ArticlePDF Available

Peracetylated α-D-glucopyranosyl fluoride and Peracetylated α-maltosyl fluoride

Authors:

Abstract and Figures

The X-ray analyses of 2,3,4,6-tetra-O-acetyl-alpha-D-glucopyranosyl fluoride, C(14)H(19)FO(9), (I), and the corresponding maltose derivative 2,3,4,6-tetra-O-acetyl-alpha-D-glucopyranosyl-(1-->4)-2,3,6-tri-O-acetyl-alpha-D-glucopyranosyl fluoride, C(26)H(35)FO(17), (II), are reported. These add to the series of published alpha-glycosyl halide structures; those of the peracetylated alpha-glucosyl chloride [James & Hall (1969). Acta Cryst. A25, S196] and bromide [Takai, Watanabe, Hayashi & Watanabe (1976). Bull. Fac. Eng. Hokkaido Univ. 79, 101-109] have been reported already. In our structures, which have been determined at 140 K, the glycopyranosyl ring appears in a regular (4)C(1) chair conformation with all the substituents, except for the anomeric fluoride (which adopts an axial orientation), in equatorial positions. The observed bond lengths are consistent with a strong anomeric effect, viz. the C1-O5 (carbohydrate numbering) bond lengths are 1.381 (2) and 1.381 (3) A in (I) and (II), respectively, both significantly shorter than the C5-O5 bond lengths, viz. 1.448 (2) A in (I) and 1.444 (3) A in (II).
Content may be subject to copyright.
Peracetylated a-D-glucopyranosyl
fluoride and peracetylated a-maltosyl
fluoride
Simone Dedola,
a
David L. Hughes
b
* and Robert A. Field
a
a
Department of Biological Chemistry, John Innes Centre, Colney Lane, Norwich
NR4 7UH, England, and
b
School of Chemistry, University of East Anglia, Norwich
NR4 7TJ, England
Correspondence e-mail: d.l.hughes@uea.ac.uk
Received 23 November 2009
Accepted 29 January 2010
Online 3 February 2010
The X-ray analyses of 2,3,4,6-tetra-O-acetyl--d-glucopyran-
osyl fluoride, C
14
H
19
FO
9
, (I), and the corresponding maltose
derivative 2,3,4,6-tetra-O-acetyl--d-glucopyranosyl-(1!4)-
2,3,6-tri-O-acetyl--d-glucopyranosyl fluoride, C
26
H
35
FO
17
,
(II), are reported. These add to the series of published
-glycosyl halide structures; those of the peracetylated
-glucosyl chloride [James & Hall (1969). Acta Cryst. A25,
S196] and bromide [Takai, Watanabe, Hayashi & Watanabe
(1976). Bull. Fac. Eng. Hokkaido Univ. 79, 101–109] have been
reported already. In our structures, which have been
determined at 140 K, the glycopyranosyl ring appears in a
regular
4
C
1
chair conformation with all the substituents, except
for the anomeric fluoride (which adopts an axial orientation),
in equatorial positions. The observed bond lengths are
consistent with a strong anomeric effect, viz. the C1—O5
(carbohydrate numbering) bond lengths are 1.381 (2) and
1.381 (3) A
˚in (I) and (II), respectively, both significantly
shorter than the C5—O5 bond lengths, viz. 1.448 (2) A
˚in (I)
and 1.444 (3) A
˚in (II).
Comment
Glycosyl fluorides are widely used in carbohydrate chemistry
and biochemistry. The F atom is comparable in size with a
hydroxy group, hence the steric demand upon introduction of
this group is quite small (O’Hagan 2008; Howard et al., 1996).
The popularity of glycosyl fluorides in chemical synthesis is
due to their remarkable stability yet ease of chemospecific
activation in performing glycosylation reactions. One notable
advantage in using glycosyl fluorides as glycosyl donor is their
high thermal stability compared with glycosyl chlorides,
bromides or iodides. The utilization of carbohydrate fluorides
as glycosyl donors originates from the work by Mukaiyama et
al. (1981) on the synthesis of simple glucosides and disac-
charides. Progress made in the utilization of glycosyl fluorides
as donors in the synthesis of O-andC-glycosides has been
reported by Toshima (2000) and updated in the more recent
review by Carmona et al. (2008). Interest in glycosyl fluorides
has increased since Hayashi et al. (1984) developed a reliable
and safe method for the preparation of these compounds by
exposing suitably protected sugars to a 50–70% mixture of
hydrogen fluoride in pyridine. The stability of glycosyl fluor-
ides in their deprotected form also makes them important
compounds for use as mechanistic probes in the elucidation of
enzyme mechanisms and as reagents for enzymatic synthesis
(reviewed by Williams & Withers, 2000). Extending our
interest in the impact of fluorine substitution on carbohydrate
biotransformations (Errey et al., 2009) and the generation of
amylose mimetics (Marmuse et al., 2005; Nepogodiev et al.,
2007; Cle
´et al., 2008), we had cause to investigate glucosyl
fluorides. In this paper, we report the crystal structures of the
2,3,4,6-tetra-O-acetyl--d-glucopyranosyl fluoride, (I), and the
corresponding maltose derivative 2,3,4,6-tetra-O-acetyl--d-
glucopyranosyl-(1!4)-2,3,6-tri-O-acetyl--d-glucopyranosyl
fluoride, (II). The crystal structures obtained integrate with
the published series of -glycosyl halide derivatives; X-ray
structures of peracetylated -glucosyl chloride (James & Hall,
1969) and bromide (Takai et al., 1976) have been reported
previously and the members of this series show most clearly
the anomeric effect, where the preference for the axial
orientation of the halogen atom renders synthesis of the
equatorial counterpart a synthetic challenge. Results from
X-ray analyses typically allow direct evaluation of the impact
of the anomeric effect on sugar structure.
The glucosyl unit in (I) (Fig. 1) adopts a
4
C
1
chair confor-
mation. All bond lengths and angles conform with the values
found in acetylated glucose. Values for the bond lengths which
are affected by the anomeric effect, together with results from
the X-ray crystal structures of other acetylated glucosyl
halides, are summarized in Table 1. The conformational
properties of pyranosyl halides have been explored by a
number of theoretical studies using model compounds such as
2-fluorotetrahydropyran or 2-chlorotetrahydropyran. The
theoretical approaches to generate three-dimensional struc-
tures rely on experimental data to generate the necessary set
of parameters. In this context, good agreement was obtained
by Tvaroska & Carver (1994) by comparison of their theore-
tical results with experimental ones obtained for the acetyl
and benzoyl d-xylopyranose fluorides. To our knowledge, no
crystal structure of anomeric aldohexosyl fluorides has been
reported to date. The structural data reported herein are in
agreement with the theoretical data obtained by Tvaroska &
Carver (1994), supporting the theoretical methodology
reported in their study.
Influences on the bond lengths in a series of X-ray crystal
structures of glycopyranosides have been examined by Briggs
organic compounds
o124 #2010 International Union of Crystallography doi:10.1107/S0108270110003641 Acta Cryst. (2010). C66, o124–o127
Acta Crystallographica Section C
Crystal Structure
Communications
ISSN 0108-2701
et al. (1984). They concluded that there is no correlation
between the electronegativity of the substituent at the
anomeric position and the C5—O5 bond length. Comparison
of C5—O5 bond lengths in the series of halo-derivatives given
in Table 1 shows a similar lack of correlation. The C1—O5
bonds in the fluoro- and chloroglucosides have similar values
[1.381 (2) A
˚in the fluoride, (I), 1.383 A
˚in the chloride and
1.381 (3) A
˚in the maltosyl fluoride, (II)]; the same bond is
shorter in the glucosyl bromide (1.346 A
˚). Comparing the
sugar-ring bond lengths in these halides with those in penta-
acetyl--d-glucopyranose (Jones et al., 1982), it seems that the
shortening of the C1—O5 bond is accompanied by a propor-
tional lengthening of the C1—C2 and C3—C4 bonds. In
contrast, the C2—C3, C4—C5 and C5—O5 bond lengths
change little, with no apparent correlation with the C1—O5
bond lengths.
In the maltosyl fluoride structure, (II), both pyranose rings
adopt a
4
C
1
chair conformation (Fig. 2). It is interesting to
observe in (II) the orientation of the contiguous pyranose
rings, which is described by the torsion angles around the
glycosidic bonds, C4—O4 and O4—C41, denoted as confor-
mational angles and [in (II), = H4—C4—O4—C41 =
29and = C4—O4—C41—H41 = 32], and by the
valence angle = C4—O4—C41, which is 116.66 (14)in (II).
All these values are in good agreement with those in
-maltoseoctaacetate (Brisse et al., 1982) and -octapropanoate
(Johnson et al., 2007) and conform closely with those in other
maltose derivatives discussed by Johnson et al. (2007) in
respect of having short chains containing an -(1!4) inter-
sugar glycosidic linkage, and are therefore useful as models to
study starch structure. The twist of the nonreducing sugar ring
is defined by the virtual torsion angle O44—C44C41—O4;
this has a value of 4.8 (3)in compound (II), which, if
inserted in an amylose chain of starch (see, for example,
Takahashi & Nishikawa, 2003), would add to the bias of
successive residues, forming a left-handed helix (French &
Johnson, 2007).
Intermolecular interactions in crystals of both (I) and (II)
are principally through weak hydrogen bonds. In (I), there are
five contacts (four C—HO and one C—HF) in which the
HF/O distance is less than 2.55 A
˚. In (II), there are six
interactions (five C—HO and one C—HF). In all these
contacts, the angles subtended at the H atoms (in calculated
sites) are greater than 137and most are greater than 150.
Experimental
The title compounds, (I) and (II), were both obtained as single -
anomers (as judged by
1
H NMR spectroscopy). They were prepared
following known procedures (Juennemann et al., 1993), exposing the
peracetylated glucose or maltose to a 70% mixture of hydrogen
fluoride in pyridine in a Teflon bottle. The resulting products were
purified by crystallization from a mixture of ethyl acetate and hexane
(ratio ca 4:1). Crystals suitable for X-ray diffraction were obtained as
colourless blocks in both cases by slow recrystallization from the
same solvent system.
Compound (I)
Crystal data
C
14
H
19
FO
9
M
r
= 350.29
Monoclinic, P21
a= 5.35502 (11) A
˚
b= 7.96182 (14) A
˚
c= 20.1151 (5) A
˚
= 92.061 (2)
V= 857.06 (3) A
˚
3
Z=2
Mo Kradiation
= 0.12 mm
1
T= 140 K
0.55 0.31 0.11 mm
organic compounds
Acta Cryst. (2010). C66, o124–o127 Dedola et al. C
14
H
19
FO
9
and C
26
H
35
FO
17
o125
Figure 2
The molecular structure of the fully acetylated maltosyl fluoride, (II),
showing the atom-numbering scheme. Displacement ellipsoids are drawn
at the 50% probability level and H atoms are shown as white rods.
Figure 1
The molecular structure of the fully acetylated glucosyl fluoride, (I),
showing the atom-numbering scheme. Displacement ellipsoids are drawn
at the 50% probability level and H atoms are shown as white rods. The
methyl groups of three of the acetyl groups were refined as disordered in
two distinct orientations; only one arrangement for each is shown here.
Data collection
Oxford Xcalibur 3 CCD area-
detector diffractometer
Absorption correction: multi-scan
(CrysAlis RED; Oxford
Diffraction, 2008)
T
min
= 0.970, T
max
= 1.033
24426 measured reflections
2677 independent reflections
2325 reflections with I>2(I)
R
int
= 0.037
Refinement
R[F
2
>2(F
2
)] = 0.033
wR(F
2
) = 0.073
S= 1.02
2677 reflections
224 parameters
1 restraint
H-atom parameters constrained
max
= 0.22 e A
˚
3
min
=0.14 e A
˚
3
Compound (II)
Crystal data
C
26
H
35
FO
17
M
r
= 638.54
Monoclinic, P21
a= 5.63832 (9) A
˚
b= 18.2908 (3) A
˚
c= 14.8144 (2) A
˚
= 94.4966 (15)
V= 1523.09 (4) A
˚
3
Z=2
Mo Kradiation
= 0.12 mm
1
T= 140 K
0.42 0.37 0.14 mm
Data collection
Oxford Xcalibur 3 CCD area-
detector diffractometer
Absorption correction: multi-scan
(CrysAlis RED; Oxford
Diffraction, 2008)
T
min
= 0.923, T
max
= 1.070
40401 measured reflections
4545 independent reflections
3785 reflections with I>2(I)
R
int
= 0.052
Refinement
R[F
2
>2(F
2
)] = 0.048
wR(F
2
) = 0.117
S= 1.07
4545 reflections
404 parameters
1 restraint
H-atom parameters constrained
max
= 0.70 e A
˚
3
min
=0.46 e A
˚
3
Since the anomalous scattering does not allow definitive deter-
mination of the absolute configurations in either of these compounds,
the intensities of Friedel pairs were merged (using the MERG 3
command in SHELXL97; Sheldrick, 2008). The configurations were
already established since these compounds were prepared from -d-
glucose and -d-maltose.
All H atoms were included in idealized positions, with C—H =
0.96–0.98 A
˚and U
iso
(H)=1.5U
eq
(C) for methyl groups or 1.2U
eq
(C)
otherwise. The methyl groups were refined as rigid groups rotating
about the C—Me bond. In compound (I), three of the methyl groups
showed disorder over alternative orientations, all of which were
included as idealized methyl groups with two positions rotated by 60
from each other. These were allowed to rotate about the C—Me
bond, and the site-occupation factors of the two orientations refined
to 0.25 (3):0.75 (3), 0.39 (2):0.61 (2) and 0.22 (2):0.78 (2) for the H
atoms at C22, C42 and C62, respectively.
For both compounds, data collection: CrysAlis CCD (Oxford
Diffraction, 2008); cell refinement: CrysAlis RED (Oxford Diffrac-
tion, 2008); data reduction: CrysAlis RED; program(s) used to solve
structure: SHELXS97 (Sheldrick, 2008); program(s) used to refine
structure: SHELXL97 (Sheldrick, 2008); molecular graphics:
ORTEPII (Johnson, 1976) and ORTEP-3 (Farrugia, 1997); software
used to prepare material for publication: SHELXL97.
This study was supported by the BBSRC.
Supplementary data for this paper are available from the IUCr electronic
archives (Reference: JZ3170). Services for accessing these data are
described at the back of the journal.
References
Briggs, A. J., Glenn, R., Jones, P. G., Kirby, A. J. & Ramaswamy, P. (1984).
J. Am. Chem. Soc. 106, 6200–6206.
Brisse, F., Marchessault, R. H., Pe
´rez, S. & Zugenmaier, P. (1982). J. Am.
Chem. Soc. 104, 7470–7476.
Carmona, A. T., Moreno-Vargas, A. J. & Robina, I. (2008). Curr. Org. Synth. 5,
33–60.
Cle
´, C., Gunning, A. P., Syson, K., Bowater, L., Field, R. A. & Bornemann, S.
(2008). J. Am. Chem. Soc. 130, 15234–15235.
Errey, J. C., Mann, M. C., Fairhurst, S. A., Hill, L., McNeil, M. R., Naismith,
J. H., Percy, J. M., Whitfield, C. & Field, R. A. (2009). Org. Biomol. Chem. 7,
1009–1016.
Farrugia, L. J. (1997). J. Appl. Cryst. 30, 565.
French, A. D. & Johnson, G. P. (2007). Carbohydr. Res. 342, 1223–1237.
Hayashi, M., Hashimoto, S. & Noyori, R. (1984). Chem. Lett. 10, 1747–1750.
Howard, J. A. K., Hoy, V. J., O’Hagan, D. & Smith, G. T. (1996). Tetrahedron,
52, 12613–12622.
organic compounds
o126 Dedola et al. C
14
H
19
FO
9
and C
26
H
35
FO
17
Acta Cryst. (2010). C66, o124–o127
Table 1
Selected bond lengths (A
˚), including those affected by the anomeric effect, in glycosyl halide derivatives and pentaacetyl--d-gluopyranose.
.
XRC5—O5 O5—C1 C1—XC1—C2 C2—C3 C3—C4 C4—C5
Br
a
Ac 1.458 (14) 1.347 (15) 2.002
f
1.572 (16)
g
1.531 (16)
g
1.600 (16)
g
1.500 (16)
g
Cl
b
Ac 1.445
f
1.383
f
1.777
f
F
c
Ac 1.4477 (18) 1.381 (2) 1.3981 (19) 1.514 (3) 1.512 (2) 1.515 (2) 1.516 (2)
F
d
(Ac)
4
Glc 1.444 (3) 1.381 (3) 1.393 (3) 1.513 (4) 1.527 (4) 1.532 (3) 1.529 (4)
OGly
d
(Ac)
4
Glc 1.433 (3) 1.420 (3) 1.409 (3) 1.514 (4) 1.517 (4) 1.514 (3) 1.532 (4)
OAc
e
Ac 1.422 (4) 1.403 (4) 1.431 (4) 1.507 (4) 1.524 (4) 1.504 (4) 1.534 (4)
Notes: (a) Takai et al. (1976); (b) James & Hall (1969); (c) this work, compound (I); (d) this work, compound (II); (e) Jones et al. (1982); (f) s.u. values are not available; (g) s.u. values are
taken from a mean value.
James, M. & Hall, D. (1969). Acta Cryst. A25, S196.
Johnson, C. K. (1976). ORTEPII. Report ORNL-5138. Oak Ridge National
Laboratory, Tennessee, USA.
Johnson, G. P., Stevens, E. D. & French, A. D. (2007). Carbohydr. Res. 342,
1210–1222.
Jones, P. G., Sheldrick, G. M., Kirby, A. J. & Glenn, R. (1982). Z. Kristallogr.
161, 237–243.
Juennemann, J., Thiem, J. & Pedersen, C. (1993). Carbohydr. Res. 249,
91–94.
Marmuse, L., Nepogodiev, S. A. & Field, R. A. (2005). Org. Biomol. Chem. 3,
2225–2227.
Mukaiyama, T., Murai, Y. & Shoda, S. I. (1981). Chem. Lett. 3, 431–432.
Nepogodiev, S. A., Dedola, S., Marmuse, L., de Oliveira, M. T. & Field, R. A.
(2007). Carbohydr. Res. 342, 529–540.
O’Hagan, D. (2008). Chem. Soc. Rev. 37, 308–319.
Oxford Diffraction (2008). CrysAlis CCD and CrysAlis RED. Versions
1.171.32.24. Oxford Diffraction, Yarnton, Oxfordshire, England.
Sheldrick, G. M. (2008). Acta Cryst. A64, 112–122.
Takahashi, Y. & Nishikawa, S. (2003). Macromolecules,36, 8656–8661.
Takai, M., Watanabe, H., Hayashi, J. & Watanabe, S. (1976). Bull. Fac. Eng.
Hokkaido Univ. 79, 101–109.
Toshima, K. (2000). Carbohydr. Res. 327, 15–26.
Tvaroska, I. & Carver, J. C. (1994). J. Phys. Chem. 98, 6452–6458.
Williams, S. J. & Withers, S. G. (2000). Carbohydr. Res. 327, 27–46.
organic compounds
Acta Cryst. (2010). C66, o124–o127 Dedola et al. C
14
H
19
FO
9
and C
26
H
35
FO
17
o127
ResearchGate has not been able to resolve any citations for this publication.
Article
Glycosyl fluorides have considerable importance as substrates and inhibitors in enzymatic reactions. Their good combination of stability and reactivity has enabled their use as glycosyl donors with a variety of carbohydrate processing enzymes. Moreover, the installation of fluorine elsewhere on the carbohydrate scaffold commonly modifies the properties of the glycosyl fluoride such that the resultant compounds act as slow substrates or even inhibitors of enzyme action. This review covers the use of glycosyl fluorides as substrates for wild-type and mutant glycosidases and other enzymes that catalyze glycosyl transfer. The use of substituted glycosyl fluorides as inhibitors of enzymes that catalyze glycosyl transfer and as tools for investigation of their mechanism is discussed, including the labeling of active site residues. Synthetic applications in which glycosyl fluorides are used as glycosyl donors in enzymatic transglycosylation reactions for the synthesis of oligo- and polysaccharides are then covered, including the use of mutant glycosidases, the so-called glycosynthases, which are able to catalyze the formation of glycosides without competing hydrolysis. Finally, a short overview of the use of glycosyl fluorides as substrates and inhibitors of phosphorylases and phosphoglucomutase is given.
Article
Glucosides and disaccharides are prepared in good yields from 2,3,4,6-tetra-O-benzyl-β-D-glucopyranosyl fluoride and hydroxy compounds in the presence of stannous chloride and silver perchlorate. In most cases, α-glucosides are predominantly (α⁄β=80⁄20∼92⁄8) obtained.
Article
The crystallographic problem: The production, and the visibility in the published literature, of thermal ellipsoid plots for small-molecule crystallographic studies remains an important method for assessing the quality of reported results. Since the mid 1960s, the program ORTEP (Johnson, 1965) has been perhaps the most popular computer program for generating thermal ellipsoid drawings for publication. The recently released update of ORTEP-III (Johnson & Burnett, 1996) has some additional features over the earlier versions, but still relies on fixed-format input files. Many users will find this very inconvenient, and will prefer to obtain drawings directly from their crystallographic coordinate files. This new version of ORTEP-3 for Windows provides all the facilities of ORTEP-III, but with a modern Graphical User Interface (GUI). Method of solution: A Microsoft-Windows GUI has been added to ORTEP-III. All the facilities of ORTEP-III are retained, and a number of extra features have been added. The GUI is effectively an editor that writes ORTEP-III input files, but the user need not have any knowledge of the inner workings of ORTEP. The main features of this program are: (i) ORTEP-3 for Windows can directly read many of the common crystallographic ASCII file formats. Currently supported formats are SHELX (Sheldrick, 1993), GX (Mallinson & Muir, 1985), GIF (Hall, Allen & Brown, 1991), SPF (Spek, 1990), CRYSTALS (Watkin, Prout, Carruthers & Betteridge, 1996), CSD-XR and CSD-FDAT. In addition, ORTEP-3 for Windows will accept any legal ORTEP-III instruction file. (ii) Covalent radii for the first 94 elements are stored internally, and may be modified by the user. All bonds are calculated automatically, and any individual bonds may be selected for removal, or for a special representation. (iii) The graphical representations of thermal ellipsoids for any element or selected sets of atoms can be individually set. All the possible graphical representations of thermal ellipsoids in ORTEP-III are also available in ORTEP-3 for Windows. (iv) A mouse labelling routine is provided by the GUI. Any number of selected atoms may be labelled, and any available Windows font may be used for the labels. The font attributes, e.g. italic, bold, colour, point size etc. can also be selected via a standard Windows dialog box. (v) As well as HPGL and PostScript Graphics graphic metafiles, it is also possible to get high quality graphics output by printing directly to an attached printer. The screen display may be saved as BMP or PCX format metafiles, and may also be copied to the clip-board for subsequent use by other Windows programs, e.g. word processing or graphics processing programs. Colour is available for all these output modes. (vi) A simple text editor is provided, so that input files may be modified without leaving the program. (vii) Symmetry expansion of the asymmetric unit to give complete connected fragments may be carried out automatically. (viii) Unit-cell packing diagrams are produced automatically. (ix) A number of options are provided to control the view direction. The molecular view may be rotated or translated by button commands from the tool bar, and views normal to crystallographic planes may also be obtained. Software environment and program specification: The program will read several common crystallographic file formats which hold information on the anisotropic displacement parameters. The operation of the program is carried out via standard self-explanatory MS-Windows menu items and dialog boxes. Hard-copy output is either by HPGL or Encapsulated PostScript metafiles, or by directly printing the graphics screen. Hardware environment: The program is implemented for IBM PC compatible computers running MS-Windows versions 3.1x, Windows 95 or Windows NT. At least a 486-66 machine is recommended with 8 Mbytes of RAM, and at least 5 Mbytes of disk space. Documentation and availability: The executable program, together with full documentation, is available free for academic users from http://www.chem. gla.ac.uk/̃louis/ortep3. Although the program is written in Fortran77, a large number of nonstandard FTN77 calls are used to create the GUI. For this reason, the source code is not available.
Article
This study is aimed at evaluating organic fluorine as a hydrogen bonding acceptor. A review of short F…H contacts from all of the organofluorine compounds deposited in the Cambridge Structural Database System (CSDS) was carried out and in parallel a theoretical estimate of the energy of such contacts with inter nuclear distance was executed. A total of 548 structures emerged which contained 1163 unique CF bonds and only 166 of these fluorine atoms posessed short CF…HX contacts of ≤ 2.35Å. Contacts between fluorine and hydrogen bound to carbon (CF…HC) represent the major category of short contacts however these were not judged to be hydrogen bonds as they are weak with energies similar to those of van der Waals complexes. Short contacts between F and the acidic hydrogens of HO or HN are rare in the CSDS with only 12 and 28 occurring respectively. There was only one contact below 2.0Å. Ab initio calculations have evaluated the relative stability and optimum distance of CF…HO bonds between water and fluoromethane and fluoroethene. It emerges that the C(sp3)F fluorine in fluoromethane can enter into stronger hydrogen bonds than C(sp2)F of fluoroethene. The X-ray data reinforces the conclusion that C(sp3)F fluorine is a better hydrogen bond acceptor than C(sp2)F fluorine. The C(sp3)F…HO bond is less than half the strength (2.38 kcal mol−1) of a CO…HO and the C(sp2)F…HO bond (1.48 kcal mol−1) is about half as weak again. Overall however short contacts in the Database which are consistent with an optimal F…H bond are extremely rare.
Article
Treatment of 1-unprotected or -O-acetylated sugar derivatives with a hydrogen fluoride–pyridine mixture affords the corresponding 1-fluoro derivatives in high yields.
Article
The crystal structure of the title compound (1) has been determined [space group P212121, a = 5.592(3), b = 14.724(8), c = 23.966(10) Å, Z = 4] and refined to R0.046 for 2965 unique reflections. The trend towards longer exocyclic bonds at the acetal centre in compounds with strongly electronegative aglycones is maintained in (1).
Article
In this part other procedures of glycosylation reactions by direct activation are presented. We will focus on the n-pentenyl glycoside, the O-alkylation and the trichloroacetimidate methods. The use of glycosyl phosphates and glycals in glycosidation reactions are also discussed. Updated examples of these glycosylation methodologies involving total synthesis of oligosaccharides and related compounds are considered.
Article
Amylose triacetate was synthesized by using enzymatically synthesized amylose. The enzymatically synthesized amylose is a linear 1,4-linked poly-α-d-glucose, which does not contain a 1,6-linkage. X-ray crystal structure analysis was carried out for amylose triacetate I. Two left-handed (14/3) helices pass through a unit cell with parameters a = 10.92 Å, b = 18.91 Å, and c (fiber axis) = 53.91 Å and space group P212121. The molecule does not have 1411 symmetry but only has 21 symmetry in the crystal. The up- and down-pointing molecules statistically coexist at a crystal site with the ratio 0.87:0.13. The conformation of the O6 atom is almost tg, but one gg and one gt conformations are included.
Article
The crystal structure of β-maltose octaacetate, C28H38O19, has been established by direct methods from 3391 independent reflections and refined by a least-squares block-diagonal approximation to a final R value of 0.061. The crystals belong to the orthorhombic system, space group P212121, and have a unit cell of dimensions a = 5.733 (1), b = 23.771 (6), and c = 25.632 (6) Å. The two D-glucose residues have the 4C1 pyranose conformation and are α(1→4) linked. The conformational angles Φ and Ψ at the glycosidic linkage have the values of -29 and -36°, respectively. The acetate substituent at C(6) of the reducing residue is in the gg conformation, but in the nonreducing residue there is a disorder of the C(6) acetate group. Two distinct orientations are observed in equal proportions, one having the gt conformation and the other having the tg conformation. A conformational analysis using the β-maltose acetate geometry reveals that the energy map computed as a function of the rotations about Φ and Ψ angles is not significantly influenced by the presence of the C(6) acetates. A survey of the distribution of the Φ angles in all known α-linked D-glucose residues discloses a binodal distribution; each node corresponds to the occurrence or the nonoccurrence of hydrogen bonding between contiguous residues. A molecular modeling of the behavior of β-maltose octaacetate in solution is proposed and tested against the experimental H(1)-H(4′) distances derived from T1 NMR experiments.
Article
Accurate X-ray crystal structure determinations for 22 axial and equatorial tetrahydropyranyl acetals and alpha - and beta -glucopyranosides reveal systematic changes in the pattern of bond lengths at the acetal center with changing electron demand in the exocyclic ('leaving') group. Stereoelectronic effects on bonding are analyzed and related to reactivity. Linear correlations between the pK//a of the conjugate acid of the leaving group and hence the free energy of activation for cleavage of the acetal C-O bond and the length of the bond being broken appear to be the rule rather than the exception over the range of leaving group studied.