ArticlePDF Available

Difficulties to Determine the Absolute Configuration of Guaiaretic Acid

SAGE Publications Inc
Natural Product Communications
Authors:

Abstract

An account of the difficulties to determine the absolute configuration (AC) of guaiaretic acid (1a), using contemporary methodology, is described in commemoration of the century of its structure elucidation. In fact, the herein studied molecule was the derived diacetate 1b, since the natural lignan slowly decomposes upon manipulation. Single crystal X-ray diffraction of 1b demonstrated the structure, but calculation of Flack and Hooft parameters to know the AC was unsuccessful since the crystals were triclinic, P-1, which is a centro-symmetric space group. In turn, manual band to band comparison of experimental and DFT B3LYP/DGDZVP calculated VCD spectra of 1b allowed ascertaining its AC, although automatic comparison using the CompareVOA software was not very successful. This behavior is associated to the fact that the studied molecule has a sole stereogenic center located on the acyclic portion of a carbon chain possessing two quite similar substituents. The behavior is discussed in relation to cases where the molecular flexibility also generates a very large number of conformers.
Difficulties to Determine the Absolute Configuration of Guaiaretic
Acid
Alfredo R. Ortegaa, Eleuterio Burgueño-Tapiab and Pedro Joseph-Nathanc,*
aInstituto de Química, Universidad Nacional Autónoma de México, Circuito Exterior, Ciudad Universitaria,
Mexico City, 04510 Mexico
bDepartamento de Química Orgánica, Escuela Nacional de Ciencias Biológicas, Instituto Politécnico Nacional,
Prolongación de Carpio y Plan de Ayala, Col. Santo Tomás, Mexico City, 11340 Mexico
cDepartamento de Química, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional,
Apartado 14-740, Mexico City, 07000 Mexico
pjoseph@nathan.cinvestav.mx
Received: May 21st, 2018; Accepted: June 26th, 2018
An account of the difficulties to determine the absolute configuration (AC) of guaiaretic acid (1a), using contemporary methodology, is described in
commemoration of the century of its structure elucidation. In fact, the herein studied molecule was the derived diacetate 1b, since the natural lignan slowly
decomposes upon manipulation. Single crystal X-ray diffraction of 1b demonstrated the structure, but calculation of Flack and Hooft parameters to know the
AC was unsuccessful since the crystals were triclinic, P-1, which is a centro-symmetric space group. In turn, manual band to band comparison of experimental
and DFT B3LYP/DGDZVP calculated VCD spectra of 1b allowed ascertaining its AC, although automatic comparison using the CompareVOA software was
not very successful. This behavior is associated to the fact that the studied molecule has a sole stereogenic center located on the acyclic portion of a carbon
chain possessing two quite similar substituents. The behavior is discussed in relation to cases where the molecular flexibility also generates a very large
number of conformers.
Keywords: Guaiaretic acid, Absolute configuration, Lignan, Vibrational circular dichroism, X-Ray diffraction.
Guaiaretic acid (1a) is a lignan [1] which, rather than being a
carboxylic acid, derives its acidic properties from two phenol
groups, and is probably the emblematic 8,8’-lignan [2]. It is a
constituent of the resins of Guaiacum officinale and G. sanctum,
which was extensively studied in Germany during the nineteen
century [3,4] probably due to the wide use of the resin from the
former tree, growing in the island of Gonave in current Haiti, for the
treatment of syphilis (morbus gallicus) that Spanish people
confronted in Santo Domingo, current Dominican Republic, in the
sixteenth century [3]. Its structure was elucidated just a century ago
[5] and verified by synthesis of the racemic dimethylether
derivative [6], while its absolute configuration (AC) was originally
inferred from the specific rotation value [7] and later confirmed by
chemical correlation with L-dihydroxyphenylalanine [8].
Nowadays, the structure of guaiaretic acid (1a) is easily evidenced
using the 1H and 13C NMR data summarized in the experimental
section, which were secured by very detailed two-dimensional
gHSQC and gHMBC measurements since the task required the
distinction of several quite closely spaced signals owing to the two
aromatic ring systems. In turn, the E nature of the double bond is
evidenced by a NOESY plot which shows cross-peaks between the
vinyl methyl group signal at δ 1.82 and the multiplet at δ 6.65-6.70
corresponding to the four ortho hydrogen atoms of the two aromatic
rings, and between the vinyl hydrogen atom at δ 6.11 and the
multiplet at δ 2.49-2.53 corresponding to the H-8’ methine signal
overlapped with one of the methylene hydrogen atoms at C-7’. Of
relevance is to note the absence of a cross-peak for the signals
owing to the vinyl hydrogen and the vinyl methyl group, which
would be present in the Z double bond geometry.
Regarding the AC determination of the sole stereogenic center of
1a by comparison of the experimental and calculated VCD spectra,
experience dictates [9,10] that the study of molecules containing
alcohols of phenols is not the best idea due to the tendency of these
functional groups to associate at least to the solvent, thus providing
spectra that are difficult to compare with the calculated ones. An
alternative approach to the AC determination would be the study of
the molecule by single crystal X-ray diffraction to determine the
Flack and Hooft paramenters. This again resulted not quite feasible
since good quality crystals could not be generated.
Consequently, it was decided to attempt the AC determination with
either of the above methods using guaiaretic acid diacetate (1b)
which was easily won by acetylation. The molecule was quite
reluctant to provide suitable crystals for X-ray diffraction analysis,
but after many attempts reasonable crystals were obtained by slow
recrystallization from ethanol. The molecule crystallized in the
triclinic system and the crystal structure (Figure 1) could be solved
to a modest agreement factor R = 7.0%, the pertinent details being
summarized in the Experimental Section. Of course all hydrogen
atoms could be found in difference Fourier synthesis, but the crystal
resulted to have P-1 symmetry, which is a centrosymmetric space
group [11] thereby precluding the determination of Flack [12] or
Hooft [13,14] parameters to ascertain its AC.
Consequent to the X-ray results, the option for the AC
determination of 1b turned out to be a VCD study, a task that could
MeO
RO OR
OMe
1a R = H
1b R = Ac
2
4
6
7
9
8'
1'
3'
5'
9'
NPC Natural Product Communications 2018
Vol. 13
No. 8
981 - 986
982 Natural Product Communications Vol. 13 (8) 2018 Ortega et al.
Figure 1: PLUTO X-ray plot for guaiaretic acid diacetate (1b).
Figure 2: Starting models for the search of the complete conformational space of 1b.
be anticipated as quite laborious since the molecule has a sole
stereogenic center located on a carbon atoms chain, rather than
on a ring. In fact, assembly of a solid Dreiding stereomodel from
Büchi (Flawil, Switzerland) for (R)-1b revealed the molecule has a
significantly high conformational freedom. Thus a molecular model
was built using the Spartan 08 software which allowed preliminary
optimization using molecular mechanics to provide conformer A
shown in Figure 2. In order to cover the complete conformational
space, starting from conformer A, the sigma bonds between the two
aromatic rings were rotated 180 degrees to generate conformers B-
E. These five conformational arrangements, shown in Figure 2,
were used as individual starting points for molecular mechanics
conformational searches using the Monte Carlo protocol as
implemented in the Spartan 08 software package. This rendered 75
conformers starting from model A, 88 conformers starting from B,
73 conformers starting from C, 78 from D, and 87 form E, always
in 10 kcal/mol energy gaps. The five sets of conformers were
individually submitted to single point optimization using DFT at the
B3LYP/6-31G(d) level of theory with the same software. The
outcome of the five sets were combined and after elimination of
duplicate structures provided 73 conformers in a 5 kcal/mol energy
gap. The 16 conformers found in the initial 2.94 kcal/mol were
further optimized at the B3LYP/DGDZVP level of theory using the
Figure 3: Comparison of the experimental IR (b) and VCD (d) spectra of (+)-di-O-
acetylguaiaretic acid with the DFT B3LYP/DGDZVP calculated IR (a) and VCD (c)
spectra for (7E, 8’R)-guaiaretic acid diacetate (1b).
Gaussian 03 suit: The final conformational optimization, IR and
VCD calculations at the same level of theory were also performed
using the Gaussian 03 software and the pertinent thermochemical
parameters, summarized in Table 1, show 15 final conformers in a 3
kcal/mol energy gap, in agreement with the expectation that 1b is a
quite mobile molecule. These 15 conformers were weighted
according their Gibb free energies to generate the final calculated
IR and VCD spectra, which seems a quite secure calculation
methodology to find all conformers significantly populating the
conformational space. Comparison of the calculated and
experimental spectra with the CompareVOA software [15], provided
an anharmonicity factor of 0.971, which was used for further
evaluation of the vibrational data, since the overall comparison was
not very good due to the molecular flexibility which provides broad
IR and VCD absorption bands (Figure 3), thus complicating the
comparison procedures.
Visual inspection of the experimental VCD spectrum of 1b and the
calculated spectrum for the R enantiomer reveals that indeed there is
good phases agreement of the signals, thereby evidencing the R AC
for 1b. To further substantiate this it was decided to perform a peak
to peak evaluation of the experimental and calculated VCD spectra
of 1b. Consequently, the band assignment in the IR and VCD
P
0
400
800
1200
1600
000
9501100125014001550
0
200
400
600
800
1000
‐60
‐40
‐20
0
20
40
60
‐30
‐20
‐10
0
10
20
30
1
(a)
(c)
(b)
(d)
x
[M
1
cm
1
]
Molar absorptivity,
1
cm/
2
3
5
46
7
8
9
11
10
1213
14
15
16 17
18
110
12
1
235
46
78
9
11
10
12
13
14
15
16 17 18
1
23
5
4
6
78
911 14 15
16
17
18
13
23
5
46
8
11
12
13
15
16 17
18
Absolute configuration of guaiaretic acid Natural Product Communications Vol. 13 (8) 2018 983
Figure 4: Comparison of the experimental and selected DFT B3LYP/DGDZVP IR
(top) and VCD frequencies (center), as well as the rotational strengths (bottom) of 1b.
spectra (Figure 3) was made following a described methodology
[16] in which initially the vibrational normal modes are numbered
in the weighted calculated IR spectrum and are then assigned to the
experimental bands according the frequencies and relative
intensities. After that, the same numbers are used to correlate the
experimental and calculated VCD peaks. Wavenumber and
intensities of selected calculated and experimental bands are shown
in Table 2, together with some vibrational modes.
To further support the AC assignment we used a plot [17-19]
generated from a set of experimental and calculated rotational
strengths, is such a way that a fit in a line of slope +1 corresponds to
the correct enantiomer while for the opposite enantiomer the points
in the plot would lie along a line of slope 1. Considering sign and
Table 1: Thermochemical parameters of guaiaretic acid diacetate (1b).
Conformer
E
6-31G
(
d
)
a %
b
E
DGDZVPc %
b
G
DGDZVPd
%e
1 0.00 51.7 0.00 43.1 0.00 25.6
2 0.84 12.5 0.47 19.5 0.04 24.0
3 1.25 6.3 0.91 9.2 0.28 16.1
4 1.42 4.7 0.96 8.6 0.68 8.1
5 2.39 0.9 2.20 1.0 0.91 5.5
6 2.61 0.6 2.35 0.8 1.11 4.0
7 1.50 4.1 1.40 4.0 1.23 3.2
8 2.94 0.4 1.75 2.2 1.25 3.1
9 2.18 1.3 2.11 1.2 1.27 3.0
10 1.69 3.0 1.69 2.5 1.36 2.6
11 1.93 2.0 1.87 1.8 1.60 1.9
12 1.21 6.7 1.44 3.8 1.66 1.6
13 2.94 0.4 2.91 0.3 2.35 0.5
14 2.05 1.6 1.98 1.5 2.40 0.4
15 2.92 0.4 2.94 0.3 2.93 0.2
aRelative to 1 (867685.34 kcal/mol). bCalculated using
E RT In K. cRelative to 1
(867791.07 kcal/mol). dRelative to 1 (867533.33 kcal/mol). eCalculated using
G =
RT In K.
intensity of the VCD absorptions, as a measure of the average
rotational strengths, the
experimental and calculated data were
compared using bands 1, 2, 5, 6, 8, 11-13, and 15-18 for (R)-1b
(Figure 4). In combination, the IR data whith a regression
coefficient of 0.9982, the VCD data with a regression coeficient of
0.9916, and the rotational strengths data with a regression coeficient
of 0.7500 clearly reveal the AC of the sole stereogenic center for
()-1b is 8'R.
The difficulties we experienced for the AC determination of
guaiaretic acid diacetate (1b) are in line with other difficult VCD
studied molecules in which a stereogenic center is found on a
carbon atoms chain. A good deal of difficulty occurred during the
determination of the AC of the sole stereogenic center of a
phloroglucinol derivative [20] possessing an α-methylbutyryl chain,
which was isolated from Achyrocline satureioides. The molecule is
relatively complex, C25H30O7, containing 236 electrons, and 180
vibrational frequency modes that are active in the VCD and IR
spectra. Furthermore, the molecule is highly unsaturated, the sole
stereogenic center is located two atoms away from a benzene ring
on a conformational flexible chain, and the stereogenic center has
two substituents with the minimum size difference (Me and Et
groups) in a molecule that possesses additional conformational
freedom. Similar situations became evident for the AC
determination of thymol derivatives, isolated from Ageratina
cylindrica [21] and A. glabrata [22], also caused by high
conformational flexibility of the studied molecules.
In the case of sapinofuranone C, isolated from Diplodia corticola,
which possesses a secondary hydroxy group on the carbon atoms
chain, the situation could satisfactorily be addressed [23] by
increasing the size of one of the sustituents on the stereogenic
center by preparation of the corresponding p-bromobenzoate, a
strategy which is not possible for the sole stereogenic center of 1b.
A general lack of VCD studies of acyclic monoterpenes like linalol,
citronellol, and citronellal is observed, although the ocimenes from
the essential oils of Artemisia absintium could be studied [24] since
a significant molecular portion is rigid due to the presence of a
conjugated diene. Of relevance is also to note that an AC study of
conformational very flexible citronellal by molecular rotational
resonance, using microwaves in the 2-8 GHz region, was reported
recently [25], also evidencing a very high number of relevant
conformers. Another interesting case of a molecule with high
flexibility is the hydrocarbon 4-ethyl-4-methyloctane [26] which,
despite of showing cryptochirality, could be studied by VCD.
R²= 0.9982
950
1050
1150
1250
1350
1450
1550
950 1050 1150 1250 1350 1450 1550
vCalculated(cm
‐1
)
vExperimental(cm
‐1
)
R²=0.9916
950
1050
1150
1250
1350
1450
1550
950 1050 1150 1250 1350 1450 1550
vCalculated(cm
‐1
)
vExperimental(cm
‐1
)
-60
-40
-20
0
20
40
60
-60 -40 -20 0 20 40 60
exp
cal
R
2
=0.7500
984 Natural Product Communications Vol. 13 (8) 2018 Ortega et al.
Table 2: Wavenumber (v), intensities (
and
), and vibrational modes of selected bands of the experimental and calculated IR and VCD spectra of guaiaretic acid diacetate.
Ex
p
erimental Calculated
IR VCD IR VCD
Band va

b
va 
c va

b
va

c Vibrational modesd
1 1506 408.3 1504 9.5 1500 531.4 1501
4.4 (as) A and B rings
2 1462 207.2 1474 1.2 1472 253.6 1470 14.9 (b) Me-10 and Me-10’; (ab) Me-9, Me-9’, CH-8’ and CH2-7’
3 1448 133.8 - - 1443 142.1 - -
4 1414 154.6 1391 1.1 1406 210.0 1408 4.1
5 1389 39.7 1366 6.0 1379 102.4 1382
5.3 (b) Me-9; (ab) CH-8’ and CH2-7’
6 1367 255.0 1353 6.9 1365 289.8 1365 2.9
7 1329 56.1 - - 1312 115.4 - -
8 1304 126.4 1275 24.2 1300 137.5 1302
12.8 (as) A and B rings; (as) C-1–C-7, C-8–C-8’; (ab) CH-7, CH-8’ and
CH2-7’
9 1279 319.8 - - 1280 469.2 - -
10 1263 321.3 - - 1268 379.5 - -
11 1257 213.2 1242 6.9 1255 307.5 1254 9.6 (s) C-4–O-10; (as) B ring; (s) C-4’–O-10’; (as) A ring
12 1217 625.3 1221 38.9 1205 879.2 1200
22.8 (as) O-4–C-11 and C-11–C-12; (as) O-4’–C-11’ and C-11’–C-12’
13 1200 931.0 1200 45.7 1193 1503.6 1193
22.9 (as) C-4–O-4, O-4–C-11, C-11–C-12 and C-1–C-7; (as) C-4’–O-
14 1151 317.2 - - 1147 392.3 - -
15 1121 297.6 1134 8.6 1107 228.8 1118
1.8
16 1090 36.1 1067 12.8 1074 54.0 1076
4.5 (s) C-8’–C-7’; (ab) CH-8’, CH2-7’ and CH3-9’
17 1038 243.7 1042 5.6 1025 171.9 1036
4.9 (s) O-3’–C-10’; (as) B ring; (ab) CH2-7’, CH-8’, and CH3-9; (s) O-
3–C-10; (as) A ring; (ab) CH2-7’, CH-8’, and CH3-9
18 1009 108.1 1005 9.2 990 207.8 1008 14.3 (as) O-4’–C-11’, C-11’–C-12’; (b) CH3-12’; (as) O-4–C-11, C-11–
C-12; (b) CH3-12
aIn cm1. bIn M1cm1.cIn M1cm1(103). d(s) stretching, (as) asymmetric strectching. (ss) symmetric strectching. (b) bending. (ab) asymmetric bending. (sb) symmetric bending.
From our results on the study of guaiaretic acid diacetate (1b) by
VCD, and other results just summarized above, it can be concluded
the AC determination of stereogenic centers located on carbon atom
chains, or side chains in natural products, is much more demanding
than for cases where the sterogenic center is located on a cyclic
system, but at the end the VCD methodology turned out to be
capable providing the desired information.
Experimental
General: Optical rotations were measured in EtOH at 25 °C in a
Perkin-Elmer 341 polarimeter. IR and VCD measurements were
made on a BioTools dualPEM ChiralIR FT spectrophotometer.
NMR measurements were performed from CDCl3 solutions
containing TMS at 300 MHz for 1H and 75 MHz for 13C on a
Varian Mercury spectrometer. CC was carried out on Merck silica
gel 60 (230-400 mesh ASTM). TLC was developed on silica gel 60
F254 plates.
Compounds: Guaiaretic acid (1a) was isolated from Guaiacum
sanctum according to known procedures [3], while the diacetate
(1b) was obtained by routine acetylation using acetic anhydride and
sodium acetate [4].
Guaiaretic acid (1a)
White crystals that slowly turn slightly yellow (EtOH).
MP: 85-87°C.
[α]589 78.2, [α]578 82.2, [α]546 95.8, [α]436 190.2 (c 0.1, EtOH).
1H NMR (300 MHz, CDCl3) δ 6.85 (1H, d, J = 8.3 Hz, H-5), 6.82
(1H, d, J = 8.3 Hz, H-5'), 6.70 (1H, dd, J = 8.3, 1.9 Hz, H-6), 6.65
(3H, m, H-2, H-2' and H-6'), 6.11 (1H, br s, H-7), 5.60 (1H, s, OH-
4'), 5.52 (1H, s, OH-4), 3.85 (3H, s, OMe-3), 3.83 (3H, s, OMe-3),
2.74 (1H, m, H-7a'), 2.53 (1H, m, H-7b'), 2.49 (1H, m, H-8'), 1.81
(3H, br s, Me-9), 1.08 (3H, d, J = 6.6 Hz, Me-9').
13C NMR (75 MHz, CDCl3): δ 146.2 (C-3), 146.0 (C-3'), 143.8 (C-
4'), 143.6 (C-4), 141.0 (C-8), 133.1 (C-1'), 130.9 (C-1), 124.5 (C-7),
121.9 (C-6), 121.8 (C-6'), 114.0 (2C, C-5 and C-5'), 111.7 (C-2),
111.5 (C-5'), 55.9 (MeO), 55.8 (MeO), 45.6 (C-8'), 41.6 (C-7'), 18.9
(C-9'), 15.0 (C-9).
Guaiaretic acid diacetate (1b)
White needles (EtOH).
MP: 82-83°C.
[α]589 66.4, [α]578 69.1, [α]546 80.3, [α]436 156.5, [α]365 306.7
(c 0.66, EtOH).
1H NMR (300 MHz, CDCl3) δ 6.95 (1H, d, J = 8.2 Hz, H-5), 6.92
(1H, d, J = 8.4 Hz, H-5'), 6.75 (1H, m, H-6), 6.74 (1H, m, H-2'),
6.73 (1H, m, H-6'), 6.72 (1H, m, H-2), 6.14 (1H, br s, H-7), 3.80
(3H, s, OMe-3), 3.79 (3H, s, OMe-3'), 2.80 (1H, dd, J = 12.8, 6.6
Hz, H-7a'), 2.59 (1H, m, H-7b'), 2.53 (1H, m, H-8'), 2.31 (3H, s,
Ac), 2.30 (3H, s, Ac), 1.83 (3H, d, J = 1.4 Hz, Me-9), 1.11 (3H, d,
J = 6.6 Hz, Me-9').
13C NMR (75 MHz, CDCl3): δ 169.2 (2C, 2 Ac CO),150.6 (C-3'),
150.5 (C-3), 142.5 (C-8), 140.0 (C-1'), 137.8 (2C, C-4 and C-4'),
137.4 (C-1), 124.6 (C-7), 122.2 (2C, C-5 and C-5'), 121.3 (C-6'),
121.2 (C-6), 113.2 (C-2'), 113.0 (C-2), 55.8 (2C, 2MeO), 45.4
(C-8'), 41.8 (C-7'), 20.7 (2C, 2 Ac Me), 19.0 (C-9'), 15.0 (C-9).
Single crystal X-Ray diffraction analysis of guaiaretic acid
diacetate (1b): A crystal of the title compound, C24H28O6, M =
412.46 was mounted on an Agilent Xcalibur Atlas Gemini
diffractometer using the enhance Cu Kα X-ray source radiation (λ =
1.54184 Å) at 293(2) K in the scan mode. Unit cell refinements
using 1600 machine detected reflections were done with the
CrysAlisPro, Agilent Technologies, Version 1.171.34.49 software.
The crystal was triclinic, space group P-1, a = 10.3460(6) Å, b =
10.3942(6) Å, c = 11.7217(7) Å,
= 71.992(3) deg,
= 89.300(3)
deg,
= 65.172(3) deg, V = 1077.8(1) Å3, Z = 2,
= 1.27 mg/mm3,
μ = 0.742 mm-1, total reflections = 51529, unique reflections 3132
(Rint 0.01%), observed reflections 2479. The structure was solved by
direct methods using the SHELXS-97. For the structural refinement,
the non-hydrogen atoms were treated anisotropically, and the
Absolute configuration of guaiaretic acid Natural Product Communications Vol. 13 (8) 2018 985
hydrogen atoms, included in the structure factor calculation, were
refined isotropically. The final R indices were [I > 2(I)] R1 = 7.0%
and wR2 = 14.8%. Largest difference peak and hole, 0.59 and 0.32
e/Å3. Crystallographic data (excluding structure factors) have been
deposited at the Cambridge Crystallographic Data Centre under
CCDC deposition number 1861855. Copies of the data can be
obtained free of charge on application to the CCDC, 12 Union
Road, Cambridge CB2 IEZ, UK. Fax: +44-(0)1223-336033 or e-
mail: deposit@ccdc.cam.ac.uk.
VCD measurement: A sample of 6.6 mg of 1b was dissolved in
150 μL of 100% atom-D CDCl3 and placed in a BaF2 cell with a
path length of 100 μm. VCD data were acquired at a resolution of 4
cm-1 for 6 h, and the stability of the sample was verified by 1H
NMR measurements at 300 MHz immediately prior and after the
VCD determination.
VCD calculations: The conformational search for 1b was started
generating a molecular model of the R enantiomer, in the
arrangement shown in conformer A of Figure 2, using the
Spartan’08 software (Wavefunction, Inc., Irvine, CA 92612, USA).
After a preliminary optimization using the same software and
molecular mechanics at the MMFF94 level of theory, the C-2–C-
1/C-7–C-8, C-7–C-8/C-8’–C-2’, and C-8’–C-2’/C-1’–C-2’ bonds
had dihedral angles of 109, –58, and 105 degrees, respectively;
while Me-9’ and Me-9 had a dihedral angle of –115 degrees. Thus,
to cover the entire conformational space, starting from initial
conformation A, the C-1–C-7, C-8–C-8’, C-8’–C-7’, and C7’–C1’
bonds were rotated 180 degrees to generate conformations B-E,
respectively (Figure 2). Conformers A-E were then subjected to
conformational searches using the Monte Carlo protocol and the
MMFF94 force field to afford 75, 88, 73, 78, and 87 conformers,
respectively, in 10 kcal/mol energy windows. Each set of
conformers was then submitted to single point energy analysis using
DFT and the B3LYP/6-31G(d) functional and basis set. All
conformers were combined in a unique set, arranged according to
their energy values, and their conformations were evaluated. After
eliminating equals, 75 conformers remained in a 5 kcal/mol energy
gap. The 16 lower energy conformers which were found in a 3
kcal/mol gap were completely optimized using Gaussian 03W
(Gaussian, Inc., Wallingford, CT 06492, USA) at the DFT and
B3LYP/DGDZVP level of theory and their frequencies were
calculated at the same level of theory. After complete optimization,
an additional conformer was eliminated and the 15 conformers that
remained, having a Gibb free energies within a 2.93 kcal/mol gap
were used to generate the final calculated IR and VCD spectra
considering a Boltzmann distribution based on their G values. The
thermochemical analysis was made at 298 K and 1 atm, the VCD
and IR frequencies were plotted using Lorentzian bandshapes and
bandwidths of 6 cm1. Calculated vibrational frequencies were
obtained from the most stable conformer using the GaussView 4.1
software while DFT calculations required on average 12 h
computational time per conformer when using a PC operating at 3
GHz with 4 Gb RAM.
Acknowledgment - Partial financial support from CONACYT-
Mexico grant number 284194, and SIP-IPN grant 20181801 are
acknowledged.
References
[1] Hearou WM, MacGregor WS. (1955) The naturally occurring lignans. Chemical Reviews, 56, 958-1068.
[2] Gottlieb OR. (1978) Neolignans. In Progress in the Chemistry of Organic Natural Products; Herz W, Grisebach H, Kirby GW., Eds., Springer:
New York, Vol. 35; pp 1-72.
[3] Doebner O, Lücker. (1896) Ueber das Guajakharz. Archiv der Pharmazie, 234, 590-610.
[4] Herzig J, Schiff F. (1897) Studien über die Bestandtheile der Guajakharzes. Monatshefte für Chemie, 18, 714-721.
[5] Schroeter G, Lichtenstadt L, Ireneu D. (1918) Über die Konstitution der Guajacharz-Substanzen. Berichte der Deutschen Chemischen Gessellschaft,
51, 1587-1613.
[6] Haworth RD, Mavin CR, Sheldrick G. (1934) The constituents of Guaiacum resin. Part II.Sybthesis of dl-guaiaretic acid dimethyl ether. Journal of
the Chemical Society, 1423-1429.
[7] Schrecker AW, Hartwell JL. (1956) Components of podophyllin. XX. The absolute configuration of podophyllotoxin and related lignans. Journal
of Organic Chemistry, 21, 381-382.
[8] Schrecker AW, Hartwell JL. (1957) The absolute configuration of lignans. Journal of the American Chemical Society, 79, 3827-3831.
[9] Joseph-Nathan P, Gordillo-Román B. (2015) Vibrational circular dichroism absolute configuration determination of natural products. In Progress in
the Chemistry of Organic Natural Products; Kinghorn AD, Falk H, Kobayashi J., Eds., Springer International Publishing: Switzerland, Vol. 100; pp
311-451.
[10] Burgueño-Tapia E, Joseph-Nathan P. (2015) Vibrational circular dichroism: Recent advances for the assignment of the absolute configuration of
natural products. Natural Product Communications, 10, 1785-1795; (b) Burgueño-Tapia E, Joseph-Nathan P. (2017) Vibrational circular dichroism:
Recent advances for the assignment of the absolute configuration of natural products. Natural Product Communications, 12, 641-651.
[11] Thompson AL, Jenkinson SF, Fleet GWJ. (2017) Some experimental aspects of the absolute configuration determination using single crystal X-ray
diffraction. Tetrahedron: Asymmetry, 28, 1330-1336.
[12] Parsons S, Flack HD, Wagner T. Use of intensity quotients and differences in absolute structure refinement. Acta Crystallographica (2013) B69,
249-259.
[13] Hooft RWW, Straver LH, Spek AL. (2010) Using the t-distribution to improve the absolute structure assignment with likelihood calculations.
Journal of Applied Crystallography, 43, 665-668.
[14] Hooft RWW, Straver LH, Spek AL. (2008) Determination of absolute structure using Bayesian statistics on Bijvoet differences. Journal of Applied
Crystallography, 41, 96-103.
[15] Debie E, Gussem ED, Dukor RK, Herrebout W, Nafie LA, Bultinck P. (2011) A confidence level algorithm for the determination of absolute
configuration using vibrational circular dichroism or Raman optical activity, CHEMPHYSCHEM, 12, 1542-1549.
[16] Nafie LA. (2008) Vibrational circular dicroism: A new tool for the solution-state determination of the structure and absolute configuration of chiral
natural product molecules. Natural Product Communications, 3, 451-466.
[17] Stephens PJ, Pan J-J, Krohn K. (2007) Determination of the absolute configurations of pharmacological natural products via density functional
theory calculations of vibrational circular dichroism: The new cytotoxic iridoid prismatomerin. Journal of Organic Chemistry, 72, 7641-7649.
[18] Stephens PJ, Pan J-J, Devlin FJ, Urbanova M, Hajicek J. (2007) Determination of the absolute configurations of natural products via density
functional theory calculations of vibrational circular dichroism, electronic circular dichroism and optical rotation: The schizozygane alkaloid
schizozygine. Journal of Organic Chemistry, 72, 2508-2524.
[19] Stephens PJ, Pan JJ, Devlin FJ, Krohn K, Kurtan T. (2007) Determination of the absolute configurations of natural products via density functional
theory calculations of vibrational circular dichroism, electronic circular dichroism, and optical rotation: The iridoids plumericin and isoplumericin.
Journal of Organic Chemistry, 72, 3521-3536.
986 Natural Product Communications Vol. 13 (8) 2018 Ortega et al.
[20] Casero C, Machín F, Méndez-Álvarez S, Demo M, Ravelo AG, Pérez-Hernández N, Joseph-Nathan P, Estévez-Braun A. (2015) Structure and
antimicrobial activity of phloroglucinol derivatives from Achyrocline satureioides, Journal of Natural Products, 78, 93-102.
[21] Bustos-Brito C, Sánchez-Castellanos M, Esquivel B, Calderón JS, Calzada F, Yepez-Mulia L, Hernández-Barragán A, Joseph-Nathan P, Cuevas G,
Quijano L. (2014) Structure, absolute configuration, and anti-diarrheal activity of a thymol derivative from Ageratina cylindrica, Journal of Natural
Products, 77, 358-363.
[22] Arreaga-González HM, Pardo-Novoa JC, del Río RE, Rodríguez-García G, Torres-Valencia JM, Manríquez-Torres JJ, Cerda-García-Rojas CM,
Joseph-Nathan P, Gómez-Hurtado MA. (2018) Methodology for the absolute configuration determination of epoxythymols using the constituents of
Ageratina glabrata, Journal of Natural Products, 81, 63-71.
[23] Mazzeo G, Cimmino A, Masi M, Longhi G, Maddau L, Memo M, Evidente A, Abbate S. (2017) Importance and difficulties in the use of
chiroptical methods to assign the absolute configuration of natural products: The case of phytotoxic pyrones and furanones produced by Diplodia
corticola. Journal of Natural Products, 89, 2406-2415.
[24] Julio LF, Burgueño-Tapia E, Díaz CE, Pérez-Hernández N, González-Coloma A, Joseph-Nathan P. (2017) Absolute configuration of the ocimene
monoterpenoids from Artemisia absinthium, Chirality, 29, 716-725.
[25] Domingos SR, Pérez C, Medcraft C, Pinacho P, Schnell M. (2016) Flexibility unleashed in acyclic monoterpenes: conformational space of
citronellal revealed by broadband rotational spectroscopy, Physical Chemistry Chemical Physics 18, 16682-16689.
[26] Fujita T, Obata K, Kuwahara S, Miura N, Nakahashi A, Monde K, Decaturc J, Harada N. (2007) (R)-(+)-[VCD(+)945]-4-Ethyl-4-methyloctane, the
simplest chiral saturated hydrocarbon with a quaternary stereogenic center, Tetrahedron Letters, 48, 4219-4222.
... 16 These low energy modes, that can be characterized as internal rotations of functional groups, rather than as molecular vibrations, could be the main source of the difficulties found herein and in other cases. [17][18][19] Alternatively, applying the same principle used by VISSAT to obtain a set of ISF values that maximizes IR similarity, it is also possible to search for the optimal linear combination that maximizes similarity between observed and calculated VCD spectra. Since enantiomers produce antipode VCD spectra, the searches need to be performed independently for each enantiomer, producing separated similarity values for the (R) and the (S) enantiomer. ...
... It is just of further relevance to mention that they are in line with other natural products cases like perezone 17,18 and guaiaretic acid diacetate. 19 Although a single-crystal XRD study of 11-coumaryloxytremetone (1) has already been reported, 5 which revealed that the molecule provides orthorhombic P2 1 2 1 2 1 non-centro-symmetric crystals, the AC was not verified at that time. Thus, to independently determine the tridimensional structure of the molecule, another XRD study was undertaken since the AC was not tested in the original XRD study. ...
... In certain sense, this is indicative of the ISF that would be required for a perfect fit. This methodology was used for guaiaretic acid diacetate 19 after the failure of classic approaches. ...
Article
Full-text available
Although tremetone [5-acetyl-2-(1-methylvinyl)-2,3-dihydrobenzofuran] has only one stereogenic center, the absolute configuration (AC) determination of its naturally occurring 11-acyloxy derivatives 1 and 2 by vibrational circular dichroism (VCD) turned out to be difficult. Similarity-based comparison of the experimental VCD spectrum of 11-coumaryloxytremetone (1), isolated from Parastrephia quadrangularis, with spectra calculated using popular density functional theory (DFT) levels of theory, provided poor enantiomeric similarity indices ( ESI), even when the p-coumaroyl ester group of 1 was replaced by the acetyl group in 2. In search for a better understanding of these results, IR-guided individual scaling factors (ISFs), recently introduced as part of the Vibrational Spectra Similarity and Analysis Tool (VISSAT) software, were used to correct DFT frequencies, while a VCD-guided conformational analysis was developed to explore conformational preferences. These studies showed that for both molecules 72% of the individual conformations gave ESI values in favor of the ( R) enantiomer. Likewise, when conformer abundances were optimized to produce the best possible similarity for each enantiomer, the obtained ESI values were always larger for the ( R) isomer than for the ( S) isomer. These results point toward the ( R) AC in both compounds and highlight the incorrect conformer abundance prediction by DFT calculations as the potential source of the initial difficulties. In addition, the AC of 1 was independently verified using the Flack and Hooft parameters gained after a single-crystal x-ray diffraction (XRD) study.
... Calculated VCD spectra of guaiaretic acid diacetate [(−)−127] (Figure 34), when compared with the experimental spectrum allowed to complete the task, finally showing the AC is 8'R. 71 The relative and AC of caffeic acid ester derivatives 128-130 ( Figure 34), isolated from Tithonia diversifolia, was established by a combined use of experimental and calculated 13 C NMR chemical shifts, as well as by ECD and VCD spectroscopies. 72 The ( (Figure 35), isolated from Prionosciadium thapsoides, was assigned by VCDEC and verified by VCD. ...
Article
Full-text available
Although demonstrated in 1975, vibrational circular dichroism (VCD) finally started to popularize during this century as a reliable tool to determine the absolute configuration (AC) of organic molecules. This research field continues to be a very dynamic one, in particular for the study of natural products which are a unlimited source of chiral molecules. It therefore turns of interest to summarize the accomplishments published in recent years and to comment on some eventual difficulties that emerged in rare cases to complete the AC determination task. Therefore the aim of this review is to update VCD results for the AC assignment of natural products published from 2015 to 2019, a period in which VCD was reported in some 126 publications involving almost 300 molecules. They are organized according the type of studied metabolite allowing an easily search. The molecules correspond to 28 monoterpenes concerning 17 papers, to 42 sesquiterpenes in 14 papers, to 51 diterpenes in 19 publications, to 5 other terpenoids in three papers, to 48 aromatic molecules in 15 reports, to 20 polyketides in 10 publications, to 27 miscellaneous formulas also in 10 papers, and to 76 nitrogen containing compounds, which include alkaloids and their synthetic analogs, in 38 articles. The landscape of reviewed molecules is quite wide as it goes from simple monoterpenes, like borneol or camphor, to very relevant biological molecules like the alkaloid cocaine or tadalafil samples to distinguish genuine and counterfeit Cialis ® . In addition, 5 natural products and a simple derivative published outside the reviewed period, were used to illustrate some aspects of density functional theory calculations.
Article
Chiroptical spectroscopic measurements serve as routine methods to assign the absolute configuration of chiral compounds and interpret their conformational behavior in solution. One common challenge is the use of strongly hydrogen-bonding solvents, which can significantly bias the conformational ensemble and affect the vibrational circular dichroism (VCD) active bands in solution. One such solvent is dimethyl sulfoxide (DMSO)-an excellent solvent for stubborn compounds-that must be explicitly considered in VCD analysis. Explicit consideration of solvent remains a critical challenge in chiroptical spectroscopy due to the need to explore solute-solvent conformational space and the computational expense in modeling these clusters. Interested in the recent development of the Quantum Cluster Growth (QCG) program by the Grimme lab, we set out to model and interpret previously reported VCD spectra for several molecules using their efficient program. Our purposes are two-fold: (1) to investigate the applicability of the QCG program to the problem of reproducing VCD spectra in DMSO solvent and (2) to identify limitations in using this approach. We find that we can conveniently model and analyze the VCD spectra of investigated molecules in DMSO. However, the final set of conformers used for VCD calculations are functional dependent and different sets of conformers can provide satisfactory quantitative agreement between experimental and predicted VCD spectra. We hope that this study provides guidance for future chiroptical studies in the challenging DMSO solvent.
Article
Evaluation of DFT calculated vibrational parameters for the IR and VCD spectra similarity of perezone (1) and dihydroperezone (2) was undertaken. Conformational sets were obtained using different search engines, and the parameters needed for spectra prediction were obtained using several combinations of commonly employed functionals and basis sets, and then weighted spectra were generated and compared with observed traces to provide infrared similarity (SIR) and enantiomeric similarity index (ESI) values. These values evidenced a poor performance of the evaluated levels of theory that were overcome when using the individual scaling factors approach, providing 16% to 139% increases of the ESI values. The best performing level of theory was the B3LYP/DGDZVP2 with ESI values of 0.722 and 0.792 for 1 and 2. Moreover, a correlation analysis showed that the irregular DFT performance arises from rotational strength deviations, which suggests to discard conformational abundance accuracy as the main source of differences. Furthermore, a similarity guided conformational analysis showed that conformations with high ESI values prefer particular orientations of the CC bonds directly attached to the stereogenic carbon atom, with more distant dihedral angles having less influence. Additionally, folded and extended conformers appear to be equally capable to yield high individual ESI values, although abundances of folded conformers just account for 16% of the total population. Nevertheless, abundance optimization showed that a high ESI similarity value of 0.834, is possible when the population of these conformers is increased to 26%, suggesting that a larger abundance of these conformers might be present in solution.
Article
The use of IR individual scaling factors (ISF) for the correction of DFT-calculated frequencies, and its effect on IR and VCD similarity functions, has been evaluated using (+)-(R)-3-methylcyclopentanone as a probe molecule. Contrary to using a single scaling factor to improve spectra matching, this approach sequentially searches for the optimal scaling factor for each calculated transition using a computational search algorithm to maximize the overlap of the calculated and observed IR spectra expressed as the IR similarity (SIR) function. The obtained ISFs are then applied to the calculated frequencies, which are used to produce a scaled VCD spectrum for comparison with the observed trace, thereby yielding enantiomeric similarity index (ESI) values as a similarity measure. This procedure provides a significant improvement of the SIR and ESI values when compared with the use of a single scale factor, showing 15.1% and 34.1% in average increments, respectively, and values as high as 0.98 and 0.94, respectively. When a set of manually found ISFs is used, most differences in SIR and ESI performance disappear, and nearly perfect spectra matches are found throughout the levels of theory tested. This suggests that the observed differences in computed IR/VCD spectra with commonly used levels of theory are related to differences in frequency rather than to intensity accuracy. Finally, the use of ISFs is expected to enhance the ability to aide stereochemical assignments, particularly in cases where sufficiently accurate frequencies are difficult to obtain due to the system size or complexity.
Article
Full-text available
Vibrational circular dichroism (VCD) emerged during the last decade as a reliable tool for the absolute configuration (AC) determination of organic compounds. The principles, instrumentation, and methodology applied prior to early 2013 were recently reviewed by us. Since VCD is a very dynamic field, the aim of this review is to update VCD advances for the AC assignment of terpenoids, aromatic compounds, alkaloids, and other natural products for the 2013-2014 period, when VCD was applied to the AC assignment of some 70 natural products. In addition, although discovered in 2012, a brief introduction to the VCD exciton coupling approach and its applications in natural products AC assignment is presented.
Article
Full-text available
Vibrational circular dichroism (VCD) has been used in recent years to determine the absolute configuration of a number of natural product chiral molecules. In this brief review, these applications will be described and the methodology of VCD determination of absolute configuration (AC) will be explained. The principal advantages of VCD versus X-ray crystallography for absolute configuration determination are: 1) only solution-state samples are needed and therefore single crystals are not required, 2) high enantiomeric sample purity is not required, 3) high chemical purity is not required as long as impurities are not chiral and 4) solution-state conformations are obtained as an extra feature of the AC determination.
Article
Full-text available
A methodology to determine the enantiomeric excess and the absolute configuration (AC) of natural epoxythymols was developed and tested using five constituents of Ageratina glabrata. The methodology is based on enantiomeric purity determination employing 1,1′-bi-2-naphthol (BINOL) as a chiral solvating agent combined with vibrational circular dichroism (VCD) measurements and calculations. The conformational searching included an extensive Monte Carlo protocol that considered the rotational barriers to cover the whole conformational spaces. (+)-(8S)-10-Benzoyloxy-6-hydroxy-8,9-epoxythymol isobutyrate (1), (+)-(8S)-10-acetoxy-6-methoxy-8,9-epoxythymol isobutyrate (4), and (+)-(8S)-10-benzoyloxy-6-methoxy-8,9-epoxythymol isobutyrate (5) were isolated as enantiomerically pure constituents, while 10-isobutyryloxy-8,9-epoxythymol isobutyrate (2) was obtained as a 75:25 (8S)/(8R) scalemic mixture. In the case of 10-benzoyloxy-8,9-epoxythymol isobutyrate (3), the BINOL methodology revealed a 56:44 scalemic mixture and the VCD measurement was beyond the limit of sensitivity since the enantiomeric excess is only 12%. The racemization process of epoxythymol derivatives was studied using compound 1 and allowed the clarification of some stereochemical aspects of epoxythymol derivatives since their ACs have been scarcely analyzed and a particular behavior in their specific rotations was detected. In more than 30 oxygenated thymol derivatives, including some epoxythymols, the reported specific rotation values fluctuate from −1.6 to +1.4 passing through zero, suggesting the presence of scalemic and close to racemic mixtures, since enantiomerically pure natural constituents showed positive or negative specific rotations greater than 10 units.
Article
Full-text available
Conformational flexibility is intrinsically related to the functionality of biomolecules. Elucidation of the potential energy surface is thus a necessary step towards understanding the mechanisms for molecular recognition such as docking of small organic molecules to larger macromolecular systems. In this work, we use broadband rotational spectroscopy in a molecular jet experiment to unravel the complex conformational space of citronellal. We observe fifteen conformations in the experimental conditions of the molecular jet, the highest number of conformers reported to date for a chiral molecule of this size using microwave spectroscopy. Studies of relative stability using different carrier gases in the supersonic expansion reveal conformational relaxation pathways that strongly favour ground-state structures with globular conformations. This study provides a blueprint of the complex conformational space of an important biosynthetic precursor and gives insights on the relation between its structure and
Book
The volumes of this classic series, now referred to simply as "Zechmeister" after its founder, L. Zechmeister, have appeared under the Springer Imprint ever since the series' inauguration in 1938. The volumes contain contributions on various topics related to the origin, distribution, chemistry, synthesis, biochemistry, function or use of various classes of naturally occurring substances ranging from small molecules to biopolymers. Each contribution is written by a recognized authority in his field and provides a comprehensive and up-to-date review of the topic in question. Addressed to biologists, technologists, and chemists alike, the series can be used by the expert as a source of information and literature citations and by the non-expert as a means of orientation in a rapidly developing discipline.
Article
Students of single crystal X-ray diffraction are often give advice as to how best to collect their data when attempting absolute configuration determination. These 'rules' often have more grounding in gut-feeling than evidence. Thus, in an effort to provide advice and evidence that today's crystallographers can pass onto to tomorrow's young scientists, we present a systematic study of 1-deoxy-l-arabinitol, a straight chain sugar which crystalises well in the space group I41.
Article
α-Pyrones and furanones are metabolites produced by Diplodia corticola, a pathogen of cork oak. Previously, the absolute configuration (AC) of diplopyrone was defined by chiroptical methods and Mosher’s method. Using X-ray and chiroptical methods, the AC of sapinofuranone C was assigned, while that of the (4S,5S)-enantiomer of sapinofuranone B was established by enantioselective total synthesis. Diplofuranone A and diplobifuranylones A–C ACs are still unassigned. Here electronic and vibrational circular dichroism (ECD and VCD) and optical rotatory dispersion (ORD) spectra are reported and compared with density functional theory computations. The AC of the (4S,5S)-enantiomer of sapinofuranone B and sapinofuranone C is checked for completeness. The AC of diplobifuranylones A–C is assigned as (2S,2′S,5′S,6′S), (2S,2′R,5′S,6′R), and (2S,2′S,5′R,6′R), respectively, with the Mosher’s method applied to define the absolute configuration of the carbinol stereogenic carbon. The AC assignment of sapinofuranones is problematic: while diplofuranone A is (4S,9R), sapinofuranones B and C are (4S,5S) according to ORD and VCD, but not to ECD. To eliminate these ambiguities, ECD and VCD spectra of a di-p-bromobenzoate derivative of sapinofuranone C are measured and calculated. For phytotoxicity studies, it is relevant that all six compounds share the S configuration for the stereogenic carbon atom of the lactone moiety.
Article
The absolute configuration (AC) of the naturally occurring ocimenes (−)-(3S,5Z)-2,6-dimethyl-2,3-epoxyocta-5,7-diene (1) and (−)-(3S,5Z)-2,6-dimethylocta-5,7-dien-2,3-diol (2), isolated from the essential oils of domesticated specimens of Artemisia absinthium, followed by vibrational circular dichroism (VCD) studies of 1, as well as from the acetonide 3 and the monoacetate 4, both derived from 2, since secondary alcohols are not the best functional groups to be present during VCD studies in solution due to intermolecular associations. The AC follows from comparison of experimental and calculated VCD spectra that were obtained by Density Functional Theory computation at the B3LYP/DGDZVP level of theory. Careful nuclear magnetic resonance (NMR) measurements were compared with literature values, providing for the first time systematic 1H and 13C chemical shift data. Regarding homonuclear 1H coupling constants, after performing a few irradiation experiments that showed the presence of several small long-range interactions, the complete set of coupling constants for 3, which is representative of the four studied molecules, was determined by iterations using the PERCH software. This procedure even allowed assigning the pro-R and pro-S methyl group signals of the two gem-dimethyl groups present in 3.