Está en la página 1de 8

Food Hydrocolloids 19 (2005) 313320 www.elsevier.

com/locate/foodhyd

Rheological behaviour of carboxymethylcellulose manufactured from TCF-bleached Milox pulps


Sonia Dapaa, Clara Asuncion Tovarb, Valentn Santosa, Juan Carlos Parajoa,*
a

Department of Chemical Engineering, University of Vigo (Campus Ourense), Polytechnical Building, Ourense, Spain b Department of Applied Physics, University of Vigo (Campus Ourense), Ourense, Spain Received 20 March 2004; revised 2 June 2004; accepted 30 June 2004

Abstract Carboxymethylcellulose (CMC) samples with different degrees of substitution were obtained from Eucalyptus globulus pulps made in formic acid-peroxyformic acid medium (Milox delignication) and subjected to Totally Chlorine Free bleaching. In order to assess the rheological behaviour of the synthesized materials, their properties (rheological behaviour, solubility and molar mass) were determined and examined in relation to their total degree of substitution. Additional discussion on the structure and purity of pulps, alkali celluloses, and CMC is provided. q 2004 Elsevier Ltd. All rights reserved.
Keywords: Carboxymethylcellulose; Eucalyptus wood; Milox pulping; Rheological properties

1. Introduction Carboxymethylcellulose (CMC) is the most important cellulose ether, with an annual world production of about 3!105 metric tons (Heinze & Liebert, 2001). A large number of commercial CMCs exist, with a variety of substitution degrees (in the range 0.51.4), purity and rheological properties in aqueous solutions. Over 300 different types of CMC are produced all over the world (Keller, 1997). CMCs are usually employed in the food industry as viscosity-increasing, thickening and stabilizing agents. Preparation and use of CMC aqueous solutions involves an extensive range of shear conditions due to the wide variety of processes where they are employed (such as mixing, storing, pumping, heating, cooling, separation, etc). Aqueous solutions of CMC may have different rheological properties depending on the synthesis conditions, total degree of substitution (DS), molar mass, concentration and temperature (Ghannam & Esmail, 1997; Kulicke,
* Corresponding author. Tel.: C34-9-8838-7033; fax: C34-9-88387001. E-mail address: jcparajo@uvigo.es (J.C. Parajo). 0268-005X/$ - see front matter q 2004 Elsevier Ltd. All rights reserved. doi:10.1016/j.foodhyd.2004.06.007

Kull, Kull, & Thielking, 1996; Barba, Reguant, Farriol & Montane, 2002; Vais, Palazoglu, Sandeep, & Daubert, 2002). Many studies on aqueous CMC systems have been reported in literature: DeButts, Hudy and Elliott (1957) characterised the ow properties of CMCs with DS in the range 0.650.85 at 13% concentration in aqueous solution, and concluded that non-uniform substitution along the molecular chain is a key parameter for the developing thixotropic behaviour. Abdelrahim and Ramaswamy (1995) studied the effects of concentration (0.52.0%) and temperature (60140 8C) on the rheological behaviour of CMC solutions, and developed empirical equations to predict the consistency and ow behaviour index from both variables. Ghannam and Esmail (1997) reported a comprehensive study on the rheological properties of CMCs at different concentrations (15%), including steady-state ow, transient shear stress responses and dynamic test measurements. These authors found that aqueous samples with increased CMC concentrations resulted in stronger time-dependent ow and increased viscoelastic properties. CMCs are usually manufactured from cotton linters or from wood after pulping and subsequent bleaching with chlorine-containing chemicals. In this eld, pulping

314

S. Dapa et al. / Food Hydrocolloids 19 (2005) 313320

processes based on the utilization of organic solvents (organosolv processes) can play a signicant role. Milox pulping (nomenclature derived from Milieu Pure Oxidative), based on the utilization of both formic acid and peroxyformic acid (formed in situ by reacting formic acid with hydrogen peroxide) is one of the most promising organosolv processes. Milox pulping shows interesting features, including: (i) environmentally hazardous chemicals are not used at all, (ii) pulping by-products are suitable for further chemical utilisation, (iii) the process is suitable for wide range of raw materials (Obrocea and Cimpoesu, 1998; Dapa, Santos, & Parajo, 2000) and the pulps show satisfactory brightness, intrinsic viscosity and ease of bleaching (Dapa, Sixta, Borgards, Harms, & Parajo, 2002, 2003). This work deals with the physicochemical characterisation and rheological properties of CMCs made from Miloxdelignied, TCF-bleached Eucalyptus globulus pulps in aqueous solutions. The effects of the DS and concentration on the rheological properties are reported.

aqueous ethanol (70% v/v), washed again with pure methanol and dried at 60 8C under vacuum for 24 h. 2.4. Analytical characterisation of cellulose, alkali cellulose and CMC The chemical composition of CMC samples was determined by HPLC according to Heinze and Pfeiffer (1999) and by 1H-NMR according to Ho and Klosiewicz (1980). X-ray diffraction of cellulose, alkali cellulose and CMC samples was performed with a Bruker Smart-CCD 1000 in the angular range 10708 (2q). The cristallinity index (CI) was calculated according to Buschler-Diller and Zeroniam (1992): CI Z 1 K Imin =Imax (1)

2. Materials and methods 2.1. Materials Eucalyptus globulus wood samples were kindly provided by a local mill (ENCE, Pontevedra, Spain). Wood chips were screened to select a fraction of particles with size below 2 mm, homogenised in a single lot to avoid compositional differences and then stored in polyethylene bags. 2.2. Milox delignication and TCF bleaching sequences Delignication treatments were carried out in glass reactors with temperature control under reported operational conditions (Abad, Santos, & Parajo, 2001). A TCF bleaching sequence of Milox pulps (EOZQP) was carried out according to Abad, Saake, Puls, and Parajo (2002). 2.3. Carboxymethylation of cellulose in isopropanol/water/NaOH medium Carboxymethylation was carried out by a standard slurry method according to Dapa, Santos, and Parajo (2003). Bleached pulp (3 g) was suspended in isopropyl alcohol (180 ml) under mechanical stirring at room temperature, and 8 ml of aqueous NaOH were added dropwise over 8 min. The mixture was stirred for 60 min to yield alkali pulp. The desired amount of monochloroacetic acid (MCA) was dissolved in 15 ml of isopropanol and added to the mixture, and then temperature was raised to 60 8C and kept for 1 h under stirring. When the reaction time was elapsed, the mixture was ltered, suspended in 150 ml aqueous methanol (80% v/v) and neutralised with acetic acid (90% v/v). The nal product was washed three times with 150 ml

where Imin is the minimum intensity (at 2qz188) and Imax is the maximum intensity (at 2qz228). Intrinsic viscosity of CMC was measured 25 8C after dissolution in 0.1 M NaCl to prevent the polyelectrolyte expansion from triplicate data obtained for at least four CMC concentrations. Samples of dry CMC were dissolved in water at different concentrations and placed in a viscometer kept at 25 8C. The purity of the different CMC samples was determined according to a standard method (ASTM D1439-72). 2.5. Rheological measurements Rheological properties were determined using a controlled stress rheometer (Bohlin CVO from Bohlin Instruments Inc., Cranbury, NJ) (Collyer & Clegg, 1998) with cone-plate geometry CP 4/40 (diameter 4 cm with an angle of 48 and gap 0.150 mm). In order to assure thermal equilibrium before testing, solutions were equilibrated at 25 8C for 5 min after the measuring system reached its testing position. Calibration of the apparatus was performed using a Newtonian oil (2.5 Pa s at 25 8C) from Infra Scientic. In stress-controlled ow viscometry, viscosities varied less than G5%. Temperature calibration of the lower plate was carried out at 25G0.1 8C by computer-controlled uid circulation. All the reported data are the averages of at least three replicate measurements. The linear viscoelastic range was determined by performing 25 steps in a continuous operation at a constant frequency of 0.01 Hz and 25 8C for all samples. Stress sweep tests were carried out from 0.1 Pa to a given maximum stress (smax) within the linear viscoelastic range. Thixotropy tests were carried out by gradual modication of the shear stress within the linear viscoelastic region from a given initial value to an upper limit, and then from the upper limit to the initial value. The value of shear

S. Dapa et al. / Food Hydrocolloids 19 (2005) 313320

315

Table 1 Samples employed in this study, experimental conditions used in the synthesis of the various CMC samples (dened by CNaOH and CMCA), and results achieved for the total average DS and partial DS (glucose, mono-, di- and tri-substituted mole fractions obtained by HPLC and x2, x3, and x6 obtained by 1H-NMR) Sample 1 2 3 4 5 6 7
a

CNaOH 4.8 5.4 4.2 4.2 3.2 3.0 5.4

CMCA 2.0 2.0 2.0 2.5 2.0 1.0 3.0

Glua 0.18 (0.1756) 0.21 (0.1918) 0.26 (0.2196) 0.29 (0.2746) 0.35 (0.3098) 0.36 (0.3285) 0.38 (0.3579)

Monoa 0.41 (0.4139) 0.40 (0.4223) 0.39 (0.4332) 0.43 (0.4436) 0.40 (0.4441) 0.40 (0.4428) 0.40 (0.4386)

Dia 0.29 (0.3252) 0.28 (0.3100) 0.24 (0.2848) 0.23 (0.2389) 0.20 (0.2122) 0.19 (0.1989) 0.18 (0.1791)

Tria 0.11 (0.0851) 0.10 (0.0759) 0.10 (0.0624) 0.05 (0.0429) 0.06 (0.0338) 0.05 (0.0298) 0.04 (0.0244)

DSHPLC 1.32 1.27 1.19 1.05 0.97 0.93 0.87

x2 0.60 0.58 0.51 0.49 0.42 0.44 0.34

x3 0.29 0.31 0.30 0.25 0.15 0.13 0.15

x6 0.43 0.37 0.42 0.35 0.41 0.37 0.32

DSHRMN 1.34 1.26 1.23 1.09 0.98 0.94 0.84

In brackets: results calculated from the Spurlins model.

stress and shear rate corresponding to each curve (up-ow and down-ow curves) were obtained by tting the experimental results to the Ostwald-de-Waele power-law equation: _ s Z k$gn (2)

_ where s (Pa) is shear stress, g (sK1) is shear rate, k (Pa sn) is consistency, and n is the ow behaviour index. Creep and recovery tests were carried out by applying a sudden stress (s0, within linear viscoelastic region), which was maintained for 10 s, then nally removed, allowing the sample to relax for 300 s. All measurements were made at 25 8C. The results were interpreted in terms of the shear compliance function: Jt Z gt=s0 (3)

In mechanical spectrometry tests, solutions were subjected to harmonical stresses at variable frequencies. The target strain was xed to 1 and oscillatory frequency sweeps were run from 0.9 to 0.005 Hz (25 steps per cycle) at 25 8C. Complex viscosity (m*) and complex shear modulus (G*) were determined from the results obtained for storage modulus (G 0 ) and loss modulus (G 00 ). Loss modulus was adjusted to a power law equation: G 00 Z Go 00 $nn
00
00

(4)

where Go is the viscous modulus at frequency 1 Hz, n is the frequency in Hz, and n 00 is the exponent of the power law equation. The data obtained for each frequency were checked to conrm that the resulting s was within the linear viscoelastic region.

of unsubstituted (glucose) and partially substituted units (mono-, di-, and tri-substituted mole fractions). The respective mole fractions were compared to theoretical values calculated by a reported statistical method (Spurlin, 1939). The determination of the chemical structure of the CMC samples by 1H-NMR spectroscopy provides again information on the total average DS (DSHNMR), but also on the partial degrees of substitution at positions C2 (denoted x2), C3 (denoted x3), and C6 (denoted x6). Table 1 lists the NaOH/cellulose mole ratios (variable denoted CNaOH) and the MCA/cellulose mole ratios (variable denoted CMCA) employed in the synthesis of the CMCs considered in this work, as well as experimental results concerning the total and partial degrees of substitution determined by the methods cited above. The same Table includes the theoretical results calculated from the Spurlins model. The standard deviations found for DS determined by HPLC and 1H-NMR methods (0.0070.042) were in the range reported in literature (Heinze & Pfeiffer, 1999; Kauper et al., 1998). On the other hand, the experimental results for the relative amounts of glucose, mono-, di-, and tri-substituted mole fractions obtained by HPLC were in good agreement with the results predicted by the statistical model. Table 2 presents the purity degrees and the viscosityaverage molecular weights (denoted Mw) calculated from the intrinsic viscosity [h] determined in the Newtonian regime using the Mark-HouwinkSakurada equation: hml=g Z 1:65 !10K2 $Mw (5)

3. Results and discussion 3.1. Chemical structure of CMC samples The chemical structure of the CMC samples was studied after removing the water insoluble fraction of CMC and subsequent freeze drying of the soluble fraction. HPLC analysis provided information on the total average DS (DSHPLC), as well as on the relative amounts

Table 2 Purity degree and viscosity-average molecular weight (Mw) determined in the Newtonian regime using the Mark-HouwinkSakurada equation Sample 1 2 3 4 5 6 7 Purity degree (%) 99.7 99.5 99.3 99.5 99.8 99.7 99.2 Mw (g/mol) 1.63!105 6.21!104 2.51!105 2.53!105 2.35!105 2.60!105 2.59!105

316

S. Dapa et al. / Food Hydrocolloids 19 (2005) 313320

The experimental results showed that higher DS resulted in decreased Mw. The purity of CMC samples was higher than 99.5% (the purity required for the pure grade CMCs applications such as cosmetics, food and pharmaceuticals, see Greenway, 1994), except for samples 3 and 7 (99.3 and 99.2%, respectively). Consequently, all the samples were suitable for rened grade CMC applications (paints, adhesives, ceramics, textiles, etc.), for which purity O97% is required. The DS and the uniformity of substitution are key factors affecting CMC solubility. CMC samples with DS below 0.3 are insoluble in water but soluble in alkali (Feddersen and Thorp, 1993), whereas solubility increases with DS (Keller, 1997). Fig. 1 shows the DSHPLC of the carboxymethylated samples (before dissolution) as a function of the insoluble CMC fraction (expressed as weight percent referred to the total CMC). The insoluble CMC fraction found when carboxymethylation was carried out under the mildest operational conditions (43.4%) was remarkably higher than the ones determined for the sample with the highest DSHPLC (6.4%) insoluble CMC. In order to get further insight on the structure of the bleached pulp and carboxymethylation products, Fig. 2a shows the X-ray diffractograms obtained for the initial cellulose pulp, for the aqueous insoluble CMC fraction of the sample with DSHPLCZ0.87, and for samples derivatised under the conditions of experiments 7 and 1 (leading to the minimum and maximum DSHPLC, 1.32 and 0.87, respectively). Additional X-ray diffractograms were also recorded for alkali-celluloses manufactured with 30, 40 and 50% NaOH (see Fig. 2b), which did not show signicant differences. The diffractogram of the bleached Milox pulp showed a characteristic prole of cell-I type, with the principal peak at a Bragg angle of 22.58 (002 plane) corresponding to the cellulose crystalline structure and secondary peaks at 158 (101 planes) and at 34.58. The CI determined for the Milox pulp by Eq. (1) (77.4%) was in the range reported for other celluloses (Awadel-Karim, Nazhad, & Paszner, 1999; El Seoud, Marson, Ciacco, & Frollini, 2000; Zhan, Cheng, Dongli, Guigeng, & Xiahong, 1993).

Fig. 2. Diffractograms of: (a) Cellulose pulp; (b) Insoluble fraction of sample 7 in Table 1; (c) Soluble fraction of sample 7 in Table 1; (d) Soluble fraction of sample 1 in Table 1. (b) Diffractograms of alkali-celluloses obtained with various NaOH concentrations.

A comparison between the CMC diffractograms of samples with the minimum and maximum DSHPLC showed lower peak intensity at the Bragg angle of 22.58 for the second case, corresponding to a reduction in the crystallinity degree. Higher crystallinity was inferred from the diffractogram for the water insoluble fraction of sample 7, which was formed by glucose and a small portion of monosubstituted fraction (checked by HPLC analysis). 3.2. Study of the inuence of solution concentration in the rheological properties The main factor responsible for the pseudoplastic behaviour is the lack of uniformity in the DS (Feddersen and Thorp, 1993; Vais et al., 2002), which lead to carboxymethylated samples having residual crystalline

Fig. 1. Interrelationship between the degrees of substitution (DSHPLC) and content of insoluble matter for samples 17 of Table 1.

S. Dapa et al. / Food Hydrocolloids 19 (2005) 313320

317

zones causing cross-linking and entrapping soluble CMC molecules by electrostatic hydrogen bonding and/or van der Waals forces (DeButts et al., 1957). When systems are subjected to shear, a fraction of cross-linking centers and entrapped molecules are dispersed. When increasing shear rates are applied, chain type CMC molecules are disentangled, stretched, and reoriented parallel to the direction of the shearing force. Molecular aligning allows molecules to slip past each other more easily and reduce the viscosity of the CMC solution (El Ghzaoui, Trompette, Cassanas, Bardet, & Fabregue, 2001). Stress viscosimetry has been employed for ow characterization. In this part of the work, the rheological behaviours of samples 1 and 7 in Table 1 were characterised at different solution concentrations (from 1.4 to 3.6% and from 1.4 to 2.5%, respectively). For this purpose, the stress range of the linear viscoelastic region (smax) was rst determined at 25 8C for constant frequency. smax increased with concentration from 17 to 150 Pa at DSHPLCZ1.32, and from 23 to 160 Pa at DSHPLCZ0.87 (Table 3). The same Table lists the results obtained for consistency (k) and ow behaviour index (n) from the Ostwaldde-Waele equation, as well as the respective regression coefcients for both upand down- ow curves. The regression coefcients obtained for all the cases were near 1, conrming the close interpretation of the experimental data provided by equations. When the CMC concentrations of both samples were increased from the lowest to the highest values considered, the up-ow behaviour index decreased from 0.9596 to 0.852 for sample 1 and from 0.939 to 0.800 for sample 7. This corresponds to higher shear thinning behaviour for increased concentrations, a tendency more marked in the case of sample with the minimum DSHPLC
Table 3 Linear viscoelastic range (smax) and parameters of the Ostwald-de-Waele equation and regression coefcients determined for samples 1 and 7 at 25 8C CMC (%) smax (Pa) k (Pa sn) n r2

Fig. 3. Dependence of shear stress on shear rate in experiments carried out at 25 8C.

Sample 1 (DSHPLCZ1.32) 1.4 17 0.0967G0.0003a 0.153G0.003b 2.5 60 0.501G0.012a 0.79G0.02b 3.0 96 0.86G0.02a 1.13G0.04b 3.6 150 1.59G0.03a 2.69G0.08b Sample 7 (DSHPLCZ0.87) 1.4 23 0.1509G0.0011a 0.249G0.005b 2.0 80 0.849G0.017a 2.5 160 1.80G0.06a 2.00G0.07b
a b

0.9596G0.0015a 0.884G0.011b 0.866G0.013a 0.781G0.013b 0.861G0.013a 0.818G0.019b 0.852G0.012a 0.75G0.019b 0.939G0.004a 0.848G0.010b 0.839G0.010a 0.800G0.017a 0.784G0.018b

1 0.9983 0.9981 0.9979 0.9978 0.9951 0.9981 0.9943 0.9999 0.9989 0.9986 0.9960 0.9953

Upow curve. Downow curve.

(sample 7). Figs. 3a and b show the shear stress-shear rate up-ow curves for different concentrations of both samples in a loglog scale. For a given concentration, the sample with lower DSHPLC presented a higher pseudoplastic behaviour, as it can be conrmed by comparing the ranges determined for n from both up- and down- ow curves (see Table 3). Ghannam and Esmail (1997), working with CMC solutions with DSZ0.7, found k in the range 0.052.30 Pa sn and in the range 0.950.73 Pa sn for 13% solutions. Related results have been reported by Vais et al. (2002), who found kZ17.09 Pa sn and nZ0.41 at 25 8C for 1.5% aqueous CMCs with DS between 0.7 and 0.85. Pseudoplastic behaviour (k in the range 0.618.95 and n in the range 0.7450.506) was also found for 2% aqueous solutions of CMC with DS between 0.74 and 0.89 made from pine, poplar or wheat straw pulped in fast soda/AQ media and bleached with sodium chlorite in acidic media (Barba, Reguant, & Farriol, 2000). In order to compare the apparent viscosities obtained for solutions with various CMC concentrations, the viscosities (denoted m, and calculated as the ratios between shear stress and shear rate, Eq. (2)) were determined at a shear rate of 10 sK1 (see Table 4). For the sample with DSHPLCZ1.32,

318

S. Dapa et al. / Food Hydrocolloids 19 (2005) 313320

Table 4 Apparent viscosity (m) and complex viscosity (m*) determined for samples 1 and 7 at 25 8C CMC (%) Sample 1 (DSHPLCZ1.32) 1.4 2.5 3.0 3.6 Sample 7 (DSHPLCZ0.87) 1.4 2.0 2.5
a b c

ma (mPa s) 88b 117c 368b 479c 627b 742c 1131b 1517c 131b 175c 585b 1136b 1209c
K1

m* (mPa s)

90G2 340G8 530G9 900G70

117G2 480G7 930G2

Calculated at shear rate 10.0 s Upow curve. Downow curve.

m increased with concentration from 88 to 1131 mPa s in the up-ow curves and from 117 to 1517 in the down-ow curves when CMC concentration increased from 1.4 to 3.6%, whereas for the sample with DSHPLCZ0.87, viscosity increased from 131 to 1136 mPa s in the up-ow curves and from 175 to 1209 mPa s in the down-ow curves when CMC concentration increased from 1.4 to 2.5%. Heinze and Pfeiffer (1999), operating at a shear rate of 10 sK1, determined viscosities in the range 211193 mPa s for 1.5% aqueous solutions of CMCs having DS in the range 0.701.21. According to literature (Keller, 1997), our CMC samples can be classied as medium-viscosity type. The viscoelastic properties of CMC solutions are key factors controlling their industrial applications. To obtain information in this point, stress sweep tests were carried out for samples 1 and 7 at different concentrations in order to nd the stress range leading to linear viscoelastic behaviour (see Fig. 4a and b). At high CMC concentrations, the tendency of NaC to move out the inuence charge zone situated over the CMC polymeric molecules is low, but when dilution increases, ions move out the region between polymers, thus increasing the net charge of molecules (Feddersen and Thorp, 1993). The experimental results determined in this work for the complex viscosity (m*) and the complex shear modulus (G*) of samples with DSHPLC 1.32 and 0.87 are listed in Table 4. When concentration increased from 1.4 to 2.5% for both samples, the higher density of intermolecular contacts resulted in increased values of m* (from 90 to 340 mPa s for the sample with the higher DS and from 117 to 930 mPa s in case the sample with the lower DS in up-ow curves). The higher rigidity of solutions prepared for the sample with the lower DS are caused by the lower electric charge density of molecular chains, which lead to a higher sample structuration, and so

Fig. 4. Linear viscoelastic range for experiments carried out at 25 8C.

to increased strain resistance and higher smax (Fig. 4a and b and Table 4). 3.3. Study of the inuence of DS in the rheological properties at the same solution concentration The inuence of DS on the rheological properties of CMC solutions was studied for a constant solution concentration (1.4%) using all the samples listed in Table 1 (with DSHPLC from 0.87 to 1.32). In order to describe the ow behavior of CMC solutions, a steady-state ow study (including up- and down- ow curves) was performed. As before, the stress range of the linear viscoelastic region (smax) was rst determined at 25 8C for a constant frequency. smax decreased with DSHPLC (from 23.0 Pa at DSHPLCZ0.87) to reach a minimum value (between 13.0 13.5 Pa at DSHPLC between 0.97 and 1.19), and then increased slightly (see Table 5). The values determined for the Ostwald-de-Waele parameters and corresponding regression coefcients are listed in Table 6. For the up-ow curves, when DSHPLC increased, k rst decreased to reach a minimum value (0.1509 for sample 7 to 0.06851 Pa sn for sample 3) and then increased to maximum value (0.06581 for sample 3 and 0.0967 for sample 1). Oppositely, n rst increased to maximum value (0.939 for

S. Dapa et al. / Food Hydrocolloids 19 (2005) 313320 Table 5 Linear viscoelastic range (smax) and complex viscosity (m*) determined for samples 17 at 25 8C Sample 1 2 3 4 5 6 7 smax (Pa) 17.0 15.0 13.5 13.0 13.3 19.0 23.0 m* (mPa s) 90G2 74G1 65G3 69G2 60G1 118G1 117G4

319

sample 7 and 0.9785 for samples 2), and then decreased to 0.9596 for sample 1 (see Table 6). A related behaviour was again observed (but with even more marked relative variations) when the down- ow curve data were considered (see Table 6). It can be pointed out that the sample presenting the lowest DSHPLC (0.87) and one of the highest Mw (2.59!105 g molK1) exhibited the most marked pseudoplastic response (see Table 6). A related variation pattern was determined experimentally for m* (complex viscosity) for each sample along the whole linear viscoelastic range (see Table 5). The results from dynamic oscillatory and stress viscometry tests exhibited a related dependence on DS (see Tables 5 and 6). A creep-recovery test was carried out to assess the viscoelastic behaviour of CMC solutions. Samples were tested for their time response to a constant stress within the linear viscoelastic region. Fig. 5 shows the shear compliance J(t) obtained for all the samples as a function of time. The abrupt and constant increase in J(t) for t!10 s corresponds to the viscose response during creep, whereas the constant values in J(t) for tO10 s (after stress is removed) conrmed the purely viscous character of solutions. When DS increased, J(t) in the recovery phase raised from 93 PaK1 for sample 7 (with the lowest DSHPLC) to about
Table 6 Parameters of the Ostwald-de-Waele equation and regression coefcients for samples 17 operating at 25 8C Sample 1 2 3 4 5 6 7
a b

Fig. 5. Creep and recovery curves for samples 17 of Table 1 (assays performed at 25 8C).

k (Pa sn) 0.0967G0.0003 0.153G0.003b 0.07365G0.00009a 0.109G0.002b 0.06851G0.00010a 0.1059G0.0017b 0.06860G0.00007a 0.1045G0.0016b 0.0591G0.0002a 0.0758G0.0013b 0.1269G0.0012a 0.233G0.002b 0.1509G0.0011a 0.249G0.005b
a

n 0.9596G0.0015 0.884G0.011b 0.9785G0.0006a 0.916G0.010b 0.9763G0.0007a 0.904G0.008b 0.9785G0.0005a 0.907G0.008b 0.9676G0.0017a 0.926G0.008b 0.932G0.005a 0.83G0.05b 0.939G0.004a 0.848G0.010b
a

r2 1 0.9983 1 0.9987 1 0.9993 1 0.9991 1 0.9994 0.9998 0.9755 0.9998 0.9990

180 PaK1 for samples with DSHPLC in the range 0.971.19. A slight decrease of J(t) up to 137 PaK1 was observed for the sample 1 (with the highest DS), showing that the solutions with higher pseudoplastic response (samples 6 and 7) presented less deformability. The decreased concentration of polar groups in solution resulted in lower interactions, causing increased resistance towards deformation (minimal J(t) values). This fact is in agreement with the Mw data, because both samples exhibited the highest molecular weights (2.60!105 and 2.59!105 g molK1, respectively), and with the ndings from both the steady state ow studies and oscillatory tests at a constant frequency. In order to assess the time dependence of the viscoelastic moduli, mechanical spectra measurements were carried out at 0.80.005 Hz frequency sweep keeping the stress within the linear viscoelastic region. As expected, the samples did not present elasticity at 1.4% concentration. The storage modulus G 0 was remarkably smaller than G 00 in the frequency range tested. For example, at 0.1 Hz, G 0 were in the range 1.8!10K34!10K3 Pa, while G 00 were in the range 4.2!10K2K8.7!10K2 Pa. Table 7 shows the values of the power law parameters (Go 00 and n 00 ), as well as the regression coefcients. The values of n 00 were close to 1, conrming the predominantly viscous character of the solutions, whereas Go 00 showed a similar dependence on DS that the one discussed before (see Table 7 and Fig. 5). Samples 6 and 7 (with the lowest DS) showed the higher resistance towards viscous deformation (higher Go 00 values),
Table 7 Results determined for the par ameters of equation 4 for samples 17 from experiments carried out at 25 8C Sample 1 2 3 4 5 6 7 Go 00 (Pa) 0.5552G0.0011 0.4613G0.0008 0.4234G0.0011 0.439G0.007 0.4094G0.0012 0.6797G0.0015 0.818G0.003 n 00 0.9894G0.0014 0.9927G0.0012 0.9896G0.0017 1.0084G0.0011 0.993G0.002 0.9838G0.0015 0.971G0.002 r2 1 0.9999 0.9999 0.9999 0.9999 0.9999 0.9999

Upow curve. Downow curve.

320

S. Dapa et al. / Food Hydrocolloids 19 (2005) 313320 Barba, C., Reguant, J., Farriol, X., & Montane, D. (2000). Carboxymethyl cellulose from waste lignocellulosic pulps produced by a fast soda/AQ process. Journal of Wood Chemistry and Technology, 20(2), 185204. Buschler-Diller, G., & Zeronian, S. H. (1992). Enhancing the reactivity and strength of cotton bers. Journal of Applied Polymer Science, 45(6), 967979. Collyer, A. A., & Clegg, D. W. (1998). Rheological measurement (2nd ed). London: Chapman & Hall. Dapa, S., Santos, V., & Parajo, J. C. (2000). Formic acid-peroxyformic acid pulping of Fagus sylvatica: An engineering assesment. Journal of Wood Chemistry and Technology, 20(4), 395413. Dapa, S., Sixta, H., Borgards, A., Harms, H., & Parajo, J. C. (2002). Production of acetate-grade pulps by organic acid pulping and TCF bleaching. TAPPI Journal, 1(2), 2126. Dapa, S., Sixta, H., Borgards, A., Harms, H., & Parajo, J. C. (2003). TCF bleaching of hardwood pulps obtained in organic acid media: Production of viscose-grade pulps. Holz als Roh und Werkstoff, 61(5), 363368. Dapa, S., Santos, V., & Parajo, J. C. (2003). Carboxymethylcellulose from totally chlorine free bleached Milox pulps. Bioresourse Technology, 89, 289293. DeButts, E. H., Hudy, J. A., & Elliott, J. H. (1957). Rheology of sodium carboxymethylcellulose solutions. Industrial and Engineering Chemistry, 49(1), 9498. El Ghzaoui, A., Trompette, J. L., Cassanas, G., Bardet, L., & Fabregue, E. (2001). Comparative rheological behavior of some cellulosic ether derivatives. Langmuir, 17, 14531456. El Seoud, O. A., Marson, G. A., Ciacco, G. T., & Frollini, E. (2000). An efcient, one-pot acylation of cellulose under homogeneous reaction conditions. Macromolecular Chemical Physics, 201(8), 882889. Ghannam, M. T., & Esmail, M. N. (1997). Rheological properties of carboxymethylcellulose. Journal of Applied Polymer Science, 64(2), 289301. Greenway, T. M. (1994). Water-soluble cellulose derivatives and their commercial use. In R. D. Gilbert (Ed.), Cellulosic polymers. Blends and composites (pp. 173188). Hanser Gardner Publications. Heinze, T., & Pfeiffer, K. (1999). Studies on the synthesis and characterization of carboxymethylcellulose. Angewandte Makromoleculare Chemie, 266, 3745. Heinze, T., & Liebert, T. (2001). Unconventional methods in cellulose functionalization. Progress in Polymer Science, 26, 16891762. Ho, F. F. L., & Klosiewicz, D. W. (1980). Proton nuclear magnetic resonance spectrometry for determination of substituents and their distribution in carboxymethylcellulose. Analytical Chemistry, 52(6), 916920. Kauper, P., Werner-Michael, K., Horner, S., Saake, B., Puls, J., Kunze, J., Fink, H. P., Heinze, U., Heinze, T., Klohr, E. A., Thielking, H., & Koch, W. (1998). Development and evaluation of methods for determining the pattern of functionalization in sodium carboxymethylcellulose. Angew Makromolecular Chemistry, 260, 5363. Keller, J. D. (1997). Sodium Carboxymethylcellulose (CMC). In A. Imeson , Thickening and gelling agents for food (Vol. III). Kulicke, W. M., Kull, A. H., Kull, W., & Thielking, H. (1996). Characterization of aqueous carboxymethylcellulose solutions in terms of their molecular structure and its inuence on rheological behaviour. Polymer, 37, 27232731. Spurlin, H. M. (1939). Arrangement of substituents in cellulose derivatives. Journal of American Chemical Society, 61, 22222227. Vais, A. E., Palazoglu, T. K., Sandeep, K. P., & Daubert, C. R. (2002). Rheological characterisation of carboxymethylcellulose solutions under aseptic processing conditions. Journal of Food Processing Engineering, 25, 4161. Zhan, J., Cheng, F., Dongli, L., Guigeng, L., & Xiaohong, Z. (1993). Solvent effect on carboxymethylation of cellulose. Journal of Applied Polymer Science, 49, 741746.

conrming the higher entanglements between CMC chains in agreement with the results obtained in the transient and stress viscometry tests.

4. Conclusions CMCs were synthesized in heterogeneous medium from Milox-delignied, TCF-bleached Eucalyptus globulus pulps. This type of CMC, obtained by reaction in media made up of 2-propanol and mixtures of NaOH:MCA:cellulose, can be classied as medium-viscosity type. The purity of CMC samples was higher than 99.2% which means that these samples are suitable for rened grade applications (paints, ceramics, textiles, adhesives,.) and in some cases for pure grade applications such as cosmetics, foods and pharmaceuticals. In order to assess the rheological behaviour of the synthesized materials, their properties (rheological behaviour, solubility and molar mass) were determined and examined in relation to their total DS. The different tests applied for this purpose (dynamic and transient tests) were coherent in their results. Higher shear thinning was observed for increasing concentrations, this effect being more marked for the sample with the lowest DS. For a given concentration, samples with lower DS showed more resistance to both ow and viscous deformation and lower shear compliance resulting from a higher structuration degree.

Acknowledgements Authors are grateful to the Ministry of Education of Spain for the nancial support of this work (project Desarrollo de procesos con bajo impacto ambiental para la produccion de pastas de celulosa de alta calidad, ref. QUI99-0346). The authors wish to thank to C.A.C.T.I. (Centro de Apoio Cientco e Tecnoloxico a Investigacion) Institute (University of Vigo, Spain), as well as to Mr Antonio Casares and Ms Rosa Padn for their excellent technical assistance.

References
Abad, S., Saake, B., Puls, J., & Parajo, J. C. (2002). Totally chlorine free bleaching of Eucalyptus globulus dissolving pulps delignied with peroxyformic acid and formic acid. Holzforschung, 56(1), 6066. Abad, S., Santos, V., & Parajo, J. C. (2001). Evaluation of Eucalyptus globulus wood processing in media made up of formic acid, water, and hydrogen peroxide for dissolving pulp production. Industrial and Engineering Chemistry Research, 40(1), 413419. Abdelrahim, K. A., & Ramaswamy, H. S. (1995). High temperature/pressure rheology of carboxymethyl cellulose (CMC). Food Research International, 28(3), 285290. Awadel-Karim, S., Nazhad, M. M., & Paszner, L. (1999). Factors affecting crystalline structure of cellulose during solvent purication treatment. Holzforschung, 53, 18.

También podría gustarte