Progress in Structural Characterization of Functional Polysaccharides

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 40

1

Progress in Structural Characterization of Functional


Polysaccharides
Kanji Kajiwara
Otsuma Women’s University, Chiyoda-ku, Tokyo, Japan

Takeaki Miyamoto
National Matsue Polytechnic College, Matsue, Japan

I. INTRODUCTION (Fig. 1). In later sections, it will be shown that these basic
conformations of amylose and cellulose are supposed to
Oligosaccharides and polysaccharides are biopolymers be retained to some extent in aqueous solutions. The
commonly found in living organisms, and are known to difference in the structure is reflected by the respective
reveal the physiological functions by forming a specific physiological functions of edible amylose and nonedible
conformation. However, our understanding of polysaccha- cellulose.
ride chains is still in its premature state with respect to their There are some evidences that the higher-order struc-
structure in solid and in solution. Structural analysis may ture of polysaccharide chains is related to their physiolog-
offer the most fundamental knowledge to understand the ical function as exemplified by the triple-stranded helix of
functions of polysaccharides, but the diversity and irregu- scleroglucan, which is known to possess an antitumor
larity of polysaccharide chains make it a formidable task. activity. Many polysaccharide chains are able to assume
Polysaccharide chains are partly organized but are consid- an ordered or quasi-ordered structure such as a double-
ered to be mostly amorphous. No single crystal was made stranded helix, but the ordered structure is interrupted by
from polysaccharides up to now. Thus the crystallographic the irregularity of the primary structure in the polysaccha-
analysis of polysaccharide chains has been performed by ride chains. Many polysaccharide chains form gel in solu-
either using the small oligosaccharide single crystals or the tions by assuming an ordered or quasi- ordered chain
x-ray fiber pattern diffraction from drawn polysaccharide structure, which constitutes a cross-linking domain.
gels. The conformational analysis of polysaccharide chains
Although a monosaccharide unit is common to many involves two aspects: (1) the characterization of a single
polysaccharides, its linkage mode varies and characteristic chain conformation and (2) the analysis of the chain
functions/properties will appear accordingly. A good assembly (suprastructure) of polysaccharides. A single
example is demonstrated by simple poly-D-glucans—wa- chain conformation of polysaccharides is primarily deter-
ter-soluble, digestible amylose and non-water-soluble, mined by the chemical structure specified by the types of
nondigestible cellulose. Both amylose and cellulose are sugar residues, sugar linkages, and side groups. A single
homopolymers composed of glucosidic residues, but they chain conformation accounts, to some extent, for the
differ in the mode of linkage. Amylose is a (1!4)-a-D- formation of suprastructures such as the complexing ca-
linked polyglucan, whereas cellulose is a (1!4)-a-D- pability of amylose and the fringed micelle formation of
linked polyglucan. The (1!4)-a linkage (amylose) and cellulose.
the (1!4)-a linkage (cellulose) of D-glucosidic residues Unlike cellulose and amylose, most polysaccharides
yield a wobbled helix and a stretched zigzag chain, have no regular homopolymeric structure, where the reg-
respectively, by joining the D-glucosidic residues in a ularity is interrupted by the random intrusion of different
simple manner so as to place the chain on a plane [1] types of linkage and/or sugar units. The introduction of

Copyright 2005 by Marcel Dekker


2 Kajiwara and Miyamoto
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 1 Wobbled helical conformation (a) and stretched zigzag conformation (b), representing the basic conformations of
amylose and cellulose, respectively.

such an irregularity hampers crystallization and promotes potential application of the crystallographic approach,
the formation of a suprastructure that is characteristic of which has been a main method of the structural analysis
the polysaccharide species. The interchain interaction of in protein science. Here we will describe two methods that
polysaccharides seems to be specific as exemplified by the are currently applied to the structural and conformational
suprastructure depending on the chemical structure and analysis of oligosaccharides and polysaccharides: small-
counterions (in the case of polysaccharides possessing angle x-ray scattering (SAXS) [2] and nuclear magnetic
carboxyl or sulfate groups). The formation of the supra- resonance (NMR) [3].
structure often results to gelation. Molecular modeling by computer is considered to
The complexity in characterizing polysaccharide chain supplement the analysis by small-angle x-ray scattering
conformation is due to the fact that the interchain interac- and NMR. Although an initial intention of molecular
tion of polysaccharides is so specific that polysaccharide modeling is to predict physical properties of carbohydrates
chains are seldom dispersed in solvent as a single chain. a priori [4], the ab initio calculation is limited to a small
Thus a first task to understand the structure–function monosaccharide and the semiempirical quantum method
relationship of polysaccharides is to evaluate the intrinsic can be applied for the structural characterization of mol-
chain (single chain) characteristics free from interchain ecules up to the size of disaccharides. Molecular mechanics
interaction. Once the intrinsic chain conformation is spec- or molecular dynamics is an alternative method applied to
ified, the interchain interaction can be analyzed in terms of the computer modeling of larger carbohydrate molecules,
the mode of suprastructure composed of several polysac- where the motion of constituent atoms is assumed to be
charide chains. described in terms of classical mechanics.
This review is intended to demonstrate the recent In the final chapter, the structural and conformational
strategy in the structural and conformational characteriza- aspect is discussed from the chemical point of view. Here
tion of oligosaccharides and polysaccharides. Although the controlled chemical modification of cellulose is treated
various techniques are applied for the structural and and the physicochemical characteristics are discussed by
conformational analysis of oligosaccharides and polysac- taking into account the structural change due to chemical
charides, the general inability to crystallize excludes the modification of cellulose.

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 3

II. STRATEGY AND METHODS OF ANALYSIS found in nature. The structural analysis of disaccharide
implies the identification, the linking order, the link posi-
Because many monosaccharides have a single, well-estab- tion, and the link mode of constituent monosaccharides.
lished conformation, the conformational analysis of oligo- Chemical and optical methods are available to determine
saccharides and polysaccharides starts from understanding the structure of disaccharide, but the recent development in
the energetic relationship when the monosaccharide resi- NMR has facilitated the assignment of specific protons or
dues are linked in a specific way. The entire geometry of carbons as well as the conformational determination of the
oligosaccharides and polysaccharide chains can be de- glucosidic linkage as shown in the next section. Fourier-
scribed in terms of a set of the pairs of dihedral angles of transform infrared (FTIR) and laser Raman spectroscopy
rotation about the monosaccharide links. If the rotation is are also useful tools for the characterization of glycosidic
independent at each monosaccharide link, the chain should bonds [5].
assume a random coil conformation. However, the con- X-ray and neutron diffraction can be applied to deter-
formation of saccharide chains is found in most cases to mine the crystal structure and hydrogen bonding of mono-
assume nonrandom conformations due to intra- and inter- saccharides and disaccharides that form a single crystal. A
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

chain interactions that suppress the conformational space classic example will be found in the crystal structure
available for the chains linked by independent rotation. analysis of h-maltose monohydrate [6] (Fig. 4), where the
Even the crystal structure is partly retained in solution as in earlier structure determination using x-ray diffraction [7]
the case of protein. Thus a single chain conformation may was refined to give a more accurate description of the
account, to some extent, for the mode of interactions and hydrogen bond structure. The x-ray diffraction analysis
the formation of suprastructures. provides the most explicit information on structure in
This section gives a brief introduction on the structure terms of the precise atomic coordinates. The Cambridge
of monosaccharides and disaccharides as the basis of the Structural Database lists the crystal structure of about 40
structural and conformational analysis of oligosaccharides small oligosaccharides (cyclodextrins are omitted) includ-
and polysaccharides. The fundamentals of SAXS and ing about 10 trisaccharides, 2 tetrasaccharides, and 1
NMR together with the molecular modeling are also hexasaccharide. Here a number of crystal structures of
described. mono-, di-, and trisaccharides were determined from the
acetate derivatives because the acetylated derivatives are
A. Structure of Monosaccharide found to crystallize more easily than original (untreated)
and Disaccharide oligosaccharides. (1!3)-h-D-glucopyranosyl residues con-
sist of a main chain of a medically important class of
A monosaccharide is given by the chemical formula polysaccharides including curdlan, lentinan, schizophyl-
CnH2nOn, where n = 3–10. Pentose (n = 5) and hexose lan, scleroglucan, and grifolan, which possess branches at
(n = 6) are the most abundant in nature, and are C6 (except for curdlan). Glcph 1!3 Glc disaccharides are
composed of a pyranose or a furanose (Fig. 2) as a basic systematically synthesized and the crystal structures are
ring structure. A pyranose ring has two stable chair form determined. A first attempt was made by Takeda et al. [8]
(C) conformers C1 and 1C where four atoms of O, C2, on 3-O-h-D-glucopyranosyl-h-D-glucopyranose (h-D-lam-
C3, and C5 are on the same plane. Fig. 3 lists some of inarabiose) ethyl hepta-O-acetyl-h-D-laminarabioside [9],
pyranose-type pentose and hexose, which appear in later followed by Perez et al. [10] on octa-O-acetyl-h-D-laminar-
sections, where the abbreviated description is given for abiose), and by Lamba et al. [11] on (methyl hepta-O-
each monosaccharide. acetyl-h-D-laminarabiose). Recently, 3-O-h-D-glucopyra-
A disaccharide is composed of two monosaccharides nosyl-h-D-glucopyranoside (methyl h-D-laminarabioside)
linked by any of the four modes of glycosidic linkages a,aV, [12] and methyl hepta-O-acetyl-h-D-laminarabioside [13]
a,hV, h,aV-, or h,hV-. Table 1 shows some disaccharides were prepared, and the crystal structures were determined
by x-ray diffraction (Fig. 5). Table 2 summarizes two
dihedral angles, / and w, with respect to the glycosidic
bond for (1!3)-h-linked disaccharides, evaluated from the
crystallographic data of laminarabiose and laminarabio-
side derivatives. Here the dihedral angles are taken as / =
h[H(C1) ,C1 ,O1, C3V] and w = h[C1, O1, C3V, H(C3V)].
(See Sec. II.D for the definition of dihedral angles / and w.)
The angle / is almost invariant around 45j regardless of
the substituents, while the angle w is classified in two
groups of around 45j and 8j. When the intramolecular
hydrogen bond is formed between 04V and 05, the angle y
assumes a negative value. The introduction of acetyl
groups prevents the formation of intramolecular hydrogen
bonds as seen from the stereoview of the molecular struc-
tures of methyl h-D-laminarabioside and methyl hepta-O-
Figure 2 Pyranose and furanose. acetyl-h-D-laminarabioside in Fig. 5. The invariance of

Copyright 2005 by Marcel Dekker


4 Kajiwara and Miyamoto
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 3 Pyranose-type pentose and hexose.

the angle 4) is attributed to the exo-anomeric effect that Here the magnitude of the scattering vector is given by (4p /
restricts rotation around the bond between an anomeric k) sin(h / 2) with k and h being the wavelength and the
carbon atom and a glycosidic oxygen atom [14]. scattering angle, respectively. c(r) is a correlation function
representing the average of the product of two electron
B. Fundamentals of Small-Angle X-Ray density fluctuations at a distance r. The distance distribu-
Scattering [2] tion function p(r) is defined as

Small-angle x-ray scattering is characterized by its small pðrÞ ¼ Vr2  cðrÞ ð2Þ
scattering angle. A scattering process obeys a reciprocal
law that relates the distance r in an ordinary (real) space which is characteristic of the shape of the scattering object.
with the scattering vector q in a Fourier (scattering) space The phase difference between scattered rays becomes
by the phase factor defined by exp(q  r); that is, the more prominent as the scattering angle increases. Thus the
scattered intensity I( q) is given by the Fourier transforma- scattered intensity is maximum at zero scattering angle and
tion of the electron density distribution in the object: proportional to the number of electrons in the object where
ðl the scattered rays are all in phase. The scattered intensity
IðqÞ ¼ V ¼ 4pr2 cðrÞ  expðiq rÞdr ð1Þ decreases with increasing scattering angle and diminishes
0 at a scattering angle of the order of k / D, where k and D

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 5

Table 1 Disaccharides in Nature


Mode of linkage Common name Structure Origin
(1!4) Linkage maltose Glcpa 1 ! 4 Glc starch
cellobiose Glcph 1 ! 4 Glc cellulose
lactose Galph 1 ! 4 Glc mammal milk
xylobiose Xylph 1 ! 4 Xyl xylan
chitobiose GlcNh 1 ! 4 GlcN chitin
cellobiouronic acid GlcUAph 1 ! 4 Glc D. pneumoniae
(1!6) Linkage isomaltose Glcpa 1 ! 6 Glc amylopectin, etc.
gentiobiose Glcph 1 ! 6 Glc gentianose
melibiose Galpa 1 ! 6 Glc raffinose
planteobiose Galpa 1 ! 6 Fruf planteose
(1!3) Linkage nigerose Glcpa 1 ! 3 Glc mutan
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

laminaribiose Glcph 1 ! 3 Glc laminaran


turanose Glcpa 1 ! 3 Fruf meleziose (honey)
hyalobiuronic acid GlcUAph 1 ! 3 GlcN hyaluronic acid
chondrosine GlcUAph 1 ! 3 GalN chondroitin sulfate
(1!2) Linkage kojibiose Glcpa 1 ! 2 Glc Aspergillus orryzae
sophorose Glcph 1 ! 2 Glc Sophora japonica
sucrose Frufh 1 ! 2 aGlcp beet sugar
(1!1) Linkage a,a-trehalose Glcpa 1 ! 1 aGlcp yeast
p and f denote pyranose and furanose, respectively.

denote the wavelength of an incident beam and the average (Rt) of a flat particle by describing approximately the
diameter of scattering objects. When x-ray is used as an scattering from the cross-section or the thickness in terms
incident beam (k = 0.154 nm), the limiting scattering angle of the exponential form. The scattering factor of a rod-like
to be observed is approximately equal to 0.450 when D = particle (a cylinder) consists of two components of the
10 nm, or to 0.0450 when D = 100 nm. height and the cross-section as
Because the phase factor exp(q  r) can be replaced by p  
its space average sin qr/qr for the statistically isotropic Pcylinder ðqÞc  exp q2 R2c =2 ð6Þ
2Hq
system according to Debye [15], Eq. (1) can be expanded in
the series of q2 at very small angles by expanding the sine where 2H denotes the height of the cylinder. The scattering
term to yield the particle scattering factor as factor of a flat particle (a disk) is given by the product of
ðl two terms of the cross-sectional area and the thickness as
1
PðqÞuIðqÞ=Ið0Þ ¼ 1  q2 4pr2 cðrÞ  r2 dr=2 2p  
3 0 Pdisk ðqÞc  exp q2 R2t ð7Þ
ð ð3Þ Aq2
l
 4pr2 cðrÞdr þ Oðq4 Þ
0
where A denotes the cross-sectional area. Equations (6) and
(7) suggest that the cross-sectional radius of gyration Rc
where the second term on the right side represents the
radius of gyration RG, that is
ðl ðl
R2G ¼ pðrÞ  r2 dr=2 pðrÞdr ð4Þ
0 0

in terms of the distance distribution function Eq. (2). The


sine expansion of Eq. (3) is approximately closed in the
exponential form, and the particle scattering factor is
reduced to the Guinier approximation [2,16]:

PðqÞcexpðq2 R2G =3Þ ð5Þ

suggesting that the radius of gyration can be evaluated


from the initial slope by plotting ln P( q) against q2 (the
Guinier plot). A similar argument can be applied to
evaluate the radius of gyration corresponding to the Figure 4 Stereoview of h-maltose monohydrate. (From
cross-section of a rod-like particle (Rc) or the thickness Ref. 6.)

Copyright 2005 by Marcel Dekker


6 Kajiwara and Miyamoto
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 5 Stereoview of methyl h-D-laminarabioside (top) and methyl hepta-O-acetyl-a-D-laminarabioside (bottom).

Table 2 Dihedral Angles of the Glycosidic Linkage for Glcph 1!3 Glc Disaccharides

Compound / = h[H(C1), C1, O1, C3V] w = h[C1, O1, C3V, H(C3V)]


Methyl h-D-laminarabioside 43 52
h-D-laminarabiose 28 38
Methyl hepta-O-acetyl-h-D-laminarabioside 43 5
Methyl hepta-O-acetyl-a-D-laminarabioside 43 2
Octa-O-acetyl-h-D-laminarabioside 42 14
Octa-O-acetyl-a-D-laminarabioside 54 11

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 7

and the thickness radius of gyration Rt are evaluated from equivalent to the particle scattering factor of such a mol-
the initial slope of the corresponding Guinier plots: ln ecule that freely moves in space. If a molecule (e.g., a
qPcylinder ( q) plotted against q2 or ln q2Pdisk( q) plotted flexible polymer molecule) has a large internal freedom,
against q2. the distance dij fluctuates with time due to the internal
Although the polymeric chain has an approximate motion of such a molecule. In this case, the particle
shape as represented by a sphere or an ellipsoid as a whole scattering factor should be calculated as an average over
in solution, the density distribution is not homogeneous a statistical ensemble generated by the Monte Carlo pro-
but decays exponentially from the center to the circumfer- cedure [18,19] according to the conditional bond confor-
ence. A simple Ornstein–Zemike type is generally applied mation probability [20].
to the density correlation function for a Gaussian chain: When no molecular model is available, the scattering
profile can be analyzed in terms of a triaxial body model of
n
cðrÞic expðr=nÞ ð8Þ homogeneous density representing the shape of the object
r [21] or by assuming a suitable pair correlation function for
where c is a concentration of polymer chains and n is a the electron density distribution in the object [22]. The
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

correlation length specifying the range of effective density scattering factor is explicitly calculated for some homoge-
fluctuation. Introducing in Eq. (1) for a statistically isotro- neous triaxial bodies including a sphere, an ellipsoid, a
pic system, Eq. (8) yields the scattering profile as cylinder, and a prism. For example, the scattering factor
for a sphere is given by Eq. (14) as
cn3
IðqÞc ð9Þ " #2
1 þ n2 q2 2 3ðsin Rq  Rq cos RqÞ
PðqÞ ¼ / ðqRÞ ¼ ð14Þ
The volume term V in Eq. (1) is replaced by cn3, which ðRqÞ3
corresponds to the number of units in the correlated
density fluctuation. Debye and Beuche [17] proposed a where R denotes the radius of a sphere. The observed
correlation function that specifies the density correlation scattering profile is compared with that calculated from
for a randomly associated system: an assumed triaxial model of a suitable dimension, which is
supposed to be composed of associated oligosaccharides or
cðrÞiexpðr=nÞ ð10Þ polysaccharide chains.
No interdomain (interparticular) interaction is con-
Equation (10) yields the scattering profile as sidered in the above argument, and the scattering is con-
sidered to be due solely to an isolated domain (or an
cn3 isolated particle). When the interdomain (interparticular)
IðqÞc ð11Þ
ð1 þ n2 q2 Þ2 interaction becomes dominant, an interference peak will
appear at the q range corresponding to the interaction
which exhibits a faster decay of the scattered intensity with distance in the scattering profile. If the interdomain (inter-
q. particular) interaction is isotropic and spherically symmet-
The particle scattering from a single molecule is in ric, the scattering profile is decomposed into the product of
principle calculated from the coordinates of the constituent two terms of the particle scattering factor P( q) and the
atoms interference SI( q) [16]:

X
n n1 X
X n
IðqÞcPðqÞ  SI ðqÞ ð15Þ
IðqÞ ¼ g2i /2i ðqÞ þ 2 gi gj /i ðqÞ/j ðqÞ
i¼1 t¼1 j¼iþ1
ð12Þ where the interference term is written as
sindij q
 1
dij q SI ðqÞc 3=2
ð16Þ
1  ð2pÞ ðe=m1 ÞbðqÞ
where q denotes the magnitude of the scattering vector with e and m1 being a constant close to unity and the average
given by (4p / k) sin(h / 2) with k and h being the wavelength volume allocated to each interacting domain, respectively.
of the incident beam and scattering angle, respectively, and b( q) represents the interaction potential in the Fourier
gi is an atomic scattering factor. dij is the distance between (scattering) space. When the interdomain interaction is
the ith and jth atoms, and the form factor for a single atom given in terms of a hard-sphere repulsion, b( q) is repre-
/i( q) is assumed to be given by the form factor for a sphere sented by the scattering amplitude of a sphere, Eqs. (13)
with a van deer Walls radius of the ith atom and (16) reduce to

3½sinðRi qÞ  ðRi qÞcosðRi qÞ


1
/i ¼ ð13Þ SI ðqÞc ð17Þ
1 þ 8ðm0 =m1 Þe/ð2qRÞ
ðRi qÞ3
where m0 is the volume of the sphere and the hard-sphere
where RI is the van deer Walls radius of the ith atom. If a interaction is represented by the sphere of a uniform radius
molecule is rigid, the distance dij is fixed and Eq. (1) is 2R. The interaction potential b( q) is approximately given

Copyright 2005 by Marcel Dekker


8 Kajiwara and Miyamoto

by the Gaussian function when the interaction is softer 6 ppm for protons on the ring. The anomeric protons (Hi)
[22,23], and Eq. (16) is rewritten as have peaks in the region between 4.5 and 5.5 ppm, whereas
the chemical shifts for other protons (H2–H6) ranges from
1
SI ðqÞc   ð18Þ 2 to 4.5 ppm. The H1 chemical shift database will provide a
1 þ 2A2 Mw c exp n2 q2 starting key to assign the chemical shifts of unknown
samples, although the chemical shift database for oligo-
where the Gaussian-type interaction potential is specified
saccharides and polysaccharides are still far from comple-
by the correlation length n of interaction.
tion with respect to the accumulation and systematization.
As the chemical shifts are also sensitive to the conforma-
C. Fundamentals of Nuclear Magnetic tional change, solvent, and temperature, it requires expe-
Resonance Spectroscopy Applied rience and skill to identify the 1H NMR peaks for unknown
to the Conformational Analysis oligosaccharides and polysaccharide samples. Various
two-dimensional NMR techniques have been developed
Nuclear magnetic resonance (NMR) spectroscopy has to facilitate the assignment and identification of the chem-
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

been widely employed in the structural analysis and the ical shifts as described in a later section.
conformational dynamics of polymers in solution, gel, or The 1H NMR chemical shift data are summarized for
solid states. However, its application is limited to the monosaccharides in Table 3 [24]. The data are shown for
polymers that are not entirely crystalline in general. It monosaccharides as the components of oligosaccharides in
provides information on microscopic chemical structures, which each is linked to an adjacent monosaccharide via a
including the primary structure, stereoregularity, confor- glycosidic bond oriented either below (a) or above (b) the
mation, and secondary structure of synthetic polymers, plane of the ring. The chemical shift values of monosac-
proteins, and polysaccharides. Various NMR techniques charides will assist the identification of oligosaccharides
have also been developed to investigate molecular motion and polysaccharides, but the values vary considerably with
through relaxation times, correlation times, and self-diffu- the configuration and conformation of samples.
sion coefficients. One of the advantages of NMR in the
structure analysis is its sensitivity to a microscopic struc- 2. Relaxation Time
ture within a short-range order in comparison with small-
The spin-lattice relaxation time (T1) and the spin–spin
angle x-ray scattering. The other advantages of NMR are
relaxation time (T2) reflect the conformational change
that (1) it is a noninvasive method where no probes are
and the local tumbling motion of oligosaccharides and
needed; (2) the sample for measurement can be liquid,
polysaccharide chains. The relaxation process has been
solid, or gel; (3) the NMR signals can be assigned individ-
observed to understand the structure-dependent molecular
ually to the main chain, the side chain, or the functional
motion, the helix-coil transition, the sol–gel transition, the
group of a sample and yield the structural information on a
crystalline structure, the amorphous structure, the aggre-
specific site; and (4) the molecular motion and dynamic
gation structure, and the hydration structure.
structure (time-dependent structure) can be observed.
The spin-lattice relaxation time T1 is measured with
However, NMR has some disadvantages: (1) the spatial
the repeated p–s–2/p radio frequency (RF) pulse sequence
position of atomic groups is not determined accurately; (2)
by the inversion recovery method [25]. T1 follows Eq. (19)
the information on the long-range and higher-order struc-
derived from Bloch’s equation:
ture will be lost; and (3) the duration time is long to observe
NMR peaks from polymer samples with a reasonable S/N
lnðAl  As Þ ¼ ln 2Al  s=T1 ð19Þ
ratio and high resolution. Thus NMR spectroscopy com-
pliments other methods of the structural and conforma- where Al and As are the magnitude of the recovering
tional analysis of polymers, including x-ray diffraction, vector of magnetization evolved by a p/2 RF pulse at time t
light scattering, and small-angle x-ray (neutron) scattering. = l and s, respectively. T1 is evaluated from the plot of
A variety of NMR techniques are available for the ln(Al  As) against s. T1 is given in terms of the viscosity g
structure analysis of oligosaccharides and polysaccharides. and temperature T [26] as
The one-dimensional pulse NMR technique is mainly
  
applied for the analysis of the saccharide primary structure 1 128p3 l4 a3 g

in solution state and the determination of relaxation times. ¼ ð20Þ


T1 h2 r6 kT
The solid state, high-resolution NIVIR technique can be
applied for the structure analysis of oligosaccharides and where l denotes a nuclear moment, a is the effective radius
polysaccharides in viscose solution, gel, and solid state. The of a spherical molecule, and r is the distance from the
two- or three-dimensional techniques are used to determine observed nucleus to its magnetic neighbor. T1 decreases in
the primary and secondary structures and the conforma- proportion to g/T and a3 increases with r6. The effective
tion of oligosaccharides and polysaccharides. volume a3 is replaced with the molar volume in the case of
oligosaccharides and polysaccharides in solution. T1 as a
1. Chemical Shift function of the correlation time indicates the degree of
Oligosaccharides and polysaccharides show several 1H molecular motion, and T1 takes a minimum at the temper-
NMR signal peaks in the spectrum region between 2 and ature when the relaxation occurs according to the dipole–

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 9

Table 3 Chemical Shifts (ppm) of Monosaccharides from Acetone at 2.225 ppm in D2O at 22–27jC
Protons

Monosaccharidea H1 H2 H3 H4 H5 H6 H7 CH3 NAC


a-D-Glc-(1! 5.1 3.56 3.72 3.42 3.77 3.77 3.87 – –
h-D-Glc-(1! 4.4 3.31 3.51 3.41 3.45 3.74 3.92 – –
a-D-Man-(1! 1.9 3.98 3.83 3.70 3.70 3.78 3.89 – –
h-D-Man-(1! 4.7 4.04 3.63 3.58 3.37 3.76 3.93 – –
a-D-Gal-(1! 5.2 3.84 3.90 4.02 4.34 3.69 3.71 – –
h-D-Gal-(1! 4.5 3.52 3.67 3.92 3.71 3.78 3.75 – –
h-D-GlcNAc-(1! 4.7 3.75 3.56 3.48 3.45 3.90 3.67 – 2.04
a-D-GalNAc-(1! 5.2 4.24 3.92 4.00 4.07 3.79 3.68 – 2.04
h-D-GalNAc-(1! 4.7 3.96 3.87 3.92 3.65 3.80 3.75 1.23 2.01
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

a-L-Fuc-(1! 5.1 3.69 3.90 3.79 4.1–4.9b – – 1.28 –


a-L-Rha-(1! 4.9 4.06 3.80 3.46 3.74 – – – –
c
h-D-Xyl-(1! 4.5 3.27 3.43 3.61 – – 1.32 –
3-u-Me-a-L-Fuc-(1! 4.8 3.70 3.40 – 3.89 – – 1.32 –
3-u-Me-a-L-Rha-(1! 5.0 4.24 3.59 3.52 3.77 – – 1.32 –
2,3-di-u-Me-a-L-Rha-(1! 5.1 3.94 3.52 3.41 3.73 – – – –
3,6-di-u-Me-h-D-Glc-(1! 4.7 3.34 3.31 3.51 3.51 3.66 3.78 –
a
These are average values for nonreducing terminal sugars linked by a glycosidic linkage to the adjacent monosaccharides. Signals for protons at
the ring carbons are shifted downfield when linked by another monosaccharide at the hydroxyl group of that carbon.
b
These signals considerably vary more than other signals due to conformational features.
c
H5ax 3.29; H5eq 3.93.
Source: From Ref. 24.

dipole interaction [27]. The correlation time, sc, is given power proton dipolar decoupling (DD) to improve the
approximately by resolution.
A magic angle spinning (MAS) method is employed to
sc ¼ 4p3 a3 g=3kT ð21Þ diminish the chemical shift anisotropy [32]. A sample
1
placed in a cylindrical rotor is rotated about an axis making
H T1 varies with the spin diffusion [28] and the value of T1 an angle a with the magnetic field, H0, at 800–5000 Hz by
is much influenced by O2 gas. air. The chemical shift Hamiltonian is composed of a time-
The T2 experiments are performed to observe the independent term and a time-dependent term [33]. The
molecular motion in an extreme narrowing condition [27] time-dependent term yields side bands at the multiples of
where the viscosity of a sample solution is low and the the rotation rate in the spectrum, but the side bands
motion is fast. The T2 measurements are suitable especially disappear at a spinning rate faster than a half of the width
for 1H nuclei because the problem resulting from the spin of the chemical shift anisotropy powder pattern observed
diffusion can be avoided in the T2 experiments. The T2 in the viscose solution or solid samples. When the sample is
value is determined by the Carr–Purcell [29]/Meiboom– rotated at the fixed angle a being equal to 54.74j (magic
Gill [30] (CPMG) method. Here the pulse sequence (p/2)– angle) with respect to the magnetic field, the chemical shift
s–py–2s–py–2s–py–p. . .. (s is the pulse interval) is used to anisotropy vanishes and the time-independent term con-
avoid the cumulative error due to incorrect pulse lengths. tains only the isotropic chemical shift.
Due to long 13C T1, a long repetition time is needed to
3. High-Resolution Solid State Nuclear Magnetic observe an NMR spectrum with a sufficient S/N ratio and a
Resonance high resolution in solid state experiments. The reduction of
A rapid isotropic tumbling molecular motion is restrained T1 can be achieved by transferring the energy of 13C spins in
in the viscose solution state or in the solid state of oligo- the excited state (at a high-spin temperature) to the NMR
saccharides and polysaccharides. The NMR spectrum lattice. The energy is transferred from 13C spins at a high-
shows a proton dipolar broadening of many kilohertz spin temperature to 1H spins in the cross-polarization (CP)
due to strong dipole–dipole interaction and a chemical technique [34,35] where the Hartmann–Hahn condition is
shift anisotropy as a result of the restraint of the molecular satisfied. The RF pulse sequence of the CP technique for
motion. A high-power, proton-decoupling field [31] is measuring 13C nuclei is shown in Fig. 6a. SL denotes a
found to be effective to remove a proton dipolar broaden- pulse for spin locking and DD is a pulse for heteronuclear
ing. 13C–1H scalar coupling can be removed by the high- dipolar decoupling. The Hartmann–Hahn condition is

Copyright 2005 by Marcel Dekker


10 Kajiwara and Miyamoto

response of the nuclear spin system becomes a two-dimen-


sional function of two independent times t1 and t2. When
FID is two-dimensionally Fourier-transformed, a two-
dimensional spectroscopy is obtained as a function of
two independent frequencies.
The 2-D shift correlated spectroscopy (COSY) in-
forms the connection of nuclei. The pulse sequence of
COSY for 1H nuclei is shown in Fig. 6b. (p/2)/1 and
(p/2)/2 are the first and second pulses, which differ in
phase. The time interval between two pulses, t1, is an
evolution time and t2 corresponds to an acquisition time.
A 1H–1H COSY spectrum is represented by a square, where
both axes correspond to 1H chemical shifts. The signals in
the spectrum are classified in diagonal peaks and cross-
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

peaks. The diagonal peaks are equivalent to the 1-D NMR


spectroscopy. The cross-peaks appear symmetrically with-
respect to the diagonal peaks and correspond to the dif-
ference of the chemical shifts of two sites specified on the
diagonal line by the two coordinates of respective peak
position.
The 2-D homonuclear Hartman–Hahn spectroscopy
(HOHAHA) reveals a spin–spin interaction network as a
totally correlated spectroscopy that is obtained by chang-
ing the duration of the spin-locking application [36]. When
the Hartman–Hahn condition is satisfied by spin locking,
the magnetization transfer takes place by the spin–spin
coupling between I and S spins and its degree can be
adjusted by the duration of spin locking. Homonuclear
Hartman–Hahn spectroscopy is more sensitive than COSY
with respect to the line resolution, and facilitates the
assignment of 1H signals along covalent bonds. To satisfy
Figure 6 Timing diagrams for the NMR pulse sequence: (a)
the Hartman–Hahn condition over a wide range, a proton
CP, (b) COSY, (c) HOHAHA, and (d) NOESY. broadband decoupling is introduced by a specially
designed pulse sequence. Fig. 6c shows the pulse sequence
of HOHAHA, where SLy is a spin-locking pulse and sm a
mixing time. The MALEV-17 composite pulse [37] applied
satisfied by the pulse applied to 13C while applying the SL during the mixing time to lock spins over a wide frequency
pulse. The signal created by CP is four times the original range.
magnetization in an ideal condition. The DD and the MAS The nuclear Overhauser effect correlated spectroscopy
are usually combined with the CP technique to obtain high- (NOESY) observes the nuclear Overhauser effect due to the
resolution spectra (CP/MAS). magnetic dipole–dipole interaction between nuclei in a
short distance, and reveals the conformation, configura-
4. Two-Dimensional Nuclear Magnetic Resonance tion, and chemical exchange of large molecules [38]. The
In interpreting the NMR spectra, the first step is to identify pulse sequence of NOESY is basically the same as COSY
signal peaks. As mentioned above, the spectra for oligo- except for the additional p/2 pulse after a fixed time sm as
saccharides and polysaccharides are complicated and two- shown in Fig. 6d, where sm denotes a mixing time. The
dimensional (2-D) NMR technique is commonly applied to distance between 1H nuclei is determined from the intensity
separate the NMR signals on the basis of J coupling. The 2- of cross-peaks, and offers a mean of investigating spatial
D NMR technique yields information on the spin–spin relationships between nuclei through NOE. The cross-
coupling between heteronuclei, chemical exchange, and the relaxation rate for an I and S spin system, rIS, is a function
nuclear Overhauser effect (NOE). of the distance between the I and S spin:
The 2-D NMR technique involves several spectro-  
scopic methods classified by the mode of pulse sequence c4 t2 6sc
(Fig. 6). The response of the nuclear spin system to the RF rIS ¼  sc ð22Þ
10r6 1 þ 4x2 s2c
pulse is observed as FID (free induction decay) as a
function of a time t2, which is Fourier transformed to yield where x is the Larmor frequency and the sc is the correla-
an NMR spectrum in the conventional (1-D) spectroscopy. tion time of reorientation [39]. rIS is evaluated from the sm
By applying two RF pulses with a time interval t1, a second dependence of cross-peak intensities. The spatial informa-
time axis t1 (an evolution time) can be introduced where the tion obtained by NOESY is restricted within the distance of

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 11

about 0.5 nm. sc depends on the motility of molecules. The evaluation of a total conformational energy as a function
cross-peaks show negative and positive values for xsc < 1 of a pair / and w. / and w can take any value between 180j
and xsc > 1, respectively. When xsc c 1, the cross-peaks and +180j. The most likely conformation is expected to
of NOE are not observed. By applying spin locking, a have the lowest potential energy. For example, 38 pairs of /
positive NOB is observed over the wide time scale of and w evaluated from the crystallographic data of maltose
molecular motion. The rotating frame nuclear Overhauser Glcpa 1!4 Glc are found to lay within the low-energy
effect spectroscopy (ROESY) is developed [39,40] to ob- range of 2 kcal/mol above the absolute energy minimum on
serve NOB of the sample whose molecular weight ranges the energy map provided by molecular mechanical calcu-
from 1000 to 2000 and xsc c 1. lation, proving the validity of computer modeling. Here
molecular modeling permits to evaluate the range of attain-
D. Molecular Modeling able conformations in terms of the potential energy at each
point specified by a pair of / and w. The observed value of /
1. Monte Carlo Method and w will vary among the attainable conformations
Two dihedral angles / and w with respect to the glycosidic according to the crystal packing (in the solid state) or the
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

bond determine the conformation of a disaccharide, pro- type of solvent (in the solution). Fig. 7 shows the confor-
vided that a pyranose ring is rigid (Fig. 7). The conforma- mational energy map of maltose, cellobiose, xylobiose,
tional analysis of a disaccharide thus comprises the chitobiose, laminaribiose, and sphorobiose calculated by

Figure 7 Definition of two dihedral angles, / and w, to determine the conformation of a disaccharide, and 2-D contour energy
map (the potential energy as a function of two dihedral angles / and w) of (a) maltose, (b) cellobiose, (c) xylobiose, (d)
chitobiose, (e) laminaribiose (s = 112.5j), and (f ) sophorose.

Copyright 2005 by Marcel Dekker


Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Copyright 2005 by Marcel Dekker


Figure 7 Continued.
Progress in Structural Characterization of Functional Polysaccharides 13

determined by a set of / and w when the bond angle s is


fixed. The effect of excluded volume can be taken into
account by excluding the step that places a unit in a
specified vicinity of the space occupied already by the unit
in a previous step.
Because the polysaccharide chain undergoes thermal
fluctuation in solution, the scattering from the solution is
observed as an average over space and time. Assuming the
ergodicity, several chains are independently generated to
constitute a microcanonical ensemble, and the particle
scattering function is then given by an ensemble average
over the scattering calculated from the atomic coordinates
of each generated chain according to Eq. (12). Fig. 9 shows
the scattering profiles averaged over 500 chains of two 1,4-
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

glucans, (1!4)-a-D-glucan (amylose), and (1!4)-h-D-glu-


can (cellulose), generated by the scheme represented by Fig.
8 with varying the number of glucosidic residues in terms of
the Kratky plots by plotting q2I( q) against q. Here the
occurrence probability for a pair of / and w is provided by

Figure 7 Continued.

molecular mechanics (MM2 or MM3) with the force-field


including bond vibration, bond stretching, angular torsion,
and van der Waals interaction (see Table 1 for the termi-
nology). Here the glucose residue is assumed to be rigid and
replaced with a virtual bond connecting the neighboring
oxygen atoms of the glycosidic linkage (Fig. 7). The bond
angle s is fixed, for example, to 110j in the case of amylose so
as to yield a consistent value for the radius of gyration as
observed for high molecular weight amylose.
Longer chains are generated by the Monte Carlo
method according to the scheme summarized in Fig. 8,
where the assumption is made that the short-range inter-
action between two adjacent residues determines the range
of permissible values of a pair of the dihedral angles / and
w. That is, a pair of / and w is provided from the energy
map (Fig. 7) according to the occurrence probability P(/,
w) specified by the Boltzmann factor associated with the
potential energy E(/, w) of a disaccharide with a set of /
and w. Here P(/, w) is given as

Pð/; wÞ ¼ c exp½Eð/; wÞ=kB T


ð23Þ

with kB being a Boltzmann constant, T an absolute tem-


perature, and c a normalization constant. A chain is con- Figure 8 Flow chart to generate polysaccharide chains and
structed step by step as the geometry of each unit is calculate the scattering factor.

Copyright 2005 by Marcel Dekker


14 Kajiwara and Miyamoto
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 9 Simulated scattering profiles of (1!4)-h-D-glucan (a) and (1!4)-h-D-glucan (b) as a function of the number of
glucosidic residues with snapshots of a simulated structure of respective glucan chains composed of 40 glucosidic residues (a
stereo figure) on the right.

the energy map of Fig. 7. The snapshots of simulated chains chain conformation of (1!4)-a-D-glucan, which yields a
(DP = 40) are also shown on the right side of the Kratky thicker cross-sectional radius of gyration evaluated as
plots. The simulated amylose chain reveals the wobble approximately 5 Å from the cross-sectional Guinier plot
helical conformation with localized highly ordered helical [Eq. (6)] of the simulated scattering profiles for (1!4)-a-D-
regions, whereas the cellulose chain seems to have a rather glucan chains of over 20 glucosidic residues. The cross-
extended chain structure as expected from the primary sectional radius of gyration remains as small as 2.1 Å in the
structure. The calculated scattering profiles reveal a pro- case of (1!4)-h-D-glucan chains, whose extended chain
nounced peak in Kratky plots at q = 0.2 Å1 for (1!4)-a- conformation promotes to assume the intermolecular hy-
D-glucan of higher degrees of polymerization, whereas drogen bonding to form non-water-soluble aggregates.
(1!4)-h-D-glucan exhibits a scattering profile typical to a Table 4 summarizes the radius of gyration of (1!4)-a-D-
rigid rod-like molecule. The intramolecular hydrogen glucan and (1!4)-h-D-glucan, each calculated from the
bonding is responsible for stabilizing the quasi-helical simulated profiles and/or estimated from the observed

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 15

SAXS profiles. Although some refinement of the probabil-


ity map is necessary, the simulation accounts, at least
qualitatively, for the DP dependence of the radius of
gyration and the difference in the radius of gyration due
to the glucosidic linkage mode.
When the saccharide chain is longer, the excluded
volume effect becomes more serious. The excluded volume
effect can be taken into account by considering the inter-
action between the nonbonded units. Conventionally, the
repulsive interaction is dealt with in the Monte Carlo
simulation by replacing the chain units (segments) with
hard spheres of a finite radius. Fig. 10 demonstrates the
snapshots of amylose chain generated by the Monte Carlo
method with and without excluded volume, which is rep-
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

resented by a sphere of a radius 4 Å at the position of each


glycosidic oxygen. The excluded volume effect is seen to
expand the chain, but the helical nature of amylosic chains
is retained in both unperturbed and perturbed states.

2. Molecular Dynamics
The Monte Carlo method described in the preceding
section is based on the disaccharide conformation energy
map, and no water molecules are taken into account in the
model. Although the Monte Carlo method is capable of
simulating longer polysaccharide chains, it does not allow
including the solvation effect directly through water-medi-
ated hydrogen bonds. Molecular dynamics (MD) simula- Figure 10 Snapshots of amylose chain generated by the
tions [41] can be applied to the structural studies of various Monte Carlo method with (b) and without (a) excluded volume.
polysaccharides, where water molecules can be explicitly Here the excluded volume is taken into account by replacing the
included in the simulation. However, the MD simulation is glycosidic oxygen (denoted by a circle in the Figure) with the
restricted to relatively shorter chains owing to a present hard sphere of radius 4 Å.
computational capacity. The results of the MD simulation
depend on the employed force-filled models such as Gro-
mos [42], Glycam93/99 [43], and Cff91/Cff [44], as well as on the glycosidic linkages. Here Gromos ignores the exo-
on the starting conformation. Among the force fields anomeric effect, while Glycam incorporates the effect
mentioned above, Gromos and Glycam is composed of a through the torsional terms determined from the ab initio
set of parameters specifically developed for amylose; how- geometry optimization at the HF 6-31G* level. Cff91 is a
ever, they differ in the treatment of the exoanomeric effect general-purpose force field for biomolecules, and Cff is
expanded from Cff91 to include the parameters with a
proper account of the anomeric effect on the glycosidic
linkages. The example of the MD simulations will be
Table 4 Radius of Gyration of (1!4)-a-D-Glucan and
(1!4)-h-D-Glucan Oligomers shown in the later section.

(1!4)-h-D-glucan (Å) (1!4)-a-D-glucan (Å)


DP (obs.) (cal.) (obs.) (cal.) III. STRUCTURAL AND CONFORMATIONAL
ANALYSIS OF OLIGOSACCHARIDES
1 3.47 –
AND POLYSACCHARIDES
2 3.53
3 4.85 4.55 4.43 The structural characterization of simple homoglucans is
4 6.22 5.28 5.32 mainly introduced in this section. The structural and
5 7.61 6.15 6.04 mechanical properties of the gels from various marine
6 9.00 6.69 6.56 polysaccharides, plant polysaccharides, microbial polysac-
7 10.40 7.26 7.06 charides, and animal polysaccharides are reviewed by
8 11.84 8.08 7.54 Clark and Ross-Murphy [45]. This chapter intends to
9 13.23 10.21 7.77
demonstrate the advanced methods applied for the struc-
10 14.70 10.11 8.83
tural and conformational characterization of oligosaccha-
20 28.53 13.89
rides and polysaccharides, particularly in solution, by
40 55.18 23.18
50 28.12 taking the examples of the homoglucans composed of
different modes of glucosidic linkage.

Copyright 2005 by Marcel Dekker


16 Kajiwara and Miyamoto

A. (1!4)-A-D-Glucan Represented by Amylose a single chain conformation of amylose. Here no adjustable


parameter is involved, except for the normalization with
The oligomers of (1!4)-a-D-glucans dissolve well in water. respect to the scattered intensity at q = 0.
The observed SAXS profiles from maltohexaose and mal- Amylose is known to assume a double-stranded (B-
tooctaose are shown in Fig. 11, where no effect of associ- form) [46,47] or single-stranded helical (V-form) [48] con-
ation was observed. The simulated SAXS profiles are also formation in a solid state from the analysis of the x-ray
shown in Fig. 11 to examine the consistency of simulation fiber diffraction, the x-ray powder, and the electron dif-
with the observed profiles. The characteristics of wobbled fraction pattern of single crystals. Amylose aqueous solu-
helix represented by a pronounced peak in the Kratky plots tion forms gel by cooling. Gelation takes place through the
become more distinct in maltooctaose than in maltohexaose formation of nanocrystallites that serve as cross-linking
as expected from the molecular weight dependence of domains. Particle scattering from model nanocrystallites is
simulated SAXS profiles (Fig. 9). A good agreement be- calculated [49] by assuming nanocrystallites composed of
tween simulated and observed SAXS profiles assures that B-form double helices or single-stranded V helices. The
the simulation can be extended to a longer chain to elucidate model nanocrystallite is approximately represented by an
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 11 Simulated and observed scattering profiles from maltohexaose (a) and maltooctaose (b). The Figures on the right
show a snapshot conformation of simulated maltohexaose and maltooctaose, respectively (a stereoview).

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 17

elliptical cylinder of 8.32-nm thickness (contains 42–222


duplexes composed of 24 glucosidic residues per strand) or
a parallelepiped of 6.44-nm thickness (contains 120 helices
composed of 24 glucosidic residues per strand) for the B-
form or the V-form, respectively.
The SAXS profile from amylose aqueous solution
reveals a sharp upturn at q!0 in the Kratky plots (Fig.
12) [ln q2I( q) plotted against q] according to the sol–gel
transition. This pronounced upturn is ascribed on the
formation of an infinite structure (gel) as expected by the
cascade theory of gelation [50]. At higher q regions, two
scattering profiles from sol and gel coincide, indicating that
the local conformation is identical in the sol and gel states.
The local conformation is probably represented by a single-
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

stranded chain simulated by the Monte Carlo method


shown in Fig. 9a, considering that single-stranded chains
are present in amorphous region of gel or in solution [51].
The observed profiles are fit to the scattering profile from
Figure 13 Scattering profile of amylose gel decomposed into
two components. Iexcess denotes the excess scattering from
amylose gel with respect to amylose sol (= IGEL  ISOL).
Imodel is the scattering calculated from the oblate ellipsoid of
revolution (12.9  13.1  4.3 nm), and Ical = ISOL + Imodel.

simulated (1!4)-a-D-glucan chains (DP = 40) in Fig. 12.


The Guinier plots for the cross section [Eq. (6)] yields the
cross-sectional radius of gyration as 0.45 nm in both gel
and sol. The value of 0.45 nm (close to 0.5 nm), which is
evaluated for the cross-sectional radius of gyration from
the model double-stranded helix [52], also corresponds to
an apparent cross-sectional radius of gyration of a single
(1!4)-a-D-glucan chain. Here the deviation at lower q
ranges is considered to be due to the presence of double-
stranded helices formed by the coupling of two neighboring
single-stranded helices without significantly disturbing the
conformation. The SAXS profile from amylose gel was
analyzed in terms of two components representing nano-
crystallites and amorphous region [53], respectively, by
assuming that no interference would take place between
two components. The structure of the amorphous region in
amylose gel should be identical to that in the sol state. Thus
the excess scattering in the gel state with respect to the sol
state mainly resulted from the formation of nanocrystal-
lites that function as the cross-linking domain ( junction
zone). The oblate ellipsoid of revolution was found to yield
the scattering profile fit to the excess scattering (Fig. 13),
and its dimension (three semiaxes 12.9  13.1  4–3 nm)
approximately corresponds to the nanocrystallite com-
posed of 42 B-form duplexes with 24 glucosidic residues
per strand.
The molecular dynamics simulation was performed on
maltopentaose with currently available force fields [54].
Figure 12 Small-angle x-ray scattering profile from amylose The results are compared with the small-angle x-ray scat-
gel and sol, where closed and open circles denote gel and sol, tering observed from maltopentaose in aqueous solution.
respectively. Solid lines represent the calculated scattering Fig. 14 compares the simulated profiles with the observed
profile from simulated (1!4)-h-D-glucan chains of DP = SAXS profiles, where Fig. 14a provides a series of results
40. simulated with available force fields, and Fig. 14b the

Copyright 2005 by Marcel Dekker


18 Kajiwara and Miyamoto

B. (1!4)-B-D-Glucan Represented by Cellulose


Cellopentaose cannot be completely dissolved in water
because of a strong intermolecular interaction by hydrogen
bonding through OH groups on C6. The SAXS from the
aqueous solution of cellopentaose (30 mg/mL) exhibits a
sharp upturn toward lower q due to the formation of large
aggregates (Fig. 16). If the aggregation is caused by inter-
molecular hydrogen bonding, the aggregates are con-
sidered to be formed by the side-by-side stacking of
cellopentaose chains. The simulated profile (a solid line in
Fig. 16) reflects a chain stiffness of a (1!4)-h-D-glucan
chain, but the observed profile significantly deviates from
the simulated profile at a lower q region. The cross-sectional
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

radius of gyration is estimated as 3.5 Å at the inter-


mediate q range and as over 70 Å at the smaller q range. A
single (1!4)-h-D-glucan chain has the cross-sectional ra-
dius of gyration of 2.1 Å, so that two cellopentaose chains
are considered to form a stable aggregate and some further
aggregate into a larger cluster. The intermolecular hydro-
gen bonds can be broken by adding urea in the aqueous
solution of cellopentaose. Fig. 16b observes a good agree-
ment between the observed and simulated SAXS profiles,
and the cross-sectional radius of gyration is estimated as 2.1
Å, which is expected for a single (1!4)-Å-D-glucan chain.
Regarding the local conformation of (1!4)-h-D-glu-
can chain in the presence of water, the potential use of
multiple-RELAY-COSY is suggested from the analysis
of complex spin networks of 1H NMR spectra of cello-
oligosaccharides where the complete assignment of 1H
NMR resonance was achieved for cellotriose [56]. Solvent-
suppression COSY provides also a useful method to eluci-
Figure 14 Small-angle x-ray scattering profiles observed
date the interaction of the hydroxyl groups with water [57].
from maltopentaose aqueous solution (open circles) of 20.13
The 1H NMR of methyl h-cellobioside in H2O-acetone-
mg/mL at 25jC with simulated profiles (respective curves).
(a) MD results (the radius gyration and force field are shown
d6 (85 :15) yields sharp signals due to the seven hydroxyl
in the Figure) and (b) Monte Carlo results and profiles cal- groups at 20jC (Fig. 17), where all signals are identified
culated from crystalline regular helices (the radius of gyration [57].
and the source of other data are shown in the Figure).

Monte Carlo results with two probability maps (Monte


Carlo K denotes a rigid map employed in Fig. 9a) and the
profiles calculated from the atomic coordinates of three
regular helices. Both Monte Carlo results show a satisfac-
tory agreement with observed SAXS profiles, where a small
difference due to the glucose geometry was observed at
higher q. The results of MD simulations vary with the force
fields, and the Cff91 seems to yield the best fit to an
observed profile. Because the helix model of Goldsmith et Figure 15 Stereoviews of the snapshot conformations of
al. [55] fits satisfactorily well to the observed SAXS profile, maltopentaose as simulated by the Monte Carlo method and
maltopentaose seems to assume a quasi-helical conforma- MD, including regular amylose helices. (a) Regular helix (8.3
tion specified by a radius of 5.38 Å, a rise of 2.44 Å, a pitch Å) [47], (b) regular helix (7.4 Å) [55], (c) regular helix (5.9 Å)
of 17.60 Å, a repeat of 7.2 Å, / = 105j, and w = 135j. A [48], (d) Monte Carlo/MM3 (7.57 Å), (e) Glycam93 (8.85 Å),
typical conformation of maltopentaose is shown in Fig. 15 (f ) Glycam99 (8.08 Å), (g) modified Glycam93 (7.72 Å), (h)
as simulated with various force fields. In fact, the confor- Cff91 (7.83 Å), (i) Cff (8.30 Å), ( j ) Gromos (8.32 Å). The
mation observed by the MS simulation with the Cff91 is values in each bracket denote the radius of gyration, which
similar to the helix model of Goldsmith et al. was evaluated as 7.4 F 0.2 Å from the SAXS profile.

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 19
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 16 Small-angle x-ray scattering profile from cellopentaose in water (a) and in 1 M urea aqueous solution (observed and
simulated as indicated in the Figures). A stereoview of a simulated cellopentaose chain is shown on the right.

The crystal structure of cellulose has been a subject of a found to vary its pattern, implying that the ratio of two
long-standing argument. Cellulose is known to have four allomorphs Ia/Ih differs by the origin of native cellulose
different polymorphic crystalline forms classified as cellu- [62–64]. It is interesting to note that a single microfibril of
lose I, II, III, and IV. Parallel chain packing is proposed for native cellulose is a composite of two crystalline phases, Ia
native cellulose I [58], and regenerated cellulose II is and Ih [65,66].
supposed to assume antiparallel chain packing [59,60] as The crystal structure of cellulose II is considered to
analyzed from the results of x-ray fiber diffraction pattern. consist of two antiparallel chains of almost identical con-
Because CP/MAS 13C NMR revealed cellulose I as being formation packed in a monoclinic unit cell, where the
composed of the allomorphic mixture of triclinic Ia and hydroxymethyl group at C6 assumes a tg or a gt confor-
monoclinic Ih, the refinement of cellulose crystal structure mation in the ‘‘up’’ or ‘‘down’’ chain, respectively. How-
again became a main issue in the cellulose science [61]. Here ever, CP/MAS 13C NMR exhibits a singlet at 64 ppm for
the multiplicity at C4, C1, or C6 is due to magnetically the C6 resonance from cellulose II polymorph against the
nonequivalent sites present in crystalline domain, and is expected doublet to be observed at 64 and 66 ppm from the

Copyright 2005 by Marcel Dekker


20 Kajiwara and Miyamoto

tg and gt conformations [67]. The cellulose II crystal


structure is re-examined [68] from the crystal structure of
cellodextrin oligomers, including h-D-cellotetraose (Fig.
18) [69,70] and methyl h-cellotrio side [71]. In those cello-
dextrin oligomers, all the hydroxymethyl groups (C6–O6
bonds) are in the gt position, but the two antiparallel chains
assume a different glucose ring conformation. This finding
accounts, at least qualitatively, a singlet for C6 and a
doublet for C1 and C4 observed for cellulose II by CP/
MAS 13C NMR.

C. (1!3)-B-D-Glucan
(1!3)-h-D-glucan consists of a backbone of a group
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

of extracellular plant/fungal glucans such as cinerean,


curdlan, krestin, laminaran, lentinan, schizophyllan, and
scleroglucan, which are known to affect the immune sys-
tem as an unspecific modulator [72]. Except for curdlan,
which is linear (1!3)-h-D-glucan, the (1!3)-h-D-glucan
family contains some amount of h(1!6) branched D-
glucose residues, and assumes a triple-helical conforma-
Figure 17 1H NMR spectra of methyl h-cellobioside in tion. Although the structural requirement is not expli-
H2O–acetone–d6 (85:15) at 20jC. citly understood, the antitumor activity is said to be more

Figure 18 Molecular structure (stereoview) of two h-D-cellotetraose chains.

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 21

pronounced in lower h(1!6) branched (1!3)-h-D-glucans over the ensemble was calculated to compare with the
with a relatively high molecular mass [73]. Those (1!3)-h- observed SAXS profiles. The simulated scattering profiles
D-glucans form triple-stranded helices of high rigidity (in terms of the Kratky plots) exhibit characteristic maxima
in aqueous solution [74,75], and the TEM image revealed of helical conformation with increasing degree of polymer-
the macrocyclic species made of multiple triple-stranded ization (Fig. 19). Fig. 20 shows the observed and calculated
(1!3)-h-D-glucan chains in some cases after a cycle of de- SAXS profiles of laminarahexaose together with a snap-
naturation-renaturation process [76]. shot of a simulated chain. Although laminarahexaose is not
Laminaran is produced by Laminaria seaweeds, and long enough to show the characteristics of helical confor-
contains a small amount of h(1!6)-branched D-glucose mation, the observed SAXS profile is in good agreement
residues and alkyl groups at reductive ends. The confor- with the simulated scattering profile. The observed profile
mation of laminara oligosaccharides was characterized in has a smooth shoulder at q = 0.2–0.25 Å1, whereas a
aqueous solution by the combined method of small-angle simulated profile shows a slight peak at q = 0.2 Å1 due to
x-ray scattering and Monte Carlo simulation [77]. The a quasi-helical structure. The deviation of the observed
conformational energy map of laminarabiose (Fig. 7e) profile from the simulated one is probably due to hydra-
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

shows four local minima including two global minima tion, which is not properly taken into account in the
around (/, w) = (0j, 50j) and (/, w) = (30j, 0j). The simulation. The radius of gyration RG and the cross-
crystallographic data of laminarabiose and laminarabio- sectional radius of gyration RG,c are consistent with the
side derivatives (except for methyl b-D-laminarabioside respective values evaluated from observed and simulated
and h-D-laminarabiose) confirm that two dihedral angles profiles—7.8 Å (RG) and 3.0 Å (RG,c) from the observed
/ and w with respect to the glycosidic bond fall in one of the profile for laminarahexaose, and 7.7 Å (RG) and 3.4 Å
global minima in the conformational energy map of lam- (RG,c) from the simulation.
inarabiose (see Table 2). w is twisted by the formation of A triple-helical structure has been proposed for (1!3)-
intramolecular hydrogen bond between O4V and O5, which h-D-glucan [78,79]. For example, the molecular and crystal
is prevented by introducing acetate substituents. The glob- structure of the anhydrous curdlan and its hydrated form
al minima indicate the helical conformation of laminaran, was determined by combined x-ray diffraction from ori-
which will be interrupted by the other local minima at (20j, ented curdlan fibers and stereochemical model refinement
170j) and (160j, 10j). [80]. Here both hydrous and anhydrous forms assume a
Over 500 chains were generated to constitute a statis- right-handed triplex (sixfold triple-helical conformation)
tical ensemble of laminara-oligomers according to the crystallized in a hexagonal unit cell with the interstrand
scheme shown in Fig. 8, and the average scattering factor O2: : : O2 hydrogen bonds. Curdlan is believed to assume a

Figure 19 Scattering profiles calculated for laminara-oligosaccharides as a function of the degree of polymerization.

Copyright 2005 by Marcel Dekker


22 Kajiwara and Miyamoto
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 20 Small-angle x-ray scattering profile from laminarahexaose in water (observed and simulated as indicated in the
Figure). A stereoview of a simulated laminarahexaose chain is shown on the right.

single- or triple-helical sevenfold conformation by swelling, Gel is formed in the aqueous solution of (1!3)-h-D-
where a chain is expanded along the chain direction to glucans, but its mechanism seems to differ from linear and
increase the helix repeat distance to 22.7 Å from 17.6 Å (in branched species. Curdlan low-set gel is prepared by
anhydrous form) or 18.8 Å (in hydrous form). Regular or heating a slurry (>0.5% w/v) to above 60jC, and will be
irregular short-branch substitutions on the main chain O6 high-set with annealing at 95jC [84]. Gelation is suggested
hydroxyls seem not to affect the triplex structure as exem- to proceed with breaking hydrogen bonds to solubilize
plified by scleroglucan [81], schizophyllan [75], and len- curdlan and reforming intermolecular hydrogen bonds
tinan [82], which retain a triplex structure even in aqueous subsequently to consist the junction zones. Hydrophobic
solution. It is interesting to note that the dihedral angles / interaction promotes the intermolecular association of
and w of curdlan polymorphs are similar to those of the curdlan at elevated temperatures to form stronger high-
acetylated derivatives of laminarabiose or laminarabioside set gel. Thus curdlan gel is supposed to contain both
(Table 2). Similar / and w values are also evaluated from liquid-like (composed of flexible chains) and solid-like
the molecular structure of the tetrasacharride (1!6) (composed of associated chains) domains. 13C NMR was
branched (1!3)-h-D-glucan [83]. applied to curdlan gel where various methods (including

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 23

CP/MAS, broadband coupling, and MAS) were employed of the anhydrous form appears at the early stage of
to obtain the signals from the domains of different molec- annealing, and eventually the transformation takes place
ular motions [85]. Fig. 21 shows the 13C NMR spectra of from the anhydrous to the hydrous form.
curdlan hydrate and gel recorded by various methods. The 13C NMR of branched (1!3)-h-D-glucan gel such
Here a conventional high-resolution NMR coupled with as lentinan and schizophyllan shows the characteristic
broadband decoupling confirms that the liquid-like do- peaks of the triple helix [87], but the peaks corresponding
main is composed of single-helical chains that are flexible to the liquid-like domain disappear by gelation [82]. Thus
and undergo free molecular motion. The intermediate the gelation of h(1!6) branched (1!3)-h-D-glucans is
domain is also composed of single chains as indicated by mainly due to partial association of triple-helical chains.
high-power dipolar decoupling with magic angle spinning The gelation of schizophyllan is promoted by the
(MAS). The CP/MAS spectrum reveals a small amount of presence of sorbitol [88] where thermoreversible optically
triple helices visible in the solid-like domain as shown by transparent gel is formed by lowering the temperature.
an arrow (a C5 signal from the triple helix) in Fig. 21, but However, the small-angle x-ray scattering (SAXS) profile
otherwise gives the characteristics of the swollen sevenfold from the schizophyllan/sorbitol system shows less-marked
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

helical form of solid curdlan with C3 at 87 ppm and no change by the sol–gel transition. The 1.5% aqueous solu-
peak at 79 ppm. Annealing at elevated temperatures results tion of schizophyllan containing 4 M sorbitol is sol at 60jC
in the increase of the fraction of anhydrous (in a later stage but forms transparent gel at 5jC. The SAXS profiles from
hydrous) sixfold helical domains and the decrease of the the solution at respective temperature were analyzed in
swollen form portion [86]. The NMR observation indicates terms of a modified broken rod model [89], which reads
that curdlan undergoes gelation by forming quasi-cross-
linking domain composed of single helical chains associ-
ated hydrophobically after swelling at lower temperatures, X 4J12 ðqRci Þ
and subsequently, by increasing the triple-helical fraction q2 IðqÞc pqwi MLi  þ const: ð24Þ
at elevated temperatures. The triple-helical conformation i ðqRci Þ2

Figure 21 13C NMR spectra of curdlan hydrate (A) and gel (B–D), observed by CP/MAS (A, D), by broadband decoupling (B),
and by MAS (C).

Copyright 2005 by Marcel Dekker


24 Kajiwara and Miyamoto

where wi, MLi, and Rci denote the weight fraction, the broken rod model [Eq. (24)] where each component is
linear mass density, and the cross-sectional radius of the replaced with a triple helix and a single coil of the
rod-like component i, respectively. J1(x) is the first-order schizophyllan molecular model. The atomic radius is
Bessel function, and the constant term accounts that the reduced to half of the van der Waals radius to account
rod-like components are connected by a free joint. The for the smaller cross-sectional radius. The inclusion of a
model takes into account the heterogeneity with respect to constant term is necessary, so that schizophyllan triple
the cross-section. The results are shown in Fig. 22 in two helices are speculated to disentangle into single chains that
types of Guinier plots. Schizophyllan assumes a triple- act as a free joint. Sorbitol breaks intramolecular hydro-
helical conformation in water, and undergoes no confor- gen bonds of schizophyllan triple helices, and solvates the
mational change by decreasing the temperature from 60jC broken parts to form a cross-linking junction by intermo-
to 5jC as shown in Fig. 22a,b, where the scattering profile lecular hydrogen bonding through sorbitol. An apparent
was calculated from the molecular model of schizophyllan smaller atomic radius observed at 5jC is probably due to
triple helix. The cross-sectional radius of gyration of solvated sorbitol reducing the contrast between solvent
schizophyllan is estimated as 6.4 Å. When sorbitol is and solute.
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

added, the cross-sectional Guinier plots yield a smaller


apparent cross-sectional radius (5.1 Å), which becomes
even smaller (4.6 Å) by gel formation at a lower temper- D. Cyclic and Linear (1!2)-B-D-Glucan
ature (5jC). Here the SAXS scattering profile at 60jC was
fitted with the molecular model of (1!3)-h-D-glucan triple Gram-negative bacteria such as Agrobacterium and Rhizo-
helices with no side chain (i.e., curdlan-type triple helix). bium [90,91] are known to produce a cyclic (1!2)-h-D-
The SAXS profile at 5jC can be fitted by a modified glucan referred to as cyclosophoran. The DP value (the

Figure 22 Small-angle x-ray scattering profiles from the 1.5% aqueous solution of schizophyllan and schizophyllan/4 M
sorbitol at 60jC (a) and 5jC (b). The solid lines represent the scattering profiles calculated according to the molecular model (c)
and Eq. (23).

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 25

degree of polymerization) of cyclosophoran varies from 17


to 24 depending on the bacterial strain; the largest DP
reported is 40. Cyclosophoran is thought to act as a
regulator of the osmotic balance between the cytoplasm
and the periplasmic space for bacteria to adapt the change
in environmental osmolality [92] or to mediate the bacte-
rium–plant host [93] interactions during the infection of the
host. Although the exact physiological role of cyclic (1!2)-
h-D-glucan is a matter of speculation, its physiological
function is assumed to be closely related to its conforma-
tion [94]. The conformation of (1!2)-h-D-glucan has been
extensively investigated by computer modeling and 13C/1H
NMR [95], but the homopolymeric nature and conforma-
tional identity of the glucose residues obscure the structure
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

determination by NMR.
The conformation of cyclic and linear (1!2)-h-D-
glucans was investigated by the combination of the Monte
Carlo simulation and SAXS [96]. Cyclosophoran mixtures
produced by Agrobacterium radiobactor and Rhizobium
fphaseoli were fractionated into nine fractions from DP
= 17 to 25 (each designated as CS17 to CS25) by high-
performance liquid crystallography (HPLC). Linear
(1!2)-h-D-glucans (designated as LS 19 and LS2 1 accord-
ing to DP) was prepared by acid hydrolysis of CS21 and
subsequent fractionation by HPLC.
Small-angle x-ray scattering (SAXS) was observed
from the aqueous solutions of cyclic glucans (CS17 to
CS24) and linear glucans (LS19 and LS21) at 25jC where
the concentration was varied from 10 to 40 mg/mL (for the
cyclic glucans) or from 12.5 to 25 mg/mL (for the linear
glucans). The observed range of the magnitude of the
scattering vector was from q = 2.50  102 Å1 to q =
0.375 Å1, which is equivalent to the Bragg spacing from
251 to 16.8 Å. The observed SAXS profiles reveal the
structural difference of cyclic and linear (1!2)-h-D-glucan
chains in aqueous solution in the region of q > 0.15 Å1
(Fig. 23). The radius of gyration, RG, was evaluated from
the initial slope of the Guinier plots as summarized in Table
5. The cross-sectional radius of gyration Rc was evaluated
from the Guinier plots for cross section [Eq. (6)] in the case
of linear (1!2)-h-D-glucan chains or the thickness T from
the Guinier plots for thickness [Eq. (7)] in the case of cyclic
(1!2)-h-D-glucan chains.
The results indicate that a cyclic (1!2)-h-D-glucan
chain assumes the shape of a flat disk and a linear homolog
the shape of a cylinder. The compact conformation of a cy-
clic (1!2)-h-D-glucan chain is confirmed from the smaller
radius of gyration in comparison with a corresponding
linear (1!2)-h-D-glucan chain.
(1!2)-h-D-glucan chains were generated by the Mon-
te Carlo method, consistent with the disaccharide confor-
mational energy map (Fig. 7f ). The region of the energy
well is specified as the conformation +A for w > 20j, or
A for w < 20j according to York [95]. The glucose
residue is assumed to be rigid, and the conformational
energy map for h-sophorobiose is calculated by the mo-
lecular mechanics as a function of the dihedral angles /
Figure 22 Continued.
and w defined as / = h[H1, C1, O, C2V] and w = h[C1, O,
C2V, H2V]. Nonbonded van der Waals interactions and

Copyright 2005 by Marcel Dekker


26 Kajiwara and Miyamoto
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 23 Small-angle x-ray scattering profiles of cyclic and linear (1!2)-h-D-glucan chains in (a) Guinier plots [ln P( q) plotted
against q2] and (b) Kratky plots [ q2P( q) plotted against q].

electrostatic interactions due to partial charges are taken The scattering factors are calculated according to Eq.
into account in the calculation. The occurrence probability (12) from the atomic coordinates of generated chains in an
is given by the Boltzmann factor for a pair of (/, w) ensemble, with the radii of carbon and oxygen atoms being
normalized with respect to the sum of the Boltzmann taken to be 1.67 and 1.50 Å, respectively. Here all the O6
factors for all pairs of (/, w), whereas the bond angle s atoms of the glucose unit are affixed to the pyranose ring at
at the glycosidic oxygen is fixed at 113.6j. Among the a gauche–trans (gt) position with respect to the torsion
chains generated by the Monte Carlo method, those with angle h[O5, C5, C6, O6] and the torsion angle h[C4, C5, C6,
the end-to-end distance less than 1.5 Å are collected to O6], respectively.
compose an ensemble of cyclic (1!2)-h-D-glucan chains. Fig. 24 shows a reasonable agreement of the simulated
An ensemble of linear (1!2)-h-D-glucan is composed of scattering profile for cyclic (1!2)-h-D-glucans with that
500 chains. observed by SAXS, where the scattering profiles calculated

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 27

Table 5 The Radius of Gyration RG, the Cross-Sectional large molecule for the formation of an inclusion complex.
Radius of Gyration Rc, and the Thickness T of Cyclic and All the glucosidic linkage torsion angles are found within
Linear (1!2)-h-D-Glucan Chains Evaluated from the the region A of the conformational energy map (Fig. 7f )
Corresponding Guinier Plots of SAXS Data with 13 linkages in the region +A and 7 linkages in the
region A where no special mode is observed for arranging
Sample
code RG (Å) Rc (Å) T (Å)
+A and A.
The Monte Carlo simulation for linear (1!2)-h-D-
CS17 7.8 – 10.0 glucans yields less satisfactory results with respect to the
CS18 8.1 – 10.0 scattering profile (Fig. 24). Although the Monte Carlo
CS19 8.5 – 10.0 simulation yields a consistent value of the radius of gyra-
CS20 8.3 – 10.0 tion with an observed one, the deviation in the scattering
CS21 8.6 – 10.5 profile becomes apparent at u (u qRG) > 1.3. A good
CS22 8.4 – 10.7 linearity observed in the cross-sectional Guinier plots [Eq.
CS23 8.9 – 10.8 (6)] indicates a cylindrical shape of a linear (1!2)-h-D-
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

CS24 10.6 – 9.8 glucan chain as shown in Fig. 26 with space filling models.
LS19 11.1 5.9 – The cross-sectional diameter is evaluated as 11.8 Å (LS 19)
LS21 12.0 6.6 –

from two elementary models (a rigid ring [97] and a flexible


Gaussian ring [98]) are shown for comparison. The particle
scattering factors of two elementary models are analytical-
ly given, respectively, as
X
N
sin½ð2uÞsinðpn=NÞ

Prigid ðqÞ ¼ N1 ð25Þ


n¼1
ð2uÞsinðpn=NÞ

and
 2

u u
Pflexible ðqÞ ¼ ð2=uÞexp  D ð26Þ
4 2
where u and D(x) denote the reduced scattering vector and
the Dawson integral defined as
uuqRG ð27Þ
ðx
DðxÞ ¼ expðt2 Þdt ð28Þ
0

The observed SAXS profiles from cyclic (1!2)-h-D-


glucans exhibit a certain chain stiffness in comparison with
the profiles calculated from the elementary models, where
the conformational freedom is suppressed by linking end-
to-end. The Monte Carlo simulated scattering profiles
reproduce the observed SAXS profiles reasonably well
except for the deviation at the higher q region. The devia-
tion at the higher q region may indicate that the effect of
hydration should be taken into account, as no interference
term due to the solvent–solute interaction is incorporated
in the scattering profile calculation. However, the intro-
duction of an apparent scattering unit smaller than 1.67 or
1.50 Å (for a carbon atom or an oxygen atom, respectively)
reduces the deviation from the observed profile at u (u
qRG) > 3 without any physical significance. Figure 24 Monte Carlo simulated and observed small-angle
A typical conformation is shown in Fig. 25 with space X-ray scattering profiles (the Kratky plots) of CS17 (a) and
filling models, which reveal an irregular doughnut-like ring CS21 (b). The open circles denote the SAXS data, the solid
with the thickness of 10 Å. The diameter of the CS21 ring lines the Monte Carlo simulated profile, the dotted lines the
annulus is about 4–5 Å; that is, the cavity in a cyclic (1!2)- calculated profile for a rigid ring [Eq. (24)], and the broken
h-D-glucan chain seems too small to embrace a relatively lines the calculated profile for a Gaussian ring [Eq. (25)].

Copyright 2005 by Marcel Dekker


28 Kajiwara and Miyamoto
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 25 Space filling models of cyclic and linear (1!2)-h-D-glucan chains generated by the Monte Carlo method. CS21 (left)
and LS21 (right) are seen from the top and the side. Hydrogen atoms are not included.

and 13.2 Å (LS21) from the SAXS data, or as 10.6 Å (LS considered to account for the interference effect at the
19) and 11.3 Å (LS21) from the Monte Carlo simulation. higher q region. We have observed an opposite tendency
The discrepancy between the two sets of corresponding of the interference effect at the higher q region in cyclic and
cross-sectional diameters evaluated independently ac- linear (1!2)-h-D-glucan aqueous solutions. Although no
counts to some extent for the deviation of the scattering physical significance is known at this stage, the apparent
profiles at the larger q region. When the observed thicken- difference in the size of the scattering units may explain the
ing is compensated by introducing larger radii for C and O mechanism of the physiological function found only in
atoms than the equivalent van der Waals radii, the consis- cyclic (1!2)-h-D-glucans.
tency of the scattering profiles at the larger q region also
improves. In Fig. 24, the scattering profile calculated from
larger apparent scattering units (C and O atoms) shows a IV. SUPRAMOLECULAR STRUCTURE OF
better fitting to the observed SAXS profile. The cross- POLYSACCHARIDES IN SOLUTION
sectional diameter becomes approximately 12% larger by AND GEL
doubling the radius of the scattering units.
Although the Monte Carlo simulation yields reason- Polysaccharides assume not a completely random a but
ably consistent results as a whole, more detail inspection quasi-ordered conformation in solution as shown in the
reveals that the interaction with water (solvent) needs to be preceding sections. This particular characteristic results in

Figure 26 Monte Carlo simulated and observed small-angle X-ray scattering profiles (the Kratky plots) of LS19 (a) and LS21
(b). The open circles denote the SAXS data, the solid lines the Monte Carlo simulated profiles with the radii of scatterers 1.67 Å
(C) and 1.50 Å (O), the broken lines the Monte Carlo simulated profiles with the radii of scatterers 3.34 Å (C) and 3.00 Å (O).

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 29

the formation of quasi-ordered domains in polysaccharide


solutions. As a consequence, many polysaccharides form
physical gel where the cross-linking domain is constituted
of the quasi-ordered assembly of polysaccharide chains.
The size of the quasi-ordered domain varies from a few
nanometers to a few hundred nanometers, and the overall
appearance of polysaccharide solutions and gels is deter-
mined by its size, structure, and the mode of its connection.
This section deals with the structural characterization of
quasi-ordered domains formed by oligo- and polysaccha-
rides in solution.
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

A. Thermotropic Liquid Crystal of Cellulose


Derivatives
Polysaccharide is hydrophilic and water-soluble in most
cases as one of the most fundamental molecular character-
istics. However, for example, cellulose is not water-soluble.
Figure 27 Isotropization temperature Ti and melting The lack of solubility of cellulose in water is caused by
temperature Tm as a function of side-chain length index for numerous intra- and intermolecular hydrogen bonds. Cellu-
fully substituted cellulose derivatives; tri-O-alkyl cellulose (.) lose is a linear polysaccharide consisting of anhydroglucose
and cellulose trialkanoate (E). units linked by (1!4)-h-glucosidic bonds. The equatorial

Figure 28 Representative structures of poly- and oligosaccharide-based liquid crystals: (a) chiral nematic (cholesteric), (b)
hexagonal columnar, (c) discotic hexagonal columnar, and (d) smectic A phase. Here (d) is supposed to be a structure for 1-O-
alkyl-h-D-cellobiosides.

Copyright 2005 by Marcel Dekker


30 Kajiwara and Miyamoto

other hand, MC samples prepared in a homogeneous phase


with a nonaqueous solvent system shows no sol–gel tran-
sition.
In general, the polymers possessing polar hydrophilic
and nonpolar hydrophobic groups can dissolve in water, if
water is a good solvent for the hydrophilic groups and any
neighboring hydrophobic group is hydrated. However,
when the temperature is increased, hydrogen bonds are
weakened and hydration is reduced in the aqueous solu-
tion; that is, solvated hydrophobic groups lose their weakly
bound water at higher temperatures. Consequently, they
coalesce into a water-insoluble phase and the polymer
precipitates at a certain temperature termed as a lower
critical solution temperature (LCST) [112,113]. The cloud
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 29 Transition temperature Ti (.) and Tm (o) for point is defined as the temperature at which turbidity is
narrow fractions of fully decanoated cellulose. observed while heating a dilute solution slowly.
Cellulose and chitosan are rich sources of lyotropic
and thermotropic liquid crystals (LCs) [114,115]. As al-
ready described, both are linear, stereoregular polymers of
configuration of the (1!4)-h-glucosidic bond predestines the D-glucose and D-glucosamine, respectively, linked by a
stretched chain conformation of cellulose, which in turn (1!4)-h bond. This bonding together with the bulky
promotes the intra- and intermolecular hydrogen bonds
among their chains. The conformation of (1!4)-h-glucans
is discussed in Sec. III.B in some detail. Another character-
istic of cellulose is that almost all cellulose derivatives can
form lyotropic liquid crystals because of the semirigidity of
cellulose main chains. Chitin and its deacetylated derivative,
chitosan, are also (1!4)-h-linked homopolysaccharides, and
they can form lyotropic liquid crystals. Both cellulose and
chitin can form fibers having good mechanical properties.
Besides these, there occurs (1!4)-h-linked linear homopoly-
saccharide in nature.
The properties of cellulose derivatives depend not only
on the total degree of substitution (DS) and the molar
substitution (MS) but also on (1) the distribution of sub-
stituents in the anhydroglucose (AHG) units (i.e., the
relative DS and MS at three different types of hydroxyl
groups), (2) the distribution of substituents along the
cellulose chain, and (3) the distribution of DS and MS.
The distribution within glucose unit arises because the
three hydroxyl groups of the glucose residue generally
differ in reactivity. This distribution can be estimated by
1
H and 13C NMR methods. On the other hand, the
nonuniformity of the distribution along the chain is caused
by heterogeneous reaction. In the case of copolymers, the
control of compositional distribution is known to be very
important to control their physical properties. The prob-
lem is equivalent to the control of the distribution of DS
and MS values in the cellulose derivatives, although at
present it is not easy to estimate their distributions. At all
event, the substituent distribution control play a major role
for the higher functionalization of cellulose [96–109].
An example is shown in the water solubility of the
derivatives. Commercial methyl cellulose (MC) is water-
soluble and shows a thermally reversible sol–gel transition
in aqueous solution [103,110–112]. Commercial products
are usually prepared by the so-called alkali cellulose pro-
cess, which is based on a heterogeneous reaction. On the Figure 30 Four types of monomer units of xyloglucan.

Copyright 2005 by Marcel Dekker


Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 31 Snapshots of xyloglucan monomer units with (1!4)-h-D-glucan spines on the right.

Copyright 2005 by Marcel Dekker


32 Kajiwara and Miyamoto
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 32 Small-angle X-ray scattering observed (black circles) and calculated (solid or dotted lines) from four types of
tamarind seed xyloglucan monomer units. Since octasaccharide has two isomers (i.e., XXLG and XLXG), two dotted lines
correspond to the respective scattering profiles from XXLG (above) and XLXG (below) and the solid line represents the
calculated scattering profile as an average from the two isomers (30% XXLG and 70% XLXG in this instance).

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 33

monomeric units forces the molecules to assume an essen- the polymeric phase, it is parallel to the column axis. The
tially flat and extended conformation, affording these transition from the perpendicular to the parallel orienta-
polymers and their oligomers mesogenic characters. Sev- tion of the molecular axis is expected to occur at a DP of
eral studies have been reported on lyotropic and thermo- around 10, as Fig. 29 suggests, but it cannot be actually
tropic cellulose derivatives. For a recent literature review observed, as the transition temperature will be well below
on the lyotropic and thermotropic systems of these deriv- the melting temperature Tm.
atives, the readers are referred to Ref. [116]. However, This behavior of the alkyl ester derivatives forms a
these studies are mostly concerned with HIPC-related remarkable contrast to that of the alkyl ether derivatives,
derivatives. Those are chemically disordered polymers, which, as already described, form a chiral nematic phase
the molecular structure–property relationships of which when the chain lengths are sufficiently large. (Short alkyl
are difficult to establish. Here we describe the main ethers show no liquid crystallinity.)
features of thermotropic mesophases exhibited by fully Acylated derivatives of chito-oligosaccharides also
substituted derivatives, which are chemically ordered poly- form a discotic hexagonal phase [121]. Owing to the
mers [115]. hydrogen bonding interaction of the amide group, their
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Fig. 27 demonstrates the phase behavior of two types phases have a higher stability than those of the cello-
of fully substituted cellulose derivatives, trialkyl celluloses counterparts. The stability of the discotic hexagonal phase
and cellulose trialkanoates [117]. The abscissa scale N of the chito-compounds decreases with increasing DP of
denotes the side chain length (i.e., the number of the C the main chain, and the derivative with a DP of 6 and a side
and O atoms forming the side chain skeleton). As N chain carbon number of 14 is likely to form a discotic
increases, the melting temperature Tm decreases drastically nematic phase [122].
at first and seems to level off or slightly increases for N z 10,
in all cases. Thus it is evident that the introduction of alkyl B. Supramolecular Structure in Xyloglucan Gel
side chains effectively lowers the melting temperature of
cellulose, but as the side chains become longer and the side Xyloglucan is a general term applied to nonstarch plant
chain fraction becomes larger, the melting behavior of the polysaccharides composed of a (1!4)-h-D-glucan with
systems becomes more governed by that of the side chain (1!6)-a-branched xylose, which is partially substituted
components. by (1!2)-h-galacto-xylose [123]. Xyloglucan is normally
Although the melting behaviors of the two mentioned contained in plant seed, and its flour has been traditionally
systems are similar to each other, their mesomorphic used as a food additive in everyday life. Four types of
properties are very different [118]. All cellulosic LCs, monomer units are allocated to xyloglucan as designated as
lyotropic or thermotropic, that were known in former XXXG, XXLG, XLXG, and XLLG. Each unit is com-
times were cholesteric (or nematic). A cholesteric (chiral posed of a sequence of four (1!4)-h-D-glucans but differs
nematic) phase is characterized by the director field prop- in the number of galactose side chain (Fig. 30), so that the
agating in one direction, forming a helix of pitch P (Fig. total number of sugar residues is 7 (in XXXG), 8 (in XXLG
28a). The ether derivatives form a cholesteric phase in the and XLXG), or 9 (in XLLG). Tamarind indica (TSP) and
vertically hatched region in Fig. 27, while the ester deriv-
atives form quite a different phase in the horizontally
hatched region. This phase is of a columnar hexagonal
type [117].
The N dependence of the isotropization temperature
Ti is also different for the ether and ester derivatives. The Ti
of the ethers decreases with N rather monotonically,
whereas that of the esters goes through a small maximum
(Fig. 27). To be emphasized here is that though the
difference in the chemical structures of these polymers is
rather small, the observed differences in their mesophase
properties may be surprisingly large.
Fig. 29 shows the chain length dependence of transition
temperatures Ti and Tm obtained for narrow fractions of
fully decanoated cellulose and its oligomers [115,118,119].
This phase diagram consists of four regions—a crystalline
solid region K, an isotropic liquid region I, and two
mesomorphic regions D and C. The oligomeric phase D,
which is relevant to homologs with DP < 5, is a discotic
hexagonal columnar phase, as illustrated in Fig. 28c [120].
On the other hand, the polymeric phase C, relevant to
homologs with DPw > 20 corresponds to the structure
given in Fig. 28b [117]. Thus in the oligomeric phase, the Figure 33 Sol-gel transition temperature diagram of enzy-
molecular axis is perpendicular to the column axis, while in matically degraded tamarind seed xyloglucan solution (1%).

Copyright 2005 by Marcel Dekker


34 Kajiwara and Miyamoto

Detarium senegalense are two common xyloglucans com- are seen in Fig. 31, together with respective (1!4)-h-D-
mercially available as a food additive, which have the same glucan spines (without branches) shown on the right side.
fundamental (1!4)-h-D-glucan spine with the different Here (1!4)-h-D-glucan spines are seen to assume rather
content of galactose with respect xylose and glucose. flat zigzag conformation, and xylose and galacto-xylose
Four types of xyloglucan monomers were simulated by branches to extend and fold upright on the (1!4)-h-D-
molecular dynamics, and the particle scattering function glucan flat surface. In detail, nonasaccharide (xyloglucan
was calculated according to Eq. (12) from the atomic monomer composed of nine sugar residues) exhibits a flat-
coordinates of the simulated xyloglucan monomer units arched spin conformation, while octasaccharide (xyloglu-
[124]. The snapshots of simulated xyloglucan monomers can composed of eight sugar residues) and heptasaccharide
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 34 Molecular model for xyloglucan aggregate.

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 35
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Figure 35 Scattering profile from the model xyloglucan aggregates. The number of aggregated chains is indicated in the Figure.

Copyright 2005 by Marcel Dekker


36 Kajiwara and Miyamoto

Table 6 Evaluated Parameters for the Tree-Like Model as a Function of Reaction Time
Weight fraction of
Reaction time
(min) ( f1)a b (nm) Single chain 14-Chain aggregate

0 0.45 6.0 1.0 0.0


40 0.5 11.0 0.49 0.51
57 1.0 13.5 0.24 0.76
74 1.04 13.7 0.13 0.87
91 1.06 14.0 0.07 0.93
108 1.07 14.0 0.05 0.95

(xyloglucan composed of seven sugar residues) assume a course of the enzymatic degradation. As the reaction time
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

slightly twisted conformation. The loss of galactose side proceeds, the scattering profile at the medium q range
chains results in the increase of hydrophobicity, and seems becomes flat in the Kratky plots, while the profile at a
to twist the backbone. A similar conformation of galacto- higher q region remains almost invariant, exhibiting the
xylose side chains is observed by Levy et al. [125] from the characteristics of the rod-like scattering. The scattering
simulation by assuming a fixed flat (cellulose-like) or curve at q ! 0 indicates to upturn after 57 min. This
twisted (cellobiose-like) backbone conformation. Simu- symptom is a typical behavior of the scattering from gelling
lated scattering profiles are compared with the observed systems [50]. Here gelation is assumed to take place accord-
small-angle x-ray scattering profiles from tamarind seed ing to the classic Flory-Stockmayer polyfunctional poly-
xyloglucan monomer units in Fig. 32. The consistency of condensation scheme [127], and the scattering intensity
the simulated and observed results confirms the reality of from such a system is given as
the chain conformation visualized in Fig. 31.
As gelation is prevented by the steric hindrance and IðqÞ ¼ A2 ðaÞð1 þ a/Þ=½1  ðf  1Þa/

hydrophilicity of (1!2)-h-D-galacto-xylose branches, the ð29Þ


 
enzymatic degradation by h-galactosidase promotes gela- / ¼ exp b2 q2 =6
tion of xyloglucan aqueous solution [126]. At room tem-
perature, tamarind seed xyloglucan aqueous solution Here f denotes the functionality of the cross-linking
forms gel at about 45% release of galactose residues, but domain (the number of branches from a domain), a the
this gel will melt at an elevated temperature. The resulted
gel is opaque and has a unique property to have two
melting points at lower and higher temperatures as shown
in Fig. 33. The loss of (1!2)-h-galactose proceeds with a
reaction time and more aggregation will take place to form
cross-linking domains during the course of enzymatic
degradation. Here the aggregation will take place laterally
at the portion of xyloglucan chains lacking in terminal
galactose. The aggregation is expected to form a quasi-
ordered domain composed of laterally arranged xyloglu-
can chains. Because the conventional analysis of the ob-
served small-angle x-ray scattering profiles indicate the
formation of flat objects with 1.1-nm thickness upon
gelation [50], the molecular model of a quasi-order domain
was constructed by stacking cellulose-like (1!4)-h-D-glu-
can chains in parallel as shown in Fig. 34, and the scattering
profile was calculated according to Eq. (12). Here the
model consists of 14 xyloglucan chains each composed of
40 (1!4)-h-D-glucans with 30 (1!6)-a-xylose branches in
the sequence of XXXG. (1!2)-h-galactose terminal
groups are eliminated because the quasi-ordered domains
are formed by the loss of these terminal groups. The Figure 36 Observed and calculated scattering profiles at
scattering profile calculated from the model aggregate various reaction times (indicated in the Figure). Symbols
reveals distinguished peaks in the Kratky plots as the represent the observed small-angle x-ray scattering intensities
number of chains in the aggregate increases (Fig. 35). and solid lines the scattering profiles calculated from Eq. (29)
The small-angle x-ray scattering was observed from with A2( q) corresponding to the domain composed of 14
1% tamarind seed xyloglucan aqueous solution during the aligned xyloglucan chains.

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 37

stacked in parallel to constitute a cross-linking domain),


( f  1)a should be regarded as a parameter to specify the
average branching degree, where ( f  1)a = 1 indicates a
gel point. The analysis involves three parameters—b, ( f 
1)a, and the weight fraction of the domain composed of 14
aligned xyloglucan chains. The evaluated parameters are
summarized in Table 6, and indicate that gelation takes
place after 57 min of the reaction time in the present system.
Table 1 also indicates that about 3/4 of chains are involved
in the quasi-ordered domain at a gel point, and more single
Figure 37 Chemical structure of lactose-carrying polystyrene. chains are incorporated into the quasi-ordered domains as
further reaction takes place mainly on single chains after
gelation. The thickness of the aggregated domain does not
grow from 1.1 nm, and thus the domain seems to be
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

composed of a single layer of stacked xyloglucan chains.


conversion (the probability that an arbitrary chosen unit is At the end of reaction, most of the chains are incorporated
reacted), b2 the mean square average of the distance
between the neighboring scattering units, and A2( q) the
scattering amplitude of each scattering unit [i.e., A2( q) = 1
in the case of a point]. The scattering amplitude A2( q) in the
gelling system could be represented by the scattering
factors of the domain composed of aggregated chains or
a single chain of a certain length (Fig. 35). For simplicity,
the gelling system of enzymatically degrading tamarind
seed xyloglucan is assumed to consist of two phases of
single chains (a dilute phase) and the domains of 14 parallel
stacked chains (a condensed phase), and the observed
scattering profiles are analyzed according to Eq. (29).
The results are summarized in Table 6 and Fig. 36. The
calculated scattering profiles are consistent with the ob-
served profiles over the entire time course of enzymatically
degrading reaction, although the condensed phase is not
necessarily composed of 14 xyloglucan chains. Because no
explicit number is known for the functionality f of a
domain ( f should be equal to 24 if exactly 14 chains are

Figure 39 (a) Scattering profiles for three degrees of po-


lymerization calculated from the molecular model of lactose-
Figure 38 Small-angle x-ray scattering profile of VLA29, carrying polystyrene and (b) the fitting example for VLA92
VLA92, and PVLA. The concentrations of each solution are where the symbols denote the observed SAXS intensities and a
the same (2 wt.%). sold line represents the calculated profile for DP = 92.

Copyright 2005 by Marcel Dekker


38 Kajiwara and Miyamoto

in the thin domains, and the gel seems to be constituted of


the cell-like network.

C. Glycoconjugate Synthetic Polymer


Recent advances in the precise polymerization technique
has resulted in synthesizing novel functional polymers
mimicking biopolymers. Hybrids of synthetic polymers
and biopolymers are of a particular interest, as the hybrid
may enhance the characteristics of parent polymers. A
series of glycoconjugate polystyrene derivatives have been
synthesized with varying the types of pendant oligosac-
charides [128]. The synthesized glycoconjugate polystyrene
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

derivatives are amphiphilic with hydrophilic pendant oli-


gosaccharides densely grafted on hydrophobic polystyrene
main chain. Highly concentrated multiantennary glycol
signals along hydrophobic main chain were in fact found
to enhance the interaction with various types of carbohy-
drate-binding proteins, and the synthesized glycoconjugate
polymers to function as a highly sensitive ligand [129]. For
example, lactose-carrying polystyrene is suitable for the
incubation of liver cells and the drug delivery systems [130].
Amphiphilic glycoconjugate polystyrene derivatives
are water-soluble, as glycoconjugate polymers will form a
single-molecule micelle in water to prevent precipitation.
Three lactose-carrying polystyrene derivatives (the chemi-
cal structure is shown in Fig. 37) were prepared by radical Figure 40 Simulated molecular model of lactose-carrying
homopolymerization or living radical polymerization of polystyrene.
vinylbenzyl lactose amide [131]. The one prepared by
radical homopolymerization (PVLA) has a high degree of
polymerization with a broad molecular weight distribu-
use of Cerius2 ver 3.5, and the particle scattering factor was
tion, while the other two (VLA29 and VLA92) prepared by
calculated according to Eq. (12) for three lactose-carrying
living radical polymerization have the degrees of polymer-
polystyrenes from the atomic coordinates of simulated
ization of 29 and 92, respectively, with a narrow molecular
molecular models. The results are shown in Fig. 39 includ-
weight distribution around 1.2. Small-angle x-ray scatter-
ing the fitting example for VLA92. The consistency of the
ing from the aqueous solutions of those samples (Fig. 38)
calculated profile to the observed SAXS is satisfactory in
reveals an identical profile at a high q region ( q > 0.1 Å1)
all three cases, where VLA29 (a low degree of polymeriza-
[132], indicating that the conformation of lactose-carrying
tion) can be represented by a shape of an ellipsoid rather
polystyrene is almost the same regardless of the molecular
than a cylinder. The simulated molecular model consists of
weight and the SAXS profile difference at a low q region is
a pseudohelical polystyrene backbone covered with lactose
due to the size of a whole molecule. The shape of both
side chains (Fig. 40). The simulated molecular model
VLA92 and PVLA is approximately represented by a
confirms that the pseudohelical conformation of polysty-
cylinder of the same cross-sectional radius as conformed
rene backbone is retained even at DP = 29. Because
from the cross-sectional Guinier plots of respective SAXS
polystyrene backbone is atactic, its conformation is ran-
intensities, whereas VLA29 is not long enough to reveal the
dom in principle but the backbone conformation is obliged
characteristics of a cylinder.
to assume pseudohelical by the amphiphilic character of
The cylindrical shape of VLA 92 and PLLA could be
the backbone (hydrophobic) and side chains (hydrophilic).
accounted for by a polystyrene spiral backbone with
In this context, lactose-carrying polystyrene forms a single-
protruding lactose side chains. Based on this conjecture,
molecule cylindrical micelle.
the molecular model of lactose-carrying polystyrene was
constructed first by assuming an arbitrary sequence of
trans–gauche (TG observed in the crystalline phase of
isotactic polystyrene), trans–trans (TT observed in the ACKNOWLEDGMENTS
crystalline phase of syndiotactic polystyrene), and trans–
trans–gauche–gauche (TTGG observed in syndiotactic The authors are indebted to Dr. Isao Wataoka, Dr.
polystyrene) for a polystyrene spiral backbone, and sec- Hidekazu Yasunaga, and Dr. Mitsuru Mimura for their
ondly by linking lactose side chains as shown in Fig. 37 valuable comments during the preparation of the manu-
[132,133]. Then, the molecular model was simulated by the script.

Copyright 2005 by Marcel Dekker


Progress in Structural Characterization of Functional Polysaccharides 39

REFERENCES 33. Andrew, E.R. Philos. Trans. R. Soc. Lond. 1981, A299, 505.
34. Hartmann, S.R.; Hahn, B.L. Phys. Rev. 1962, 128, 2042.
1. Burchard, W.; Meuser, F., In Plant Polymeric Carbohy- 35. Pines, A.; Gibby, M.G.; Waugh, J.S. J. Chem. Phys. 1973,
drates; Manners, D.L., Seibel, W., Eds.; The Royal Soci- 59, 569.
ety of Chemistry: Cambridge, 1993. 36. Bax, A.; Davis, D.G.; Sarkar, S.K. J. Magn. Reson. 1985,
2. Glatter, O.; Kratky, O., Eds. Small Angle X-ray Scattering; 63, 230.
Academic Press: London, 1982. 37. Bax, A.; Davis, D.G. J. Magn. Reson. 1985, 65, 355.
3. Freeman, R. A Handbook of Nuclear Magnetic Resonance; 38. Macra, S.; Ernst, R.R. Mol. Phys. 1980, 41, 95.
Longman Scientific Technical: Harlow, U.K., 1988. 39. Bax, A.; Davis, D.G. J. Magn. Reson. 1985, 63, 207.
4. French, A.D.; Brady, J.W., Eds. Computer Modeling of 40. Bothner-By, A.A.; Stephens, R.L.; Lee, J.-M.; Warren,
Carbohydrate Molecules, ACS Symposium Series 430; C.D.; Jeanloz, R.W. J. Am. Chem. Soc. 1984, 106, 811.
American Chemical Society: Washington, DC, 1990. 41. Kollman, P. Chem. Rev. 1997, 93, 2395; van Gunsteren,
5. Kacuráková, M.; Mathlouthi, M. Carbohydr. Res. 1996, W.F. Weiner, P.K., Wilkinson, A.J., Computer Simu-
284, 145. lation of Biomolecular Systems: Theoretical and Exper-
6. Gress, M.E.; Jeffrey, G.A. Acta Crystallogr. 1977, B33, imental Applications; Kluwer Academic: Dordrecht, 1997.
2490. 42. Koehler, S.E.H.; Saenger, W.; van Gunsteren, W.F. Eur.
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

7. Quigley, G.J.; Sarko, A.; Marchessault, R.H. J. Am. Biophys. J. 1988, 15, 197; J. Mol. Biol. 1987, 203, 241.
Chem. Soc. 1970, 92, 5834. 43. Woods, R.J.; Dwek, R.A.; Edge, C.J.; Fraser-Reid, B. J.
8. Takeda, H.; Yasuoka, N.; Kasai, N. Carbohydr. Res. Phys. Chem. 1995, 99, 3832; Woods, R.J. Reviews in
1977, 53, 137. Computational Chemistry; Lipkowitz, K.B. Boyd, D.B.,
9. Takeda, H.; Kaiya, T.; Yasuoka, N.; Kasai, N. Carbo- Eds.; VCH: New York, 1995.
hydr. Res. 1978, 62, 27. 44. Maple, J.R.; Dinur, V.; Hagler, A.T. Proc. Natl. Acad.
10. Perez, S.; Vergelati, C.; Tran, V.H. Acta Crystallogr. Sci., U. S. A. 1998, 85, 5350; Hwang, M.-J.; Ni, X.;
1985, B41, 262. Waldman, M.; Ewig, C.S.; Hagler, A.T. Biopolymers
11. Lamba, D.; Burden, C.; Mackie, W.; Sheidrick, B. 1988, 45, 435.
Carbohydr. Res. 1986, 153, 205. 45. Clark, A.H.; Ross-Murphy, S.B. Adv. Polym. Sci. 1987,
12. Noguchi, K.; Okuyama, K.; Kitamura, S.; Takeo, K. 83, 57.
Carbohydr. Res. 1992, 237, 33. 46. Wu, H.-C.; Sarko, A. Carbohydr. Res. 1978, 61, 7.
13. Ikegami, M.; Noguchi, K.; Okuyama, K.; Kitamura, S.; 47. Imberty, A.; Perez, S. Biopolymers 1988, 27, 1205.
Takeo, K.; Ohno, S. Carbohydr. Res. 1994, 253, 29. 48. Rappenecker, G.; Zugenmaier, P. Carbohydr. Res. 1981,
14. Perez, S.; Marchessault, R.H. Carbohydr. Res. 1978, 65, 89, 11.
114. 49. Muller, J.J.; Gemet, C.; Schulz, W.; Muller, E.-C.;
15. Debye, P. Ann. Phys. (Leipz.) 1915, 28, 809. Vorweg, W.; Damaschum, G. Biopolymers 1995, 35, 271.
16. Guinier, A.; Foumet, G. Small-Angle Scattering of X-rays; 50. Kajiwara, K.; Kohjiya, S.; Shibayama, M.; Urakawa, H.
Wiley: New York, 1955. Polymer Gels; Fundamentals and Medical Applications;
17. Debye, P.; Bueche, A.M. J. Appl. Phys. 1949, 20, 518. DeRossi, D., Kajiwara, K., Osada, Y., Yamauchi, A.,
18. Zierenberg, B.; Carpenter, D.K.; Hsieh, J.H. J. Polymer Eds.; Plenum: New York, 1991.
Sci., Polym. Symp. 1976, 54, 145. 51. Leloup, V.M.; Colonna, P.; Ring, S.G.; Roberts, K.;
19. Kajiwara, K.; Burchard, W. Macromolecules 1984, 17, Wells, B. Carbohydr. Polym. 1992, 18, 189.
2669. 52. Miles, M.J.; Morris, V.J.; Ring, S.G. Carbohydr. Polym.
20. Flory, P.J. Statistical Mechanics of Chain Molecules; 1984, 4, 73.
Interscience: New York, 1969. 53. Reuther, F.; Pleitz, G.; Damaschun, G.; Purschel, H.-V.;
21. Kajiwara, K.; Hiragi, Y. In Applications of Synchrotron Krober, R.; Schijerbaum, F. Colloid Polym. Sci. 1983,
Radiation to Materials Analysis; Saisho, H., Goshi, Y., 261, 271.
Eds.; Elsevier: Amsterdam, 1996. 54. Shimada, J.; Kaneko, H.; Takaha, T.; Kitamura, S.;
22. Kajiwara, K. The Method of Small-Angle X-ray Scattering Kajiwara, K. J. Phys. Chem., B 2000, 104, 2136.
and Its Application to the Structure Analysis of Oligosac- 55. Goldsmith, E.; Sprang, S.; Fletterich, R. J. Mol. Biol.
charides and Polysaccharides in Solution; Rome Univer- 1982, 156, 411.
sity: Rome, 1997. 56. Ikura, M.; Hikichi, K. Carbohydr. Res. 1987, 163, 1.
23. Shimode, M.; Urakawa, H.; Yamanaka, S.; Hoshino, H.; 57. Hanessian, S.; Hod, H.; Tu, Y.; Boulanger, Y. Tetrahe-
Harada, N.; Kajiwara, K. Sen’i Gakkaishi 1996, 52, 293. dron 1994, 50, 77.
24. Hounsell, E.F. Progr. Nucl. Magn. Reson. Spectrosc. 58. Sarko, A.; Maggli, R. Macromolecules 1974, 7, 486.
1995, 27, 445. 59. Kolpak, F.J.; Blackwell, J. Macromolecules 1976, 9, 273.
25. Farrar, T.C.; Becker, E.D. Pulse and Fourier Transform 60. Stipanovic, A.J.; Sarko, A. Macromolecules 1976, 9, 851.
NMR: Introduction to Theory and Methods; Academic 61. Atalla, R.H.; VanderHart, D.L. Science 1984, 223, 283.
Press: New York, 1971. 62. Horii, F.; Hirai, A.; Kitamaru, R. ACS Symp. Ser. 1984,
26. Bovey, F.A. High Resolution NMR of Macromolecules; 260, 29.
Academic Press: New York, 1972. 63. Cael, J.J.; Kwoh, D.L.W.; Bhattacharjee, S.S.; Patt, S.L.
27. Bloembergen, N.; Purcell, E.M.; Pound, R.V. Phys. Rev. Macromolecules 1985, 18, 821.
1948, 73, 679. 64. Horii, F.; Hirai, A.; Kitamaru, R. Macromolecules 1987,
28. Bloembergen, N. Physica 1949, 15, 386. 20, 2117.
29. Carr, H.Y.; Purcell, B.M. Phys. Rev. 1954, 94, 630. 65. Sugiyama, J.; Vuong, R.; Chanzy, H. Macromolecules
30. Meiboom, S.; Gill, D. Rev. Sci. Instrum. 1958, 29, 688. 1991, 24, 4168.
31. Komoroski, R.A., Eds. High Resolution NMR Spectro- 66. Sugiyama, J. Cell. Commun. 1994, 1, 6.
scopy of Synthetic Polymers in Bulk; VCH, 1986. 67. Dudley, R.L.; Fyfe, C.A.; Stephenson, P.J.; Deslandes,
32. Andrew, B.R. In Progress in Nuclear Magnetic Resonance Y.; Hamer, G.K.; Marchessault, R.H. J. Am. Chem. Soc.
Spectroscopy; Emsley, J.W., Feeney, J., Sutcliffe, L.H., 1983, 105, 2469.
Eds.; Pergamon Press: Oxford, 1971; Vol. 8, Part 1, 1 p. 68. Raymond, S.; Chanzy, H. Cell. Commun. 1995, 2, 13.

Copyright 2005 by Marcel Dekker


40 Kajiwara and Miyamoto

69. Geissler, K.; Krauss, N.; Steiner, T.; Betzel, C.; Sand- 100. Miyamoto, T.; Sato, Y.; Shibata, T.; Tanahashi, M.;
mann, V.; Saenger, W. Science 1994, 266, 1027. Inagaki, H. J. Polym. Sci., Polym. Chem. Ed. 1985, 23, 1373.
70. Raymond, S.; Heyraud, A.; Tran Qui, D.; Kvick, A.; 101. Kamide, K.; Saito, M.; Akedo, T. Polymer Int. 1992, 27, 35.
Chanzy, H. Macromolecules 1995, 28, 2096. 102. Aiba, S. Makromol. Chem. 1993, 194, 65.
71. Raymond, S.; Henrissat, B.; Tran Qui, D.; Kvick, A.; 103. Takahashi, S.; Fujimoto, T.; Miyamoto, T.; Inagak, H. J.
Chanzy, H. Carbohydr. Res. 1995, 277, 209. Polym. Sci., Polym. Chem. Ed. 1987, 25, 987.
72. Tabata, K. Carbohydr. Res. 1981, 89, 121. 104. Daicel Chem. md., Ltd., Japanese Open Patent Sho 60–
73. Gomaa, K.; Kraus, J.; Roßkoph, F.; Roper, H.; Franz, 42241, 1985, Taguchi, A.; Omiya, T.; Shimizu, K. Cell.
G.; Cane, I. Res. Clin. Oncol. 1992, 118, 136. Commun. Jpn. 1995, 2, 29.
74. Yanaki, T.; Norisue, T. Polym. J. 1983, 15, 389. 105. Sikkema, D.J.; Jannsen, H. Macromolecules 1989, 22, 364.
75. Yanaki, T.; Norisue, T.; Fujita, H. Macromolecules 1980, 106. Fukuda, T.; Sato, T.; Miyamoto, T. J. Soc. Fiber Sci.
13, 1482. Technol. Jpn. 1992, 48, 320.
76. Stokke, B.T.; Elgsæter, A.; Brant, D.A.; Kitamura, S. 107. Klemm, D.; Heinze, T.; Stein, A.; Liebert, T. Macromol.
Macromolecules 1991, 24, 6349. Symp. 1995, 99, 129.
77. Kitamura, S.; Minami, T.; Nakamura, Y.; Isoda, H.; 108. Miyamoto, T.; Long, M.; Donkai, N. Macromol. Symp.
Takeo, K.; Kobayashi, H.; Mimura, M.; Urakawa, H.; 1995, 99, 141.
Downloaded by [University of California, San Diego (CDL)] at 04:33 11 December 2016

Kajiwara, K.; Ohno, S. J. Mol. Struct. 1997, 395–396, 425. 109. Philipp, B.; Wagenknecht, W.; Nehls, I.; Klemm, D.;
78. Marchessault, R.H.; Deslandes, Y.; Ogawa, K.; Sundar- Stein, A.; Heinze, T. Polym. News 1996, 21, 155.
ajan, P.R. Can. J. Chem. 1977, 55, 300. 110. Kato, T.; Yokoyama, M.; Tukahashi, A. Colloid Polym.
79. Bluhm, T.L.; Sarko, A. Can. J. Chem. 1977, 55, 293. Sci. 1978, 256, 15.
80. Deslandes, Y.; Marchessault, R.H.; Sarko, A. Macro- 111. Donges, R. Br. Polym. J. 1990, 23, 315.
molecules 1980, 13, 1466. 112. Doelker, E. Adv. Polym. Sci. 1993, 107, 199.
81. Bluhm, T.M.; Deslandes, Y.; Marchessault, R.H. Carbo- 113. Mueller, K.F. Polymer 1992, 33, 3470.
hydr. Res. 1982, 100, 114. 114. Guo, J.-X.; Gray, D.G. In Cellulosic Polymers, Blends and
82. Sato, H.; Ohki, T.; Takasuka, N.; Sasaki, T. Carbohydr. Composites; Gilbert, R.D., Ed.; Hanser Verlag: Munich,
Res. 1977, 58, 293. Chap. 2, 1994.
83. Noguchi, K.; Kobayashi, E.; Okuyama, K.; Kitamura, S.; 115. Fukuda, T.; Takada, A.; Miyamoto, T. In Cellulosic
Takeo, K.; Ohno, S. Carbohydr. Res. 1994, 258, 35. Polymers, Blends and Composites; Gilbert, R.D., Ed.;
84. Harada, T. In Extracellular Microbial Polysaccharides; Hanser Verlag: Munich, Chap. 3, 1994.
ACS Symposium Series; Sanford, P.A., Laskin, A., Eds.; 116. Gilbert, R.D. Agricultural and Synthetic Polymers, ACS
Washington, DC: American Chemical Society, 1977; Vol. Symp. Ser.; American Chemical Society: Washington,
45, 265 pp. DC, 1990; Vol. 433.
85. Saito, H. Annu. Rep. NMR Spectrosc. 1995, 31, 157. 117. Yamagishi, T.; Fukuda, T.; Miyamoto, T.; Yakoh, Y.;
86. Stipanovic, A.J.; Giammatteo, P.J. In Industrial Poly- Takashina, Y.; Watanabe, J. J. Liq. Cryst. 1991, 10, 467.
saccharides: Genetic Engineering, Structure Property 118. Fukuda, T.; Tsujii, Y.; Miyamoto, T. Macromol. Symp.
Relations and Applications, Yalpani, M., Eds.; Elsevier: 1995, 99, 257.
Amsterdam, 1987; 281 pp. 119. Takada, A.; Fujii, K.; Watanabe, J.; Fukuda, T.;
87. Saito, H.; Yoshioka, Y.; Yokoi, M.; Yamada, J. Miyamoto, T. Macromolecules 1994, 27, 1651.
Biopolymers 1990, 29, 1689. 120. Itoh, T.; Takada, A.; Fukuda, T.; Miyamoto, T.; Yakoh,
88. Fuchs, T.; Richtering, W.; Burchard, W. Macromol. Y.; Watanabe, J. J. Liq. Cryst. 1991, 9, 211.
Symp. 1995, 99, 227. 121. Sugiura, M.; Minoda, M.; Fukuda, T.; Miyamoto, T.;
89. Guenet, J.-M. Thermoreversible Gelation of Polymers and Watanabe, J. J. Liq. Ciyst. 1992, 12, 603.
Bioploymers; Academic Press: London, 1992. 122. Sugiura, M.; Minoda, M.; Watanabe, J.; Miyamoto, T.
90. Amemura, A.; Hisamatsu, M.; Mitani, H.; Harada, H. Polym. J. 1994, 26, 1236.
Carbohydr. Res. 1983, 114, 277. 123. York, W.S.; van Halbeek, H.; Darvill, A.G.; Albersheim,
91. Dell, A.; York, W.S.; McNeil, M.; Darvill, A.G.; P. Carbohydr. Res. 1990, 200, 9.
Albersheim, P. Carbohydr. Res. 1983, 117, 185. 124. Yamanaka, S.; Mimura, M.; Urakawa, H.; Kajiwara, K.;
92. Miller, K.J.; Kennedy, E.P.; Rheinhold, V.N. Science Shirakawa, M. Sen’i Gakkaishi 1999, 55, 590.
1994, 231, Breedveld, M.W.; Miller, K.J. Microbiol. Rev. 125. Levy, S.; York, W.S.; Stuike-Prill, R.; Meyer, B.;
1986, 58, 145. Staehelin, A.L. Plant J. 1991, 1, 195.
93. Abe, M.; Amemura, A.; Higashi, S. Plant Soil 1988, 64, 126. Shirakawa, M.; Yamatoya, K.; Nishinari, K. Food
Stanfield, S.W.; Jelpi, L.; O’Brochla, D.; Helinski, D.R.; Hydrocoll. 1998, 12, 25.
Ditta, G.S. J. Bacteriol. 1982, 170, 3523. 127. Kajiwara, K.; Kohjiya, S.; Shibayama, M.; Urakawa, H.
94. Brant, D.A.; Mclntire, T.M. In Cyclic Polyssacharides in In Polymer Gels: Fundamentals and Biomedical Applica-
Large Ring Molecules; Semlyen, J.A., Ed.; Wiley: London, tions; DeRossi, D., Kajiwara, K., Osada, Y., Yamauchi,
1996. A., Eds.; Plenum: New York, 1991.
95. Palleschi, A.; Crescenzi, V. Gazz. Chim. Ital. 1985, 115, 128. See, for example Flory, P.J. Principles of Polymer
York, W.S.; Thomsen, J.U.; Meyer, B. Carbohydr. Res. Chemistry; Cornell UP: Ithaca, 1953.
1995, 248, 55; York, W.S. Carbohydr. Res. 1995, 278, 129. Kobayashi, K.; Sumitomo, H.; Ina, Y. Polym. J. 1985, 17,
205. 567.
96. Mimura, M.; Kitamura, S.; Gotoh, S.; Takeo, K.; 130. Kobayashi, K.; Tsuchida, A.; Usui, T.; Akike, T. Macro-
Urakawa, H.; Kajiwara, K. Carbohydr. Res. 1996, 289, 25. molecules 1997, 30, 2016.
97. Burchard, W. Theory of Cyclic Macromolecules. In Cyclic 131. Goto, M.; Yura, H.; Chang, C.-W.; Kobayashi, A.;
Polymers; Semylen, J.A., Ed.; London: Elsevier Applied Shinoda, T.; Maeda, A.; Kojima, S.; Kobayashi, K.;
Science Pub., 1986. Akaike, T. J. Control. Release 1993, 28, 223.
98. Casassa, E.F. J. Polym. Sci. 1980, A3, Burchard, W.; 132. Ohno, K.; Tusjii, Y.; Miyamoto, T.; Fukuda, T.; Goto, T.;
Schmidt, M. Polymer 1965, 21, 745. Kobayashi, K.; Akaike, T. Macromolecules 1998, 31, 1064.
99. Lukanoff, B.; Philipp, B.; Schleicher, H. Cellul. Chem. 133. Wataoka, I.; Urakawa, H.; Kobayashi, K.; Ohno, K.;
Technol. 1979, 13, 417. Fukuda, T.; Akaike, T.; Kajiwara, K. Polym. J. 1999, 31, 590.

Copyright 2005 by Marcel Dekker

You might also like