throbber
Cellulose (2007) 14:409–417
`DOI 10.1007/s10570-007-9137-9
`
`Viscosity properties of sodium carboxymethylcellulose
`solutions
`
`Xiao Hong Yang Æ Wei Ling Zhu
`
`Received: 2 April 2007 / Accepted: 27 May 2007 / Published online: 29 June 2007
`Ó Springer Science+Business Media B.V. 2007
`
`Abstract Through viscosity measurements, con-
`centration and temperature dependences of viscosity
`of sodium carboxymethylcellulose (CMC) solution
`were recorded. Effects of glycerin, mechanical
`shearing and several electrolytes on the CMC solu-
`tion were also determined. Results showed that the
`viscosity dependence on concentration obeyed the
`Huggins and Kramer equation, the dependence on
`temperature complied with the Arrhenius equation.
`CMC chain could synergize with glycerin, konjac
`glucomannan (KGM), and aluminum sulfate 18-
`hydrate. Sodium chloride, hydrochloric acid, and
`calcium dichloride reduced the viscosity of the CMC
`solution. By suggesting the ion-binding and hydrogen
`bond as the major form of the electrostatic interaction
`in the CMC solution, the synergistic and pseudoplas-
`tic phenomena as well as the maximum over stirring
`time were reasonably explained.
`
`Keywords Electrostatic interaction Ion-binding
`Hydrogen bond Polyelectrolyte Sodium
`carboxymethylcellulose Viscosity
`
`X. H. Yang (&) W. L. Zhu
`
`Department of Chemistry and Environmental
`Engineering, Wuhan Polytechnic University, Hankou,
`Wuhan 430023, P.R. China
`e-mail: yangxh88@yahoo.com.cn
`
`Introduction
`
`Sodium carboxymethylcellulose (CMC) is a deriva-
`tive of cellulose formed by its reaction with alkali and
`chloroacetic acid. Purified CMC is a white to off-
`white, non-toxic, odorless, biodegradable powder,
`which can be dissolved in hot or cold water. CMC is
`used for a variety of applications in a number of
`industries, including the food, personal care, phar-
`maceutical, oilfield and paper industries due to the
`superior properties as a binding,
`thickening, and
`stabilizing agent
`in these end uses. The most
`important character that makes them useful in these
`applications is high viscosity in low concentration.
`Literature has reported some progress in this basic
`concern about both CMC solution and mixture. After
`the apparent viscosity of CMC solutions as a function
`of shear rate,
`temperature, and concentration was
`modeled (Andriana et al. 2002), the steady shear
`viscosity of aqueous CMC solution was measured
`within the power-law range over temperature and
`concentration (Lin and Ko 1995). By studying the
`Newtonian behavior of aqueous CMC solution, effect
`of shear-induced recombination of CMC macromo-
`lecular crystallites was proposed (Jayabalan 1989).
`Recent years,
`rheological behavior of CMC in
`aqueous solution from non-wood pulps was synthe-
`sized and characterized (Barba et al. 2002). Mixtures
`of CMC with natural or synthesized polymers behave
`more complicated viscous properties and attract more
`attentions. In order to evaluate synergistic/non-syn-
`
`123
`
`ALKERMES Exh. 2035
`Luye v. Alkermes
`IPR2016-1096
`
`

`

`410
`
`Cellulose (2007) 14:409–417
`
`ergistic effects of mixed polysaccharide systems,
`rheological character of the mixture of CMC with
`xanthan was studied under destructive and non-
`destructive shear conditions (Florjancic et al. 2002).
`By means of viscosity measurement,
`the CMC
`interaction with mucin (Rosi et al. 1996) and with
`5-fluorouracil (Nishida et al. 1982), an anticancer
`drug, were reported. Although these works extended
`the knowledge about
`the viscosity of
`the CMC
`solution, aspects of explanation are not sufficient to
`guide the CMC application.
`The sodium CMC molecular structure is based on
`the b-(1?4)-D-glucopyranose polymer of cellulose.
`Different preparations may have different degrees of
`ionizable group substitution, but it is generally in
`the range 0.6–0.95 derivatives per monomer unit.
`When it dissolves in water, an electrolytic process
`takes place to separate a CMC molecule into sodium
`cations and a polymer anion. In this sense, the CMC
`belongs to polyelectrolyte. These ions in solution
`interact with each other through electrostatic forces.
`In addition to this,
`the water molecule and OH
`groups on the CMC molecule exhibit electric dipole
`that performs considerable electrostatic interacting
`force (the so called hydrogen bond). These electro-
`static interactions play a key role to understand the
`viscosity of
`the CMC polyelectrolyte solution.
`Regarding this important issue, investigations have
`been carried out with experiment on the CMC
`solution filled with low molecular weight salt and
`with theoretical model
`for polyelectrolyte. The
`effect of polyion charge on specific viscosity of
`CMC (Trivedi et al. 1987), and viscosity behavior of
`multivalent metal
`ion-containing carboxymethyl
`cellulose solutions (Thomas et al. 2003) were
`studied. Rheological properties of CMC aqueous
`solution with sodium and chromium chloride salts
`were measured (Matsumoto and Mashiko 1988;
`Andreeva et al. 1992). Polyelectrolyte effects in
`the CMC water-cadoxene solution was studied by
`translational diffusion and viscometry methods
`(Okatova et al. 1990). Although several theoretical
`models (Markus et al. 1997; Kunimasa et al. 2004;
`Dobrynin and Rubinstein 2005)
`for
`the reduced
`viscosity of polyelectrolyte solution may be suitable
`to the CMC solution, correlation between electro-
`static interaction and the viscosity properties of
`the CMC solution is not understood well. For
`example,
`the influence of
`ion-binding on the
`
`123
`
`viscosity properties of the CMC solution has seldom
`been reported. Because of the wide applications of
`CMC and its typical representation for polyelectro-
`lyte, investigation into the electrostatic interaction
`by viscosity measurement has practical and theoret-
`ical values.
`In this paper, the viscosities of CMC solutions will
`be measured over concentration and temperature.
`And then, the influences of several added salt ions,
`glycerin, konjac glucomannan (KGM) and shear rate
`will be determined. Recorded curves will be respec-
`tively fitted to Huggins and Kramer equation (Mothe´
`and Rao 1999), Arrhenius equation (Rubinstein and
`Colby 2003), the equation for polyelectrolyte solution
`and Williamson model. Obtained result will be
`discussed with electrostatic and dynamic interactions
`in the solution.
`
`Experimental
`
`Materials
`
`Sodium CMC in this study was made by Shanghai
`Reagent Co., Ltd., of Sinopharm. It was a white to
`off-white, non-toxic, odorless, biodegradable powder,
`which obeyed the Q/CYDZ-03-92 specification.
`Sodium content located in 6.5–8.5%, chloride com-
`pound  3.0%, water  1.0% in weight. The CMC
`were directly dissolved in distilled hot water to obtain
`the solution samples. KGM sample used in our
`experiment was supplied by Hubei Enshi Hongye
`Konjac Development Co., Ltd., which had principal
`properties: granularity  220 mm, glucomannan
`content  95%, protein content <0.6%, ash con-
`tent <3%. The other reagents were chemical grades.
`
`Methods
`
`For the samples, a NDJ-1 viscometer, equipped with
`column geometries of four scales numbered 1–4 and
`working in steady shear mode, was used to measure
`the viscosities. An electromantle water bath was used
`to control the temperature. To measure the shear rate
`dependence, a rheometer AR500 by TA instruments
`was set to steady state flow. A standard steel parallel
`plate with 1,000 mm gap and 40 mm diameter was
`chosen for the steady flow measurement.
`
`

`

`Cellulose (2007) 14:409–417
`
`411
`
`2
`
`The fitting results were evaluated by a standard
`error defined in Eq. (1).
`ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
`r
`P
`ðxmÿxcÞ
`n
`ð1Þ
`range
`where xm was the measured value and xc was the
`calculated value of x for each data point, n was the
`number of data points, and the range was the
`maximum value of xm minus the minimum value.
`Generally speaking, a reasonable fit gives a standard
`error  3%.
`
`Results and discussion
`
`Dissolving process with mechanical stirring was the
`necessary step to make sodium CMC solution,
`therefore it was important to measure the viscosity
`during this process. The measured viscosity data at
`50 rounds/min stirring angular velocity was presented
`in Fig. 1 as hollow diamonds with smoothly lined
`curve for clarity. It showed that viscosity initially
`increased to a maximum, and then decreased to a
`stable value from the moment when CMC was added
`in water to a sufficient long period. Firstly, a part of
`solid CMC dissolving in water resulted in a sharp
`viscosity increase, while the others remained as
`suspended powder, neither swelling nor dissolving.
`With continuous stirring, the majority of the solid
`CMC had dissolved into the water reaching a point of
`
`maximum viscosity without complete disaggregation
`of CMC molecule. Finally,
`the CMC molecules
`reached maximum disaggregation corresponding to
`an unchangeable viscosity value. These phenomena
`in the dissolving process proposed that the CMC
`solution had its maximum viscosity at an intermedi-
`ate state at which the CMC polymer remained some
`molecular binding.
`In order to obtain a basic knowledge into the flow
`properties of the sodium CMC solution, the depen-
`dences of viscosity upon concentration in weight
`percent and temperature were showed in Figs. 2 and 3
`as smoothly lined points of various shape (The
`following figures mean the same). Viewing Fig. 2, it
`was clear that the viscosity exponentially increased
`with concentration. A lot of polymer solution had this
`tendency that was classically described by the
`Huggins and Kramer equation (Mothe´ and Rao 1999),
`
`ln gr
`c ¼ g½ Š þ k2 g½ Š2c
`ð2Þ
`where gr = g/gs was the relative viscosity, g, gs the
`viscosities of the solution and the solvent, g in a
`[g] was a variable traditionally
`square bracket
`denoting the intrinsic viscosity, c the concentration,
`k2 the constant. After curve in the Fig. 2 was fitted to
`Eq. (2), good agreement was obtained. The corre-
`sponding parameters were numerically determined as
`[g] = 115, gs = 31.6 mPa.s, k2 = 0. Standard error
`between the experimental data and the theoretical
`data calculated by Eq. (2) with this set of constants
`was 1.8%.
`
`10000
`
`1000
`
`100
`
`Viscosity (mPa.s)
`
`0
`
`5
`
`15
`10
`Stirring Time (min)
`
`20
`
`25
`
`10
`
`0
`
`1
`
`4
`3
`2
`Concentration (%)
`
`5
`
`6
`
`1500
`1400
`1300
`1200
`1100
`1000
`900
`800
`700
`600
`
`Viscosity (mPa.s)
`
`
`
`Fig. 1 Viscosity variation during the dissolving process of
`CMC in water with a final concentration of 3% by mechanical
`stirring at 50 rounds/min angular velocity
`
`Fig. 2 Viscosity of CMC solution against concentration in
`weight percent
`
`123
`
`

`

`Cellulose (2007) 14:409–417
`
`full of electrostatic interactions, by which ion-binding
`and hydrogen bond might connect
`two or more
`ployions as an effective larger polyion. Just like the
`neutral polymer solution, the temperature stochastic
`process in the solution tended to separate some bound
`ions to an equilibrium state. There was thus a balance
`at which the entire interacting mechanism in the
`solution reached. Each balance corresponded to an
`activation energy value in Eq. (3), which determined
`the temperature dependence of viscosity.
`Beyond these basic properties, sodium CMC
`solution performed much more interesting flow
`properties in the presence of low molecular weight
`electrolytes, which introduced cation and anion into
`the solution, such as NaCl. The properties of the
`CMC solution could be obviously altered by these
`ions. Figure 4 illustrated the influences of various
`ions from added salts on the viscosity of 1% CMC
`solution. It showed that sodium chloride, hydrochlo-
`ric acid, calcium dichloride and ferric chloride
`decreased the viscosity when increasing their molar
`concentration. However, aluminum sulfate 18-hy-
`drate increased the viscosity while increasing its
`molar concentration to reduce the pH value of
`solution from 7 to 6. For aluminum ion, a similar
`result was previously reported for Al3+ to form CMC
`solution as rigid elastic gel (Elliot and Ganz 1974).
`Mechanically, viscosity is the ratio of stress and
`shear rate, which describes the ability for faster fluid
`layer to draw slower fluid layer. The more expanded
`polymer in solution can connect more distant layers,
`so enhances the ability to transfer drawing force from
`
`Al2(SO4)3*18H2O
`NaCl
`HCl
`CaCl2*2H2O
`FeCl3
`
`200
`
`150
`
`100
`
`50
`
`Viscosity (mPa.s)
`
`0
`
`0
`
`0.04
`0.03
`0.02
`0.01
`Molar concentration (mol/L)
`
`0.05
`
`Fig. 4 Effects of various salts on the viscosity of 1% CMC
`solution
`
`1%
`3%
`5%
`
`20
`
`60
`40
`Temperature (°C)
`
`80
`
`100
`
`10000
`
`1000
`
`100
`
`10
`
`1
`
`0
`
`412
`
`Viscosity (mPa.s)
`
`Fig. 3 Viscosities of CMC solutions at 1, 3, 5% concentrations
`versus temperature
`
`the viscosities of CMC
`Figure 3 showed that
`solutions behaved like most water-soluble polymers,
`whose viscosity decreased when temperature in-
`creased. The measured data could be satisfactorily
`fitted to Arrhenius equation given by
`
`ER
`
`gðTÞ ¼ Aexp
`ð3Þ
`T
`where E was the activation energy and R = 8.31 J/
`8K.mole was the gas constant, T was absolute
`temperature, and A was a pre-exponential constant.
`The fitted parameters A and E for 1, 3, 5% CMC
`solutions were 2.29 · 10ÿ4 mPa.s and 3.17 · 104 J/
`mole, 6.50 · 10ÿ4 mPa.s and 3.46 · 104 J/mole,
`3.72 · 10ÿ3 mPa.s and 3.61 · 104 J/mole, respec-
`tively. Standard error between experimental and
`theoretical data for these sets of constants were
`within 2.1%.
`A number of literatures have reported similar
`results for
`the viscosity of neutral hydrocarbon
`polymer solution varying with concentration and
`temperature. They were commonly interpreted with
`both the dynamics of the neutral polymer chains in
`solution and the stochastic effect of the temperature.
`In the case of sodium CMC solution, which belonged
`to a polyelectrolyte solution, electrostatic interaction
`among ions globally existed in the solution. Not only
`the polyions electrolysed from the CMC interacted
`with the small counterions which rendered the system
`electroneutral, but also the electric dipole of the OH
`group on the CMC molecule joined in the interaction.
`As a matter of fact, the CMC solution was a system
`
`123
`
`

`

`Cellulose (2007) 14:409–417
`
`413
`
`Table 1 Fitted parameters to Hess and Klein formula
`c0 (mol/L)
`
`Added salt
`
`g0 (mPa.s)
`
`Standard
`error (%)
`
`NaCl
`
`HCl
`CaCl2 2H2O
`FeCl3
`
`105
`
`112
`
`113
`
`113
`
`0.0958
`
`0.0358
`
`0.0453
`
`0.0309
`
`3.2
`
`3.9
`
`4.2
`
`4.6
`
`here, ca was a constant, and ga was the viscosity
`amount in the absence of added salt. It was clear that
`the variation trend of this equation qualitatively
`agreed with measured data in Fig. 4 for sodium
`chloride, hydrochloric acid, calcium dichloride and
`ferric chloride. Quantitative numerical fitting results
`were listed in Table 1 for these salts. Standard error
`between the experimental data and the theoretical
`data calculated with Eq. (5) were within 5%.
`As an example of the importance of order of
`addition for the small ions, Fig. 5 was a graph of
`viscosity versus salt concentration at 2% CMC
`concentration. In one case, the CMC was dissolved
`in the water before the sodium chloride salt, and the
`salt had a minimal effect on the viscosity of the
`solution. In the other case, the CMC was dissolved
`after the salt, and the resulting final viscosity was
`much lower, especially as the salt concentration
`increased. As pointed out above,
`if the salt was
`dissolved in water before CMC, the salt ion promptly
`shielded the dissolved CMC chains stopping the
`expansion. In contradiction, the salt ion could not
`depress all the expansion by shielding if the salt were
`
`After CMC
`Before CMC
`
`0.01
`
`0.1
`Molar concentration of cation
`
`1
`
`300
`250
`200
`150
`100
`50
`0
`
`Viscosity (mPa.s)
`
`faster to slower fluid layers, and results in higher
`viscosity in the solution. In the presence of low
`molecular weight ions, the ionized –COO groups on
`CMC chain was electrostatically shielded and the
`CMC chains then adapted a less expanded structure.
`The viscosities were thus reduced by the small ions,
`such as sodium cation, chloride anion, hydrogen
`cation, and so on. For gel
`formation, a higher
`concentration was thus required in the presence of
`salts than for a small ion free medium. However,
`when the aluminum sulfate dissolved in the CMC
`solution, hydrolyzing and polymerizing reactions
`occurred in the solution to produce colorless and
`viscous Al(OH)3 colloid. This colloid itself increased
`the viscosity of the CMC solution. Futhermore, the
`OH groups on the colloid had the steric capability of
`binding CMC polyions in solution through hydrogen
`bond. Therefore, the viscosity of CMC solution was
`actually increased by two mechanism. The hydrolyz-
`ing and polymerizing reactions are as follows,
`
`Hydrolyzing reaction 2Al2(SO4)3 + 2n H2O =
`2Al2(OH)n (SO4)3ÿn/2 + nH2SO4
`Polymerizing reaction mAl2(OH)n (SO4)3ÿn/2 =
`[Al2(OH)n (SO4)3ÿn/2]m
`The electrostatic shielding effect commonly exits in
`polyelectrolyte solution. By considering the influence
`of charges on the single hydrodynamics of a poly-
`electrolyte and the cooperative coupling of all parti-
`cles, a formula for viscosity of polyelectrolyte solution
`at low polymer concentration was calculated (Hess and
`Klein 1983). It was expressed as a reduced viscosity,
`cpZ4
`ð2MsMp cs þ ZeffcpÞ3=2
`whereZeff was the effective charge number per polymer,
`Mp andMs the molecular weights of the polymer and low
`molecular weight salt in solution, cp and cs the concen-
`trations of the polymer and salt. When considering the
`dependence of viscosity on the molar concentration of
`added salt, it was beneficial to assume this formula
`could be extended to the solution at high concentration of
`this paper by fixing the other variables at constant.
`On this assumption, a formula describing the effect of
`added salt on the CMC viscosity was proposed as,
`
`gred /
`
`eff
`
`ð4Þ
`
`g ¼ ga
`
`c3=2a
`ðcs þ caÞ3=2
`
`ð5Þ
`
`Fig. 5 The viscosities of CMC solution with sodium chloride
`addition when the salt added after and before CMC
`
`123
`
`

`

`414
`
`Cellulose (2007) 14:409–417
`
`mixture of water and glycerin. At higher than 70%
`concentrations of glycerin, the CMC was not fully
`dissolved in solution and thus did not give as much
`viscosity. There were three OH groups on a glycerin
`molecule. Unlike molecules prefer to interact with
`each other. So it was more reasonable to assume
`hydrogen-bonds on a glycerin molecule bound three
`CMC molecules as an effective larger molecule.
`According to Mark–Houwink equation, this effect
`(the so called synergy) increased the viscosity of the
`solution.
`Other hydrocolloids can also give a synergistic
`viscosity increase with sodium CMC. Figure 8 pre-
`sented an example. If one were to mix a 2% KGM
`solution of 340 mPa.s with a 2% CMC solution of
`380 mPa.s, the net result was not the 360 mPa.s
`average of the two;
`it would be higher than the
`average value. With the ratio of CMC solution to
`total solution rising from 0 to 100%, the viscosity
`went up to a maximum, and then gradually dropped
`down to the average value. As is well known that
`KGM molecule contains glucose and mannose on
`which there are rich hydroxyls. The hydroxyl would
`form hydrogen bonding between the KGM and CMC
`molecules in the solution. There were more average
`electrostatic attractions (hydrogen bonding) between
`unlike molecules, which resulted in this synergistic
`viscosity increase. The occurrence of the interactions
`between KGM and CMC molecular chains through
`hydrogen bond was previously found in blend films
`of KGM and CMC by using FT-infrared, wide-angle
`X-ray diffraction, and differential thermal analysis
`(Xiao et al. 2001). The ratio of synergistic viscosity
`
`dissolved after CMC. The higher ion concentration
`the larger stoppage in the solution leaded to more
`viscosity drops. This effect moderately reduced the
`viscosity when the molar concentration of cation
`increased. Furthermore, a fraction of the CMC chains
`might have bound with each other before the salt
`dissolved. This might be the other reason for the
`viscosity of the solution had much higher viscosity
`value when the salt was added after CMC than before
`CMC.
`Figures 6 and 7 showed examples of the effect of
`water/non-solvent mixtures on the viscosity of CMC
`solution at various concentrations and temperatures.
`In this test,
`the non-solvent was glycerin. With
`increasing glycerin concentration, the viscosity went
`up. The maximum value was reached with a 30/70
`
`5%
`3%
`1%
`
`100000
`
`10000
`
`1000
`
`100
`
`Viscosity (mPa.s)
`
`10
`
`0
`
`20
`
`60
`40
`% water
`
`80
`
`100
`
`Fig. 6 CMC viscosity in water–glycerin solution at 1, 3, 5%
`concentration
`
`Synergistic
`value
`Average value
`
`0
`
`20
`
`60
`40
`% CMC
`
`80
`
`100
`
`700
`
`600
`
`500
`
`400
`
`300
`
`Viscosity (mPa.s)
`
`
`
`20°C
`40°C
`60°C
`60
`40
`% water
`
`80
`
`100
`
`100000
`
`10000
`
`1000
`
`100
`
`10
`
`0
`
`20
`
`Viscosity (mPa.s)
`
`Fig. 7 CMC viscosity in water–glycerin solution at 20, 40,
`60 8C
`
`Fig. 8 Synergistic viscosity with konjac glucomannan
`
`123
`
`

`

`415
`
`20°C
`40°C
`60°C
`
`Cellulose (2007) 14:409–417
`
`10000
`
`1000
`
`100
`
`Viscosity (mPa.s)
`
`to average value could be roughly estimated by using
`Mark–Houwink equation. Considering a simple
`binary interaction, a hydrogen bond connected a
`KGM and a CMC chain as an effective chain that had
`the molecular weight of their sum. Keeping concen-
`trations of CMC and KGM at constant and ignoring
`the small effect of solvent viscosity, the ratio of
`synergistic viscosity to average value could be
`derived as
`
`10
`
`0.1
`
`1
`
`10
`Shear rate (s-1)
`
`100
`
`1000
`
`Fig. 10 Dependence of viscosity of CMC solution upon shear
`rate at 20, 40, 60 8C and 3% concentration
`
`viscosity was higher because random molecular
`orientations exhibited increased resistance to flow.
`In addition to this orientation effect,
`there were
`enough OH groups on a CMC chain, which could
`cause hydrogen bonding among CMC polymers. The
`electrostatically bound state of the CMC chain in the
`solution was not stable. Strong mechanical shearing
`could break some hydrogen bonds in the solution
`leading to smaller effective molecular weight.
`According the Mark–Houwink equation, smaller
`molecule weight corresponded to lower viscosity.
`Due to the relatively high bonding energy,
`it
`suggested that
`this mechanism preferred to cause
`pseudoplastic property in higher shear rate regime.
`The reversible pseudoplastic property of CMC solu-
`tion reflected the break and recover of both hydrogen
`bond and random molecular orientation.
`Due to the behavior of the CMC solution was
`pseudoplastic,
`the most suitable model
`to fit
`the
`experimental curves of viscosity versus shear rate in
`Figs. 9 and 10 might be the Williamson model given
`by
`
`g0
`
`n
`
`g ¼
`1 þ ðk_cÞ
`where k was the consistence index, n was the flow
`behavior index and g0 was the zero-rate viscosity.
`Numerically fitting the data in Figs. 9 and 10 to this
`equation, parameters were determined in Table 2.
`Standard errors between the experimental data and
`the theoretical data calculated with Eq. (7) were
`consistently within 2%. At the same temperature, all
`
`123
`
`a
`
`ratio  ðMCMC þ MKGMÞ
`ð6Þ
`
`
`ðMaCMC þ MaKGMÞ
`Here, MCMC and MKGM respectively denoted the
`molecular weight of the CMC and KGM polymers, a
`the Mark–Houwink constant that usually takes the
`in 0.5*1. No matter what amounts the
`amount
`molecular weights took,
`the numerical value of
`Eq. (6) was bigger than 1 to show a viscosity synergy.
`Figures 9, 10 demonstrated changes of viscosity
`with shear rate at different sodium CMC concentra-
`tions and temperatures. Measured results showed
`totally reversible pseudoplastic (or shear-thinning)
`properties, which meant that the viscosity decreased
`at increasing shear rate. As soon as the shear stopped,
`the viscosity returned to its original value. These
`pseudoplastic properties could be attributed to
`CMC’s long chain molecules that tended to orient
`themselves in the direction of flow. As the shear
`stress was increased,
`the more chains rearranged
`themselves along the shear direction, then the shear
`resistance to flow (viscosity) decreased. When a
`lower stress was imposed on the same solution, the
`
`12
`
`1%
`3%
`5%
`
`100000
`
`10000
`
`1000
`
`100
`
`Viscosity (mPa.s)
`
`
`
`100.1
`
`1
`
`10
`Shear rate (s-1)
`
`100
`
`1000
`
`Fig. 9 Dependence of the viscosity of CMC solution upon
`shear rate at 1, 3, 5% concentration and 20 8C
`
`

`

`416
`
`Cellulose (2007) 14:409–417
`
`Table 2 Fitted parameters to Williamson model
`n
`
`g0 (mPa.s)
`
`K (s)
`
`Condition
`
`Standard
`error (%)
`
`pseudoplastic property of the CMC solution in
`the high shear rate regime.
`
`Investigation into the complicated interaction due
`to Coulomb force in the polyelectrolyte solution is
`one of the most important theoretical problems that
`interest the polymer chemist. CMC not only has the
`properties of both polymer and electrolyte, but also is
`a typical example of the polyelectrolyte. Appending
`the wide variety of application of CMC, the presented
`results in this article have theoretical and practical
`values.
`
`References
`
`Andreeva IA, Romanova LG, Sushko VA (1992) Investigation
`of rheological properties: viscosity and nuclear-relax-
`ational characteristics of gels on the basis of Na-CMC
`solutions with chromium chloride. Kolloidnyi Zhurnal
`54(2):21–27
`Andriana EV, Tunc KP, Sandeep KP et al (2002) Rheological
`characterization of carboxymethylcellulose solution under
`aseptic processing conditions.
`J Food Process Eng
`25(1):41–61
`Barba C, Montane D, Farriol X et al (2002) Synthesis and
`characterization of carboxymethylcelluloses from non-
`wood pulps II. Rheological behavior of CMC in aqueous
`solution. Cellulose 9(3–4):327–335
`Dobrynin AV, Rubinstein M (2005) Theory of polyelectrolytes
`in solutions and at surfaces. Prog Polym Sci 30:1049–
`1118
`Elliot JH, Ganz AJ (1974) Some rheological properties of so-
`dium carboxymethylcellulose solutions and gels. Rheol-
`ogica Acta 13(4–5):670–674
`Florjancic U, Zupancic A, Zumer M (2002) Rheological
`characterization of aqueous polysaccharide mixtures
`undergoing shear. Chemical
`and Biochemical Eng
`16(3):105–118
`Hess W, Klein R (1983) Generalized hydrodynamics of sys-
`tems of Brownian particles. Adv Phys 32:173–283
`Jayabalan M (1989) Newtonian behaviour of sheared aqueous
`carboxymethylcellulose solution on aging. Br Polym
`21(3):233–235
`Kunimasa M, Biman B, Arun Y (2004) Self-consistent mode-
`coupling theory for the viscosity of rodlike polyelectrolyte
`solutions. J Chem Phys 121(16):8120–8127
`Lin CX, Ko SY (1995) Effects of temperature and concentra-
`tion on the steady shear properties of aqueous solutions of
`Carbopol and CMC. Int Commun in Heat and Mass
`Transfer 22(2):157–166
`Markus A, Andreas B, Stephan F (1997) Quantitative
`description of the intrinsic viscosity of branched poly-
`electrolytes. Macromolecules 30:2700–2704
`Matsumoto T, Mashiko K (1988) Influence of added salt on
`dynamic viscoelasticity of carboxymethylcellulose aque-
`ous systems. Polym Eng and Sci 28(6):393–402
`
`5% at 20 8C
`3% at 20 8C
`1% at 20 8C
`3% at 40 8C
`3% at 60 8C
`
`1.12 · 104
`1.12 · 103
`1.20 · 102
`5.00 · 102
`1.66 · 102
`
`0.198
`
`0.0879
`
`0.0741
`
`0.0917
`
`0.0129
`
`0.742
`
`0.593
`
`0.409
`
`0.438
`
`0.410
`
`1.2
`
`1.6
`
`1.5
`
`1.9
`
`1.4
`
`the three parameters increased when concentration
`rose. Among the three parameters the most sensitive
`one was the zero-rate viscosity g0.
`
`Conclusion
`
`the viscosities of sodium CMC
`In this paper,
`solutions over concentration, temperature, and the
`influences of several added salt ions, glycerin, KGM
`and shear rate were determined. Recorded curves
`were respectively fitted to Huggins and Kramer
`equation, Arrhenius equation, the equation for poly-
`electrolyte solution and the Williamson model. By
`suggesting the electrostatic and dynamic interactions
`as well as temperature stochastic process as the
`fundamental mechanism in the CMC solution, the
`measured data were reasonably interpreted. It was
`concluded that:
`
`1. The concentration and temperature dependences
`of viscosity of CMC solution obeyed the Huggins
`and Kramer equation and the Arrhenius equation,
`respectively.
`2. CMC chain could synergize with aluminum ion,
`glycerin and KGM. This character provided a
`useful method to make CMC product of special
`properties, which was required by the increasing
`industrials.
`3. Electrostatic interaction globally existed in the
`CMC solution. It was the key factor to explain
`the viscous properties of CMC solution. Ion
`binding and hydrogen bonding were the most
`important interactive forms.
`4. The balance of hydrogen bonding in CMC
`solution was at quasi-static, which could be
`broken by strong mechanical shearing. This
`mechanism might be one of the reasons of the
`
`123
`
`

`

`Cellulose (2007) 14:409–417
`
`417
`
`Mothe´ CG, Rao MA (1999) Rheological behavior of aqueous
`dispersions of cashew gum and gum arabic: effect of
`concentration and blending. Food Hydrocolloids 13:501–
`506
`Nishida K, Ando Y, Enomoto M (1982) Interaction of 5-fluo-
`rouracil with sodium carboxymethylcellulose. Colloid and
`Polymer Sci 206(5):511–513
`Okatova OV, Lavrenko PN, Dautzenberg H (1990) Polyelec-
`trolyte effects in diffusion and viscosity phenomena in
`water-cadoxene
`solutions of
`carboxymethylcellulose.
`Polymer Science USSR (English Translation of Vy-
`sokomolekulyarnye Soyedineniya Series A) 32(3):533–539
`Rossi S, Bonferoni MC, Ferrari F et al (1996) Characterization
`of mucin interaction with three viscosity grades of sodium
`carboxymethylcellulose. Comparison between rheological
`and tensile testing. Eur J Pharm Sci 4(3):189–199
`
`Rubinstein M, Colby RH (2003) Polymer physics. Oxford
`University Press, New York, p 337
`Thomas H, Ute H, Dieter K (2003) Viscosity behaviour of
`multivalent metal ion-containing carboxymethyl cellulose
`solutions. Angewandte Makromolekulare
`Chemie
`220(1):123–132
`Trivedi HC, Patel RD (1987) Studies on carboxymethylcellu-
`lose: 3. Effect of polyion charge on specific viscosity.
`Polymer
`Communications
`(Guildford,
`England)
`28(5):158–159
`Xiao CB, Lu YS, Liu HJ, Zhang LN (2001) Preparation and
`characterization of konjac glucomannan and sodium car-
`boxymethylcellulose blend films. J Appl Polym Sci
`80(1):26–31
`
`123
`
`

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