throbber
13
`Rheological Behavior of Polysaccharides Aqueous Systems
`
`Jacques Lefebvre and Jean-Louis Doublier
`INRA-Laboratoire de Physico-Chimie des Macromole´cules, Nantes, France
`
`I.
`
`INTRODUCTION
`
`The relevance of rheology to polysaccharide studies con-
`cerns both their practical and their fundamental aspects.
`The capacity of polysaccharides to extensively modify
`the rheology of aqueous media into which they are intro-
`duced, even at fairly low concentrations, is the basis of their
`‘‘functional properties’’ as thickening and gelling agents. It
`is also involved in many other types of applications, such as
`encapsulation, controlled release, etc. Some degree of
`rheological characterization is essential in particular to
`evaluate the potential uses of a polysaccharide as extracted
`from a natural source or subsequently modified.
`On the other hand, rheology provides precious tools to
`explore and understand the properties of polysaccharides in
`aqueous systems. The rheological behavior of polymer
`systems manifests the underlying structure of the systems.
`In the simplest case, that of polymer solutions, viscosity is
`directly related to fundamental molecular properties (mo-
`lecular conformation, molecular weight and molecular
`weight distribution, intramolecular and intermolecular inter-
`actions). In the case of more structured polymer systems,
`gels, for example, their viscoelastic properties are related to
`supramolecular organization. Although the relation between
`structure and rheological properties subsequently becomes
`more complex and less direct, rheology allows us to probe
`the structure of the systems at different scales, in conditions
`where other physical methods are impossible or difficult to
`use. Rheological techniques are especially useful to monitor
`and to probe structural changes in the systems, such as
`gelation or phase separation processes.
`Polymer rheology, which has been thoroughly studied
`and set up on firm theoretical foundations, provides a
`general frame for the investigation and the interpretation
`of the rheological behavior of polysaccharide systems.
`However, usually the parallel cannot be drawn very far
`into the details, particularly on quantitative grounds, as the
`following short discussion will make clear.
`
`A remarkable variety of physical chemical properties
`reflect the structural diversity of polysaccharides. Even
`chemically related polysaccharides can behave quite differ-
`ently in aqueous media, and, furthermore, in a way which
`strongly depends on solvent conditions. The same applies
`by way of consequence to their rheological properties.
`Polysaccharides are seldom homopolymers; in most
`cases, their backbones comprise several types of sugar
`monomers linked in sequences which, when indeed not
`totally unknown, are not characterized in detail. The
`‘‘heteropolymeric’’ character is responsible for the capacity
`of many polysaccharides to form gels in certain conditions.
`Moreover, polysaccharides are often branched polymers;
`the degree and pattern of branching, the lengths of the side
`chains, their composition itself in many cases, are generally
`ill defined. Among the polysaccharides that behave on the
`whole as linear homopolymers from a rheological point of
`view, many can be nevertheless grafted with short side
`chains, in which the number, length, and distribution along
`the backbone do play an important role in their properties.
`On the other hand, the presence of a few of sugar hetero-
`monomers inserted along an otherwise uniform linear
`sequence has a strong effect on the conformation of the
`chain and its physical chemical behavior. Such limited
`structural differences often result in a large variety of
`rheological properties, which can be observed among
`polysaccharides as belonging to the same chemical class
`but arising from different biological origins.
`Extraction, purification, and fractionation processes
`are another source of structural diversity. First, some
`polysaccharide or protein impurities, the nature and quan-
`tity of which depend on the purification procedure, are
`likely to remain in the sample. Their presence has been
`found to deeply affect the properties of the polysaccharide
`under consideration, or can be suspected to do so. Sup-
`posing this problem is solved, because in the initial material
`the polysaccharide of interest certainly presents some
`degree of structural heterogeneity of natural origin, the
`
`Copyright 2005 by Marcel Dekker
`
`357
`
`ALL 2032
`PROLLENIUM V. ALLERGAN
`IPR2019-01505 et al.
`
`

`

`358
`
`Lefebvre and Doublier
`
`exact composition and structure of the purified sample will
`depend on the way it has been prepared. This is the very
`consequence of the selectivity of the extraction and frac-
`tionation operations themselves. Finally, chemical modifi-
`cations of the polysaccharide are likely to occur during the
`preparation of the sample, giving rise to an artifactual
`heterogeneity; an example is backbone hydrolysis and
`partial demethylation of pectins during their extraction
`and purification. Indeed, many different samples may be
`obtained with as many preparation procedures! As a result,
`purified polysaccharide samples generally show at the same
`time very large polydispersity, i.e., they contain macro-
`molecules with widely different molecular weights and
`polymolecularity—they contain chains with qualitatively
`similar compositions, but differing in their quantitative
`compositions and in their structures. With much work,
`polysaccharide fractions with relatively narrow polydis-
`persity can be obtained (to be used as molecular weight
`standards, for example), but in most cases, it is practically
`impossible to avoid some degree of polymolecularity.
`One more source of complexity in the behavior of
`polysaccharide systems is the special nature of the sol-
`vent—water—and the delicate balance between chain–
`chain and chain–solvent interactions (hydrogen bonding,
`hydrophobic interaction,
`ionic interactions). Small
`changes in ionic strength or ionic composition, and limited
`shifts of temperature can induce a change in the physical
`nature of the system, e.g., a transition from the state of a
`macromolecular solution to that of a gel, causing a drastic
`modification of its rheological behavior.
`The polydisperse and polymolecular character of poly-
`saccharides, as they are available in practice, and the fact
`that their macromolecular structure is generally only quite
`vaguely known put severe limitations to the quantitative
`application of polymer rheology theories, whereas their
`structural diversity and complexity give rise to an extremely
`rich phenomenology, which largely gives rise to their
`functional versatility. However, within a defined physical
`state, the rheological behaviors of polysaccharide/(aque-
`ous) solvent systems are amenable to a general pattern, and
`they essentially differ on quantitative grounds. Structural
`differences mainly show in the conditions controlling the
`shift from one physical state to the other (from the solution
`to the gel and conversely, or from the solution to the
`dispersion or to the precipitate in phase separating sys-
`tems). Accordingly, on one hand, theoretical considera-
`tions will be kept to a minimum in this chapter. On the
`other hand, the rheology-relevant specific features perti-
`nent to the different classes of polysaccharides will not be
`surveyed. Our intention is neither to scan the different
`classes of polysaccharides, nor to draw a panorama of the
`application of different rheological techniques in this field;
`several valuable books and extensive reviews, written along
`these lines, are available and will be quoted in due place. We
`shall instead consider the main types of polysaccharide
`aqueous systems, and show schematically how rheological
`studies can provide an insight into their organization,
`focusing more on the common features than on the specific
`properties of each polysaccharide. The main types of
`polysaccharide systems that are encountered in the appli-
`
`cations can be distributed schematically in three classes:
`solutions, gels, and polysaccharide/polysaccharide (or
`polysaccharide/protein) mixtures in aqueous media. The
`last class comprises an extremely broad spectrum of sys-
`tems ranging from mixed solutions to complex structured
`multiphase systems. Biopolymers, even differing little in
`composition or structure, are in effect generally incompat-
`ible in aqueous media. As a consequence, the simultaneous
`presence of two polysaccharides (or of a polysaccharide
`and a protein) in the system results in a variety of micro-
`structures through phase separation processes. These pro-
`cesses can interfere or combine with gelation processes in a
`way that is highly dependent on the details of the structure
`of the biopolymers and on the experimental conditions.
`This gives rise to a fascinating variety of morphologies at
`different spatial scales, and, among other applications,
`allows ‘‘tuning’’ the rheological behavior of the systems
`to suit specific requirements. In spite of the theoretical and
`practical interest of polysaccharide mixtures, we shall
`restrict this chapter to simple solutions and gels, the
`rheology of which marks out the field of the behaviors of
`polysaccharide aqueous systems. Only shear deformation
`will be considered. Finally, the discussion of viscoelasticity
`will be limited to the linear domain.
`
`II. POLYSACCHARIDE SOLUTIONS:
`GENERAL REMARKS
`
`True polymer solutions are defined by the conjunction of
`the two following characteristics: (1) they are thermody-
`namically stable systems; (2) only physical interactions
`exist between the coils—hydrodynamic interactions and
`topological constrains which develop above a critical con-
`centration (coil overlap concentration).
`Rheology of polymer solutions has been extensively
`studied and thorough theoretical treatments are available,
`at least for linear neutral chains. This provides a frame to
`understand the rheology of polysaccharide solutions.
`However, because of polydispersity, polymolecularity,
`and molecular interactions, departures from polymer laws
`are often observed.
`In particular, obtaining true polysaccharide solutions
`is often not trivial. Polysaccharides are well-known to
`manifest a strong propensity to associate via hydrogen
`bonds, because of the abundance of hydroxyl groups. This
`is indeed the basis of the gelling properties of polysaccha-
`rides such as amylose and amylopectin, agarose, etc. but
`reversible and/or irreversible association occurs also in
`many other cases, although it does not lead to gel forma-
`tion in usual observation conditions. In any event, solubi-
`lization of polysaccharides is always difficult when the
`concentration is not low. Aggregation and incomplete
`solubilization result in the presence of microgels of more
`or less swollen particles in the system. The system is then
`actually no longer a solution, but a suspension in a polymer
`solution. These problems increase with polysaccharide
`concentration, with the consequence that polysaccharide
`solutions can be prepared in practice only up to rather
`limited concentrations. In many cases, the range of con-
`
`Copyright 2005 by Marcel Dekker
`
`358
`
`

`

`Rheological Behavior of Polysaccharides
`
`359
`
`distant along the chain cannot occupy the same volume
`element at the same time when approaching each other, as a
`consequence of their finite volume. This is the so-called
`‘‘excluded volume interaction.’’ On the other hand, there
`are always attractive and repulsive monomer/monomer
`and monomer/solvent interactions. The balance between
`these interactions will result in coil contraction or coil
`expansion. Whereas backbone rigidity and molecular
`weight are intrinsic characteristics of the polymer chain
`considered, this balance depends on solvent ‘‘quality’’ (the
`chemical nature of the solvent and the temperature). The
`effective ‘‘excluded volume’’ will therefore depend on
`solvent and temperature. In ‘‘good’’ solvent conditions,
`the chain configuration is more expanded, because solva-
`tion of chain segments increases excluded volume; then, its
`2 and can
`average square end-to-end distance will beL 2>Lo
`be written as:
`
`2 ¼ b2N2m; with m
`L
`
`z
`
`0:5
`
`ð1Þ
`
`Parameter m is the exponent of the radius of gyration–
`molecular weight relationship of the coil (Rg~Mm), its
`value depending on chain flexibility: obviously, m=1 for
`fully extended rigid chains, and m=1/3 for compact
`spheres. For polymer coils, m lies of course in between these
`two limiting values. However, it depends not only on the
`more or less flexible character of the backbone, but also on
`polymer–solvent interactions, which govern coil expan-
`sion. Exponent m has the value of 0.5 for Gaussian chains,
`as we have seen; for typical flexible polymers in good
`solvents, it is close to 0.6. In ‘‘bad’’ solvent conditions,
`attraction between chain segments dominates and causes
`coil collapse: the coil contracts, forming a dense particle
`(m!1/3), which eventually precipitates from the solution.
`Somewhere in between these two situations, repulsion
`between chain segments can be exactly compensated for
`by attraction; such solvent conditions are called ‘‘Q con-
`ditions’’ or ‘‘Q solvent.’’ In Q conditions, the polymer coil
`behaves indeed as if it were actually Gaussian and has its
`2=b2N (m=0.5).
`‘‘unperturbed’’ dimensionLo
`It is classical to express coil dimension as:
`
`ð1bÞ
`
`Þo
`
`2 g
`
`
`
`¼ a2gðR
`
`2 g
`
`; or R
`
`2 o
`
`2 ¼ a2
`L
`LL
`
`2, called the expansion factor of the polymer, is equalaL2 or ag
`
`
`to 1 in Q conditions and >1 in good solvents. Because the
`probability that one segment comes to take the place already
`occupied by another increases with N, the expansion factor
`increases with the molecular weight of the polymer. It is a
`complicated function of L, b, m, and of the second virial
`coefficient A2 of the polymer in the solution [1].
`
`B. Solutions of Noncharged Chains at Finite
`Concentrations: The Three Concentration
`Regimes
`
`Three concentration domains can be distinguished in so-
`lutions of polymers with molecular weights above the
`critical value Mc.
`
`centrations over which the rheological behavior can be
`studied is moreover limited on the lower side because of
`relatively low MW, hence low viscosity and low viscoelas-
`ticity that would require high sensitivity instruments to be
`measured. Few systematic rheological studies have been
`carried out on polysaccharide solutions (with the exception
`of cellulose derivatives) for the above reasons, and also
`because of the tedious extraction, purification, and frac-
`tionation operations necessary to obtain polysaccharide
`samples suitable for quantitative characterization.
`
`III. SOLUTIONS OF NONCHARGED CHAINS
`
`A. The Isolated Polymer Coil
`
`A typical polymer molecule is a long, flexible or semiflex-
`ible chain that can adopt in solution any configuration
`compatible with its fixed bond lengths and angles and other
`possible steric restrictions. The set of these spatially and
`temporally fluctuating configurations defines the equilibri-
`um statistical conformation of the polymer coil in the
`solvent. The polymer coil is a swollen structure; the volume
`occupied by the monomers accounts only for a small
`fraction of the total volume pervaded by the coil; monomer
`density decreases from the center of gravity of the coil to its
`periphery. If the volume of the chain itself is neglected, and
`if sterical and physical chemical interactions between
`monomers or between monomers and solvent molecules
`are absent, the spatial distribution of the monomers (or of
`the chain segments) within the coil would be Gaussian
`(‘‘random coil’’). The volume of the Gaussian coil depends
`only on the length (or the molecular weight M) of the chain,
`and on the size bo of the monomer. In real polymer
`molecules, bond angles are fixed and there are steric and
`energetic restrictions to rotations, resulting in a less-flexible
`chain (nonfreely jointed chain). Nevertheless, polymer
`molecules can still be treated as an equivalent Gaussian
`chain by replacing as statistical units the monomers by
`segments comprising n monomers; the rigidity of the
`backbone will be reflected in the length b=nbo of the
`statistical elements the articulation of which forms the
`chain. This length b (twice the ‘‘persistence length’’ Lp)
`can be directly determined from small angle neutron or
`small-angle X-ray scattering experiments. An appropriate
`chain stiffness parameter is the ratio of the contour length L
`of the chain to the length b of the statistical element: for a
`given polymer, chain flexibility increases with the molecu-
`lar weight. A ratio L/b > 10 would be required for the
`polymer conformation to be regarded as a coil; this corre-
`sponds to M higher than some limiting value Mc. Most
`polysaccharides are relatively stiff chains, with statistical
`f3  104.
`element length of f10 nm and Mc
`ffiffiffiffiffiffi
`N, being the number of statistical elements of the
`chain, the average square end-to-end distance of the equiv-
`¼ b2N ¼ kM: The diameter of
`p
`alent Gaussian chain is L
`
`
`1/2, its radius of gyration (Rg2)o
`the coil will be equal to b
`N
`2/6; the average monomer (or segment)such as (R g2)o=L o
`
`
`density within the coil will decrease as N1/2.
`The coil volume is actually somewhat larger than that
`of the equivalent Gaussian coil, because two segments
`
`2 o
`
`Copyright 2005 by Marcel Dekker
`
`359
`
`

`

`Lefebvre and Doublier
`
`zones exhibit lifetimes much larger than that of entangle-
`ments. The system then has shifted from the state of an
`entangled solution to that of a ‘‘physical gel’’ (cf. Section
`VII). The difference between the characteristics of the two
`types of systems is primarily a difference in degree, and not
`in nature. However, the establishment of such junction
`zones generally involves a change in the conformation of
`the chains, which lose their character of ‘‘random’’ coils.
`(cf. Section VII).
`
`3. Concentrated Regime (c > c**)
`At certain concentrations c**> c *, the coils reach their Q
`dimension; the excluded volume interaction is completely
`screened off. Above c**, coils will shrink no more; the
`polymer solution becomes an entanglement network where
`the chains have completely lost their individual character.
`The only characteristic length in the system is now the mesh
`size f of the network, which continues to decrease as
`concentration increases, tending toward its limit value b
`in the melt.
`
`IV. THE CASE OF POLYELECTROLYTES
`
`A polyelectrolyte is a flexible polymer electrically charged
`because its structure includes monomers bearing ionizable
`groups with charges of the same sign. Many polysaccha-
`rides, such as alginates, low-methoxyl pectins, carra-
`geenans, etc., are anionic polyelectrolytes, negatively
`charged at pH values above the pK of ionization of their
`acid groups. The only commonly found cationic polysac-
`charide is chitosan.
`The distinctive feature of polyelectrolytes is that the
`conformation of the macromolecule depends sharply on
`the ionic strength of the solvent, because the range of the
`electrostatic interaction decreases as ionic concentration
`increases. The Debye screening length j1~I1/2, where I
`is the ionic strength, classically measures the range of the
`electrostatic interaction in simple electrolyte solutions. In
`the case of polyelectrolyte solutions, the polymer itself,
`because of its proper charge and of the counterions sur-
`rounding its charged groups, as well as the salt dissolved in
`the solvent, both contribute to electrostatic screening. The
`
`Debye length is now j2 ¼ j2p þ j2s, where the indices p and s
`
`refer to the polymer and to the small ions contributions
`(including the counterions of the polyion), respectively.
`Thus, the conformation depends on both the polymer and
`~fcp, where cp is the concentra-
`the salt concentrations; j2
`p
`tion of the polyelectrolyte and f is the fraction of charged
`monomers in its chain. However, the effective contribution
`of the polymer is generally lower than expected from its
`theoretical charge density, as a result of the binding of
`counterions on the macroion. Ion binding (‘‘condensa-
`tion’’) results from strong attraction of counterions by
`the polyelectrolyte when its charge density is high; the ionic
`atmosphere surrounding the fixed charges can then differ
`widely from that of the Debye–Hu¨ ckel approximation.
`Obviously, electrostatic repulsion between the charged
`segments of the polyelectrolyte will favor more expanded
`conformations of the chain than if excluded volume effect
`were the only long-range interaction existing between chain
`
`360
`
`1. Dilute Regime (c < c*)
`
`In a very dilute solution, the volume available to each
`polymer molecule is much higher than that of the individ-
`ual coil. The coils remain statistically far from each other,
`and encounters are infrequent. The coils maintain the
`dimensions of the isolated chain. This situation prevails
`up to the critical overlap concentration c*, at which the
`coils fill the volume of the solution. The volume of each coil
`
`Þ3=2 ¼ ð1=6Þ3=2ðL2Þ3=2,
`being proportional to ðR
`
`c ~M=ðL2Þ3=2~N13m
`
`ð2Þ
`
`2 g
`
`2. Semidilute Regime (c* < c < c**)
`
`When polymer concentration is increased above c*, there is
`a progressive interpenetration of the coils, concomitant
`with a contraction of their individual volume. Coil con-
`traction is a result of the progressive screening off of the
`‘‘excluded volume interaction’’: as a result of interpenetra-
`tion, segments of ‘‘foreign’’ chains interpose themselves
`between segments belonging to the same chain. The solu-
`tion becomes a transient network of entangled chains, an
`entanglement being the topological constraint correspond-
`ing to a point of contact between two chains, and being due
`to the fact that the chains cannot cross each other. A given
`polymer with M>Mc statistically contracts a determined
`number of entanglements at a given concentration c>c*,
`but, because of chain conformation fluctuations, these
`entanglements continuously unfasten to be reformed on
`other points along the chain contour; their lifetime is very
`short. If M<Mc, the length of the chain is shorter than the
`minimum distance required between entanglement points.
`An alternative view [2] is to consider the system as equiv-
`alent to a cage between the rails of which one given chain
`has to reptate along its own contour in order to move away;
`the chain is confined within a tube, the diameter of which is
`the mesh size of the temporary network.
`In the semidilute domain, the mesh size f of the network
`(average distance between entanglement coupling) and the
`size of the coil decrease as concentration increases. f2
`measures the mean square end-to-end distance of the chain
`segment comprised between two successive entanglement
`points along one chain. The chain segment thus defined,
`made up of g statistical elements, constitutes a ‘‘blob,’’ and a
`chain of N/g blobs forms the polymer molecule. The
`excluded volume interaction between two adjacent blobs
`along the chain being screened off by an interposed foreign
`chain, the chain of blobs is Gaussian, so that the mean
`square end-to-end distance a polymer chain can be written:
`2 ¼ f2N=g. Within the blob, on the contrary, the excluded
`L
`volume interaction is effective and f2=b2g2m. Scaling laws
`give [3]: f~cm/13m andL
`2~c2m1=13m.
`In the semidilute regime, there are two characteristic
`lengths in the system: the size of the coil—as the coils still
`retain some degree of individuality—and the entanglement
`spacing (or size of the blob).
`Looking at the system as a network can be readily
`extended to the case where low-energy physical chemical
`interactions develop between chains in the regions of
`entanglements, giving rise to junction zones. Junction
`
`Copyright 2005 by Marcel Dekker
`
`360
`
`

`

`Rheological Behavior of Polysaccharides
`
`361
`
`segments. Thus, the ionized polyelectrolyte will exhibit
`larger mean-square end-to-end length and radius of gyra-
`tion values than the uncharged chain would have in good
`solvent. This can be accounted for by introducing an
`electrostatic contribution to the persistence length of the
`chain (see, for example, Ref. [4]). In very dilute salt-free
`solutions, the macromolecule tends to adopt an extended
`rodlike conformation in order to minimize the electrostatic
`2cL~
`contribution to the free energy of the chain; then,L
`N. As I increases, electrostatic interaction between charged
`segments is progressively screened off. At moderate values
`of I the chain conformation resumes a spherical symmetry,
`but with a larger radius of gyration than the equivalent
`uncharged chain in the same solvent, and at higher values
`of I, the dimension of the coil approaches that of the
`uncharged chain.
`Even a very schematic analysis of the effect of cou-
`lombic repulsion of ionized groups on polymer conforma-
`tion would be beyond the scope of this chapter. The reader
`can refer to Refs. [5,6] for a thorough theoretical treatment.
`Many points about the behavior of polyelectrolytes remain
`indeed unclear. Therefore, we shall just point out here a few
`remarks of practical importance.
`Since its conformation strongly depends on its own
`concentration c, as well as on salt concentration cs, the
`phase diagram of a polyelectrolyte is more complex than
`that of neutral polymers. An example of such a diagram is
`shown on Fig. 1. Because the macromolecule is highly
`expanded, the coil overlap concentration c* is extremely
`low at low and intermediate salt concentrations. Below c*,
`the situation is in fact complex. At low salt concentra-
`tions, the polyelectrolyte is in its extended conformation
`(DR regime of Fig. 1), far different from that of a neutral
`polymer, and furthermore, there are strong intermolecular
`interactions. However, for large enough salt concentra-
`
`Figure 1 Theoretical phase diagram for an aqueous poly-
`electrolyte solution (molar concentration c) in presence of
`added salt (molar concentration cs). DR: dilute rodlike regime.
`DF: dilute regime of flexible conformation. SF: semi-dilute
`regime. The diagram has been calculated for a chain with
`N=3350 monomers, an average number of monomers between
`charges A=5, and Nb/A=3, where A is the actual extended
`length of the chain. Reproduced from Dobrynin et al. [5].
`
`tion, the chain becomes flexible at all polyelectrolyte
`concentrations (DF regime). The conformation is then
`analogous to that of the neutral chain in good solvent, but
`the expansion factor is controlled by the electrostatic
`screening length, which depends on both c and cs. Above
`c* (SF regime), the chain is likewise flexible, but now the
`classical excluded volume interaction screening effect of
`neutral polymers combines with the electrostatic screening
`effect in governing the expansion factor. Whereas both
`effects depend on c, the contribution of the latter
`decreases as cs increases, and at high enough salt concen-
`trations the situation becomes similar to that for the
`neutral chain. Because at low and intermediate ionic
`strength values, polyelectrolyte dimension is strongly
`dependent on its concentration, the semidilute regime is
`very wide;
`it can extend over three or four polymer
`concentration decades.
`The extremely low values of c* and the existence of
`strong intermolecular interactions for c < c* make the
`experimental characterization of the isolated macromole-
`cule difficult at low and intermediate cs values. At large salt
`concentrations, experimental problems also arise, because
`the added salts affect solvent quality, independently of their
`electrostatic screening effect; actually, many flexible neu-
`tral water-soluble polymers approach unperturbed (Q)
`dimensions as salt concentration increases; they eventually
`even precipitate (‘‘salting out’’). The density of charges
`along the chain, i.e., the relative number of monomers
`bearing ionizable groups and the degree of ionization are
`the primary intrinsic characteristics of the chain affecting
`polyelectrolyte expansion. Theories take into account an
`average value for charge density; they introduce, e.g., an
`average number of monomers between charges. But the
`distribution of the charged groups along the chain also
`plays a role: the repulsion between a pair of charges is likely
`to have a larger effect on the overall conformation when the
`charges are distant along the chain than when they are
`adjacent. In the case of polysaccharides, charge distribu-
`tion is neither even nor random along the chain, and is
`generally completely unknown; it is susceptible to consid-
`erable variation for a given polysaccharide with a given
`average charge density.
`
`V. FLOW BEHAVIOR OF POLYSACCHARIDE
`SOLUTIONS
`
`A. Origin of Rheological Properties of
`Polymer Solutions
`
`In the dilute regime, Newtonian flow behavior and absence
`of viscoelasticity are generally observable in practical
`conditions, at least for noncharged polymers,* because
`statistically, macromolecules are spatially and temporally
`noncorrelated. Nevertheless, in principle, polymer coils are
`able to deform when submitted to velocity gradients, and to
`recover their equilibrium conformation after cessation of
`
`* As we shall see later, this is not the case for dilute polyelectrolyte
`solutions in low added salt conditions, because of long-range
`electrostatic interactions.
`
`Copyright 2005 by Marcel Dekker
`
`361
`
`

`

`362
`
`Lefebvre and Doublier
`
`great theoretical interest, it has little practical importance
`and the topic will not be considered here. In this section,
`viscosity and intrinsic viscosity will refer implicitly to
`measurements performed within the Newtonian domain,
`i.e., at shear rates low enough for the dilute solutions to
`display shear rate-independent viscosity. The condition of
`Newtonian behavior, in most cases, easy to achieve for di-
`luted polymer solutions, can be in some others (when
`working on polymers with very high molecular weight, or
`with rather stiff chains, and on polyelectrolytes at low salt con-
`centrations) more difficult to accomplish experimentally.
`
`1. The Intrinsic Viscosity of Noncharged Polymers
`
`the mechanical excitation, hence intrinsic non-Newtonian
`and viscoelastic properties. However, these properties are
`very faint and are exhibited in practice only at fairly high
`deformation rates. They are, in consequence, of little direct
`concern for applications, as far as polysaccharide dilute so-
`lutions are considered. However, internal coil deformation
`modes are responsible for the short-time or high-frequency
`viscoelastic response of polymer nondilute solutions and
`gels. On the other hand, the characteristics of the polymer
`that are reflected in the viscosity of its dilute solutions
`govern, to a large extent, its viscous behavior in nondilute
`solutions. In consequence, no characterization of the semi-
`dilute and concentrated regimes of polysaccharides, which
`are the relevant issues regarding applications, is possible
`without the knowledge of the behavior of the polysacchar-
`ides in dilute solutions. This is why dilute solution proper-
`ties have to be given some development.
`In practice, non-Newtonian flow behavior and visco-
`elasticity of polymer solutions develop for c > c*. In the
`case of flexible or semiflexible chains, these properties
`originate from the disentanglement/re-entanglement pro-
`cesses resulting from the opposite actions of flow and of
`thermal agitation. They are manifestations of the transient
`network structure. Dispersions of rigid macromolecules or
`particles exhibit similar properties above some critical
`concentration; but they are mainly due, in this case, to
`the spontaneous establishment of a local order at rest, as a
`consequence of crowding; strain perturbs this order and
`thermal agitation tends to restore it. When macromolecules
`or particles are asymmetrical, orientation in the direction
`of flow plays, in addition, an important role (liquid crystal
`structures). For solutions of flexible chains as well as for
`dispersions of rigid particles, non-Newtonian flow and
`viscoelastic properties are therefore of entropic origin.
`
`B. Viscosity of Dilute Solutions
`
`Only a very schematic and short account will be provided
`here on viscosity of dilute solutions of polymers in general,
`and of polysaccharides in particular. The matter, which has
`direct bearing on macromolecular conformation studies,
`has been the subject of an enormous amount of theoretical
`and of experimental work, and books or extensive reviews
`are available on it (for example, Refs. [7–9]). The main
`information that can be extracted from viscosity measure-
`ments on dilute macromolecular solutions is contained in
`the intrinsic viscosity of the macromolecule. Intrinsic vis-
`cosity is not a viscosity at all, but actually a measure of the
`hydrodynamic volume of the coil in the case of noncharged
`polymeric chains, or of the asymmetry of the particle in the
`case of rigid macromolecules. The case of polyelectrolytes
`is specific, and remains poorly understood.
`As already alluded to, for most polyme

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket