [2].
in the enolic forms of b-diketones [6]. Larger substituents Furthermore, the steric effects in C are an oxygen perturbing
like t-butyl reduce the O/O distance, as is also found in a methyl group in both situations, whereas in A and B the
indandiones [57,58]. Recently, calculated RO/O distances steric effect are also on the acetyl groups (Fig. 6).
have been correlated to two-bond isotope effects, 2DC(D),

O O 5. Conclusions
H H
HO O O O
H For the benzene derivatives having multiple OH and acyl
O groups (compounds like 1–3) nDC(OD) isotope effects and
the hydrogen bond strength are seen to increase due to an
O O O inductive effect (small) combined with steric compression
H H
A B effects. Steric compression effects lead to an increase of both
2
DC(OD) and dOH, whereas steric twist leads to an increase
O of the former and a decrease of the latter. Isotope effects such
H
as 4DC]O(OD) and nDO1H(CD3) are shown to be useful in
O OH
distinguishing the two types of steric strain. Motifs leading to
the two types of steric effects are pinpointed. The strong
hydrogen bonds found in these types of compounds lead to a
O O decreased electron density of the aromatic ring and to distinct
H
C changes in bond lengths and angles.
The AIM analyses showed that two of key bonds of the
Fig. 6. Rotamers of compound 1. RAHB system the OH and the C]O bonds show rather
P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168 167

regular features. For the OH bond we find that both the [11] N.N. Shapet’ko, Yu.S. Bogachev, L.V. Radushnova, D.N. Shigorin,
charge density, the Laplacian behave in a parallel fashion Dokl. Akad. Nauk SSSR 231 (1976) 409.
[12] D.H. O’Brien, R.D. Stipanovich, J. Org. Chem. 43 (1978) 1105.
(see Table 1S in the Supplementary Information) to the
[13] P.E. Hansen, Prog. NMR Spectrosc. 20 (1988) 207.
hydrogen bond strength as defined by, e.g. the 2D(OD) [14] J. Reuben, J. Am. Chem. Soc. 108 (1986) 1735.
isotope effects or the OH chemical shifts. A similar feature [15] P.E. Hansen, Org. Magn. Reson. 24 (1986) 903.
is seen for the Laplacian of the C]O bond. [16] E. Liepins, M.V. Petrova, E. Gudriniece, J. Paulins, S.L. Kuznetsov,
For sterically twisted compounds like 4 it is not a simple Magn. Reson. Chem. 27 (1989) 907.
turning out of the ring plane of the acetyl group, but a [17] J. Reuben, J. Am. Chem. Soc. 109 (1987) 316.
[18] P.E. Hansen, Magn. Reson. Chem. 31 (1993) 23.
combination of a tilt of the C–C]O bond out of the ring [19] P.E. Hansen, J. Mol. Struct. 321 (1994) 79.
plane and a twist of the C]O group towards the O–H group, [20] S. Ng, H.-H. Lee, G.J. Bennett, Magn. Reson. Chem. 28 (1990) 337.
leaving the latter in the ring plane. [21] P.E. Hansen, in: S. Patai (Ed.), The Chemistry of Double Bonded
Functional Groups, Wiley, New York, 1989, p. 83.
[22] P.E. Hansen, S. Bolvig, F. Duus, M.V. Petrova, R. Kawecki,
P. Krajewski, L. Kozerski, Magn. Reson. Chem. 33 (1995) 621.
Acknowledgements [23] P.E. Hansen, S. Bolvig, T. Kappes, J.C.S. Perkin Trans. 2 1995; 1901.
[24] P.E. Hansen, M. Christoffersen, S. Bolvig, Magn. Reson. Chem. 31
The authors wish to thank Professor E. Grech for help (1993) 893.
with recrystallisation of compound 3 and Anne Lise [25] S.A. Khatipov, N.N. Shapet’ko, Yu.S. Bogavev, Yu.S. Andreichikov,
Gudmundsson for valuable help in the recording of NMR Russ. J. Phys. Chem. 59 (1985) 2097.
[26] P.E. Hansen, S. Bolvig, Magn. Reson. Chem. 35 (1997) 520.
spectra and preparation of some of the compounds. The
[27] J. Bordner, P.D. Hammen, E.B. Whipple, J. Am. Chem. Soc. 11
Carlsberg Foundation is thanked for support to SB and the (1989) 6572.
Danish Natural Science Research Council for support to [28] S. Bolvig, P.E. Hansen, Magn. Reson. Chem. 34 (1996) 467.
purchase NMR instruments and for support of our H-bond [29] P.E. Hansen, S.N. Ibsen, T. Kristensen, S. Bolvig, Magn. Reson.
research. The Polish State KBN Committee is thanked for Chem. 32 (1994) 399.
support to purchase the CCD diffractometer. The X-ray [30] R.F.W. Bader, Atoms in molecules: a quantum theory, Oxford
University Press, Oxford, UK, 1990.
measurements were undertaken in the Crystallographic Unit [31] K.K. Baldridge, J.S. Siegel, J. Am. Chem. Soc. 115 (1993) 10782.
of the Physical Chemistry Lab. at the Chemistry Department [32] A. Filarowski, A. Koll, T. Glowiak, Monatsch. Chem. 130 (1999) 1097.
of the University of Warsaw (Poland). [33] H. Kessler, C. Griesinger, J. Zarbock, H.R. Loosli, J. Magn. Reson. 57
(1984) 331.
[34] E.V. Borisov, W. Zhang, S. Bolvig, P.E. Hansen, Magn. Reson.
Chem. 36 (1998) S104.
Supplementary data [35] G.M. Sheldrick, Acta Crystallogr. A46 (1990) 467.
[36] G.M. Sheldrick, SHELXL93. Program for the Refinement of Crystal
Supplementary data associated with this article can be Structures. University of Göttingen, Germany.
found, in the online version, at doi:10.1016/j.molstruc.2005. [37] A.J.C. Wilson (Ed.), International Tables for Crystallography vol.
03.034 C, Kluwer, Dordrecht, 1992.
[38] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb,
J.R. Cheeseman, V.G. Zakrzewski, J.A. Montgomery, Jr., R.E.
Stratmann, J.C. Burant, S. Dapprich, J.M. Millam, A.D. Daniels, K.N.
Kudin, M.C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R.
References Cammi, B. Mennucci, C. Pomelli, C. Adamo, S. Clifford, J. Ochterski,
G.A. Petersson, P.Y. Ayala, Q. Cui, K. Morokuma, D.K. Malick, A.D.
[1] J. Emsley, Struct. Bonding (Berlin) 57 (1984) 152. Rabuck, K. Raghavachari, J.B. Foresman, J. Cioslowski, J.V. Ortiz,
[2] J. Abildgaard, S. Bolvig, P.E. Hansen, J. Am. Chem. Soc. 120 (1998) A.G. Baboul, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I.
9063. Komaromi, R. Gomperts, R.L. Martin, D.J. Fox, T. Keith, M.A.
[3] P.E. Hansen, R. Kawecki, A. Krowczynski, L. Kozerski, Acta Chem. Al-Laham, C.Y. Peng, A. Nanayakkara, C. Gonzalez, M. Challacombe,
Scand. 44 (1990) 826. P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong, J.L. Andres, C.
[4] G. Gilli, F. Bertulucci, V. Ferretti, V. Bertolasi, J. Am. Chem. Soc. Gonzalez, M. Head-Gordon, E.S.F. GAUSSIAN98, Revision A.7.
111 (1989) 102355. [39] D. Becke, Phys. Rev. A 38 (1988) 3098.
[5] P. Gilli, V. Bertolasi, L. Pretto, V. Ferretti, G. Gilli, J. Am. Chem. Soc. [40] P. Perdew, Y. Wang, Phys. Rev. B 45 (1992) 13244.
126 (2004) 3845. [41] P.L.A. Morphy, in: R.G.A. Bone (Ed.), Popelier with a Contribution,
[6] M.J.T. Robinson, K.M. Rosen, J.D.B. Workman, Tetrahedron 33 UMIST, Manchester, England, 1998.
(1977) 16. [42] F.K. Larsen, Private communication.
[7] H. Lampert, W. Mikenda, A. Karpfen, J. Phys. Chem. 100 (1996) [43] Y. Yaninmoto, H. Kabayashi, S. Nagakura, Y. Saito, Acta Cryst. B29
7418. (1973) 1822.
[8] K.B. Borisenko, C.W. Bock, I. Hargittai, J. Phys. Chem. 100 (1996) [44] M.K. Nirmals Kohila, K.A. Puttaraja, N. Shamala, Acta Cryst. C48
7426. (1992) 1133.
[9] G.L. Hofacker, Y. Marchal, M.A. Ratner, in: P. Schuster, G. Zundel, [45] R.F.W. Bader, J. Phys. Chem. A102 (1998) 7314.
C. Sandorfy (Eds.), The Hydrogen Bond, North-Holland, Amsterdam, [46] U. Koch, P.L. Popelier, J. Phys. Chem. 99 (1995) 9747.
1976, p. 1. [47] E. Espinosa, E. Molins, C. Lecomte, Chem. Phys. Lett. 285 (1998)
[10] D.K. Zheglova, D.G. Genov, S. Bolvig, P.E. Hansen, Acta Chem. 170.
Scand. 51 (1997) 1016. [48] E. Espinosa, C. Lecomte, E. Molins, Chem. Phys. Lett. 300 (1999) 745.
168 P.E. Hansen et al. / Journal of Molecular Structure 749 (2005) 155–168

[49] E. Espinosa, M. Souhassou, H. Lachekar, C. Lecomte, Acta Crystal- [54] P.R. Mallinson, G.T. Smith, C.C. Wilson, E. Grech, K. Wozniak,
logr. B55 (1999) 563. J. Am. Chem. Soc. 125 (2003) 4259.
[50] E. Espinosa, E. Molins, J. Chem. Phys. 113 (2000) 5686. [55] K. Wozniak, P.R. Mallinson, G.T. Smith, C.C. Wilson, E. Grech,
[51] E. Espinosa, I. Alkorta, I. Rozas, J. Elguero, E. Molins, Chem. Phys. J. Phys. Org. Chem. 16 (2003) 764.
Lett. 336 (2001) 457. [56] B. Schoitt, B. Brummerstedt Iversen, G.K. Hellerup Madsen,
[52] E. Espinosa, I. Alkorta, J. Elguero, E. Molins, J. Chem. Phys. 117 T.C. Bruice, J. Am Chem. Soc. 120 (1998) 12117.
(2002) 5529. [57] J.D. Korp, I. Bernal, T.L. Lemke, Acta Cryst. B36 (1980) 428.
[53] P.R. Mallinson, K. Wozniak, G.T. Smith, K. McCormack, J. Am. [58] I. Csöregh, R. Norrestam, Acta Cryst. B32 (1976) 2450.
Chem. Soc. 119 (1997) 11502.

Вам также может понравиться