throbber
29
`Macromolecular Complexes of Chitosan
`
`Naoji Kubota and Kei Shimoda
`Oita University, Oita, Japan
`
`I.
`
`INTRODUCTION
`
`A. Macromolecular Complexes
`
`Macromolecular complexes are molecular aggregates of
`two or more complementary polymers, which have definite
`composition and characteristics and arise from intermo-
`lecular interactions, such as Coulomb forces, hydrogen-
`bonding forces, charge-transfer forces, and van der Waals
`forces. The complexes can be divided into the following
`four classes according to the nature of their interactions [1]:
`polyelectrolyte complexes, hydrogen-bonding complexes,
`charge-transfer complexes, and stereocomplexes.
`It is rare that these intermolecular interactions take
`place singly, and hydrophobic interactions are also signif-
`icant factors in aqueous media. Although these secondary
`binding forces, among which Coulomb forces are stron-
`gest, are weaker than primary binding forces, they play
`very important roles in the biological systems (e.g., molec-
`ular recognition, accumulation and interpretation of ge-
`netic information, and formation of higher-order structure
`of proteins).
`Biological systems comprise different kinds of macro-
`molecules, such as nucleic acids, enzymes, proteins, and
`polysaccharides, and most of them have definite structure
`and specific functions. Antigen–antibody reactions and
`enzymic reactions, for example, are very specific. In these
`cases, intermacromolecular interactions play important
`roles. Many bioreactions proceed via formation of macro-
`molecular complexes in the beginning of the reaction, and
`the concerted interactions are included. As a simplified
`model of these bioreactions, it is rational to investigate the
`macromolecular complexes. The majority of researches on
`macromolecular complexes have been directed toward a
`better understanding of biological systems and of structure
`and properties of functional units. However, macromolec-
`ular complexes have recently found their way into practical
`applications, in particular, as biomaterials.
`
`B. Structure and Properties of Chitosan
`Chitin, poly[h(1!4)-2-acetoamido-2-deoxy-D-glucopyra-
`nose], is one of the most abundant natural polysaccharides
`and is present in crustacea, insects, fungi, and yeasts. The
`total annual global estimates of accessible chitin amount
`to 1  109 tons [2]. It is obtained primarily as a by-product
`of the seafood industry. Deacetylation of chitin by alkali
`readily affords chitosan (Ch), poly[h(1!4)-2-amino-2-de-
`oxy-D-glucopyranose]. Ch is also found in various fungi.
`However, the molecular structure of Ch is believed to be a
`copolymer of N-acetyl-glucosamine and glucosamine; usu-
`ally the glucosamine content is more than 90%. It is
`known that 50% N-deacetylated chitin (50% N-acetylated
`Ch) is water soluble [3]. Kubota and Eguchi showed that
`the water solubility of half N-acetylated Ch increased
`with decreasing the molecular weight, as shown in Fig. 1
`[4]. In addition, the solubility of half N-acetylated Ch
`with the molecular weight lower than 10,000 is rather high
`even in aqueous dimethylacetamide and aqueous dimethyl
`sulfoxide (DMSO) [5].
`Because Ch has an amino group in the repeating unit,
`it affords ammonium group in aqueous acidic media.
`Owing to its cationic nature, Ch spontaneously forms
`water-insoluble complexes with anionic polyelectrolytes.
`Therefore, Ch has been used mainly as a flocculant for the
`treatment of wastewater. However, it has recently been
`used in biomedical and pharmaceutical fields because of its
`favorable properties of good biocompatibility, low toxici-
`ty, and biodegradability.
`The intrinsic viscosity [g] of Ch depends on the pH and
`the ionic strength, as shown in Fig. 2 [6]. If the pH of the
`solution is increased, the intermolecular and intramolecu-
`lar electrostatic repulsions between cationic charges are
`reduced. This allows the Ch chains to come closer and thus
`lowers the hydrodynamic volume of the Ch molecules. This
`effect may enhance the interchain and intrachain hydrogen
`bonding. Similarly, as the ionic strength increases, [g]
`
`Copyright 2005 by Marcel Dekker
`
`679
`
`ALL 2022
`PROLLENIUM V. ALLERGAN
`IPR2019-01505 et al.
`
`

`

`680
`
`Kubota and Shimoda
`
`decreases due to the shielding effect of the counterions.
`Although Co¨ lfen et al. suggested that the hydrodynamic
`behavior of Ch was consistent with wormlike chain model
`[7], the Ch molecule is rather stiff. The flexibility of some
`polysaccharides has been investigated in terms of a ‘‘stiff-
`ness parameter’’ B (i.e., the dependence of the intrinsic
`viscosity on the ionic strength). The B values for a variety
`of polymers are compiled in Table 1. The B value of Ch was
`estimated as 0.08 [8]. Terbojevich et al. obtained B values
`from 0.043 to 0.091 for Ch with degrees of acetylation
`ranging from 52.2% to 12.1% [9]. These values are essen-
`tially the same as those for carboxymethyl cellulose and
`hyaluronic acid, greater than DNA, and less than poly-
`acrylate. These solution properties of Ch greatly affect the
`formation of macromolecular complexes.
`
`II. FORMATION OF MACROMOLECULAR
`COMPLEXES OF CHITOSAN
`
`The formation of macromolecular complexes in solution,
`depending on the intensity of polymer–polymer and poly-
`mer–solvent interactions, leads to the separation of com-
`plexes as a solid or liquid phase. Because Ch is a cationic
`polyelectrolyte, most studies on the macromolecular com-
`plexes of Ch are concerned with polyelectrolyte complexes
`(PECs). However, only a few articles dealt with the poly-
`ionic interaction between polysaccharides, until Kikuchi
`
`Figure 1 Dependence of water solubility of half N-acety-
`lated Ch on the molecular weight. (From Ref. 4.)
`
`Figure 2 Effects of pH and ionic strength on intrinsic viscosity [g] of Ch solution:
`L; n, 0.200 mol/L; ., 0.300 mol/L; z, 0.500 mol/L NaCl. (From Ref. 6.)
`
`x
`
`, 0.050 mol/L; Ez, 0.075 mol/L; E, 0.100 mol/
`
`Copyright 2005 by Marcel Dekker
`
`680
`
`

`

`Macromolecular Complexes of Chitosan
`
`681
`
`Table 1 Stiffness Parameter, B, for Ch and Some Polymers
`
`Polymer
`
`Ch
`
`Polyphosphate
`Polyacrylate
`Amylose xanthate
`Carboxymethylamylose
`Carboxymethylcellulose
`Hyaluronic acid
`Sodium pectinate
`DNA
`
`Degree of acetylation
`(%)
`
`12.1
`20.1
`42.1
`52.2
`—
`—
`—
`—
`—
`—
`—
`—
`
`B
`
`0.091
`0.060
`0.061
`0.043
`0.44
`0.23
`0.22
`0.20
`0.065
`0.07
`0.044
`0.006
`
`Source: Refs. 9 and 8. Reproduced by permission of Routledge, Inc.,
`part of The Taylor & Francis Group.
`
`first reported the PEC containing Ch as a component [10].
`Several reviews on PEC have been published so far [11,12].
`The formation of PEC is governed not only by the nature of
`the individual polyelectrolyte components, such as charac-
`teristics of polyions (strong or weak), position of ionic sites,
`charge density, molecular weight, flexibility, functional
`group structure, hydrophilicity and hydrophobicity, and
`stereoregularity, but also by the reaction conditions, such
`as pH, ionic strength, polymer concentration, mixing ratio,
`and temperature. This, therefore, may lead to a diversity of
`physical and chemical properties of the complexes. PECs
`arise due to interactions between oppositely charged poly-
`mers and can be additionally stabilized through short-
`
`range interactions such as hydrophobic interactions and
`hydrogen bonds.
`
`A. Thermodynamics and Stoichiometry
`of Complex Formation
`
`The complexation reaction of macromolecules is signifi-
`cantly different from that of low-molecular weight mole-
`cules due to the polymer effects by which a relatively stable
`complex occurs, although the force of each bound pair is
`small. This implies that the apparent equilibrium constant
`of the polymer–polymer complex is extremely large and the
`reaction is apparently irreversible. The free-energy change
`DGo for macromolecular complexation is a function of the
`degree of polymerization, and the equilibrium constant
`abruptly increases at the critical chain length [13]. Pe´ rez-
`Gramatges et al. revealed that complexation of Ch with
`poly(acrylic acid) (PAA) proceeded cooperatively and the
`larger degree of conversion, h, was obtained for the higher-
`molecular weight Ch [14]. The relationship between the
`stability constant, K, and h (Fig. 3) indicates that this
`complex consists of a long sequence of bound pairs of
`repeating units. The slope of this curve shows the influence
`of neighboring functional groups on their reactivity.
`Kanbayashi and Arai determined thermodynamic
`parameters for PEC formation in the methyl glycol Ch
`(MGC)–carboxymethyl cellulose (CMC) system [15]. The
`formation of PEC gained a large negative DGo, but the
`factors were dependent on the charge density of CMC. The
`PEC formation using CMC with low charge density
`showed a marked exothermic tendency and was classified
`as an enthalpy-dominating reaction. On the other hand, the
`reaction using CMC with high charge density showed a
`smaller exothermic tendency and gained large positive
`
`Figure 3 Linear relationship between the stability constant, K, and the degree of conversion h for Ch–PAA complexes: D, Mv =
`o
`8.5  104;
`, Mv = 1.1  105; 5, Mv = 2.3  105. (From Ref. 14.) Copyright 1996 Springer-Verlag.
`
`Copyright 2005 by Marcel Dekker
`
`681
`
`

`

`682
`
`Kubota and Shimoda
`
`entropy, indicating an entropy-dominating type. They also
`investigated the dependence of thermodynamic parameters
`on ionic strength using the MGC–partially hydrolyzed
`poly(acrylamide) system [16]. The degree of complexation
`did not change much with the ionic strength (0.001–0.2),
`and a gain of a large negative DGo was accompanied by the
`PEC formation. However, DHo became a large positive
`value with increasing ionic strength. Therefore, the higher
`ionic strength suggested the entropy-dominating-type PEC
`formation reaction. A much higher ionic strength affects
`PEC formation; for example, when Ch and n-carrageenan
`(Car) were mixed in the presence of 5.7% NaCl at pH 2.8
`on a boiling water bath, such a phase separation did not
`occur, but a viscous and macroscopically homogeneous
`PEC mixture was obtained [17]. The presence of Na+ and
`Cl reduces the electrostatic attraction between oppositely
`charged polyelectrolytes. This mixture gelled as its temper-
`ature or ionic strength decreased.
`In general, the composition of the reaction mixture, Z,
`is unity at the equivalence point, suggesting a 1:1 stoichi-
`ometry for the complex, no matter what order of mixing is
`chosen. However, h varied with Z in the case of Ch–
`polygalacturonate (PGal) complex [18]. It fell from unity
`to 0.8 as Z increased from 0.2 to 0.5 and then rose again up
`to 0.85 for Z = 1. If the complexation reaction is termi-
`nated before a 1:1 stoichiometry of charge neutralization is
`reached, water-soluble complexes are possible. In addition,
`when the PEC involves two polyelectrolytes of different
`molecular size, soluble nonstoichiometric PEC can be
`formed. In such a complex, the larger polymer chain
`behaves as a host macromolecule to the shorter one, and
`the host polymer should be a strong polyelectrolyte. For
`weak polyelectrolytes, not all of the repeating units need to
`have a 1:1 stoichiometry.
`
`B. Complexes with Polysaccharides
`
`The interaction between oppositely charged polyelectro-
`lytes involving weak polyacid and/or weak polybase is
`affected by the solution pH because of the change in the
`degree of dissociation. In this sense, it is of great interest to
`examine the properties of complexes involving weak poly-
`electrolytes under various pH conditions. When Ch, a
`weak polybase, is reacted with a weak polyacid, the insol-
`uble complex formation occurs only in the narrow pH
`range. Argu¨ elles-Monal et al. reported the complex forma-
`tion reaction of the Ch–CMC system [19]. At pH 3.6, the
`PEC was rich in CMC, whereas at pH 4.8 the excess
`component was Ch. At pH 4, the yield of PEC increased
`with the addition of one polyelectrolyte solution to the
`other, and the maximum yield corresponded to the ratio
`[CMC]/[Ch] = 1. Beyond this point, the yield remained
`constant, and the composition of the PEC obtained showed
`the stoichiometry. Similar results were obtained in the case
`using PAA, a synthetic polymer, as a polyacid [20,21]. The
`mixing ratio Ch/(Ch + PAA) for maximum insoluble
`complex formation, Rmax, increased with increasing solu-
`tion pH. Interestingly, at the initial pH = 6, the superna-
`tant pH of the reaction mixture increased as the complex
`was formed. At the initial pH=3, the opposite behavior
`
`was observed. As shown in Fig. 4, pKa of Ch increased as
`the addition of the polyanion increased, suggesting that the
`charged carboxylate group induced the ionization of the
`amino group of Ch [22]. Therefore, the supernatant pH
`increased at pH 6 as the yield of the complex increased:
`
`NH2 þ OOC þ Hþ ! NHþ3 OOC
`The supernatant pH decreased at pH 3 as the yield of
`complex increased:
`3 þ HOOC ! NHþ 3 OOC þ Hþ
`
`NHþ
`However, changes in the supernatant pH by complexation
`was dependent on the type of polyanions [23]. For Ch–Car,
`Ch–alginate (Alg), and Ch–pectin (Pec) complexes, no
`significant pH change occurred at the initial pH = 3. At
`pH 4.5, there was a slight increase in pH as a result of
`complexation of Ch with Alg and Pec, and a more signif-
`icant increase with Car. At pH 5.4, there was an increase in
`pH analogous to that at pH 6 for the Ch–PAA system.
`However, unlike PAA, a maximum pH change was main-
`tained in the wide mixing ratio from 0.2 to 0.4. Differences
`in polyanion conformation are responsible for the differ-
`ences in pH changes between Ch–PAA and Ch–Car, Ch–
`Alg, Ch–Pec complex formation. For example, structural
`similarity between Ch and Car, Alg, and Pec or flexibility of
`PAA is possible.
`Takahashi et al. observed the difference in basic
`properties between Ch–Alg and Ch–PAA complexes [24].
`In the Ch–Alg system, the insoluble complexes were
`formed at a constant unit molecular ratio of 1:1.3 (Ch:Alg)
`at pH 3.7–4.7. However, the unit molecular ratio of the
`Ch–PAA system was greatly affected by pH, showing a
`change from 1:2.4 to 1:1.7 (Ch:PAA) with an increase in
`pH from 3.7 to 4.7. They concluded that this was due to
`the higher flexibility of the polymer chain of PAA than
`that of Alg.
`On the other hand, owing to the ease of interfacial
`PEC formation, the fitness of the structures of the back-
`bones of Ch and some polyanions was estimated as fol-
`lows [25]:
`heparin ðHepÞ > Alg > carboxymethyl
`chitin ðCMChitinÞ > PAA
`
`Nakajima and Shinoda showed that the backbone chain
`conformations of component polymers together with the
`kind and location of ionizable groups were important
`factors to discuss the formation and structure of PEC
`[26]. They used glycol Ch (GC), which is water soluble at
`all pHs, as a polycation and hyaluronic acid (HA), chon-
`droitin sulfate (CS), Hep, and sulfated cellulose (SCS) as
`polyanions. In the GC–HA and GC–CS systems, the
`experimental curves of complex composition, R, to pH
`crossed the theoretical curves at R = 0.5 (Fig. 5 shows the
`GC–HA system as an example.) In both complexes, the
`neutral complex appears only at R = 0.5. The positive and
`negative charges may remain in the regions R < 0.5 and
`R > 0.5, respectively. This result shows that the dominant
`factor in these cases is the pyranose structure itself rather
`
`Copyright 2005 by Marcel Dekker
`
`682
`
`

`

`Macromolecular Complexes of Chitosan
`
`683
`
`Figure 4 Degree of ionization for amino groups of Ch in the presence of xanthan. The mole ratio of carboxyl groups to amino
`groups: ., 0.30; E, 0.60; n, 1.19. (From Ref. 22.)
`
`than the kind and location of the ionizable groups on the
`pyranose ring. In the case of the GC–Hep system, the
`discrepancy between theoretical and experimental curves
`was rather small, and complex formation seemed to pro-
`ceed almost stoichiometrically. However, the complex
`composition was GC/(GC + Hep) = 0.65 at the lower
`
`pH region. They suggested that one of the possible struc-
`tures having these compositions was the ladder form
`sandwiching a Hep molecule between two GC molecules.
`Moreover, the experimental curve of the GC–SCS system
`was remarkably different from the theoretical curve, as
`shown in Fig. 6. However, the complex composition at the
`
`Figure 5 Composition of complex plotted against pH for GC–HA system. (From Ref. 26.) Copyright 1976, reprinted with
`permission from Elsevier Science.
`
`Copyright 2005 by Marcel Dekker
`
`683
`
`

`

`684
`
`Kubota and Shimoda
`
`Figure 6 Composition of complex plotted against pH for GC–SCS system. (From Ref. 26.) Copyright 1976, reprinted with
`permission from Elsevier Science.
`
`lower pH region was the same value as that of the GC–Hep
`system. In this case, it is thought that the GC–SCS system is
`similar to the GC–Hep system in structure, but negative
`charges are always conserved. The reason for such differ-
` groups on SCS and
`ence may be the location of the –OSO3
`Hep molecules. Similarly, the theoretical and experimental
`curves of the mixing ratio intersect at 0.46, 0.52, and 0.53,
`for the GC–CMC, GC–Alg, and GC–PGal complexes,
`respectively [27]. On the contrary, in the case of the GC–
`dextran sulfate (DS) complex, the stoichiometric compo-
`sition occurs at R = 0.62 at pH < 3.5. It was concluded
`that the GC–DS complex had a different structure from the
`other three [28].
`In the case of n-Car, K+ ions induce helix conforma-
`tion and promote helix–helix aggregation. The decisive
`factor of the Ch–Car complex formation seems to be
`whether n-Car is in a helix–helix aggregated state [29].
`The interaction between Ch and n-, L-, and E-Car in their
`coil or nonaggregating helical conformation normally
`resulted in the formation of PEC with stoichiometric
`charge ratios of unity. Nevertheless, the formation of
`PEC was significantly affected by the fraction of n-Car in
`K+-induced helical conformation. If the n-Car exists in the
`helix–helix aggregated state then the interaction with Ch
`produces PEC with a charge ratio below unity, thereby
`providing a means of complexes with a surplus of negative
`charge. In this case, Ch molecules act as binding elements
`between helix–helix aggregated n-Car.
`Partially N-acylated Ch samples were used to examine
`the effect of the N-acyl groups on the complex formation
`with CS [30]. The Rmax values became larger with increas-
`ing substitution degree of N-acyl groups but were not
`dependent on the kind of N-acyl group. Broad turbidity
`
`curves appeared at pH 4.5 at the higher substitution degree
`of N-propionyl, N-hexanoyl, and N-myristoyl. The higher
`the acyl group, the stronger the effect, due to the insolu-
`bility of the complexes and the hydrophobic interactions of
`N-fatty acyl groups. It is difficult to form a ladderlike
`structure in this system owing to bulky acyl groups.
`Therefore, some of the free groups that are not involved
`in the PEC formation may exist inside the complexes.
`
`C. Complexes with Proteins
`
`Remunˇ a´ n-Lo´ pez and Bodmeier investigated optimal con-
`ditions for the complexation between Ch and type B
`gelatin (Gel) [31]. All of the optimal preparation condi-
`tions were essentially coincident with those for Ch–poly-
`saccharide complexes. In addition, complexation was
`found to depend on the molecular weight of Ch and Gel;
`higher-molecular weight Ch resulted in higher amounts of
`complex formed, whereas higher-molecular weight Gel
`solubilized the complexes.
`When a protein reacts with an oppositely charged
`polyelectrolyte to form PEC, a conformational change
`occurs. Park et al. reported the conformational change of
`a-keratose (Ker) by complexing with Ch at pH 5.2 [32].
`The a-helical structure in Ker was transformed into a
`random structure by complexing with Ch at Rmax, whereas
`the h-sheet was not affected. It seemed that helix-
`favoring amino acid residues, aspartic and glutamic acids,
`participated in forming electrostatic linkages. As shown in
`Fig. 7, the lower the molecular weight of Ch, the higher
`the destruction of the a-helix in Ker.
`low-molecular
`weight Ch seems to cause easier or complete complexa-
`tion reaction.
`
`Copyright 2005 by Marcel Dekker
`
`684
`
`

`

`Macromolecular Complexes of Chitosan
`
`685
`
`Figure 7 Effect of molecular weight of Ch on the secondary structure of keratose at Rmax: (A) Mv = 5.91  105; (B) Mv =
`4.83  105; (C) Mv = 1.91  105; (D) Mv = 1.08  105; (E) Mv = 0.31  105; n, a-helix; m, h-sheet; n, random structure.
`(From Ref. 32.)
`
`Taravel and Domard investigated the interaction be-
`tween Ch and bovine atelocollagen (Col) in detail [33,34].
`When a solution of Ch HCl was added to a Col solution of
`pH 7.8, a pure PEC was formed. Interestingly, the aggre-
`gation of the triple helices of Col was not influenced by this
`interaction. During the formation of PEC, Col behaves like
`dispersions of encapsulated microgels. Accordingly, only
`some of the negative charges of Col can participate in the
`PEC. The weight proportion of Ch in these complexes was
`14.2% and 10.2% for low-molecular weight and high-
`molecular weight Col, respectively. These values are much
`lower than that of the theoretical value (28.5%). This
`limitation is attributed to a competition between the for-
`mation of PEC and Col triple helices. Moreover, they
`attempted to prepare a 1:1 complex by two different
`methods [35]. The first method was to form the complex
`at a temperature higher than the denaturation temperature
`of Col to avoid gelation of Col. Denaturation allowed the
`formation of a pure PEC with much higher Ch/Col ratio
`than the values obtained at lower temperature. However, a
`theoretical complex was not formed, as denaturation was
`not entirely completed. The presence of a large excess of Ch
`was the second method. The solution of Ch HCl was added
`to a Col solution of pH 3.6. The solutions thus prepared
`could contain a great excess of Ch up to 1200%. By in-
`creasing the pH of these mixture solutions to pH 5.8, the
`complexation between the two polymers was achieved. A
`large excess of Ch could destabilize the gelation of Col, and
`Col did not precipitate in the intermolecular associations
`
`of triple helices. The presence of an infrared (IR) absorp-
`tion band at 1600 cm1, the shift of the amide I band, and
`the insolubility at pH 5.6–5.8 of this complex suggested the
`formation of a hydrogen-bonding complex (HBC). HBC is
`often found in polymer blends, such as Ch–poly(viny
`alcohol) (PVA) [36] and N-acetylated Ch–PVA systems
`[37].
`
`Hydrophobic interactions are also important in the
`complexes with proteins. a-, h-,and n-caseins (Cas), which
`are phosphoproteins in milk, are precipitated by Ch [38].
`NaCl up to 1 mol/L was ineffective to prevent interaction
`between Ch and Cas, and a nonionic detergent, Tween 20,
`up to 2% was unable to prevent the Ch–Cas interactions.
`However, the complex could be dissolved in a mixture of
`increased NaCl and Tween 20. In this case, hydrophobic
`and electrostatic interactions participate in the association
`and coagulation of Cas with Ch. The fraction of hydro-
`phobic surface on a Ch molecule was estimated as
`51.5% at pH 6.3, and it increased slightly to 52.4% on
`increasing the ionization degree to 100% [39]. Therefore,
`an increase in the ionic strength would reinforce the
`hydrophobic interactions between Ch and hydrophobic
`residues of Cas. The complexation with Ch also imparts to
`faba bean legumin (FBL), a seed storage protein, the
`solubility at the isoelectric point, pI = 5.2, and higher
`pH [40]. Only at pH > 7.0 did the Ch–FBL complex
`precipitate from the solution due to the insolubility of Ch,
`and the complex was stable even at high ionic strength as
`1 M NaCl. This result points to a substantial contribution
`
`Copyright 2005 by Marcel Dekker
`
`685
`
`

`

`686
`
`Kubota and Shimoda
`
`of noncoulombic interactions to the Ch–FBL complex.
`The increase in negative deviations of limited viscosity
`numbers from additive values with increasing NaCl con-
`centration may be evidence for a substantial role of hydro-
`phobic interactions.
`
`D. Complexes with DNA
`
`As expected, Ch interacts very securely with DNA in
`solution and causes DNA to be precipitated from solution.
`This complexation is very similar to the Ch–polyphosphate
`(PP) [41], GC–metaphosphate (MP) [42], and MGC–MP
`[42] systems. Phosphoric acid is a slightly stronger acid
`than carboxylic acids. Therefore, Ch–DNA complex is
`stable in 0.1 mol/L HCl, but unstable in 0.1 mol/L NaOH.
`Kendra and Hadwiger showed that the Ch must be 7 or
`more sugar units in length both to optimally induce plant
`genes and to inhibit fungal growth [43]. This length re-
`quirement suggests that a series of positive charges match
`up with phosphate negative charges in the grooves of the
`DNA helix in the B form.
`Hayatsu et al. added Ch solution (pH 5) to the solu-
`tion of DNA, RNA, and homopolynucleotides (pH 7.2)
`to form insoluble complexes [44]. The double strandedness
`of DNA was retained in the complex because the PEC
`preparation process is mild enough not to inflict any
`damage on DNA. In this system, DNA molecule was
`accessible to enzymes and reagents having an affinity to
`
`DNA. For example, DNA in the complex could be
`digested with a mixture of DNase I and phosphodiester-
`ase, and cytosine residues in the DNA (denatured DNA)
`could be deaminated by treatment with sodium bisulfate.
`In addition, carcinogenic heterocyclic amines showed ad-
`sorption to DNA and RNA in the complex. On the
`contrary, it was reported that DNA complexed with Ch,
`at all charge ratios, resulted in a significant decrease in
`degradation by DNase II; the lower the molecular weight
`of Ch, the higher the inhibition effect [45].
`N-Dodecylated Ch (Ch12) also reacted with DNA to
`form a PEC [46]. Although DNA in the Ch–DNA complex
`was not sufficiently protected when it was exposed to
`DNase, DNA in the Ch12–DNA remained intact due to
`the protection from nuclease offered by Ch12. In addition,
`complexation with Ch12 enhanced the thermal stability of
`DNA. However, the complex was dissociated by the addi-
`tion of microions; the ability of Mg2+ to break the PEC
`was greater than that of Na+ and K+. This is related to the
`different affinity of ions to DNA; Mg2+ has a much higher
`affinity to DNA compared with Na+ and K+ [47].
`After complexation between Ch and DNA, phase
`separation occurs to yield coacervates that represent the
`aggregated colloidal complexes. The Ch–DNA particles
`had a negative surface charge when the complexes were
`made at N/P ratio below 2, and became positively charged at
`N/P ratio above 2 through neutral at N/P = 2, as shown in
`Fig. 8 [48]. The pH of the solution at 5–5.8 and a temperature
`
`Figure 8 N/P ratio dependence of zeta potential of Ch–DNA complex. Measurements were performed with 20 mg of DNA in
`1 mL of 0.15 mol/L NaCl. (From Ref. 48.)
`
`Copyright 2005 by Marcel Dekker
`
`686
`
`

`

`Macromolecular Complexes of Chitosan
`
`687
`
`Figure 9 Effect of pH on zeta potential of Ch–DNA nanoparticles. (From Ref. 49.) Copyright 2001, reprinted with permission
`from Elsevier Science.
`
`of the solution above 50jC resulted in Ch–DNA nano-
`particles [49]. The size of Ch–DNA nanoparticles was
`optimized to 150–250 nm with a narrow distribution at N/
`P ratio between 3 and 8 and Ch concentration of 100 Ag/mL.
`The zeta potential of the Ch–DNA nanoparticles was +22
`to +18 mV at pH < 6.5, and decreased dramatically to20
`mV at pH 8–8.5, as shown in Fig. 9. At pH 7.0–7.4, the
`nanoparticles appeared to be electrostatically neutral and
`may offer an effective protection to the encapsulated DNA
`from nuclease degradation.
`
`E. Ternary Complexes
`
`Kikuchi and his coworkers reported complexes consisting
`of three polyelectrolyte components, strong polybase–
`weak polybase–strong polyacid system and strong poly-
`base–strong polyacid–weak polyacid system.
`In the MGC–GC–poly(vinyl sulfate) (PVS) system,
`the molar ratio, S(PVS)/N(MGC+GC), to form insoluble
`complexes decreased with increasing solution pH [50,51].
`Experimental conditions and results of elemental analyses
`
`Table 2 Reaction Conditions and Elemental Analyses of MGCGCPVS Complexes
`
`Samplea
`
`1-A
`1-B
`1-C
`1-D
`1-E
`1-F
`2-A
`2-B
`2-C
`2-D
`2-E
`2-F
`
`[H+]
`
`7% HCl
`4% HCl
`1% HCl
`pH 2.0
`pH 6.5
`pH 13.0
`7% HCl
`4% HCl
`1% HCl
`pH 2.0
`pH 6.5
`pH 13.0
`
`Reacting conditions
`
`Molar ratio
`in mixture
`(S/N)
`
`Sulfur
`(%)
`
`Nitrogen
`(%)
`
`Molar ratio
`in PEC
`(S/N)
`
`2.00
`1.60
`1.20
`1.00
`0.80
`0.70
`2.00
`1.60
`1.20
`1.00
`0.50
`0.50
`
`6.70
`7.02
`8.15
`8.12
`8.10
`6.82
`6.83
`6.93
`7.68
`8.17
`6.51
`8.19
`
`2.63
`2.98
`3.13
`3.18
`3.04
`2.88
`2.08
`2.75
`3.10
`3.22
`3.19
`3.13
`
`1.11
`1.03
`1.14
`1.12
`1.16
`1.03
`1.43
`1.10
`1.08
`1.11
`0.89
`1.14
`
`a Series 1: PVS solution was added dropwise to MGC + GC solution. Series 2: MGC + GC solution was added dropwise to PVS solution.
`Source: Ref. 51. Copyright n 1988 Wiley. Reprinted by permission of John Wiley & Sons, Inc.
`
`Copyright 2005 by Marcel Dekker
`
`687
`
`

`

`688
`
`Kubota and Shimoda
`
`for each PEC prepared are given in Table 2. The compo-
`sitions of sulfur and nitrogen are almost the same. Because
`coagulation did not occur at pH > 6.5 in the GC–PVS
`system [52,53], only MGC was expected to react with PVS
`in the solution of pH > 6.5. However, in IR spectra, the
`absorption band at 1540 cm1 assigned to –NH3
`+ in GC
`was present in the PEC except for the PEC prepared at pH
`13.0, and the absorption band assigned to –NH2 that
`should appear at 1600 cm1 was absent. This is due to
`the strong induction effect by charged PVS as described
`above, because the absorption band at 1230 cm1 assigned
` in PVS was present in each PEC.
`to –OSO3
`The MGC–poly(L-glutamate) (PLG)–PVS system is
`more convenient for investigating the composition of the
`complexes by IR spectrometry, as each component in this
`system has a different reactive group [54]. The intensity of
`the IR absorption based on the –COO group increased
`with increasing pH from 2.0 to 5.0, and was constant at pH
`6.0–13.0. These results inferred that PLG was absent in the
`PEC at pH < 2, increased with increasing pH from 2.0 to
`5.0, and was constant at pH > 6.0. The intensities of the
`
`absorption band around 1240 cm1 assigned to the –OSO3
`groups of PVS for the PEC prepared at pH < 2.0 were
`medium, and those for the PEC prepared at 2.0 < pH <
`5.0 were weak. The absorption band around 1740 cm1
`assigned to the –COOH group of PLG was present in the
`PEC prepared at pH 4.0 and 5.0. These results were
`identical with those of the elemental analyses. This kind
`of PEC was also prepared in the MGC–carboxymethyl
`dextran (CMD)–PVS system under various pH conditions
`and in a different order of mixing, and similar results were
`obtained [55]. The PECs obtained are classified into three
`groups by means of elemental analysis, IR spectroscopy,
`and solubilities. The first is the group of PECs prepared at
`pH < 3, which are composed of MGC and PVS connected
`by ionic bond with each other. The second is the group of
`PECs formed at pH > 6.5, which are made up of MGC,
`CMD, and PVS linked by ionic bond with one another. The
`last is the group of PECs prepared at intermediate pH,
`which also constitute MGC, CMD, and PVS, but have a
`more complex three-dimensional network structure.
`
`F. Complexes by Template Polymerization
`
`Water-insoluble PECs of Ch–PAA and Ch–poly(styrene
`sulfonate) (PSS) can be obtained from the radical polymer-
`
`ization of acrylic acid (AA) and sodium 4-styrene sulfonate
`(NaSS), respectively, in the presence of Ch as a template at
`pH 4–4.5 [56]. It was concluded that the template polymer-
`ization technique appeared advantageous only for the
`synthesis of the Ch–PAA complex. The molecular weight
`of PAA obtained increased with increasing molecular
`weight of Ch, as shown in Table 3 [57].
`However, the structure and conformation of Ch mol-
`+ groups are separated from each
`ecule, in which –NH3
`other by a sequence of four atoms and oriented towards
`opposite directions in space, are disadvantages for the
`coordination of AA and NaSS molecules. Therefore, the
`effect of Ch as a template seems to be quite poor in
`comparison with that of vinyl polymer such as polyallyl-
`amine. At least the first step of this template polymeriza-
`tion is not an ionic preadsorption of monomer molecules
`onto Ch template, but the formation of a growing oligomer
`in the solution, which complexes with the template only
`beyond a critical chain length. In other words, an oligomer
`formation seems more favored than the polymerization of
`monomer molecules adsorbed onto Ch molecules.
`Ch–PAA complex nanoparticles were also prepared
`by template polymerization of AA in Ch solution [57]. The
`diameter of the nanoparticles was 200–300 nm, and the
`surface of the nanoparticles had positive charges. At pH
`7.4, PAA was highly extended while Ch was insoluble,
`resulting in the phase separation of nanoparticles where
`Ch was coated on the nanoparticles. These nanoparticles
`seem to form a sort of core–shell structure.
`
`III. PROPERTIES OF MACROMOLECULAR
`COMPLEXES OF CHITOSAN
`
`The properties of macromolecular complexes depend not
`only on the component polymers but also preparation
`conditions as mentioned above. The stoichiometric com-
`plexes ar

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