. III u .l‘ {-1\\£95} . ...N.Iff..fi. tel...F..I‘i.(i«!.f . VEPIIVLKAI sc‘o...) 1...... .. z}; ‘ré‘JflI‘? . o ‘ 1d ‘ . .1‘U .fiu .Y. . ' . \ . ‘. F; L93 . . .a’ \l‘ 1". ‘th‘u‘d" . 7 ' ‘ . . - Cafe... 1 .. t" . ’H . uS.‘ I). ._.!l.ya§. .. ‘.\$~ . Ilh. . $1,. ,3 .3 . . .1?“- ‘1 O. 2h.“ . . . huh “Omkwrlfl: “13:14.1... Earn. ($54.3... \. 1.... n . .‘. AI . ( n lvymb.‘ .21. aka; L 40‘ ..... o! :f‘; I . .1!!! _ . 11’). u '13., ‘21:th .LLI§,.IF\t-,f.. .0!‘.L.«1 .n\| u‘.-.!.vf..: (E 13:... $ 5. 1.304...‘ ~|1\\ ’.¢.. .1 . ,_ 1‘ '1 T;.,. 5. . $2.! .1 . 1.ij I; .Iolvl‘vgr. 3:9: p. 3. :5lnlu rllu§réuffrtca 10 (3-. u€o)vH&.. )tlt? a r . 10p . 12¢- icl I. .1- infam‘flu’ I. I» It). It!» . .fil... LN! ..... at" $3152)... .1... . Srrt:1)if‘uf.vu I) if 5'“! 3.3;... t!!! II. .E . :1. LI 3...... . In.” gargb .luflvltflrl. viii-.13.! I. 50.. .1 gigs? . in. the: l :9. [rig ‘3?" Pill i] .3" It! I (I .1511..- I fail i"l‘l""‘~:a) a: ll: . Igjflii§¢g. .. Onfl' In . . u v lei .tll Cs!!! . . :i D .. . «I! . .eHiFthrrpfl ”“43; 1 I’ll . 09b}!!- !In In... Vifll A? Ojratbrlfio ‘x‘lltnvpu‘lzl {Ill ‘ v .J.n .ylnx. iiil‘r‘ . 'I'-" 919,093 5W3 ~ w ““1 miiiifiim 1 11111311 111111 flmiversity This is to certify that the dissertation entitled Evaluating the Rheological Properties of Power-Law Fluids Using Mixer Viscometry presented by Maria Elena Castell Perez has been accepted towards fulfillment of the requirements for PhD. Agricultural Engineering degree in fw .F f/éAL Major professor Date 2/3] (70 MS U i: an Affirmative Action/Equal Opportunity Institution 0-12771 _—_—————__ PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE MSU Is An Alfirmetlve ActionlEqual Opportunity Institution EVALUATING THE WWICAL PROPERTIES OF POWER-LAP FIDIDS USING MIXER VISCOHETRY By Maria Elena Castell—Perez A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPY IN Agricultural Engineering Department of Agricultural Engineering 1990 . I :. :wn.r d‘- "‘56‘c lei... ‘ ...r.e :‘o g. \: qr um um MUSICAL PROPERTIES GP in (a .. ,4 roan-um FLUIDS usnic mum ' t '- .‘L. prod-1,1.- , . _|.V “ANA“: 1 ,7. “Zinc...“ Maria Elena Castell Perez . v ' . ' psi-so) 3:. BY . ,Jlil v ‘Lf‘ntl 1L1" . fl } :‘V.‘72‘»3-‘ ‘ . I ".V' . é 1"}!éilmprocedure was developed to determine the flow properties of ' ..:"?‘7 ‘ mayor-lawfluids using mixer viecometry data. The procedure is I?” the. mixing system: (1) Model 1, representing the analogy attic cylinders systems with negligible end effects; (2) resenting the same analogy with the addition of a term which #10 and effects. In the case of a flag impeller, an alter- .. rm evaluated, i_.e. , Model 3 representing a cylinder of 11th two blades attached. fr viscometer (Haake Rotovisko) over the same shear #40 l/s) with standard power-law fluids (aqueous CMC 2"? ——ve ‘n. _ .... r:.. ‘ > v. y x —-. .,1 o ‘1- ., i u. . P‘- ..v . I .11: Il' solutions). The best approximation of the average shear stress of stan- dard power—law fluids was obtained with the Model 2 (concentric cylinders analogy with end effects) for the paddle impellers. Model 3 (d - de ) gave the best results for the flag impeller. The procedure was evaluated with an actual power-law food product, creamy salad dressing. In this case Model 2 also gave the best results for the paddle impellers. For the flag impeller, Model 3 yielded higher values of the average shear stress, with Model 1 (concentric cylinders analogy with negligible end effects) being the best model for approximation of the average shear stress of the tested food material. Established mixer viscometry methods, the viscosity matching and the slope method, were evaluated for determination of average shear rates when agitating time—independent non-Newtonian fluids. Results indicate that the use of a constant value of the mixer proportionality constant, k', is not valid for all ranges of fluid rheological properties, system geometry and operating conditions. The relationship between the average shear rate and the impeller rotational speed was found to be different for different rheological properties and system geometries. The effect of the fluid properties on the value of k' is not clearly understood and needs further investigation. Approved Major Professor Date Approved Department Chairperson Date fill ‘l09lv Inua‘tm’ a L 1 T mic and Mu 1 t :he (‘-_ 1‘ ‘ . 31‘) Pickle. Imml.“ ‘.' la: (Mode: 1‘11 . [7&1" . . m leech". 1H! Haws-(33.. ' '. 3:123 : u "I. and cup kam- nn 1! .4. 5‘11”)“ ~‘4h:ee-a.n.‘.. . v‘ 541' t“ .1" .g Wr1uui‘l 1:- ,' 1|;.' ‘ ' Mm: R1: 83231342.»)? . -.¢. w M1. Imil.t‘ H 25‘ -A.:\;7..-" ‘*V.I~Ollvtse...u- . -. Met Bunion-on; t-w— given- . .we lmllere at», fan-or”: ’ 'tiz..a...... .. . 7 - Met Relau‘mga‘e ‘... y........‘_.. , “m1. Imllexe in; 1'.wa ~ ‘ V‘O‘Jv‘ ;'~O“-e. mun...“ 1’ .111 '7 j. I" ween “served '3. . ”5,. (me-1r; ‘Ia,r.r..‘..a\.‘,3!,. a.” Meme to ; t: new 6.6". 1.1-1 ier 269 LIST OF FIGURES FIGURE 3.1 Typical Impellers Used In The Determination Of Rheological Properties With A Mixer (Castell-Perez and Steffe, 1989) ...................................... 3 . 2 Mixe :r: Viscometry Techniques ............................ I-‘lo‘ilr patterns in Agitated Vessels. A) Newtonian Fluid. B) Elastic Fluid (Rod-climbing) ........................ 3.4 8) Simplified Flow Model for Power Correlation; b) Relative Velocity Distribution Between Paddle and Liquid (Nagata, 1975) .......................................... 3.5 Fluid Motion in Agitated Vessels (Nagata, 1975) ......... 3.5 SChematic Diagram Showing The Relation Between Power Number and Reynolds Number for a Paddle Impeller (Nagata, 1975) .......................................... 4-1 M"€161 Systems. A) Paddle Impeller; B) Flag Impeller; C) Flag Impeller (Model 3 .............................. 5'1 Padtile and Flag Impellers (all dimensions are in cm.). . 5'2 :zn‘ple containers and cup base (all dimensions are in 5'3 nix1ng System ........................................... 5-1‘ schematic Diagram of Experimental System ................ 6'1 Power Number/Reynolds Number Relationship for Newtonian :I‘Slids Using the Paddle Impellers (Cup diameter - 5.5 'l ..................................................... 6'2 POVer Number/Reynolds Number Relationship for Newtonian F:leuids Using the Paddle Impellers (Cup diameter - 3.5 cll) ..................................................... 6'3 POver Number/Reynolds Number Relationship for Newtonian uids Using the Paddle Impellers (Cup diameter - 2.54 ch). .................................................... 6'4 I) Predicted Po Versus Observed Po ; b) Power Correlation For The Paddle Impellers ................................ 6.5 Q) Predicted Po Versus Observed Po ; b) Power Correlation xiii Page 10 51 54 61 62 81 95 97 100 104 110 111 112 117 6.12 6.18 6.19 6.20 6.21 6.22 For The Flag Impeller ................................... a) Shear Stress Versus Shear Rate; b) Apparent Viscosity Versus Shear Rate (Hydroxypropyl Methylcellulose 1%) . . .. Average Mixing Shear Rate as a Function of Rotational Speed of the Paddle Impellers. a) System 3; b) System 4. Average Mixing Shear Rate as a Function of Rotational Speed of the Paddle Impellers. a) System 5; b) System 0‘ Average Mixing Shear Rate as a Function of Rotational Speed of the Paddle Impellers. a) System 7; b) System 8. Average Mixing Shear Rate as a Function of Rotational Speed of the Flag Impeller. a) CMC 1.5%; b) CMC 1% ...... Average Mixing Shear Rate as a Function of Rotational Speed of the Flag Impeller. a) CMC 2%; b) System 3 ...... Average Mixing Shear Rate as a Function of Rotational Speed of the Paddle Impellers (d/D-O.327). a) CMC 2%; b) cuc 1% ............................................... Plot of Mixer Proportionality Constant as a Function of IIIlibeller Size for the Non-Newtonian Fluids (d/D-0.327) .. "titer Torque Versus Viscosity Times Speed for Calculating “1X01: Coefficient, k2 (Paddle Impellers) ................ PlTe noxwz "N m ouswam .llaefl DIIIII. in: 5.59.3. .- 3.53:3. _ 3...: 8.8m .- .5... 38: I... a gram :2: .1 t 855. .32. 3.8 e 8.. _ A a e .30: .1 «0 3.51.32 _A S... .. stux... .. 3.385.: are. $3: .1 8 .18. I Mz... asfllasnsob 2.2. .e .- Use-8n E garners... ensure. x .181:on 2.2:. .e t .325 $3; #38 e 3c A82. 0:3“ e .5... 38a: 898 e .131 as: .e .o .38..“ SS: .1 3 nee-B 39am e!» 5.2. :8 e :8; >0 concoaoo loser Amp::m acoucoaoocTocehv :8: 25 a .25.: 38: 8c 2.. I .6... 8.5-: 2...! 2...: as... .- eta , 11 the following sections. 3 2.1) Aaalxsis_af_Iima;Isdsneadeat_§shaxi2r 3.2. 1 .1) Eiscosity Matching Method Based on the earlier work of Magnusson (1952), who proposed a Procedure for calculation of the apparent viscosity of a non-Newtonian fluid in mixing vessels from the power curve for Newtonian fluids, the Work of Metzner and Otto (1957) provided engineers with a method to determine the apparent viscosity of non-Newtonian fluids using a mixing system. Their approach is based on the theory for power consumption in agitated vessels . The generally accepted empirical relationship between the Power number, Po' and the mixing Reynolds number, Re, in the laminar flow region (Re < 10) for a Newtonian fluid is —-A-— (3.1) Where A is a constant dependent on the geometry of the system. The dimensionless numbers in Eqn. (3.1) are defined as p0 - ——2;—; (3.2) p d N and 2 R __LM (3.3) e ’7 12 where: d - impeller diameter, m N - impeller rotational speed, rev/s p - fluid density, kg/m3 7, - Newtonian viscosity, Pa 5 and the power input, P, is related to the torque M exerted on the im- peller by P - 21rNM (3.4) Combining the equations, the following relationship is obtained, 5 3 ' 2 (3.5) pd N pd N 3 2 and P-Ad N ,, (3.6) Thus, power measurements can be used to characterize the vis- cosity of a Newtonian fluid in a mixing system by measuring the torque required to turn the shaft of the impeller at a certain rotational speed. The geometry of the system and fluid density must also be known. When mixing a non-Newtonian fluid, especially one obeying the I) (“oar-law model , a-m-‘yn (3.7) 13 where: a - shear stress, Pa ’1 - shear rate, 1/5 In - consistency coefficient, Pa 5n n - flow behavior index, dimensionless the viscosity increases from a minimum value closer to the impeller to a maximum value in regions far away from the impeller (Ulbrecht and Carreau, 1985). Metzner and Otto suggested that Eqn. (3.1) could be Valid for a non—Newtonian fluid if an apparent viscosity, 77a, evaluated at an average shear rate given by "y - k’N (3.8) were used. Equation (3.8) assumes that the average shear rate around the 1mPellet is proportional to the rotational speed of the impeller, N, being k' the impeller proportionality constant. Once the average shear rate, lav , has been calculated, an &Verage apparent viscosity can then be obtained (Nagata, 1975). In their ex1)‘"‘5'3l.mental procedure, Metzner and Otto utilized two identical sets of mixing equipment, one containing a Newtonian fluid and the other a non- Newt'~')l'lian fluid. Using the same impeller speed and varying the viscosity of the Newtonian fluid so that the power measured at each impeller is the Same, the apparent viscosities should be the same in both instru- ments. Thus, they were able to determine the apparent viscosity, "a , of t he non-Newtonian fluid by knowing the viscosity of the Newtonian under 1 dentical experimental conditions in the laminar region. at ‘r 14 The procedure for data collection and analysis was: 1. Measurement of the torque required to rotate the impeller at a fixed rotational speed (Newtonian and non-Newtonian fluids). 2. Determination of Power Number- Reynolds Number curve for Newtonian fluids. The value of the constant A is determined from the slope of this curve. 3. Determination of power input and Power number, Po’ [Eqn. (3.2)] at each impeller speed N using non-Newtonian measurements in the mixing system. 4. From the curve obtained in (2), read the corresponding Re and find the viscosity from the expression for the Reynolds number JN—d (3_9) 5‘ Let n - "a . When the apparent viscosity of the non-Newtonian fluid 18 obtained, the corresponding average shear rate can be determined from the viscometric curve (0 versus ’7 data, obtained with a conventional viscometer), based on the assumption that n - 0/7. 6‘ calculate the value of the mixer impeller proportionality constant k’ at a specific value of the rotational speed N using Eqn. (3.8) , 1e. k' - +/N Holland and Chapman (1966) outlined a more complete description of the technique developed by Metzner and Otto (1957): 1 ' w1th "a obtained from Eqn. (3.9), plot "a versus impeller speed N. ERperimentally determine the apparent viscosity of the non-Newtonian fluid in a conventional viscometer at various shear rates. Plot log "a versus log ’7 . L 15 ~ From the plot of (1), determine the value of "a at a specified N b 5- From the plot of (3), read the value of "1 which gives the same "a of (4). 6- Plot these values of "7 versus the values of impeller speed N. 7- Obtain the value of k’ by measuring the slope of the plot made in step (6). The "viscosity matching" method assumes the value of k’ a con- Stem; which depends only on the geometry of the impeller. Even though this assumption has been used by numerous investigators, Metzner and Otto recommended further analysis to determine the effect of the power- 13" parameters (m and n) in the values of the proportionality constant 1"- Wood and Goff (1973) applied the matching viscosities method to eStimate the average shear rate in a Brabender Viscograph. The values of iav for the impeller were obtained from a plot of shear rate versus the viscosity (or apparent viscosity) of the Newtonian and non-Newtonian fluids- The magnitude of the shear rate where the two viscosities inter- sect is the average shear rate, fiav (Wood and Goff, 1973). An alternative procedure to obtain the average shear rate is by e quating the two expressions for the viscosities. Hence, "lav - m #2,, (3.10) '“'l - -" (3.11) and av m 1 . n - 1 1,“,- [ n/m 1 (3.12) 31 This procedure was used by Rao and Cooley (1984) and Mackey et (1987) for determination of rheological properties of fluids 16 characterized by the power-law model. Mackey et al. (1987) utilized the matching technique to inves- tigate the effect of some parameters on the assumed constant value of k’ using a Brookfield Viscometer. From power requirements: "a _ ill:— - — (3.13) Ad N N where k2 — 27"", the mixer coefficient, is a constant dependent on d A the geometry of the system. To find k2, the torque required to turn the impeller agitating a Newtonian fluid of known viscosity is measured as a function of the rotational speed. Torque M is plotted versus 1N and the 31°Pe is equal to l/kz. Thus, the value of constant A from Eqn. (3.13) is determined. It is important to check that Re <10 (laminar flow assumption), The rheological properties of the investigated power-law fluids are det'erlllined with a rotational cylinder viscometer. Torque versus rotational speed data are also collected with the mixer viscometer. Using Metzner and Otto’s approach of matching viscosities, n - "a - macro”1 (3.14) Thus, 1 1‘: H “'1 k"+[um] (3.15) 17 where k' is a function of the geometry of the system, the rheological properties of the fluid and the operational conditions. The authors observed that the value of k' was significantly affected by the values of the flow behavior index and the rotational speed of the impeller. (At rotational speeds less than 1.05 rad/s (10 rpm), the value of k’ was a strong function of n and N). Extensive work should be done to quantify the interaction of these parameters on the proportionality constant and the critical limits for the impeller-to-cup diameter ratio (d/D) must be identified . 3-2-1-2) Linea; §heag Stress Method Bongenaar et a1. (1973) developed a technique for characterizing the rheological properties of mold suspensions using a mixer (turbine) viscometer_ The instrument consisted of a standard 6-blade Rushton turbine impeller connected to a rotational viscometer and the torque on the impeller was measured as a function of the rotational speed. Data analy513 considered that the shear rate was given by Metzner and Otto’s (1957) ESSumption ( iav - k'N ) and that the shear stress was directly proportional to the torque M, aav - k"M (3.16) where k" is a constant independent of the value of n. A constant value of k' (- 10) was selected based on the work done by Calderbank and M00- Young (1959) for turbine impellers. The rheological parameters of the Pour-law fluid were determined as follows: 1. The value of the flow behavior index of the unknown suspension, nx, 18 is obtained from the slope of the log-log plot of torque,M, versus rotational speed N. , the 2. To calculate the value of the consistency coefficient, mx properties,ny and my, of a calibration fluid are measured in a conventional rotational viscometer and the following expression is written n M a m x x J; _ x _ x N k’ M —a "Ir—n n (3.17) Y Y y N Y k' 37’ Thus, M n _x_ M m (3.18) m - y x M nx 3’ (mm In the case of n - (valid when In — n |< 0.1), Eqn. (3.18) Y X X y simplifies to _?5_ (3.19) with "x and My as torque values for corresponding fluids measured in the mixer system. The authors also worked with the "a’ defined as 19 a "a - [‘7‘]. . (3.20) where 70 is an arbitrary standard shear rate. Thus, for a power-law fluid m _ "a (3.21) .n-l 7 and “x _ "fix (3.22) M n Y ay The procedure was also applied to fluids obeying the Casson model, a - a - "c "1 (3.23) with the Casson viscosity, no , obtained from the slope of a linear plot of 00.5 versus 70.5 . The yield stress, 00 , is obtained from extrapola- tion of the M versus N plot, assuming that k' and k" are independent of the Casson rheological parameters. According to Charles (1978), it is not clear if the calculated viscosity is an intrinsic property of the fluid. However, analysis of data is based on well-proven and widely accepted empirical correlations and the method is useful for determining the rheological properties of viscous fermentation broths. The mixer Viscometry technique developed by Bongenaar et al. (1973) was applied by Roels et a1. (1974) on their investigation of 20 fermentation broths. They observed that the use of a mixer prevented phase separation and settling of the particles. Their derivation for data analysis is as follows: From power requirement theory in the laminar region, 2 3 P - A "aN d (3.24) with A - 64 for a turbine impeller (using Newtonian fluids). Since the power P is related to the torque M, g3 3 (3.25) M - 2" naNd From measurement of torque M as a function of rotational speed N, the value of the apparent viscosity "a can be obtained from Eqn. (3.25) - -—!ZE-— (3.26) 0 3 a 64 N d Following Metzner and Otto's suggestion that fiav- k’N, with k’ a constant and with aav - "a 7av , or taking advantage of Eqn. (3.25), a v. Zfiflk (3.27) 5 64d Equation (3.27) can also be written as u - ci a (3.28) where : m' Ci- an' , (3.29) 21 is viewed as an instrument constant. The measured torque as a function of N and the use of Eqns. (3.8) and (3.28) allows for determination of rheograms for the investigated fluid in the laminar flow region. Also, the apparent viscosity can be calculated from Eqn. (3.26). Different rheological models were analyzed and the authors concluded that the Casson model better described the rheology of a fermentation broth. This technique requires further testing and results to be com- pared with those obtained with conventional viscometry methods. It must be emphasized that the impeller proportionality constant, k', was as- sumed a constant for the particular turbine impeller. When the torque is measured as a function of N, then the constant k' can be found from - m 7 “'1 - m (k’N) “'1 (3.30) i-"a Kemblowski and Kristiansen (1986) adapted the method developed by Roels et al. (1974) to design a suitable impeller-cylinder system for continuous on-line measurements in fermentation technology. The analysis assumed a two-cylinder system. Thus, in a rotational viscometer, the shear stress a is given by a _ z a (3.31) where: z - a constant of the instrument, Pa/reading a — instrument reading 22 The torque is then determined from the measured shear stress from the following expression: (concentric-cylinders with narrow gap) _ M (3.32) Zillr m a where: 11 - inner cylinder length, m r - (r1 + r,)/2 and r - arithmetic mean of the radii of inner and outer cylinders, m r1 - inner cylinder radius, m r2 - outer cylinder radius, m Combining Eqns. (3.31) and (3.32), M - a a (3.33) where a - instrument constant, N m/reading = 2w11r; z (3.34) According to Eqn. (3.33), the value of the constant a, obtained from a concentric cylinder system, is used to evaluate the torque for the mixer system on the basis of the reading from the instrument, a. The average shear stress is then obtained from Eqn. (3.32) as a 1rk' (3-35) av or aav - Z a (3.36) 23 where Z - a constant for a given geometry of the impeller system, Pa/reading with I 'a (3.37) The procedure to determine the value of k' for the particular impeller is as follows: 1. For power-law fluids, the shear dependent viscosity is given by Eqn. (3.30) and 3L": - III1(1<'N)“'1 (3.38) ANd or (3.39) where B is found from experimental curves of torque versus rotational speed. 2. The value of the proportionality constant k' is now a function of the geometry of the impeller and rheological properties of the fluid, since T1 k' - [115,] (3.40) with A determined from Newtonian Po versus Re curves according to Eqn. (3.1). 24 Eqns. (3.8) and (3.36) allow determination of the curve flow for the investigated fluid. The technique proved to be sensitive and able to detect the variation of rheological parameters, especially the consis- tency coefficient, m, resulting from the differences in processs parameters. Modifications of the instrument are being made for on-line application. It is important to recognize that this method assumes a narrow gap between the rotating unit and the cup. Hence, care must be taken not to violate this assumption when using a system with geometri- cal characteristics other than the ones used in this study (six—blade turbine impeller, H/d - 2.7 and 1.8, d/D - 0.93). 3 .2 . l . 3) Slope Method Based on the initial work of Metzner and Otto (1957), Rieger and Novak (1973) developed a method to determine the value of the impeller proportionality constant for the agitation of highly viscous fluids characterized by the power-law model, in the laminar region of flow. It is called the "slope method" in this study. From dimensional analysis, the following relationship is obtained __2_ _ cm) (3.41) Using the power relationship given by Eqn. (3.1): P __°_(‘L (3.42) 0 en 2 2_ with R _d N “2 (3.43) 25 and C(n) - constant, f(n, geometry) For a Newtonian fluid, P _ __A__n_ (3.44) Replacing 11 with "a for a power-law fluid, P _, 3a Lk'N)“'1 _ M (3.45) 2 R o N d p en Comparing Eqn. (3.45) with Eqn. (3.41), l C(n) - A k'“‘ (3.46) or log C - log A - (l-n) log k' (3.47) with C given by Eqn. (3.41) The authors suggested that Eqn. (3.8) is valid for a particular impeller only if a plot of log [P/(mdan+1)] versus (l-n) is a straight line [Eqn. (3.47)]. The slope of this line is equal to -log k' . If the plot were nonlinear, Eqn. (3.8) would be invalid since k' would not exist. Results suggested that the linear shear rate assumption was useful for engineering calculations of power consumption with certain mixer impellers. However, certain dependence of the value of the con- stant k' with the flow behavior index, n, was observed. This procedure has been tested to determine the average shear rates of food 26 materials by Rao (1975). 3.2.1.4) Qanhinsd Slgpg and Linear Shear Stress Method Rao (1975) combined the procedure of Rieger and Novak (1973) for determination of the impeller proportionality constant with the proce- dure for determination of the rheological properties of power-law fluids developed by Bongenaar et a1. (1973). This technique will be called the "Combined Slope and Linear Shear Stress Method" in this review, and it has been utilized with food products by several investigators (Rao and Cooley, 1984; Steffe and Ford, 1985; Ford and Steffe, 1986; Castell- Perez et a1., 1987). The procedure is as follows: 1. Collect torque versus rotational speed data of several non-Newtonian (power-law) fluids and the investigated fluid with the mixer system. 2. Determine the rheological properties of fluids with a concentric cylinder viscometer. 3. Plot log [P/(mdan+1)] versus (l-n) (Rieger and Novak, 1973) using torque data obtained from mixer system. Find value of proportionality constant k' for that particular impeller at a specific value of rotational speed. 4. Find value of flow behavior index of investigated fluid nx, from slope of log torque versus log rotational speed data from concentric cylinders. 5. Determine value of consistency coefficient of the investigated fluid mx , from mixer torque data, fluid properties and k', Eqn. (3.18) (Bongenaar et a1., 1973). Rao and Cooley (1984) compared the mixer Viscometry techniques 27 developed by Metzner and Otto (1957) (Viscosity Matching) and by Rieger and Novak (1973) (Slope) for complex geometry impellers and observed that results obtained from both techniques were in good agreement. They also observed that the ”slope method“ could lead to large errors when finding the value of k' , since k' - 10"“pe (3.48) The advantage of this method is its relative simplicity as compared to the matching of viscosities required in Metzner and Otto's. However, the "Viscosity matching" method seems to yield more consistent values of k' for a particular type of impeller. 3.2.1.5) WW 3.2.1.5.1) Egesgure Difference Method Using a helical screw, Kraynik et a1. (1984) designed an instru- ment to measure the viscosity of concentrated suspensions of coal particles in organic solvents. This Viscometry technique differs from all the Others in the fact that it relates the pressure difference in the fluid to the viscosity of the fluid. The instrument is essentially a metering screw pump operating at zero discharge. The sample can be pressurized and the pressure difference is measured across two ports Spaced at different heights along the outer wall by a series of pressure transducers (Kraynik et a1., 1984). The authors emphasize the advantages of pressure measurements OVer t°rQue measurements in high-pressure rotational instruments where fitti “3 seals are required around a rotating shaft that could affect 28 force measurements. This viscometer has potential in high-pressure Viscometry and could be applied to determination of the pressure- dependence of viscosity. It is also suitable for chemically-reacting and heterogeneous fluids. Applications of this technique in on-line vis- cometry of food products has been investigated by Tamura et a1. (1988). 3.2.1.5.2) Dirge; Detemingtign of The Flow Curve Method Kemblowski et a1. (1988) suggested that the methods of correla- tion of Power number as a function of the Reynolds number using the Metzner and Otto assumption may be suitable for engineering purposes, 9‘11: not precise enough for rheological applications. The authors developed a method which enables a direct determination of the flow curve without the need for power data obtained with Newtonian fluids. The analysis assumes a system of two concentric cylinders to model the impeller system (helical screw impeller rotaing in a draught tube). The torque on the impeller shaft is the combination of the torque resulting from the shearing in the screw channel plus the torque result- ing from the shearing in the gap between the edge of the screw flight and the innner surface of the draught tube (Kemblowski et a1., 1988)- Thus M-Ml +14, (3.49) h w ere M1 ._ torque resulting from shearing in screw channel, N m M2 ‘ torque resulting from shearing in the gap, N m 29 The technique for determination of the flow curve of the inves- tigated fluid is the following: 1. Using a helical screw impeller rotating in a draught tube as the measuring system, determine the torque M on the impeller shaft as a function of rotational speed N. 2. A log-log plot of M versus N should give a straight line for power law fluids and the slope is equal to n. 3- Calculate the parameters which characterize the geometry of the measuring system : A1 - surface of the screw channel, 111 2 A2 - surface of the edge of the screw flight, 111 de - equivalent diameter of the helical screw impeller according to Chavan et al. (1972), m 4. F01: a given value of rotational speed, the shear rate is determined from Metzner and Otto's assumption that )7 _ k'N (3.8) With 761v - shear rate on the surface of the "equivalent" cylinder and 41rCr1 k'- 2 (3.50) 1-36 The shear stress is given by 30 - 43—, (3.51) 2Hd aav where: Cr1 - correction factor - f(n, ¢) (See Calderbank and Moo-Young, 1959) H - total height of the inner cylinder, m M - experimentally determined torque on the impeller shaft during the shearing of the investigated fluid, N m p - geometric ratio, dO/de do - outer cylinder diameter, m Significant changes of k' with the value of n were observed. Comparison of experimental data with those obtained with a concentric cylinders system showed that the mixing instrument yielded reasonable results. Again, care must be taken when applying this method to dif- ferent impeller geometries due to the importance of the d/D ratio in the development of the theoretical analysis. 3.2.2) e d tr 3 Dete i ation Knowledge of the yield stress is important in handling, process- ing and transport of fluids. The presence of a yield stress can affect the settling of particles in concentrated suspensions (Nguyen and Boger, 1983). Also, agitation of such fluids often gives a well-mixed region close to the impeller and a stagnant or near-stagnant fluid in the remainder of the container if the yield stress is not exceeded (Solomon et a1., 1981). This well-mixed region around the impeller has been called a cavern (Witcherle and Wein, 1981) and the boundary of the 31 cavern is defined by the surface where the local shear stress equals the fluid yield stress (Nienow and Elson, 1988). Nguyen and Boger (1983) investigated the applicability of the use of a vane for yield stress measurements in concentrated non-food suspen- sions. Haimoni and Hannant (1988) used it on cement slurries.A vane (Figure 3.1) consists of 2-8 thin blades centered around a small cylindrical shaft. This technique has been called the "vane method". In this technique, the fluid under investigation is placed in a container and the vane (which is attached to the torsional spring- driving motor system of a concentric cylinder viscometer) is fully immersed into the sample, then rotated very slowly at a constant speed, and the torque required to maintain this motion is recorded as a func- tion of time. The technique detects the yielding moment when the torque exerted on the vane shaft reaches a maximum. The presence of such a maximum in the torque response is a characteristic of yield stress materials which can be explained by the concept of structural defamation and breaking of bonds in flocculated suspensions (Nguyen and Boger, 1985a). From a torque balance on the surface of the impeller, the yield stress can be calculated from the measured maximum torque, Tm , and the dimensions of the vane, by H l T 33L _V _ (3.52) o - 3 + at!) D 3 V V Where: To - vane yield stress, Pa 32 Tm - maximum measured torque, N m “V - vane height, m Dv - vane diameter, m The method provided satisfactory yield stress measurements only if the vane was rotated at sufficiently low speeds. At high speeds, significant viscous resistance together with instrument inertia and insufficient damping may introduce errors to the measured TIn and hence to the calculated value of yield stress. Nguyen and Boger (1983) recom- mended some operational procedures: 1. Vane should be operated at rotational speeds below 10 rpm. 2. Depth of sample and diameter of the container should be at least twice as large as the length and diameter of the vane to minimize any effects caused by the walls of the container. 3. The vane should be placed at approximately the center of the container. 4. Geometric criteria (Nguyen and Boger, 1985a) for satisfactory measurements: Hv/Dv < 3.5 ; D/Dv > 2.0 Zl/Dv > 1.0 ; 22/Dv > 0.5 where: D - container diameter, m 21 - clearance from surface to top of impeller, m 2, - clearance from bottom of impeller, m Values of yield stress obtained with the vane were compared with 33 those obtained by other methods and the agreement confirmed that the vane method is useful for measuring accurately and directly the true yield stress of concentrated suspensions (Nguyen and Boger, 1985a). The advantageous features of this technique are: 1. Introduction of the vane into the sample does not significantly disturb the sample prior to measurement. 2. Wall slip effects are eliminated and particles remain unsettled. 3. It allows measurement of the yield stress under static conditions and within the material itself. 4. Technique requires short operation time and low cost apparatus. 5. Experiments are easy to perform and of high precision. Keentok et a1. (1985) observed that the vane diameter had negli- gible effect on the ratio of the diameter of the fracture zone to the diameter of the vane. Their data supports the use of the vane for yield stress measurements if a diameter correction is applied. Leong et a1. (1987) measured the yield stress of brown-coal water suspensions using this technique. Yoshimura et a1. (1987) utilized this technique for the measurement of yield stress of oil-in-water emulsions, conducting stress-controlled rather than shear-controlled experiments. Qiu and Rao (1988) investigated the determination of yield stresses of food materials using a mixer viscometer with the vane method. The authors observed that the magnitudes of do for applesauce were higher than those obtained by extrapolation of the Herschel-Bulkey and the Mizrahi-Berk model, n a -a + [K’y J (3.53) where : 34 2 00m - Mizrahi-Berk yield stress, Pa 2 Km - constant to be determined, Pa sn nm - flow behavior index, dimensionless and very similar to those magnitudes of the Bingham yield stress ob- tained by the common procedure of extrapolation of the linear portion of the shear stress-shear rate data. Two different impellers (a star im- peller and a vane) were used and impeller geometry seemed to affect the values of do. The technique proved suitable for the measurement of yield stresses of food suspensions for a specific impeller and rotational speed. 3.2.3) Analysis of Time-Dependent fiehavior A better understanding of the flow properties of time-dependent fluids is essential in handling and process design. For instance, reduc- tion in the viscosity of the material by mechanical treatments prior to pumping may imply lower transport energy requirements and minimization of start-up problems usually associated with occasional pipeline shut- down (Nguyen and Boger, 1985b). During the mixing process, an element of thixotropic fluid ex— periences short periods of time at high shear rates close to the 1Illpeller and longer periods in the lower shear rate regions remote from the impeller (Edwards et a1., 1976). The rheological state of the material will depend upon this shear history and the instantaneous shear rate, which will affect the power requirements for the impeller. 35 3.2.3.1) Rheglogigal Mgge], 3123 Linear Shea; Rate Method Edwards et al. (1976) developed a procedure for the calculation of power consumption with time when a thixotropic fluid is agitated from rest using an impeller which rotates at constant speed. Even though a simplifying assumption, the use of Metzner and Otto's linear relation- ship ( i - k'N ), provided a simple means of predicting the power input- time behavior for thixotropic fluids using the Newtonian data [value of constant A from Eqn. (3.1)] and that for time-independent non-Newtonian fluids (value of impeller proportionality constant, k'). If the thixotropic fluid is agitated for a time t in a mixing container using an impeller at rotational speed N, this is considered to be equivalent to shearing the fluid in a viscometer, from the same starting condition, at the average shear rate for time t. Thus, using the viscometer at constant shear rate, the average apparent viscosity "a of the thixotropic fluid at time t can be measured. Experimental proce- dure consisted of: 1. Measurement of torque as a function of time as the impeller rotates at constant speed. 2. Obtaining equivalent na/time data in a concentric cylinder viscometer. 3. Calculating k' from a plot of "a versus N . Some evidence of the dependence of k' on fluid properties was present but the authors assumed an average value of the impeller con- stant. They concluded.that it was possible to extend the average shear rate ”av ) concept for time-independent non-Newtonian fluids (Metzner and Otto, 1957) to the mixing of thixotropic fluids. The procedure 36 proved satisfactory for a variety of impellers and thixotropic fluids provided that the impeller was capable of maintaining the entire fluid in motion. Nienow and Elson (1988) strongly suggested that this approach ( ”yav 0: N ) should be carefully revised when using it for time-dependent fluids, especially shear-thickening fluids due to the still unknown flow behavior of dilatant fluids in mixed vessels. The authors conclude that the method for determination of k' developed by Metzner and Otto (1957) should be carefully revised and even repeated for different fluids and mixing systems . 3.2.3.2) Com ne eo o ical And inetic Model Plus Lingr Shear Rage Meshes Sestak et a1. (1982) developed a procedure for calculating the time-dependent torque necessary for mixing inelastic thixotropic fluids by means of impellers. The relationships of the impeller (anchor agitator) torque versus time for constant rotational speeds when mixing a bentonite-water suspension, were measured. The deformation histories were expressed by means of initial values of the structural parameter, A0 , at the beginning of any mixing experiment. A stepwise change of rotational speed was engaged and time-dependent torque values were measured. A time-dependent apparent viscosity given by r) --fl£)- (3.54) 37 was used, and an expression for the time-dependency of the impeller torque was obtained, M - daa(t) (3.55) A— 2nk' The authors also compared several models of thixotropy and con- cluded that Cheng's model . n yo yl + my (3.56) —§%— - a(1-)) - b y A (3.57) where: a - yield stress, Pa yo 0 - shear stress, Pa yl A - time-dependent structural parameter (f(y)), which ranges from an initial value of 1.0 for zero shear time to an equilibrium value, Ae’ which is less than 1.0 (Tiu and Boger, 1974) a - model parameter, l/s b - model parameter, dimensionless was the best for calculations of the impeller torque-time variations for an arbitrary past deformation history of a thixotropic fluid in a mixing process. The ability of this model to include the influence of the past deformation history upon the instantaneous apparent viscosity of the fluid was also proven. 38 A more complete analysis was developed by Ford and Steffe (1986), who combined mixer viscometry techniques with a fundamental analysis CUE thixotropy to determine the basic parameters describing the time- dependent behavior of starch-thickened strained apricots. Tiu and Boger (1974) obtained a model to describe the structural breakdown of a product: no a - A [ ayo + KO 1 ] (3.58) where: no KO - consistency coefficient at time zero, Pa S n0 - flow behavior index at time zero, dimensionless A - structural parameter, accounting for time-dependent effects, dimensionless ayo - yield stress at time zero, Pa with the decay of the structural equation assumed as a second order rate equation: JA— - -k1(A - A )2 (3.59) (it e where: k1 - rate constant - f(i), l/s Ae - equilibrium structural parameter, dimensionless Q; - change in A with respect to time, 1/s dt 39 The relationship between apparent viscosity and time is 1 1 + alt (3.60) where: fie - apparent viscosity at equilibrium, Pa 5 n - apparent viscosity at time zero, Pa 3 80 t - time, 3 k1 ’7 (3.61) n ayo + K01 and the value of Ae was calculated as: n *‘1 A _ __e.___ (3.62) e . n ayo + K01 The technique developed by Ford and Steffe (1986) is as follows: 1. Use Metzner and Otto's approximation: yav- k'N (k' - 4.46 for the paddle impeller used (Steffe and Ford, 1985)). 2. Use the linear shear stress assumption: 0 - k"M (k" - 9835 for the system used (Ford, 1984)). 3. Find an approximate shear stress for the sample, aax’ given by ”g M (3.63) witil as - shear stress for standard solution, Pa 40 0x - shear stress for test sample, Pa Ms - mixer torque when agitating standard solution, N m Mx - mixer torque when agitating test sample, N m 4. Plot 1/(na - ne) versus time for each sample. 5. From linear regression obtain values of a1 (slope) and 1/("a - ne) (intercept). [See Eqn. (3.60)] 6. Plot values of a1 versus fiav’ 7. The torque at time zero, Mo, is obtained as Mo 9835 b + 9 Equation (3.64) is obtained from b - 1/("a - ne) . Since ' /' the value of the 510 e b become b — 78V "3 0 7av ’ P S, —- aao - 0e Substituting the expressions for the shear rate and the shear stress and manipulating the data yields Eqn. (3.64). 8. Calculate the value of "so from Eqn. (3.63) using the calculated value of M . o 9. Plot 0 versus 7 . ao av 10. Find the values of K0, no and ayo from linear regression using Herschel-Bulkley model: a - 0° - n 7 n 11. Determine Ae using Eqn. (3.62). 12. Determine k using Eqn. (3.61). 1 13. Obtain complete rheological characterization of sample [Eqns. (3.59) and (3.60)]. 41 This technique mathematically describes irreversible breakdown and it proved useful for suspension-type products because the slip at the wall and breakdown problems involving product loading are minimized (Ford and Steffe, 1986). 3.2.3.3) snhjeetive Aeeeeenene ef Ibinetrony Using e Vane Impeller When studying the time-dependent rheology of highly concentrated and flocculated suspensions of bauxite residue (red mud), Nguyen and Boger (1985b) found that the concentric cylinder viscometer was un- suitable since the transient data obtained were not reproducible due to the presence of slippage at the walls. A mixing system using a vane impeller was found to be particularly suited for following the time- dependent transformation of the structure of red mud suspensions without causing any significant disturbance to the material. The procedure was as follows: 1. Agitation of the suspension either in a capillary viscometer or in a separate container using an anchor impeller rotating at constant speed. 2. After a determined period of mixing, the impeller is stopped to allow for rheological measurements. 3. Using the vane method (Nguyen and Boger, 1983), the vane is slowly iJillllersed into the sample, then rotated at a speed of 0.1 rpm, and torque measurements are recorded. 4° At the end of the test, the suspension is remixed and the procedure repeated until no further changes in the flow properties are °bServed. 42 The vane method was also employed to quantify the thixotropic recovery with resting time. In the recovery experiment, the suspension was allowed to rest undisturbed in closed containers and the yield stress was determined at intervals of resting time (Nguyen and Boger, 1985b). Experimental results provided a complete description of the thixotropic behavior of highly concentrated red mud suspensions. The drastic reduction in the magnitude of the rheological properties with mixing, and the subsequent slow increase in the yield stress when at rest, may be a way to characterize irreversible thixotropic behavior. Simple thixotropic models were formulated for correlating the experimen- tal results. The same technique was followed to characterize the time- dependent behavior of brown-coal suspensions (Leong et a1. , 1987). 3.2.4) Elaetic Flnide Many fluid and semi-solid foods exhibit viscoelastic behavior, i.e. they exhibit viscous and elastic properties simultaneously. Due to their complex rheology, a complete understanding of the phenomena in- volved in mixing these fluids is important in industrial operations to ensure proper selection of process and geometrical variables (Ulbrecht, 1974). Even though the effects of fluid elasticity on agitators are not totally clear, elasticity is known to affect the power required for agitation and to produce differences in the flow fields around the mixing impeller. Generally, it is predicted that the viscoelastic nature of a fluid tends to reverse the direction of secondary flows induced by centrifugal force. White et a1. (1977) observed that three different 43 flow regimes can exist in the mixing tank depending upon the level of viscoelasticity. 3.2.4.1) W In general, mixer viscometry techniques require the measurement of the power required to turn the impeller agitating the fluid. The calculation of torque (power) requirements for mixing viscoelastic fluids is also important in the design of fermentors or processing tanks (Prud'homme and Shaqfeh, 1984). The vast majority of studies on rheology of agitated fluids have focused only on shear viscosity. However, vis- cosity alone is not sufficient for calculating the torque required to mix a viscoelastic fluid. Thus, it is important to know under what conditions power correlations for viscoelastic fluids differ sig- nificantly from those of inelastic fluids. The classical apparent viscosity approach introduced by Metzner and Otto (1957) for power consumption of non-Newtonian fluids in agitated tanks has been considered by researchers (Table 3.2) to inves- tigate agitation requirements of viscoelastic fluids. Their findings indicate that mere use of the "a of non-Newtonian fluids may not be applicable in the case of viscoelastic fluids as they exhibit different power consumptions due to their elastic nature. However, there seems to be a controversy as how viscoelasticity affects power requirements in agitated tanks. Other works (mainly theoretical) in mixing of viscoelas- tic fluids are also listed in Table 3.2. Mashelkar et al. (1975b) observed that when agitating shear- thinning liquids even having moderate elastic properties, the power 44 66.3.. 6.666.. to ma.x.: so 66.esum ”N.. oHnae :0..e5:0:00 .030: 5.353500 .28.. 8.353050 .05.. 5:95—0:00 .030... .2.... .0003 8.353500 .05.. .33. Soon. 5.2.5.500 .05.. 8.353200 330d 5.353050 3.0a 5:353:00 .030... 8:95.500 330m .00... 0500:0030! 00 0 a Amara-.5 30.00532“ 9.35.0100. :0 5.05000 .0 0.00.5 8:553:00 .05.. 5.553500 00.606. .3235... .00... 95:30.03 8.553500 .25.. 2:0... 83050.50. 5 5.553500 00.60... 8.553500 Use... .3285... .656 83866.63 .6656... .8552. 5595.0!!! :0 3.05009 .0 Sue—U 83. 12.6.. 65...... 65...... 65:3 65...... 65...... 6563. 65...... 6.3.3 62 6.66.3235... 0 .x: .8353 6.65:6 6559.8 6.23. 65...... 6.6.66... 63.8.. 65...... 6686 8.1.. 6.33.. 66.6 65.36.. 6.665.... 63:62.60 .22. .6 .6 66> .22. 6.....8» a .8» .32. .262 a .666... .R2. .863... a 638.. .32. 36.666 0 66.5.9636 .32. ..o .6 .636 .32. ..6 .6 62.6.2 .83.. .6 .6 852.66: .32. .3 .6 =38... .22. .6 .6 6.6.. .32. .3 .6 5666.. .22. 369.6260 s 8.5.6.. .32. 83.65 .32. .3 .6 6.25 372.2. .s a :65 .32. 86:6... a .6680 .32. 62665626 a 63.30 .82. 5.8.0 .22. ..6 .6 8668 .82. .0 .6 6.5.8 .82. .6 .6 6.6.68 coco Eocene 539$ aces: 00:30.3. 45 consumption is considerably less than that predicted by the classical apparent viscosity approach. The same behavior was observed by Ranade and Ulbrecht (1977) and Oliver et a1. (1984). They also found that Metzner and Otto's approach is likely to fail in scaling-up on the basis of power consumption per unit volume due to the different power require- ments. The need for extensive experimental work to evolve design procedures under these conditions using the dimensionless groups connected with the elasticity of the fluid was strongly emphasized. ConverseLy, other investigators have observed an increase on power requirements when mixing viscoelastic fluids (Nienow et a1., 1983; Prud'homme and Shaqfeh, 1984; Collias and Prud'homme, 1985).]flienow and Elson (1988) , in a review of the mixing of rheologically complex non- Newtonian fluids in mixing tanks, concluded that viscoelastic properties of the fluid may either decrease or increase the power requirements. They state the impossibility of predicting which because of the complex flow patterns developed in the mixing tank which strongly depend on the geometry of the system and impeller, the type of fluid and the scale of operation. Yap et al. (1979) assumed Metzner and Otto's method was adequate to describe the viscous properties of the fluid around duaimmeller blade only for fluids that exhibited low elasticity and for low values of rotational speed. This method assumes that the value of k' [from Eqn. (3.8)] is independent of the fluid and system characteristics. The authors developed an expression for generalized power consumption for fluids with a low degree of elasticity: o 93 o 91 1 23 -1 Po " 2“ nb[ (Rge) ° (1)/d) ° (d/l) ° ] (3-65) 46 where: nb - number of blades D - diameter of vessel, m d - diameter of impeller, m 1 - length of impeller blade, m 23 2 d NP (txlav ) Rge- Generalized Reynolds Number = fl ‘ o with t1 - fluid characteristic time, s s - fluid rheological parameter, dimensionless n - limiting viscosity at zero shear rate, Pa 5 This model was not successful with fluids showing a high degree of elastic behavior. Nienow et a1. (1983) assumed that the fiav determined by the nwthod of Metzner and Otto can be applied to parameters other than viscosity when studying the power requirements in aerated vessels. Thus, W1 - [ —- 1 <+ ) (3.66) I where A' is obtained from N1 - A' "1b correlations. Also, the ratio of elastic to inertial forces can be similarly derived as W [ 2] ( Vav) (3-67) 47 where W1 - Weissenberg Number - ¢; N $1 - first normal stress coefficient, Pa 32 The significance of this assumption was not clear from their findings and the authors suggested further work to test their results. Since most viscoelastic fluids have strong shear-thinning vis- cosity, power changes may be due to changes in fluid viscosity or elasticity. Boger (1977/1978) observed that maltose syrup-separan solu: tions were highly viscoelastic fluids which exhibited a nearly constant viscosity with high elasticity (normal stress levels) over a broad shear rate range. This model fluid, called the Boger fluid, has been used to better assess the effects of elasticity on agitated tanks (Oliver et a1., 1984; Prud'homme and Shaqfeh, 1984). Prud'homme and Shaqfeh (1984) developed a correlation that ex- plicitly includes fluid elasticity which provides a basis for assessing whether elastic effects are likely to cause significant increases in mixing torque (or power) requirements. Thus, the total torque is given by the torque that would prevail in mixing a Newtonian fluid times a contribution due to elasticity as follows: r - (1+ma1/‘) (12.71:e + 2.41 x 10'3 Res) (3.68) [ elastic] [ Newtonian effects] where: l‘ - dimensionless torque and N1 ma - 2 2 - elastic parameter, (3.69) 2 p r w 48 also called the Aberystwyth Number (Thomas and Walters, 1964) where: r - impeller radius, m N, - first normal stress function, Pa 0 - angular velocity of rotation, rad/s This correlation is based on data in the laminar flow regime and does not account for changes in the geometry of the system and fluid vis- cosity. Collias and Prud'homme (1985) found that elasticity substantially increases power requirements of turbine impellers in the viscous (laminar) regime - the torque more than tripled for the most elastic fluid. However, the magnitude of theleffect of elasticity depends on both the fluid properties and the size of the vessel. A procedure to determine the additional torque to mix an elastic fluid was developed. Secondary flow patterns are determined by the balance between inertial and elastic forces with E - 2 (3.70) and W - R . E (3.71) where: E1 - elasticity Number A dimensionless torque, T, is determined as, 49 T - “a: - £016,121) - gazes.) (3.72) '7 for a given geometry. The Torque Number is related to the Power Number by, 2 2« T - P R (3.73) o e Viscosity and elasticity data were collected in a cone-and-plate vis- cometer. A correlation for torque as a function of Re and E was 1 obtained by determining the «av in the vessel at each Reynolds Number using Metzner and Otto's relationship for turbine impellers: iav - 11 N (3.74) Finally, the torque required, at a certain Reynolds Number, to mix a viscoelastic fluid (a mixture of corn syrup, water, glycerin and polyacrylamide polymer (Boger fluid)) using a turbine impeller, was determined by adding the torque required for a Newtonian fluid the additional torque due to elasticity, given by: 3 3 2 T - 13.12 Re + 0.01167 R6 + Re (71 E - 3200 E ) (3.75) l l or, in terms of Weissenberg Number [Eqn. (3.71)]: 3 2 2 T - 13.12 R + 0.01167 R + (71 W R - 3200 W. R ) (3.76) e e i e 1 e 50 Equations (3.75) and (3.76) provide quantitative results on the effects of elasticity on mixing torque in the laminar flow regime. 3.2.4.2) WW Another manifestation of viscoelasticity is the climbing of the fluid up a rotating rod associated with nonlinear effects, the normal stress, which does not occur in Newtonian fluids (Joseph et a1. , 1984). This phenomenon is often called the Weissenberg effect. Figure 3.3 illustrates the rod-climbing phenomenon in a vessel agitated by an impeller. When the elastic force is sufficiently high, it overcomes the inertia and the fluid is pulled towards the impeller. Because of the role it may play in rheological testing and processing operations, the possibility of using the Weissenberg effect as a method of characteriz- ing viscoelastic fluids has been investigated (Table 3.2). Beavers et al. (1980) showed that the free surface deformations on a viscoelastic fluid sheared between two concentric cylinders when the Weissenberg effect occurrs, can be used to determine rheological data about the fluid. They also investigated the effect of the impeller diameter to vessel diameter ratio, using two concentric cylinders. When d/D approached unity, more complex shapes of the free surface occurred and it showed dependence on the rotational speed, N. Eitelberg (1983) numerically analyzed the influence of the finite length of a rotating cylinder upon the Weissenberg effect. Results indicate that the secondary flow influences the shape of the free sur- face and that it is affected by the ratio of the distance from the free surface of the fluid to the end of the rotating cylinder, h, to the radius of the outer cylinder, r2. The main result of this study is that Sl C IMF“ C) 3’“ ”F \ Figure 3.3: Flow Patterns In Agitated Vessels. A) Newtonian Fluid. B) Elastic Fluid (Rod-climbing). 52 secondary flow does not reach the free surface if the distance h is considerably greater than r2. Even though rod (or shaft) climbing is a sure indication of viscoelasticity, the absence of the Weissenberg effect does not imply that the fluid is inelastic. Nienow and Elson (1988) indicated that certain geometric (shaft diameter) and operational (impeller rotational speed) variables as well as the presence of a yield stress, may reduce the climbing effect. Available mixer viscometry techniques are considered unsuitable for viscoelastic foods due to the Weissenberg effect (Rao, 1977). However, the need to measure the rheological parameters (particularly elasticity) at the same time as the power data is obtained while agitat- ing the fluids suggests the potential applicability of mixers for the evaluation of rheological properties of viscoelastic fluid foods. Reliable techniques could be developed which consider the effects of geometry on the Weissenberg effect as well as power requirement deter- minat ion . 53 3.3) W 3.3.1) WWW 3.3.1.1) Relegien fieeween Elew Bettern end Power Coneumption In A Cylindrieal Vegeel. The state of flow in a cylindrical mixing vessel is complicated and there is some turbulence near the impeller blades. For simplicity, it is assumed that the tangential flow is predominant and the flow can be approximated as a type of Rankine's combined vortex (Nagata, 1975). When a low viscosity fluid is agitated in a cylindrical vessel, a cylindrically rotating zone around the central axis of the vessel is formed, where the fluid rotates with the same angular velocity as that of the impeller blade, while the flow in the outer part is similar to that of a free vortex as shown schematically in Figure 3.4a. The central area of the impeller [abcd] is assumed to have no relation to the power consumption and only the outer part (the tips of the impeller) [AadD] and [BbcC] have an important effect upon the power consumption (Nagata et a1., 1957). When a fluid of higher viscosity is agitated, the radius of the cylindrically rotating zone, rc , decreases and it approaches zero at the transition from turbulent to laminar flow. Thus, the whole impeller area [ABCD] has a relative velocity, u to the fluid and contributes rel ’ to the power consumption. Other vortices present in the low viscosity region such as V, , V1' , V2 and V2' , are weak compared with the forced vortex in the central zone (Figure 3.4a). In the range of turbulent flow, an impeller has a relative Velocity urel - AA' - AA”. Figure 3.4b illustrates the relative velocity 54 Figure 3.4: a) Simplified Flow Model for Power Correlation; b) Relative Velocity Distribution Between Paddle and Liquid (Nagata, 1975) 55 distribution in the case of a paddle impeller. The impeller power con- sumption to maintain the flow is considered as the energy per unit of time to overcome the resisting forces on the blades. The internal resisting force of fluid acting on an element of area bdl may be written as 2 dF - pCure b d1 (3.78) l where C is assumed to be a constant. When the impeller rotates opposite this resisting force at a relative velocity urel , the power consumption is expressed by dP - w dM (3.79) or dP - 0 2r dF (3.80) where M - the moment of force acting on the impeller shaft, N m w - angular velocity - 2xN, rad/s r - radial distance from the axis to any section of the impeller Thus, dP - 4«Nr dF (3.81) Substituting Eqn. (3.78) into Eqn. (3.81), 2 dP - 41rNprure r dl (3.82) l 56 By integration, the power consumption for the impeller is given 14 2 1 P - f0 «Nprurelr d (3 83) Integration of Eqn. (3.83) requires an evaluation of u which rel’ is a function of the hydrodynamics of the vessel. Thus, the fluid flow induced by a radial type impeller (paddle) rotating in an unbaffled vessel can be described by using the forced and free vortex theory: u - 0 when r < rc (3.84) rel 2 u - ZxN ( r - rc /r ) when r > rc (3.85) rel where rc is the radius of the forced vortex cylinder; it is a function of the Reynolds Number and goes to zero at small values of the Re (laminar region). An approximate equation for the power input in agitated vessels in the turbulent region is then obtained (Nagata et a1. , 1957): 3 p P _ P3 - B [10 + 0.6fRea] (3.86) a 5 3 p N d 10 + 1.6flle where a, p, f and B are the coefficient for the empirical Eqn.(3.86). In the range of laminar flow, the power consumption increases with the viscosity of the fluid, and can be characterized by the following relationship 57 A Po - R e (3.87) Following this reasoning, Nagata (1975) developed an approximate equation for the power consumption of paddle impellers in agitated vessels with free surface for the complete range of flow regime. Combining Eqns.(3.86) and (3.87), a P P P B 10 + 0.6fRe A (3.88) o - 3 5 - 3 a + R p N d 10 + 1.6fRe e [ turbulent ] [ laminar ] with B, A, f, a and p determined experimentally. Equation (3.88) can be applied to wide ranges of Reynolds numbers and to various paddle geometries. The above analysis is valid only under the following conditions: 1. The agitated fluid is Newtonian. 2. The system consists of a single impeller centered in the axis of a vertical cylindrical vessel with a flat bottom and no baffles. 3. The fluid in the cylindrically rotating zone rotates with the same angular velocity as the impeller. 4. The value of C in Eqn.(3.78) is constant. 3.3.2) 121W: Power consumption data have often been correlated using dimen- sional analysis. The variables which affect fluid motion.i111nixing are of three types (Chavan and Mashelkar, 1980): l. geometric (linear dimensions) 58 2. fluid properties [density (p), viscosity (0)] 3. kinematic and dynamic characteristics of flow [velocity (u), gravitational acceleration (g), power (P)] In mixing with rotating mechanical impellers, the velocity is defined as the linear speed of the tip of the impeller (Rushton et a1., 1950), so that u - de (3.89) where, d - impeller diameter, m N - impeller rotational speed, rev/s Power input by the impeller, P, is used to produce the forces irl the mass flow and also to overcome the force of gravity, g. The power required to rotate the shaft and blades of the impellers may be expected to be a function of many variables: P - f( d. D. H. b. C. L. p. n. g, N) (3.90) Dimensional analysis (Appendix A) gives the general equation relating the physical variables most often encountered in mixing a Newtonian fluid, +5 -. [LN—L1“ [Jim-1’33 [2]“ [if [2]“ a“ (3.91) The last five terms define the effects of system and impeller geometry. Thus, for geometrically similar systems, 59 _L 91:15.2. ’31 d N2 ’92 <3 92) . . - A [ J [ ] ' pN d ’7 8 fl; fiz Po - A (Re) (Fr) (3.93) where: P - Power Number - —% (3.94) o p N d 2 Re - Impeller Reynolds Number - —d—l:—L (3.95) 2 41-N— (3.96) F - Froude Number - r 8 White et al. (1934a) first defined the drag coefficient group now known as the Power Number, Po, which characterizes the flow pattern and represents the ratio of the power dissipated per unit volume to the increase in kinetic energy. The impeller Reynolds Number, Re’ has significance as a ratio of accelerate force to viscous force. The form [ d2Np/r) ] has come into general use for characterizing mixer operations that employ rotating agitators (Hyman, 1962). When agitaing non-Newtonian fluids, the form of the Reynolds Number may vary, as is shown in section 3.2.1. The Froude Number, Fr’ is theoretically required to account for the vortex formation as a result of the influence of gravity in an agitated system. The influence of the FI. on power consumption seems to be important only in unbaffled vessels outside the laminar flow region (Green, 1953). The addition of baffles has little effect on power re- quirements in the laminar flow region (Treybal, 1956; Nagata et a1., 60 1957; Blasinski et a1., 1970; Nagata, 1975). Also, high viscosity fluids (above 20 Pa 3) have sufficient internal resistance to show little if any vortex motion, i.e. , the surface of the fluid remains essentially horizontal (Nagata, 1975; Deak et a1., 1985). In general, the influence of the Froude Number on the mixing power requirement is considered negligible and practically non-existent in the laminar region of flow. Thus, the power consumption relationship can be expressed for each flow regions 9‘1 Laminar flow or fully baffled vessel: Po - A Re (3.97) a1 ‘12 Turbulent flow or unbaffled vessel: Po - A Re Fr (3.98) 3 . 3 . 3) Lamina; Mixing Region 3.3.3.1) Lamina: Eluig Ligtion in Agitated Vessels In the laminar flow region, the fluid around an impeller moves with the impeller rotation and the fluid distant from the impeller is almost stagnant (Figure 3.5). At very low Re, there is no turbulent flow and the secondary circulation flow is very weak, so that the momentum transfer from the fluid near the impeller to the more remote parts of the fluid depends mainly upon the molecular viscosity of the fluid and therefore the amount transferred is small and the velocity of remote fluid is low. As the Re increases, secondary circulation flow occurs and momentum transfer increases (Nagata et a1. , 1960). 6l (“1" A/ ’l b t A Transitional [D l] Baffled Reg. 4°] urb. Sm. D Stagn. Zone Reg. °f m Lam. Zone Flow U Turb. Zone L [Dz] Non- baffled Figure 3.5: Fluid Motion in Agitated Vessels (Nagata, 1975) 62 Nemax A\ It. Np max 6 (”dc C ----- '1 F (2) -‘uNp -E (3) Mpg," 1 103:102 103 104 105 106 R. Figure 3.6: Schematic Diagram Showing The Relation Between Power Number and Reynolds Number For a Paddle Impeller (Nagata, 1975). 63 3.3.3.2) MW Because of the similarity of the power correlation curves [Po versus Re (Figure 3.6)] to the friction factor plot for pipeline flow, the region where the slope is equal to -l is considered to represent a laminar (viscous) flow region. This has been experimentally verified by numerous researchers. Thus, for all impellers, the laminar flow regime is charac— terized by a linear decrease in the Power Number (P0) with Reynolds Number (Re) , (3.99) Applicability of Eqn. (3.99) is limited by the critical value of the Re which depends on the geometry of the mixed system alone for Newtonian fluids . 3.3.3.2.1) gritical Reynolds Number In the mixing system, the transition from laminar to turbulent flow proceeds gradually and no distinct critical Re exists for the flow in an agitated vessel as for other hydrodynamic processes such as pipeline flow (Re - 2100) and sedimentation (Re - 1). However, ex- perimental values of the Reynolds Number defining the limit of laminar flow for mixing by mechanical agitators have been determined (Rushton et a1., 1950; Green, 1953; Hirsekon and Miller, 1953; Nagata et a1. , 1957; Pollard and Kantyka, 1969; Nagata, 1975). Results show that the region of purely viscous (laminar) flow extends to Re numbers from ten to one hundred (10-100) and it seems to be influenced by the geometry of the system (Chavan and Mashelkar, 1980). Thus, it may be incorrect to define 64 the laminar flow region unless impeller and vessel type is defined. Laminar flow can be achieved throughout the vessel by the correct design of the system, i.e. , selecting the correct combination of impeller and vessel geometry, and diameter and rotational speed of the impeller. The laminar flow region is also a function of the type of fluid being investigated. Metzner and Otto (1957) observed that the region extends to higher Reynolds Numbers in pseudoplastic fluids than in Newtonian fluids. For turbine impellers, a value of Re - 10 was obtained for Newtonian fluids while laminar flow was observed until Re - 20 for power law fluids with 0.25 < n < 0.45. The laminar flow region can be limited quite safely by defining Re < 10. When this criterion is satisfied, baffles are not needed and it allows for maximum sensitivity when calculating the average shear rate in the agitated vessel. When agitating non-Newtonian fluids, especially pseudoplastics, the use of a generalized (modified) Reynolds Number enables the ap- proximate prediction of the power of the impeller at low Re (Metzner, 1956). The values of the apparent viscosity, "a’ which are functions not only of the fluid properties but also of the conditions under which it is flowing (Begachev et a1., 1980), are substituted into the expression for the Reynolds Number [Eqn.(3.95)] . A variety of modified Reynolds Numbers have been used by investigators in mixing studies (Table 3.3), where R - LN—E (3.100) By analogy with Newtonian fluids, an apparent viscosity is 65 Table 3.3 Modified Impeller Reynolds Numbers for Mixing of Non-Newtonian Fluids. Reynolds Number (Re) Researchers Impeller Type Magnusson (1952) Hiraoka et a1. (1979) Rushton & Oldshue (1953) Metzner & Otto (1957) Metzner et al. (1961) Godleski & Smith (1962) Nienow et al. (1983) 2 Ducla et a1. (1983) n Reher & Bohm (1970) a Hall & Godfrey (1970) Nagata et a1. (1971) Prokopec (1972) Edwards et a1. (1976) Takahashi et a1. (1984) Shamlou & Edwards (1985) Su & Holland (1967) Rieger & Novak (1974) Bourne et a1. (1981) FForesti & Liu (1959) Metzner & Taylor (1960) Wichterle & Wein (1981) d_fl___g >> Ri), (Krieger and Maron, 1954): 1 - 4 n N (3.102) where k', from Eqn. (3.8), would be equal to 4_1 . Calderbank (1958) n experimentally verified the linear relationship of the shear rate around an impeller in the laminar-flow region with the impeller speed, N. The experimental and theoretical evaluation of the impeller proportionality constant, k', and consequently the rate of shear in the 69 mixing vessel, has been the subject of many studies. The most common approach is the analysis of a non-Newtonian fluid flowing between two concentric cylinders. 3.3.4.1) Theoretieal Exereeaieae Fe; The Rate 0f Shear In A Mixing Veeeel 3.3.4.1.1) ansentris_§xlinder§ Expressions for the determination of the average shear rate in a vessel with an impeller have been determined by considering the mixing system as a two-cylinder system, with the impeller as a rotating cylinder. It is known from theoretical hydrodynamics (Bird et al. , 1960) that the generalized Newtonian law of internal friction for an incom- pressible fluid is T - 2 n D (3.103) where T - stress tensor, Pa 0 - fluid viscosity, Pa 3 D - deformation rate tensor, l/s In.cylindrica1 coordinates, equation (3.103) becomes (Bird et a1., 1960), a _ 2 3_r . a a _ r Q__ 32 + i 33; (3.104) rr " 6r r9 0r " 8r r r 60 a _ 20[ .1 330 + .3r] . a _ a _ r 321 + E 33; (3.105) on r as r 92 20 " 62 r 30 70 810 600 609 _ __z . _ _ r __g __r (3.106) azz 2" az azr arz " I: 6r + 62 :I For a rotary motion in steady-state laminar flow, the fluid moves in a circular pattern and only the tangential velocity prevails, i.e. , we = we (r) and (or - wz - 0. Then, the shear stress, a, may be written as (0 a -a -qu[-1] (3.107) to i - a/n - r a [L] (3.108) Using Eqn. (3.108) an expression for the shear rate at the sur— face of a rotating cylinder in an infinite Newtonian fluid is obtained as :1- -41rN (3.109) which relates the shear rate with the rotational speed of the impeller. Pavlushenko and Gluz (1968) referred to the use of Eqn. (3.108) to determine the average shear rate in mixing non-Newtonian fluids by mechanical impellers. To develop the analysis, the motion of the fluid caused by the rotation of a cylinder is again assumed as an approximate model of the fluid flow produced by mixing with any of the usual im- peller types and the problem of the steady motion of a non-newtonian fluid rotating between coaxial cylinders is considered. It is also 71 assumed that the internal cylinder with a radius Ri rotates at a con- stant angular velocity (w - 2rN) in a stationary cylinder with a radius of Ro - (D/d)R1. End effects are considered negligible in this analysis. At low velocities characterizing the flow of non-Newtonian fluids in vessels with impellers, the inertia forces and the pressure gradient have no marked effect on the phenomenon and may be neglected as a first approximation (Pavlushenko and Gluz, 1968). Another assumption is that the flow of the homogeneous, incompressible fluid is planar (because of the symmetry). Then, the following expression is obtained from the equations of motion in terms of stress components (cylindrical coordinates), 2 ‘i; g; (r Ora) - 0 (3.110) This equation can be rewritten as 60 _l_ 2 __rr - Integrating Eqn.(3.lll) once leads to, 2 r a - 01 (3.112) r0 For a power-law fluid, a - m in . Thus, the Ora component of the stress tensor in cylindrical coordinates is, 72 dr r (3.113) Using equation (3.112), this yields («0) n 2+n g_ _fl C1 - mr [dr r ] (3.114) After integration, we - C.‘.r(n'2)/n + Car (3.115) Using the following boundary conditions: (i) we - ZNNRi at r - Ri (3.116) (ii) we - 0 at r - R0 the expression for the angular velocity becomes [1 "R 2/1'1 ‘ ' .2 _ r w - 2er (3.117) r 1 _ P 52 2/n I. :- R1 .1 and the expression for the shear rate is 2/n . FR ‘ . _e w 1 - _r . . _ d_. [.2 ] _ 5L _ 7 r dr r r dr ZflN 1 - Ro-Z/n . .131- 1 -g 21m 1 [JJ‘Z/n (3.118) R o 73 At r - R1, the expression simplifies to i _ - 4111 (D/d)2/n (3.119) n 1 _ (D/d)2/n Considering y - k'N, then k, _ - g; (1)/<1)”n (3.120) n 1 _ (D/d)2/n Equation (3.119) determines the average shear rate in a vessel with a mixer as a function of the rheological properties of the fluid, the mixing conditions, and the geometrical characteristics of the system and should give more accurate results for mixing of non-Newtonian (pseudoplastic) fluids than the use of the equation developed for Newtonian fluids [Eqn.(3.109)]. 3.3.4.1.2) Em ir ca And eoretical Ex ressions For The Im eller Proportionality Constant, k' The suggestion of Metzner and Otto (1957) of a constant value of the impeller constant, k', which is a unique function of the geometry of the system has been questioned and other expressions for the constant have been determined (both theoretically and empirically) for the mixing of non-Newtonian (mostly pseudoplastic) fluids. Expressions for the impeller proportionality constant, k', are summarized.in Table 3.4. Looking carefully at the expressions, it seems that the value of k' can be a function of the geometry of the system (impeller shape and size, vessel size), the rheology of the fluid (values of shear-thinning index (n) and consistency coefficient (m) or 74 Table 3.4. Empirical and Theoretical Expressions for Determination of the Impeller Proportionality Constant, k'. kl Researcher Impeller Type fi/N - constant C (l-n) C - f(Hg/d) slope of a log IP/(mN“*1d>1 versus (l-n) ls ___l___ 2 v7 1- (d/D) .1. (4)1'“ :zr . 4m (1)/<02”n (a 3 2) 1 :- _ . . b n (D/d)2/n _ 1 ‘Vuflnlealx: D - cup diameter, 111 d - impeller diameter, m The previous equations assume a dependence of the shear rate at the surface of the impeller on fluid properties, impeller speed (rpm) and, system geometry, in a form identical to that of a cylindrical im- pe 11er. Thus, from Eqns. (4.3), an expression for the average shear rate in the real measuring system (a vessel with a mixer) is expected to present a similar form (but somewhat different due to the differences in geometry) to that of Eqn. (4.3.1) and (4.3.2) as follows: (1)/c1)“1 a 2, - p [b/d] 3 N (4.4) av 1 (D/d)a2 _ 1 84 where 61 is a constant and, al, a2 and as are parameters dependent on the power-law index, n. Equation (4.4) differs from Equations (4.2.1), (4.2.2), (4.3.1) and (4.3.2) in the addition of the term which takes into account the effect of impeller variation (i.e., impeller height, b). Equation (4.4) may also be written in the familiar form of the linear dependence of the average shear rate in a mixer on rotational speed, proposed by Metzner and Otto (1957), 78v - k'N (3.8) with (1)/d)“1 a k' - 9. a [b/d] 3 (4.5) (D/d) 2 - 1 Where k' is the impeller proportionality constant, dependent on the system geometry (cup and impeller) and the rheological behavior of the fluid. The average shear rate of an impeller (paddle or flag) can be E”"‘I>ected to be a function of these parameters. Also, a direct relationship between the average shear rate and the geometric dimensionless numbers [(D/d) and (b/d)] can be expressed as f0110ws: 38v - a. [ (1)/d)“1 “3 1N (4.6) Where 85 k' - fl; 1 (1)/d)“1 mEm 32> ace if ii ‘J-c 'Lr Ll4 I 4. le 2L. II! PC tE 51 31 LI' n 98 impeller and cup changes. Six impellers and three cups were used in a total of 13 treatment combinations. Tests were conducted in duplicate and the order was determined by randomization. Table 5.1 shows the different impeller/cup combinations investigated in this study. Impeller diameter, d,(Figure 5.3) was kept constant due to problems during manufacture of the paddles. The impeller blade height b, was varied to investigate the effect of impeller size. Three sample cups of different diameter, D, were used with a range of d/D ratio from 0.3 to 0.7. To not vary a large number of geometric constants at once, the length of the sample cups was maintained at L - 1.5D. The fluid level, Ii, in the cups was kept at H - 1.2D. This distance was selected to keep all impellers sufficiently immersed in the fluid to avoid surface waves, especially for the bigger impellers. Preliminary tests showed that the position of the impeller, that is, the distance between the bottom of the impeller and.the bottom of the cup, c, had no significant effect on torque readings when placed close to the top, in the middle, or close to the bottom. For practical considerations, impeller depth (c) was set at c - 0.5d since it made possible the immersion of the impellers under sufficient volume of fluid. The effect of the distance from the surface of the fluid to the top of the impeller's blade was assumed negligible (as in Nagata (1975). The effect of different impeller shape was inves- tigated by using a flag impeller (Figure 5.1). 5.2.2) m 5 . 2 . 2 . 1) Dreparation of Non-Newtonian Fluids Aqueous solutions of Hydroxypropyl Methylcellulose were prepared by heating distilled water to 70°C (158°F) and slowly pouring the Table i 99 Table 5.1: Experimental Design. GEOMETRICAL DIMENSIONS DIMENSIONLESS VARIABLES SYSTEM IQQI n d b d/D d/b Paddle Impellers l 5.55 1.8 1.0 0.327 1.8 2 5.55 1.8 1.8 0.327 1.0 3 5.55 1.8 3.0 0.327 0.6 4 5.55 1.8 4.0 0.327 0.45 5 5.55 1.8 5.0 0.327 0.36 6 3.50 l 8 1.0 0.515 1.8 7 3.50 1 8 1.8 0.515 1.0 8 3.50 1 8 3.0 0.515 0.6 9 2.54 .8 1.0 0.709 1.8 10 2.54 1.8 1.8 0.709 1.0 Flag Impellers 1 5.55 1.5 3.0 0.273 0.5 2 3.50 1.5 3.0 0.429 0.5 3 2.54 1.5 3.0 0.591 0.5 where: D - cup diameter d - impeller diameter b - impeller blade height 100 Figure 5.3: Mixing System 101 percent weight of sample into the water. Mixing was carried out with a Corning PC-351 Hot-Plate Stirrer. Solutions were cooled down and allowed to rest for a period of 24-48 hours to eliminate air bubbles. 5.2.2.2) Determinatign of gheglggigal Droperties of Non-Newtonian Fluids Rheological behavior of the materials was determined with a Haake RV-l2 concentric cylinder viscometer with M-500 head and the MVFI sensor (d/D-0.90). The viscometer is interfaced to a Hewlett-Packard 85 computer and a 3457 data acquisition system. The samples were previously agitated for a period of 10 minutes to check for thixotropic behavior. Torque was monitored as a function of time and reached an equilibrium value after the completion of the test for all samples. Triplicate replications of torque versus rotational speed data were collected for every sample at 1-120 rpm (0.105-12.57 rad/s). The values of the flow behavior index, n, and the consistency coefficient, m (Table 5.2), were obtained from shear stress-shear rate data, with shear rate evaluated using the method developed by Krieger (1968). The fluids showed power- law behavior and no elastic characteristics, such as rod climbing. The experiments were carried out at a constant temperature of 25°C i 1°C (77°C) . 5.2.2.3) leibzatign of Drogkfield Viscometers The viscosity of a Newtonian standard (Brookfield Viscosity Standards) was determined with the RVTDV-I and the HBTDV—I Brookfield Viscometers and cylindrical spindle # 7 (0.32 cm diameter, 5.37 cm height) to ensure proper instrument performance and high accuracy. 102 5.5.2.4) Data collection Figure 5u4 illustrates the overall experimental system. Once loaded into the cup, the temperature of the fluid was controlled with a constant-temperature water bath connected to the cup jacket with stan- dard tubing and fittings. Temperature of the sample was allowed to equilibrate up to 25°C (77°F). To ensure proper alignment of the im- peller and cup system, a standard base enabled proper placement of the cups. A guard leg was initially utilized to determine the proper align- ment of the impellers but removed before data collection. A selected impeller was immersed in the ution to a fixed mark with care to avoid excessive entrainment of air bubbles. The torque reading at a selected rotational speed was measured after steady state was reached (constant readings). Readings were collected during one minute at the specific value of rpm. For any given run, the rotational speed varied (a step- wise increase) and the range of rotational speed was the operating range of the Viscometers (0.5, 1.0, 2.5, 5, 10, 20, 50, and 100 rpm). The tests were done in duplicate and the reproducibility of results was very high. Neither surface waves nor vortex formation occurred in any of the impeller/cup/fluid combinations which satisfied the laminar flow condi- tion (Re<10). After completion of the test (data collected with impeller rotating at 100 rpm), the impeller was removed and the next impeller was tested. The above procedure was repeated for all systems and fluids. Results will be presented and discussed in the following chapter. Table 103 Table 5.2: Rheological Properties of Sample Fluids newtonian Fluids (Brookfield Standards) 3 Fluid n (Pa 3) p (kg/m ) Standard 1 0.093 i 0.001 928.570 Standard 2 0.923 i 0.001 930.000 Standard 3 4.840 i 0.001 966.829 Standard 4 12.200 i 0.001 969.421 anrNewtonian.F1uids 1 n 1 s Fluid m (Pa 8 ) n p (kg/m ) CMC 1% 6.492 t 0.03 0.504 i 0.01 974.800 CMC 1.5% 28.417 i 0.02 0.374 t 0.007 1025.870 CMC 2% 59.275 i 0.02 0.352 t 0.003 1144.280 1 Means of three replications for 0-2 rev/s for Haake Viscometer data. BOB Brookfield 104 firmnnuflunampd L.J T- 25' C L—i Both Wot-r T Figure 5.4: Schematic Diagram of Experimental System S Doto Acqulsltor 19 V l r4- 105 5.3) CA TIONS USING T ER VISCO Y METHO S Three commonly used mixer viscometry methods were used for deter- mination of the impeller proportionality constant, k' , (thus, average shear rates) in the mixing system. The procedures are outlined in Chapter 3. 5.3.1) W The procedure for the two matching methods: [Metzner and Otto, 1957; and Mackey et a1., 1987] is outlined in Section (3.2.1.1). The main equations used in these methods are Eqns. (3.1), (3.8) and (3.15). Expressions for the Po versus Re relationships were developed for each impeller/cup combination for both the Newtonian and the non-Newtonian (power-laMO fluids. Also, values of k' were determined as a function of fluid properties and system geometry. 5.3.2) filgpe Method The procedure is outlined in Section (3.2.1.3). The main equa- tions are Eqns. (3.41), (3.45) and (3.47). Values of the impeller proportionality constant, k', were determined as a function of system geometry . 5.4) A CU ONS US NG NEW R V COMETRY METHODS 5.4.1) Detetmination of Average Shear Rate The procedure for determination of the average shear rate in the mixing systems is as follows: 106 1) Using the values of k' obtained with traditional mixer viscometry 2) methods, a model was found by fitting the data using stepwise regression. Thus, the values of 31, a1, 02 and as from Eqns. (4.31) through (4.35) were obtained for each impeller/cup combination and fluid under study and expressions for the average shear rate were obtained. The values of k' obtained with traditional methods are plotted versus the values of k' calculated using the equations above. The equation which gives the better agreement is considered the best equation for approximation of the average shear rate in the mixing system. 5.4.2) Dttgtmingtigg 9f Axgtage Shea; Stress 1) To check the applicability of the shear stress equations [Eqns. (4.6), (4.13) and (4.24)], the values of torque calculated from the corresponding equations are compared with the experimentally measured values of torque using the impeller (mixing) system. Hence, the torque equations for the paddle impellers are, M - 2nb (d/2)2 a (4 5) 8V for the concentric cylinders analogy with negligible end effects. When end effects are considered, the equation for torque is the following: M - —£4— [ —g— + ——l— ] (4.12) f( 5.4.: 107 For the flag impellers, the above equations apply in addition to Model System 3 (d - de ), M - 21rb (de/2)2 a (4.36) av where Gav - m (iav )n , for a fluid of known rheological properties, n and m. Thus, using the values of n and m and the expression for lav , the values of an aav are obtained for use in the torque equations. 2) Plot calculated torque versus experimental torque. The equation that best represents the shear stress relationship in the mixing system will then be the one which gives better estimates of the experimental torque values. 3) Develop flow curves (average shear stress- average shear rate curves) for a set of geometric parameters (D, d and b). 5.4.3) o dur o e use 0 a ixer viscomete to directl determine e eo o ic ro erties f ower-law fluids On the basis of the considerations presented in Chapter 4, the following procedure is proposed for the determination of the flow curve (shear stress-shear rate relationship) of a power-law fluid: 1) Measure the torque on the impeller shaft, M, as a function of rotational speed, N, using the mixer viscometer. 2) Determine the value of the flow behavior index, n, from the slope of the log-log plot of M vs. N. 3) Select the appropriate equations for the average shear rate and shear 4) 5) 6) 7) 108 stress in the mixing system by following the steps outlined in Sections (5.4.1) and (5.4.2). For a given value of rotational speed, measured values of torque and known system dimensions, determine the average shear stress, a 9 av for the investigated fluid using the appropriate equation. For a given value of rotational speed, known system dimensions and the value of the flow behavior index of the fluid, n, determine the average shear rate, Iav’ using the appropriate equation. Repeat steps 4) and 5) for the complete range of rotational speeds of the viscometer. Evaluate the flow curves (rheograms) by plotting the ”av versus the fiav , for the investigated fluid in log-log coordinates. The intercept of the log-log plot is the fluid consistency coefficient, m, in Pa 3“. 1’.‘ 0V)“; 109 CHAPTER 6 RESULTS AND DISCUSSION The results of the experimental investigation are presented and analyzed in this section. In the first part, results obtained widn traditional mixer viscometry methods for estimating the average shear rate are presented and discussed. In the second part, results from the proposed method are described and its suitability determined for rheological.characterization of power-law fluids. A procedure for using the Mixer Brookfield Viscometer with Newtonian fluids is presented in Appendix C. 6.1) ESIIMAIIOM OE AVEBAQE SHEAR BATE USING MIXERS: TRADITIONAL METHODS 6.1.1) Matching Viscosities 6.1.1.1) Power Curves Methgd (Metzner and Otto, 1957) Newtonian mixing curves for the paddle impellers in terms of mixing power number (Po) versus Reynolds number (Re) are shown in Figures 6.1 to 6.3. Each figure represents power data for mixing the Newtonian fluids in the same selected sample cup. Data points are means of two replications. These plots indicate that in the viscous regime the power characteristics are in agreement with the relationship indicated by Metzner and Otto (1957) for the laminar region of flow, P -— (6.1) i.e, they follow a straight line with a slope of -1. Regression analysis 110 1 .OE+07 I’UUUU‘I I I I'U'U'l U—l UI'I'Ur r U U'UUIII I I I'U'U'] 1 .OE+00 d/D III 0.327 1 .0E+05 1 .0E+04 1 000.0 P0 = P/[ptNJatds] I 'UIUII I U rUr'I' U U'U'I' I’ TUUUIU 0.0100 0.1000 ' :.0000 10.0000 Re = parNtda/‘n 0.0001 0.001 0 Figure 6.1: Power Number/Reynolds Number Relationship for Newtonian Fluids Using The Paddle Impellers (Cup diameter - 5.5 cm). 111 1.0E+o7 I I I UUII'U‘I I I I IIIIU! I I I FUIU'I I I I IIIIII I I I VIII." I.OE+06 . - d/D - 0.515 1-—! a “-0 1.0E+05 * -. _ * . fl 2: 5: 1.0904 ‘\~ 0- 1000.0 ll 0 0- 100.0 10.0 o d/b - 0.0 o «wmu-10 x cub-IL. 1.0 I I 'I’IUI'r I I III... I U 0.000! 0.0010 ' I Trflr I I truitvr r I I U'III" 1 0.0100 0.1000 1.0000 10.0000 Re = ptNtdz/‘n Figure 6.2: Power Number/Reynolds Number Relationship for Newtonian Fluids Using the Paddle Impellers (Cup diameter - 3.5 cm). 112 1.x+o7 I IIIIII' r I IIIIIII r I I’IUI'U‘ I rIIIIItr I I IIIIIIr 1351-00 d/D - 0.700 1—-1 “-0 1.0E+05 ‘0’ LEE LOE+04 1 ‘\\ x x ‘1- 1000.0 11 0C: 100.0 - 10.0 o d/b - 1.0 I 0 at all: - 1.0 . I fiT "I'00r r 0 0.70"" r firI'U'U' fifi'U'Uli ' Tt""‘r fi 0.0001 0.001 0 0.01 00 0.1000 1 .0000 10.0000 Re = ptNtd2/n Figure 6.3: Power Number/Reynolds Number Relationship for Newtonian Fluids Using the Paddle impellers (Cup diameter - 2.54 cm). 113 on the data yielded slope values ranging from -0.98 to -1.03. Similar results were obtained with the flag impeller. The effect of the Froude Number, Fr , on power consumption was investigated and no significant effect was observed (Appendix B, Tables B1 and B2) for all systems. To account for the effect of geometry, a generalized Po versus Re relationship was preferable and Eqn. (6.1) was transformed to P0 - a0 Rea1 (d/b)°2 (d/0)°‘3 (6.2) for the paddle impellers, and Po - p, Refll (d/D) fl? (6 3) for the flag impeller. Equations (6.2) and (6.3) are nonlinear. To simplify the regression analysis, the following transformation was made, log Po - log ao + allog Re + azlog (d/b) + a3log (d/D) (6.4) log PO - log fio + fillog R8 + fizlog (d/D) (6.5) The results of the regression are presented in Table 6.1 for the paddle impellers and Table 6.2 for the flag impeller. The resultant power prediction equations are, Paddle impellers: (R2-0.990) 0 083 0 223 ,1 058 1 Po - 415.524 [ Re ° (d/b) ' (d/D) ' ] (6.6) m Ix: (V (D 114 Table 6.1: Regression Results of Eqn. (6.2) (Paddle Impellers) ‘Linear Multiple Regression.Analysis Regression Estimated Regression Estimated Standard t Coeffitient Coeffitignt Error log a0 415.524 -- -- a. -0.983 0.005 -174.50 a, -0.223 0.006 -6.52 as 1.058 0.006 17.40 Analysis of variance Sum of Degrees of Error Mean F Squates Freedom Squares Regression 516.414 3 172.138 -- Residual 5.226 310 0.017 10211.5 Iotal 521.630 314 2 R - 0.990 a - 0.05 Ta Te 115 Table 6.2: Regression Results of Eqn. (6.3) (Flag Impeller) Linear Multiple Regression.Analysis Regression Estimated Regression Estimated Standard t* Coefficient Coeffitient Error log 50 28.469 -- -- 51 -0.972 0.015 -63.23 52 0.105 0.0156 0.72 Analysis of variance Sum of Degrees of Error Mean F* Squares Freedom Squares Regression 146.532 2 73.266 -- Residual 3.188 87 0.036 1999.2 Total 149.710 90 R - 0.980 Test of hypothesis for 62 : C1: 82 - O For a level of significance of a - 0.05, t(0.975,87) - 2.00 * Since t - 0.72 < t(0.975,87), we accept C1 and conclude that 82 = 0 116 2 Flag impeller: (R -0.980) 0.972 _1 PO - 25.912 [ Re 1 (6.7) Figures 6.4 and 6.5 illustrate the validity of Eqns. (6.6) and (6.7) for the paddle and flag impellers, respectively. Figure 6.4a is the plot of predicted power numbers [Eqn. (6.6)] versus the "observed" power numbers (calculated with the measured values of torque) with the paddle. The correlation coefficient (0.990) shows the good agreement ‘between.observed and predicted values. The slope of Figure 6.4a (0.906) also indicates that Eqn. (6.6) predicts power numbers close to those observed. The closer the slope and the regression coefficient to the value of one, the better the model. Figure 6.4b presents the Po [Eqn. (6.6)] versus Re curves for Newtonian fluids using paddle impellers. Figure 6.5a is the plot of predicted [Eqn. (6.7)] versus observed power numbers when using a flag impeller. The correlation coefficient (0.980) and the slope (0.900) indicate the ability of Eqn. (6.7) to 'predict power numbers. Figure 6.5b presents the Po versus Re curves for Newtonian fluids using the flag impeller. A t-student test was performed to verify whether the ratio of diameters (d/D) has any significance on the PC for the flag impeller (Table 6.2). Results indicate that the effect of the geometric term is negligible. It is evident that Eqns. (6.6) and (6.7) are useful to indicate the effect of controllable mixing variables: cup diameter, impeller size and impeller shape. Non-Newtonian mixing data (i.e. power numbers) were based on the correlation developed for Newtonian fluids. Power numbers for the non- Newtonian (power-law) fluids were calculated and, as Metzner and Otto 117 1.01E-l-06 1* ' 1 ' F7 ' I ' J R3 = 0.990 Paddle Impellers ‘ 1.206+00-l * .1 0.009054 " 0. 4451-05 - Po (observed) 3.229054 , + I up 0.00 e a If 0 I U f t 0' 0.00 3.22905 0.44:»: 0.0090: 1.2112300 1.0115400 Po (predicted) 1.0:«1-07 Paddle Impellers Po ii i 0.0001 0.0010 0.0100 0.1000 1.0000 10.0000100.00 Re“(d/b)‘”’(d/D)“" Figure 6.4: a) Predicted Po Versus Observed PO; b) Power Correlation For The Paddle Impellers. 118 1am I r I r fi I f r I -R‘ a 0.980 Flog Impeller 1.20:1-00-4 J 0.00am -1 1 lflOE+08~ Po (observed) 300E+054 I' '7 r l f r £00£+08 Odllfldll L20£+GI L30£+00 Po (predicted) h 1300 JLDO£+00 L03+07 Flog Impeller 1.0:1-00 a ’e 1am Q. 1 “+04 ”. Po I 1000.0 100.0 ’ 10.0 b L0 r I 'r'r' I revue 0.0001 0.0010 0.01100 011000 1.0000 10.00 Rem Figure 6.5: a) Predicted Po Versus Observed Po; b) Power Correlation For The Flag Impeller. 119 (1957) proposed, the corresponding Reynolds number was determined from the plots of Figures 6.4 and 6.5. Thus, The values of an "average" apparent viscosity, "a’ were obtained from 2 Re EEC—1— (6.8) n ”8 6.1.1.1.1) Estimation of Averagg Shear Rates Figure 6.6 shows the plot of shear stress or apparent viscosity, "a , calculated as 0/1 using Haake data, versus shear rate independ- ently measured in a concentric cylinders viscometer (Haake Rotovisko) for a non-Newtonian fluid (CMC 1%, n-0.504, 111-6.4195 Pa s“). The plots for the other non-Newtonian fluids are presented in Appendix B (Figures B1 and B2). From the known shear stress (or "a ) versus shear rate relationship, a relationship between the average shear rate, ’1 , and av rotational speed, N, can be established. The values of "a obtained from Eqn. (6.8) were used to determine the corresponding lav from plots such as Figure 6.6. The fiav thus determined are shown as a function of im- peller speed and geometry in Figures 6.7 through 6.9 for the paddle impellers (Figures B3 and B4 present the results for the other systems). These plots indicate that the shear rates vary with the geometry of the system as well with the properties of the fluid being agitated. Figure 6.7 shows the relationship between the average shear rate and the rota- tional speed of two impellers in the same container. The impeller in Figure 6.7a is 1 cm smaller (impeller blade height, b) than the impeller in Figure 6.7.b. Results indicate slighlty higher shear rates at specific values of rotational speed for the smaller impeller (Figure 6.7a). a [P0] 1,. [PO S] 85. 120 68$}- 51J3- 3411‘ 17H01 (10 ' ‘T CMC 1% Tc 25 C 1 52 T— r 78 104 7"[1/8] Figure 6.6: a) Shear Stress Versus Shear Rate; b) Apparent Viscosity Versus Shear Rate (Hydroxypropyl Methylcellulose 1%). 121 30 . r—" . I r I . 1 . r . r’ . r r 1 2 "'33:? System {I 3 , n- . 24.1 A ”-0.352 (d/b-O.6. d/D-0.327) _ J 1 75' a 1 _ \ 1 ‘2 5:. 1 x . t 12~ .1 '?N x e 6" X 0 -I " e a . J a 0 1 7T’ F’ T 1 ' l r' I r* r* tr r 30 . .. . , . . , . , - , . f. , . 1 z "'33:: System i 4 4 n- . 24.1 ‘ "-0.352 (d/b'OI45. d/D-00327) _, "a: Q ‘ \ 1s ,, - v- ) '1 5.; 0 B 1 - ‘?~ 2‘4 x 1 o , e 6-l _. l 1 1 b. O ’ F7 r r I , . . . - . 0.0 0.2 0.51 016 ofa 110 112 151 1.0 1.3 N [rev/s] Figure 6.7: Average Mixing Shear Rate as a Function of Rotational Speed of the Paddle Impellers. a) System 3; b) System 4 122 Similar results are presented in Figures 6.8 and 6.9. Figure 6.8b shows the values of the average shear rate for a small impeller in a medium cup. Figure 6.9b shows the results for a bigger impeller (three times) in the same cup. Figure 6.10 presents the results for the flag impeller (Figures 6.11 and B5 show results for other conditions). The experimental data indicate that the results are dependent on the flow properties as well as on the geometry of the mixing system. In com- parison with the paddle impellers, the flag impeller usually yields lower average shear rate values. This indicates that results are depend- ent on the shape of the impeller. Regression analysis was used to determine the relationship between lav and N. Results are presented in Table 6.3 for the paddle impellers and Table 6.4 for the flag impeller. It may be seen that :yav increases linearly with N for the paddle and flag impellers. Then, the average shear rate has the same form as the expression assumed by Metzner and Otto (1957) , lav - k' N (3.8) However, when agitating the less viscous fluid (n=0.504, m=6.4915 Pa sn) with a paddle impeller (Table 6.3), a better fit of the data is given by the model lav - a + k'N (systems 3, 6, and 7) (It should be emphasized that this is only a mathematical expression with no necessary physical meaning). Table 6.4 shows that experimental data for the flag impeller always follow the relationship given by Eqn. (3.8) . It seems that the values of k' are not as highly dependent on fluid properties and system geometry as for the paddle impellers (Table 6.3), where significant 123 30 ' T ' I T l r—‘I T’ T"T l . 1 tr r x ”-3.3: System {I 5 0 1'1- . 24.1 A "-0.352 (d/b-O.36, d/D-00327) - J "a? \ 18" x —I .1 , . J .71 12- _ , .. O 6" x . —1 . ; 1 l a 0 ”T I ' r' r I r r"T’ T7 ' T7 ' r’ I I 60 ' F r I ' I ' I ' I ”I I ' I T* 1 r " "'g'gg: System f 6 , 0 1'1- . 48-1 A "-0.352 (d/b'1.8. 0/0-0.515) x _, .1 "To" \ 36* x 2 .. L .. 3 24- .. .?~ ‘ I ’ '1 12-4 " 4 2 2 b .1 0 fi r r ' T r I r T” ' r— r r r r I— ' 0.0 0.2 0.4 0.6 0.8 1.0. 1.2 1.4 1.6 1.8 N [rev/s] Figure 6.8: Average Mixing Shear Rate as a Function of Rotational Speed of the Paddle Impellers. a) System 5; b) System 6 124 35 ' [ I r fi‘ r r 1 u 1 v f y 1 f 1 J x "-3.33: System 1? 7 O n- . 28.1 A "-0.352 (d/b-1.0, d/D-0.515) _ l ,, . F3? :21. . \\7 6 r- J 4 LII—J I 14- 5 J 0% 4 X ‘ X 7- i .- 4 ° . a 0 r* r r*rut r’ r* I tr r ' I r"r r’ r r’ 50 ' I ' I I r ' I ' I ' I T’ I . x "'33:: System I 8 ' , 0 w- . J .1 71" 0 \ 3 ‘ ‘ CZ: . e . 1: 204 x .. d ‘ ‘ 1 104 O x ° ‘ e ul 8 . 1 ° b 0 e r s , . . e , . , . 0.0 0.2 0.4 016 013'110'112 1.4 1.5 1.0 N [rev/s] Figure 6.9: Average Mixing Shear Rate as a Function of Rotational Speed of the Paddle Impellers. a) System 7; b) System 8 125 20 a; 3/0'0573' . ' - ‘ ' I - ' . I 4 " - n-o.374. -2a.41 o d/D-O.429 m 6 16-4 a Ive-0.591 ‘ 4 7 12.4 \ 3 " E. 1 . B 8 oh ‘4 9 d d . 1 H g 1 < ‘ 4 a O ’f’ I I ' I ' I r I . I r’ I rWI r 20 ' I ' I ' I ' I ' I ' I ' ' ' x d/O-O.273 ' r J n-0.504. I .4 1 o d/D-O.429 m 6 9 5 J 164 A Iva-0.59: * O '7." " ' \ 12? ‘ .. 5.. ‘ I ‘ d .5 3'4 a l g . . 44 d 4 i q o D f I ' l f I f I ' l ' l r l ' r ' ’ 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 N [rev/s] Figure 6.10: Average Mixing Shear Rate as a Function of Rotational Speed of the Flag Impeller. a) CMC 1.5%; b) CMC 1% 126 20 T”I 4* I . I ' r r r‘ir’ I ' I I x d/D-O.273 “0,352, m-59.275 o d/0-0.429 ‘ 161 a Iva-0.591 ‘ J J "a? 12 - \ l 3 J z. I ” ‘ I 8-4 " ‘?~ . I 4-4 ‘ 4 Q 3 ‘ A a 0 v r ' r ' F ' I r I T I f r r F Y 2G I I V I ' T r ' I r I Y I T t ‘ x III-0.504 d /D-O.591 . o n-0.374 16-« I "0.352 ‘ F—I ‘ ‘ m \ 124 2 ‘ 2. ‘ ‘ I 8-4 ‘ ‘5 4 l ‘ ‘4 x J .. 1' I g b4 o r I r I r r I f F f ' r‘ r r I r 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 N [rev/s] Figure 6.11: Average Mixing Shear Rate as a Function of Rotational Speed of the Flag Impeller. a) CMC 2%; b) System 3 127 Table 6.3: Values of k', defined by Eqn. (3.8) (k' - fiav/N ), for Paddle Impellers. ' 2 FLUID SYSTEM 33 R 1 33.527 0.994 2 15.890 0.986 3 9.535 0.938* 4 10.855 0.979 CMC 1% 5 11.740 0.941 6 32.309 0.925** 7 16.154 0.899*** 8 23.647 0.996 9 25.372 0.995 10 20.424 0.993 1 24.375 0.996 2 13.568 0.988 3 10.305 0.977 4 9.051 0.964 CMC 1.5% 5 8.712 0.983 6 22.564 0.981 7 12.784 0.981 8 13.859 0.998 9 21.062 0.998 10 19.511 0.989 1 17.019 0.986 2 10.776 0.976 3 9.948 0.995 4 8.834 0.991 CMC 2% 5 7.395 0.978 6 20.999 0.979 7 12.215 0.998 8 9.417 0.981 9 24.199 0.989 10 13.937 0.998 2 * i - 3.5072 + 6.6217N, R -0.973 av 2 ** fiav - 13.695 + 20.9313N, R -0.966 2 *** fiav - 8.3535 + 9.2137N, R - 0.999 128 Table 6.4: Values of k', defined by Eqn. (3.8) (k' - yav /N) for Flag Impeller. 2 FLUID SYSTEM, k' R 1 8.602 0.969 CMC 1% 2 9.389 0.997 3 7.704 0.974 1 7.535 0.985 CMC 1.5% 2 7.639 0.997 3 6.952 0.994 1 6.411 0.993 CMC 2% 2 6.919 0.992 3 7.087 1.000 129 differences are obtained for each fluid. 6.1.1.1.2) actor ec in ver e ear Rates and k' 6.1.1.1.2.l) Impeller Rotational fineeg The relationship between the average shear rate and the impeller speed has been shown in Tables 6.3 and 6.4 for the paddles and flag impeller, respectively. The slope from the plots gives a value of k', independent of the value of N. However, if individual values of k' were obtained at specific values of N, a relationship between k' and N could be established. 6.1.1.1.2.2) Fluid Properties Previous research has shown a possible dependence of k' (therefore, average shear rate) on the flow behavior index, n, widn pseudoplastic fluids. However, this dependence has been found to be highly dependent on system geometry and impeller shape (Calderbank and Moo-Young, 1961; Beckner and Smith, 1966). Their results indicate a possible decrease in the value of k' with an increase in the value of n but results were not conclusive and this needs further study. It is interesting to find that, in this investigation, the less shear-thinning fluid (i.e. , larger value of n) generated higher values of k' (higher average shear rates) than that produced.by the highly shear-thinning fluid (lower n value) when agitating the fluid with the paddle impellers (Figures 6.7 to 6.9). A similar behavior was observed by Sinevic et a1. (1986) when agitating power-law fluids in a concentric cylinders system. It is reasonable to suggest that k3 is not only a function of the flow behavior index, n, but it also varies with the 130 value of the fluid consistency coefficient, m. In this investigation, the less shear-thinning fluid (CMC 1%) was also the less viscous fluid. Also, a decrease in the gap between impeller and cup (d/D) reduces the effect of the properties of the fluid. With an impeller to cup diameter ratio of 0.709 or greater, k' was not significantly affected by n (Table 6.3, Systems #9 and #10). Shear rates in a cylindrical container agitated by a flag impeller are not as significantly affected by the rheological properties of the fluid as in the case of the paddle impellers (Figures 6.10 and 6.11), with a slight variation at high rotational speeds for the less viscous fluid. 6.1.1.1.2.3) Cup Diameter The diameter of the cylindrical cup where the fhxhiis being agitated does not have a significant effect on the value of k' when using paddle impellers. Figures 6.7a and 6.9b show results for the same impeller (b-3 cm; d/b-0.6) rotating in different cups. Impeller-to-cup ratios significantly less than 1.0 (d/D-0.327; Figure 6.7a), yielded lower shear rates values (at high rotational speeds) than those deter- mined when using a larger impeller-to-cup diameter ratio, i.e. a smaller gap (d/D-0.515) (Figure 6.9b). The same behavior was observed with all the paddle impellers. However, Table 6.3 suggests that d/D may not have a significant effect on the values of k'. This was confirmed by results using multiple regression analysis. As was expected, in the case of the flag impeller, the fluid container has no significant effect (Figures 6.10a to 6.11). However, it 131 is interesting to note that the more viscous fluid (also the more shear- thinning fluid), (Figure 6.lla) generated higher values of :Yav when increasing the d/D ratio (smaller gap), a trend not exactly followed by the less viscous fluids. It is suspected that the effect of the geometric term (d/D) is highly dependent on the type of fluid being agitated. 6.1.1.1.2.4) W Values of the average shear rate determined for the five paddle impellers are shown in Figure 6.12 for two non-Newtonian fluids (CMC 1% and CMC 2%). The difference in values becomes important at high mixing speeds, with the smaller impeller (d/b-l.8) generating the higher average shear rates, especially when agitating a low viscosity fluid (Figure 6.12b). This is due to the fact that a small impeller produces low flow and thus, higher shear rates. The difference is smaller when agitating a high viscosity fluid (Figure 6.12a). In both cases, the difference in average shear rates generated by the different impellers is small at low mixing speeds. It is interesting to note that the k' values determined with the less viscous fluid (CMC 1%) show some incon- sistency, i.e., they do not correlate with the height of the impeller blade. This phenomenon was not observed with the more viscous fluids. Again, the type of fluid being agitated seems to be an important factor in the determination of the proportionality constant, k' (Figure 6.13). Thus, the correlation between k' and the height of the impeller blade was determined for each fluid. The best fit was obtained with the fol- lowing model , 132 k' - fl. (cl/m”? + p. (6.9) with 31 , 62 and B; as regression constants. Results from regression are shown in Table B3 (Appendix B). In summary, results obtained with the method developed by Metzner and Otto (1957) indicate that the value of the mixer proportionality constant, k', is not only a function of the geometry of the system as the authors suggested but is is also highly dependent on the properties of the fluid under consideration. 6.1.1.2) Mixer Torque Qurves Merhgg (Mackey et al., 1987) Following the procedure of matching Newtonian and non-Newtonian data employed by Mackey et al. (1987), the mixer torque, M, versus nN plots for the Newtonian fluids are shown in Figure 6.14 for all treat- ments with the paddle impellers. As it was expected, linear results were obtained. The slope of the lines, l/k2 (rev/m3), with k2 the mixer coefficient, seemed to be a function of the geometric variables d/b and d/D. Thus, the following model for the mixer coefficient, k2, was proposed k2 - a0 (d/b) 0. (d/D)°2 (6.10) After linearization of Eqn. (6.10), multiple linear regression analysis (Table 6.5) indicates an excellent fit of the data. The prediction equation for the mixer coefficient, k2, is given by 0 684 .0 408 k, - 6121.262 (d/b) ' (d/D) ' (6.11) 133 35 I I . I I I I I II T I I III I I r . " d/b-1.8 d/D-0.32 o d/b-1.0 284 o d/b-0.6 n-o'352 x " 4 A d/b-0.45 , ,_., v d/bi-O.36 In .. _ \ 21 I: ‘ I B 14- " c - OF 4 O ' 4 7-4 x x 6 « O s . ' a 0 I T’ I II I I I I’III I II‘T— II’I I II II SC) I I I I I I I I”“II’I I I I I I I I . " djb-LB d/D-0.327 I q o d b-1.0 48-4 o d/b-0.5 "'0'504 - , A d/b-0.45 . I-I v (Vb-0.36 In. . .. \ 35 :2 4 < t 244 " ° 1 oh . o I ‘ X ' a 12- . ~ X 1 ' ' b1 0 ”I ‘I I I I r I FI’F’ I I I I II I II I 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 N [rev/s] Figure 6.12: Average Mixing Shear Rate as a Function of Rotational Speed of the Paddle Impellers (d/D-0.327). a) CMC 2% b) CMC 1% 134 35 e,. k'=3.(d/b)fl.+p. 29-4 w [1/rev] Figure 6.13: Plot of Mixer Proportionality Constant as a Function of Impeller Size for the Non-Newtonian Fluids (d/D-0.327) Table 6.! Regressi Coeffici log (10 “I “2 N- \ Regressic Residual \ Total \ 0.0.0 135 Table 6.5: Regression results for the fit of Eqn. (6.10) Linear Multiple Regression Analysis Regression Estimated Regression Estimated Standard t* Coefficient Coefficient Error log a0 6121.262 -- -- a, 0.684 0.066 16.75 a2 -0.403 0.066 -5.51 Analysis of variance * Sum of Degrees of Error Mean F Squares Freedom Squares Regression 0.222 2 0.111 -- Residual 0.005 7 0.028 143.51 Total 0.227 10 2 R - 0.980 a - 0.05 136 2 R - 0.980 Figure 6.15 presents the predicted values of k2 using Eqn. (6.11) ‘versus the values of k2 determined from the slopes of the lines for the paddle impellers (Figure 6.14). The adequacy of the prediction equatiorI is indicated by the regression coefficient and the slope of 0.990. Figure 6.16 presents the plots of torque (M) versus viscosity times rotational speed (nN) to determine the values of the mixer coeffi- cient, kg, for the flag impeller rotating in the different cups. It is clear that the value of k2 is also a function of the geometry of the system for the flag impeller. The value of the mixer proportionality constant, k' , can then be calculated from the following equation [Section 3.2.1.1, Eqn. (3.15)], “’1 (6.12) with k2 a function of system geometry (Tables 6.6 and 6.7). Finally, the average shear rates are determined from 1 - k'N (3.8) The effect of the parameters in Eqn. (6.12) on the value of k' and, consequently, on the estimation of average mixing shear rates will be discussed in the following section. 137 00m“: . . . r .7 , . IMMKO . 1 . . , . j . J —-¢Hrlmd wN-Ll . J -" J mm-Lo I’D-om: axe-um 0.00404 “ ‘-’°"°-7°' .1 0.00404 -- a/o-cm .. o—w '1 i .4 E 2: dance a H U «I 1 3 8' 0.00204 /,- _ o x , .- 1 ''''''' . OIm‘O" """"" _ mane o m f I t v t -—»4hka§§} ' a ‘ . 1-«»dflh‘aul "'“’ . manna - H d ‘ ‘3 z: manna ‘ _ H 'o o 4 d 3- 0.00204 - L O I- i 0.00124 _ fiI ff I I I I 0 5 10 15 20 25 nI-N [Pctrev] nvN [Potrev] Figure 6.14: Mixer Torque Versus Viscosity Times Speed for Calculating Mixer Coefficient, k, (Paddle Impellers) 2E+04 IE+044 9200- 6300-4 kz (slope) 3400-4 500 R3 = 0.98 1i 138 T fl I r I 0 Paddle Impellers 500 251-04 ‘1 r 3400 r r r 6:500 9230 k: (predicted) 15:04 7 29-04 lE+04~ 9200-4 62100--l k, [rev/m5] 3400- 500 '_ r Paddle Impellers U I I l U I r r # C3 157 25 (”WWW/mm Figure 6.15: Predicted Mixer Coefficient Versus Observed Mixer Coefficient (Paddle Impellers) 2.5 Torque [NIm I 139 0.00164 d cl 0.001211 d 0.00084 if I I . I I I I I I I I I I 'T 0 d/D-0.273; Its-17443 . V d/D-0.429. Eta-15843 .. + d,/D=-0.591. Its-12488 .. l l. ,/ ’ . fiIIfIr-Irrrr.,..,..,rr,..,.. 3 6 9 12 15 18 21 24 27 30 ntN [Pot-rev] Mixer Coefficient, k2 (Flag Impeller) Figure 6.16: Mixer Torque Versus Viscosity Times Speed for Calculating 140 Table 6.6: Values of k', defined by Eqn. (6.12), for Paddle Impellers kl FLUID SYSTEM 5 rung, 20 rngi 100 rpm 1 48.594 15.176 8.206 2 36.983 14.329 8.685 3 32.451 12.772 7.823 4 26.583 11.296 6.808 CMC 1% 5 23.737 10.191 6.277 6 61.948 23.409 14.274 7 50.159 19.975 12.503 8 37.851 16.298 10.130 9 39.847 15.562 10.882 10 38.515 16.101 10.8494 1 26.094 12.872 8.325 2 27.485 13.557 8.647 3 23.803 11.939 7.636 4 21.240 10.528 6.139 CMC 1.5% 5 19.219 9.525 5.742 6 25.729 13.895 9.972 7 26.456 13.807 9.354 8 15.568 8.993 6.559 9 21.019 11.468 8.692 10 16.806 9.699 7.301 1 21.400 11.537 8.024 2 22.656 11.701 7.891 3 20.392 10.602 7.099 4 18.347 9.551 6.529 CMC 2% 5 16.347 8.683 5.665 6 22.572 12.537 9.079 7 34.282 16.127 9.605 8 18.385 10.136 7.159 9 10.903 8.914 6.099 10 18.267 10.825 8.361 141 Table 6.7: Values of k', defined by Eqn. (6.12), for Flag Impeller kl UID 00 m 1 39.852 11.246 6.402 CMC 1% 2 38.589 14.038 8.458 3 47.601 16.707 9.811 1 22.475 10.756 7.111 CMC 1.5% 2 19.045 9.846 6.625 3 21.788 11.445 7.833 1 18.135 9.558 6.801 CMC 2% 2 18.184 9.727 6.866 3 19.105 9.914 7.408 142 6.1.1.2.1) Factors Affecting k' 6.1.1.2.1.1) 1m elle otati na Equation (6.12) predicts a dependence of the proportionality constant, k', a o d/D-0.327. d/b-0.45 ‘ Z 440- e d/D-0.327. d/b-0.36 .1 75. T 0 d/D-QSIS. d/b-1.8 4 2 I d 0-0.70 . d -1, 8 4 ‘g’ 220-" - .2 E I“ j 5 110-b- i E ‘ :i 1 'i 3 0 O I I I I I I I YLI I I I I F I Wt'fi' _ 550 ’ I I I r v I I I I I I I I r I I I E £ 4' d/D-O.327. d/b-I.O "-0.504 ) o d/D-O.327. d/b-O.6 z 440.. e d/O-O.515. d/b-1.0 - E.- . o d/D-O.515. cub-0.5 , 3 I d Oil-0.709. d b-1.0 g 330-» / / " u z~ ‘ < 5 220-1: . E .. 8' a: 1 10“ cl : .3 . '5 I g . 0 . I I I I I—I I I I I I F fifih— 0.0 0.2 0.4 0.5 0.8 1.0 1.2 1.4 1.5 1.8 N [rev/s] Figure 6.17: Mixer Proportionality Constant, k', Versus Rotational Speed, N, for Paddle Impellers When Mixing a Non-Newtonian Fluid (CMC 1%) 144 — 210 I . ‘ T I ' ' F I E II d/D - 0.273 > o d/D - 0,429 Flag Impeller. n30.504 :‘ 169- o d/D - 0.591 1 as: 4 1 J S 128 .. o i . § 874 , _ i 46- ° - 0 2 3 . '5 I 5 F I’ I1 I I I IFl I a E 200 . Id/D I0273I I I I I I - " - FI - . > 131 o d/D .. 0.429 09 Impeller. n 0 352 .. t: 161- e d/D - 0.591 .. 1424 q g... . 103-JP d 64-. a 444' .- 3 25+ I. . O s . ' . TL -—-I—‘l- 0.0 0.2 0.4 015 0.3 110 1:2 1:4 1TB 1.3 2.0 N [rev/s] Figure 6.18: Mixer Proportionality Constant, k', Versus Rotational Speed, N , for Flag Impeller When Mixing Non-Newtonian Fluids. a) CMC 1%; b) CMC 2% 145 " 350 j I 1 I f r I Y I F 1 v 1 r 1 Y E .1 d/b-1.0.d/D-O.327 + “0,504 > O n20.374 2288-1 I n-0.352 E J: g 2‘15-I U 3 4 E 144-I O "E 3‘ a: 72 3' «8g 3 . . . 0 r , . r . I . I , I . r . I . I o, ~210 ' I l I I I’ I V I i I ' I t I r E .x d/b-1.o.a/o-o.515 x ”0,504 > O n-0.374 3168'“ o n-o.:552 - 45: 4 g 125-ID o 0 =2; x 3 § 84 r a X C d: ‘42-I8 E J 8 )O‘ 5. i 3 ‘ o . , . - . r r I r 1' I r IF F I I I 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 N [rev/s] Figure 6.19: Plot of Mixer Proportionality Constant, k' , Versus Rotational Speed, N, for Non-Newtonian Fluids (Paddle Impellers) 146 ~12 ' I g I d/D-0.273 { I A d/D-O.429 rpm-100 :2, 0 d D-O.591 . 10~ / .x ,.- 1% J '2 A 2 8-1 . . ° 0 (J J I i? ll ‘ ‘ 2 64 '. .2 't J o a. s .. .. b O «I .5 I I! 2 ' I 0-35 0.43 0.51 n. Flow Behavior Index Figure 6.20: Plot of Mixer Proportionality Constant, k', as a Function of Flow Behavior Index, n (Flag Impeller) 147 12 k . '3' - d/0-0.273 ‘ i . d/0-0.429 “Pm' ‘ 0° « .2. o d/D-O.591 . 10 _ 3 1g . . .2 I. g 8-1 , - 0 I u :1 5 I ‘ g 6-1 - .2 I: . , 8. s u . 3 .5 ‘ I a 2 . i f 6.00 33.00 60.00 m, Consistency Coefficient [Potsn] Figure 6.21: Plot of Mixer Proportionality Constant, k' , as a Function of Fluid Consistency Coefficient, 01 (Flag Impeller) 148 the flow behavior index. The lower the n, the lower the value of k'. However, no conclusions can be made by analyzing only the effect of the flow behavior index. Figure 6.21 shows that the more viscous fluid (m- 59.275 Pa 3“) yields the lower values of k' (consequently, average shear rates). Figures 6.22 and 6.23 show similar plots for a paddle impeller (d/b-l.8) at 20 and 100 rpm (0.333 and 1.6667 rev/s). The effect of the agitated fluid on the values of k' was the same as for the flag impeller. No final conclusions can be drawn up to this point since variations occur when the impellers operate in different geometric systems and this needs further investigation. The effect of system geometry will be discussed in the following section. 6.1.1.2.1.3) Cup Diameter Results indicate that the value of the proportionality constant, k', for the paddle impellers, is a function of the size of the cup (Figures 6.22 and 6.23). The bigger cup (i.e., big impeller-to-cup- diameter ratio, d/D-0.327) generates lower values of k' than the produced in the smaller cups when rotating at high speeds. At lower speeds (20 rpm) differences in the values of k' are only significant when agitating a low viscosity fluid (CMC 1%) (Figure 6.22). Figure 6.24b presents the changes in k' for a paddle impeller (d/b-l.0) rotating in three different cups at 2.5, 10 and 100 rpm. It may be seen that the diameter of the fluid container is an important parameter to consider for the design of a mixing system for mixer viscometry tech- niques when working at low rotational speeds. At higher speeds, the influence of cup diameter becomes almost negligible when agitating highly viscous fluids (II-0.352, Ill-59.275 Pa s“). 149 _ 20 . I . {I 0 dz’9’0-327 rpm-100, 0,’b=1.8 , > I 0/0-0.515 3‘ 17. v c/D-0.709 - ‘5 1 ‘ ‘g 14- '- 5 . 2? TE: 1 1 - v - g .. ' . g a ‘ C -1 g . .5 5 I o a _ 30 . I . E d C d/D-O.327 rpm-20' d/b-‘I 8 ‘ > I d/0-0.515 T” 25- v d/D-0.709 - If I 3 g 20- - a ‘ ‘ 3 .. I .. 5 15 . a . I g. . a: 10 ‘ g . '5 b 5 , I . 0.35 0.43 0.51 n, Flow Behavior Index Figure 6.22: Plot of Mixer Proportionality Constant, k', as a Function of Flow Behavior Index, n (Paddle Impeller); a) 100 rpm; b) 20 rpm 150 26 . duo-0.323 m-100 drb-1.8 - d/D-O.515 rp ' / 22-1 v d / D-0.709 '1 18-1 4 141-1. .1 104' I - I W fixer Propenlenellty Constant. k' [I/rev] A i L 30 O d/D-O.327 ("-20 d 1-1.8 - d/D-0.515 rp ' / 25 v Iva-0.700 - We! Proportionality con-com. III DIM] 6 III I 10-1 v 5 b r r . l 6.00 33.00 60.00 rn, Consistency Coefficient [Po n"] Figure 6.23: Plot of Mixer Proportionality Constant, k', as a Function of Fluid Consistency Coefficient, m (Paddle Impeller); a) 100 rpm; b) 20 rpm 151 Results presented in Figures 6.20 and 6.21 indicate that cup diameter becomes important only when agitating a low viscosity fluid, with changes in the values of k' ranging from 9 to 6.5 (l/rev), as compared to the values of k' of 7 to 7.5 (l/rev) for a high viscosity fluid. It is interesting to find that the value of k' is higher for the :more shearwthinning fluid (n-O.352) when the flag impeller rotates in a bigger cup (d/D-O.273) than the value of the less shear-thinning fluid (n-O.504) (Figure 6.20). The trend is reversed for smaller impeller to cup diameter ratios. These results confitm the need for a better under- standing of the influence of fluid properties on k'. 6.1.1.2.1.4) Impeller Size (Blade Height) Figure 6.24a shows the changes of the values of k' as a function of impeller size (d/b). It is clear that k' is significantly affected by the height of the impeller's blade at low impeller speeds (2.5 rpm), especially when agitating the low viscosity fluid (n-O.504). As N in- creases, the effect of (d/b) becomes less important. The values of k' for the different paddle impellers rotating in a big cup are shown in Figure 6.25. Results agree with the previous observation that k' is not significantly dependent on impeller size when rotating at high speeds, independently of the rheology of the agitated fluid. At a rotational speed of 20 rpm, k' values varied from 15.2 to 10.2 (l/rev) when agitat- ing a low viscosity fluid as compared to k' values from 11.5 to 8.7 (l/rev) for a higher viscosity fluid when using big and small paddles. It may also be noted that the bigger impeller generated the lower k' values (i.e., shear rates). The same behavior was observed when using 152 __ 20C) . , . T , . , Iv g — ”=0-504 d/D-0.515 , > ---- n-O.374 "_"" 160~ --- n-O.352 - .K g . g 12CL- - U ‘3 . ‘g‘ 80- - .2 ‘t' rpm-2.5 o T 8 i 40‘ 7; """"""""" ..-_ """" .2 3;: ‘ E 1° rt" .................. ....... . 5 100 :22: """ a o a a . , . , . , . DJ) 0.4 CLB 1.2 1.6 2n0 ti/b __ 125 . r I . r r, . é d/b-1.0 —- n-O.504 .. } .... n-0.374 1" 100- --- n-o.352 - J 2: c1 g -75_, _ u rpm-2.5 . '§ '2 50- - £2 ,_..._...-...-.... . i ‘ w- _;:l;:. ‘ 5 25d ~¢KR~u ‘ , w { -..........._ .5 4 100 M ‘ d a b 0 r r7 r I . , r. I r 0J3 CL2 0:4 0.6 CLB 1.0 Figure 6.24: Plot of Mixer Proportionality Constant, k', as a Function of System Geometry. a) Impeller Size (d/b); b) Cup Size (d/D) 153 14 fii 1 - (Vb-1.8 _ ,3 d/b-1.0 rpm 100, d/D 0.327 d/b-O.5 1 d/b-O.45 + d/b-0.36 d 12% i 10--1 IO‘I. Hirer Proportionality Constont, k' [l/rev] a: d on $4.: 6 g 21 4 e . . - a __ 25 . I T E . 0 d/b'i-B rpm=20. d/D=0.327 J > I d/b-1.0 ‘7‘ 21 4 v d/b-0.6 " 8 .: 4 o d/b-0.45 § 1. d/b-0.36 g 17- - 8 E g 14 : - r v 8 . v E 10 a: 3 § § 6 , - , b 6.00 33.00 60.00 m, Consistency Coefficient [Po 3"] Figure 6.25: Plot of Mixer Proportionality Constant, k' , as a Function of Fluid Consistency Coefficient, m (Pa s“) for Different Paddle Impellers. a) 100 rpm; b) 20 rpm. 154 the Power Curves Method. 6.1.1.2.2) Estimation 9: Average Shggz gates Using the values of the mixer proportionality constant k' deter- mined from Eqn. (6.12), the mixing average shear rates were calculated from Eqn. (3.8). Shear rates were also calculated using a constant value of k' at selected rotational speeds of 10 and 50 rpm in order to deterw mine the significance of using a constant value of k' when using mixer viscometry techniques. Results are presented in Figures 6.26 and 6.27 for the paddle impellers (Figure 86 presents other results). It is clear that the choice of a constant value of the proportionality constant at 10 rpm generates considerably higher average shear rates than when using the values of k' at 50 rpm and Eqn. (6.12). For the less viscous fluid” the larger impeller-to-cup diameter ratio (d/D-0.327) (Figure 6.26b) produces higher shear rates than the smaller gap (Figure 6.26a). Also, Eqn. (6.12) seems to produce unstable results at low values of rota— tional speeds. Similar behavior was observed when agitating a high viscosity fluid (Figure 6.27b). However, interesting results are observed in the small gap situation (Figure 6.27a). In this case, the value of 1" does not significantly lead to considerable differences in the determined shear rates. In all cases, Eqn. (6.12) presents unstability at low rotational speeds. For the flag impeller, the significant differences in shear rates when using Eqn. (6.12) and a constant value of k' (at 10 rpm) are shown in Figures 6.28 and 6.29. Variations are bigger when agitating a less viscous fluid (Figure 6.28) than for the high viscosity fluid (Figure 155 50 r r. 1 1 W T r I I {fir I F r fi r ' — k 83-) ) n-o.504, d/0-0.709 .... ' rpm 40- __ kn (10 rpm) d/b-1.8 H m \ C. I -z~ :«v a O‘Rr I ff f 1' r I f r r r r [ r r 1“— 55 ' n ' I ' * r r I ' r r r V r ' — : Egg.) ) n-0.504, d/0-0.32 .... ' rpm 44.- __ k. (‘0 rpm) d/b-1.8 7: ‘ \ 33' ’o"’ .2. - s 22- x" 0% ”” a "’ " ”.flo-‘MT: 114 "” .00. . "' b O_F1=:,p:*”:flfi; r* I rhr r— r r r*’* r *T Ti ' 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 N [rev/s] Figure 6.26: Average Shear Rate as a Function of Impeller Rotational Speed for a Paddle Impeller (CMC 1%); a) Small Cap; b) Large Cap I 25 156 I If I I I l I I r I T [ r I I I I J - 1* ("-1 n-0.352, d/D-O.709 w (50 rpm) '3? \ .2. B -z~ 0-? I I r r r I V r I r f I I r r I f 30 II I. Y, I ‘I I r ' ‘I r' I I I I r' - 1; Egg-i ) «0.352. d/0-0.32 m. ' rpm 24... -- k' (10 rpm) d/b-1'B x” H m \ :2. I -a- Figure 6.27: Average Shear Rate as a Function of Impeller Rotational for a Paddle Impeller (CMC 2%); a) Small Cap; b) Large Cap I 157 45 . _ I- (are)I I I ' T F I I I . I ' I ' k. (50 mm) n-o.so4. d/D-0.273 35" -- k' (10 rpm) (I) l” 4 ”. \ 27 ’I” hi J ”’ ’I ’I a 6 18.1 ”’ O ”’ J ””’ O. ’ OOOOOO ” wi/ 94 ”” .......... ”’ ‘ I’ ....... 04 0:0... 0000000000 l I' I'— I Y I I I r I I' r I I T— I— T Y I r I Y r f l V r ‘ v I T —v— I V J — k' («1.) n-0.504, d/D-0.591 «- k' (50 rpm) 44-1 -- k' (10 rpm) in [1/81 ’ M ’ a... 04'”. f I 1* v r r Y I r 0.0 0.2 0.4 0:6 ofa 1:0 13 1:4 1? N [rev/s] Figure 6.28: Average Shear Rate as a Function of Impeller Rotational Speed for a Flag Impeller (CMC 1%); a) Large Cap; b) Small Cap [11/81 ‘7. [1l/s] 1.. Figure 6.29: Average Shear Rate as a Function of Impeller Rotational 158 45 —T kl. (;q )I l' I V I I I 1 I I I r I . --.. k. (50 rpm) n-0.352, d/D-0.273 36-1 -- k' (10 rpm) 1 27.. 184 ,,-""” 04 r r T I T T T— r T r U j Tf r fi . I 45 T {(1 )‘I— 1 I v F v I f I v T v ‘ — ' '4- -o. 2. d -o.591 4 .... k‘ (50 rpm) n 35 /D 36- -- k' (10 rpm) J I 27- ,.”” 18- I”’ Speed for a Flag Impeller (CMC 2%); a) Large Cap; b) Small Cap 159 I. Y I I I I r x u («.3 "0,352, d/D-O.709 o k' (50 rpm) 272 o k' (10 rpm) (Vb-1'8 ‘ 9 204 ‘ o J 0. H o 136‘J . ? +0 68-4 S ‘ J 0‘. . ‘3 no 0 ' F r I 3‘ l ' r a 325 r . I f l ' I r f 1 x k (al.) n-0.352. d/D-O.327 o k' (50 rpm) 260 o k' (10 rpm) d/b-1'8 - H d 1' 195 ‘ O m ‘ ‘ H . 1:50Ja ‘ F 49 65-1 3‘ “ 4 a. x a 0 U tax o a b r f ' r ‘ F ' f—d 0.0 6.0 12.0 18.0 24.0 30.0 7~¢[1/s] Figure 6.30: Apparent Viscosity as a Function of Average Shear Rate for A Paddle Impeller (CMC 2%); a) Small Cap; b) Large Cap 160 6.29). Figure 6.30 shows the apparent viscosity-shear rate relationships when agitating the high viscosity fluid. The use of a constant k' value overestimates the values of "a while the use of Eqn. (6.12) predicts lower values of "a for the shear-thinning fluid. Figure 6.31 shows the results for a flag impeller (Other results are shown in Figures B7 and B8). In summary, results indicate that the proportionality constant k' evaluated by the Mixer Torque Curves Method is significantly affected by impeller rotational speed, fluid properties and mixing system. The conventional mixer viscometry method of using a constant value of k’ , depending only on impeller geometry, may lead to significant errors when measuring properties of shear-thinning fluids at low values of N. 6.1.2) Slope Method The Slope Method consists of the construction of plots of log [P/(mdsMn£l)] versus (l-n) [Eqn. (3.47)]. The value of the impeller proportionality constant, k', is obtained from the slope of the straight line (k' - 10$10133 ). Figure 6.32 shows a typical plot of log [P/(mdaNn+1)] versus (l-n) for the paddle impellers. Figure 6.32a is the plot for a paddle impeller with blades 3 cm high. Figure 6.32b is the plot for a paddle impeller with blade heigth equal to 1 cm. Also shown in Figure 6.32 is the line obtained by linear regression analysis (R2 = 0.980 and 0.983, respectively) of the experimental data. Similar results were obtained with the flag impeller (Figure 6.33) rotating in a cup of diameter (D) equal to 3.5 cm (d/D-O.591) (Results for the other impeller/cup combinations are presented in Figures B9 through 813). n. [Pats] n. [Pats] 161 370 . 1 ' I I I x w («.3 ”-0352, d/D-0.273 o k' (50 rpm) 296 o k' (10 rpm) " 222 ' 148 ' 0 d h ‘ x D a x o a Ja_J 0 fi I I ' I f I 350 r r I F I ‘— I x k' (04-) n-O.352. d/D-0.591 . o k' (50 rpm) 280 n k' (10 rpm) " 210 q 140 d O '1 70-4 1‘ ‘ J .0 ‘ O x " I a x n D f r T I I 0.0 6.0 12.0 18.0 24.0 Figure 6.31: Apparent Viscosity Versus Average Shear Rate for a Flag Impeller (CMC 2%); a) Large Cap; b) Small Cap 162 1.0E+04 ‘ v I ' T I l k‘ = 9.88 (1/rev) R3 = 0.980 _ 1000.01 32 j n' 2 100.01"\\'\ ,E, E \\ + Q- '1 10.0 1 0 — d/D-0.327. d/b-0.36 a , . I F 1* r l r [ ~1 LOE+04 . gr . I r I k' a 16.35 (1/rev) R2 ='- 0.983 _ 1000.0 1 2 no '2 100.0 .5. \\ O. 10.0 1.0 a . f . r r . f . 1 0.0 0.1 0.3 0.4 0.5 0.7 (1-n) 3 Figure 6.32: Plots of Dimensionless Functions [P/(md Nn+1)] Versus (l-n) For The Paddle Impellers; a) Impeller 5 (b-S cm); b) Impeller l (b-l cm) 163 LOE+04- , . r . 1 , k' = 16.04 (1/rev) R3 = 0.965 1000.0 10.0 1.0 ' -'- g/D-(Jr.591 0.0 0.1 r r I I v I v 0.3 0.4 0.6 0 7 (1 -n) . d a Figure 6.33: Plots of Dimensionless Functions [P/(md Nn+1)] Versus (l-n) For The Flag Impeller (Small Cap) 164 The scatter in the values of the function [P/(mdaNn+1 )] observed in Figures 6.32 and 6.33 is not uncommon, and it can be found for im- pellers studied by others researchers (Rieger and Novak, 1973; Rao and Cooley, 1984). Table 6.8 presents the values of the proportionality constant k' (evaluated using the Slope method) for the different geometries when agitating the fluids with the paddles and flag impeller. Considerable differences in the values of k' are observed for paddle impellers of different size (Systems 1 to 5). The effect of the geometric variables will be discussed in the following section. 6.1.2.1) Factors Affecting k' 6.1.2.1.1) System Geometry The geometry of the system is an important factor in the evalua- tion of the proportionality constant (Table 6.8). As in the matching method (Power Curves Method), the bigger impeller produces the lower values of k' , and an increase in the values of k' is observed in the smaller cups (small gap). For the flag impeller, the smaller gap produced smaller values of k'. Thus, impeller shape seems to be a factor since different results are obtained for the two impeller types. Deviations from linearity in the plots of Eqn. (3.47) (Figures 6.32 and 6.33) were not observed for the impellers investigated in this study. This suggests that the value of the impeller proportionality constant is not a function of the properties of the fluid when using this mixer viscometry method. It was suspected that the slope method may mask the effect of the impeller rotational speed on the value of k' . Since some researchers have sometimes used a single value of the impeller rotational speed for 165 Table 6.8: Values of k' (evaluated using the Slope Method) Paddle Impellers SYSTEM k' 3 d/D, d/b 1 0.327 1.80 16.354 0.983 2 0.327 1.00 12.576 0.985 3 0.327 0.60 10.407 0.981 4 0.327 0.45 10.000 0.960 5 0.327 0.36 9.880 0.980 6 0.515 1.80 16.89 0.960 7 0.515 1.00 15.99 0.960 8 0.515 0.60 10.306 0.904 9 0.709 1.80 17.730 0.972 10 0.709 1.00 9.827 0.920 Flag Impeller 2 SYSTEM k' R 1 0.273 20.070 0.980 2 0.429 15.090 0.970 3 0.591 16.04 0.965 166 1.0E+05 . I . r a I d/D=0.327. d/b=0.6 "g __ 1.0E+04 3 E o - . r a; ”13 1000.0 g. A 100 rpm 0. + 50 100.0 0 20 n 10 o 5 I 2.5 10.0 I . r . , . r . T . aj 1.0E+05 v 1 I I ' I ‘— I r 7 d/D=0.327. d/b-1.8 ,_ 1.0E+04 g . z n ”13 1000.0 g Q‘ - .3. A 100 rpm -. o. + 50 100.0 0 20 o 10 o 5 I 2.5 10.0. a ,— 1 I F bj 0 0 0 1 0.3 0 4 0 6 0.7 (1-n) 3 Figure 6.34: Plots of Dimensionless Functions [P/(md Nn+1)] Versus (l-n) For The Paddle Impellers at Different Values of N; a) Impeller 3 (b- 3cm); b) Impeller l (b-l cm) 167 the development of plots such as Figure 6.32, this effect was inves- tigated in this study. Figure 6.34 shows the changes in the slope of the lines when selecting different values of N for the paddle impellers. Similar results were obtained with the flag impeller (Figure 6.35) . Table 6.9 presents the values of the proportionality constant k' (evaluated by using the slope method) for the different system geometries when agitating the fluids with the paddles and flag impeller obtained at three different values of N (5, 20 and 100 rpm). It is clear that if only a particular value of the impeller rotational speed is used for the evaluation of k' , the selection of N becomes a critical step in the method since significant differences in the value of k' are observed at the different values of N: the lower the value of N, the higher the value of k'. Also, as N increases (100 rpm), k' becomes less dependent on system geometry. Regression analysis was conducted for the data for the paddle impellers and results indicated that the diameter ratio (d/D) was not statistically significant (Table 6.10). The following model was proposed: k' - a. (d/b)fl1 N’32 (6.13) Thus, 0 41 0 347 k' - 9.365 (d/b) ' N ' (6.14) Comparing the magnitude of k' (Table 6.8) for the flag impeller with the one obtained for the paddle impeller with same blade height (Impeller # 3: b - 3 cm; b/d - 0.6), the magnitude of k' is higher for the flag impeller. Results in Table 6.9 also indicate that higher values 168 LOE+05 - . . , . d/D= 0.273 Flog Impeller _ 1.05-+04 ' ' N a 1" "9 1000.0 5 “h ‘< ~ e. ' . 100.0 + 50 rpm I 215 10.0 , . , . . r 8| 1.00.05 , . r , r r . d/D- 0.591 Flog Impeller LOE+04 g .E g m "1: 1000.0 "' no é '- __‘ A 100 rpm E: + 50 100.0 0 20 o 10 o 5 10.0. "' .15. . . . - - b 0.0 0.1 0.3 0.4 0.6 0.7 (l-n) a Figure 6.35: Plots of Dimensionless Functions [P/(md Nn+1)] Versus (l-n) For The Flag Impeller at Different Values of N; a) Large Cap; b) Small Cap 169 Table 6.9: Values of k' (evaluated using the Slope Method) as a Function of Impeller Rotational Speed Paddle Impellers SYSTEM, k' d/D d/b 5 rpm 20 rpm 100 rpm 1 0.327 1.80 30.196 13.542 8.685 2 0.327 1.00 20.705 11.719 8.154 3 0.327 0.60 19.217 9.895 7.020 4 0.327 0.45 14.348 9.732 6.469 5 0.327 0.36 14.290 9.375 6.038 6 0.515 1.80 56.959 13.651 8.526 7 0.515 1.00 34.825 13.343 7.839 8 0.515 0.60 19.405 9.932 6.807 9 0.709 1.80 28.859 16.528 12.365 10 0.709 1.00 17.312 11.151 8.405 Flag Impeller SYSTEM k' 5 rpm 20 rpm 100 rpm 1 0.273 134.986 14.544 7.592 2 0.429 52.150 12.478 7.577 3 0.591 40.885 9.760 6.883 170 Table 6.10: Regression Results of Eqn. (6.13) (Paddle Impellers) Linear Multiple Regression Analysis Regression Estimated Regression Estimated Standard t Coefficient Coefficient Error 108 30 9.365 -- -- 51 0.410 0.073 5.96 82 0.347 0.073 -11.39 Analysis of Variance Sum of Degrees of Error Mean F* Saueres Freegem Squares Regression 1.366 2 0.683 -- Residual 0.226 27 0.008 81.6 Tgtel 1.592 30 2 R - 0.900 a - 0.05 171 of k' are encountered for the flag impeller, especially when using low rotational speeds. However, at higher values of N, the differences in the magnitude of k' become negligible and the average value of k' for the paddle impellers (k' - 8.03 i 1.78 1/rev) is higher than the value for the flag impeller (k' - 7.35 2t 0.40 l/rev). This may be due to the larger surface area of the paddle impellers. There is no published data on the value of k' for paddle impellers using this mixer viscometry method. Steffe and Ford (1985) reported k' values of 4.64 (at 60 rpm) for a pitched flag impeller. This magnitude is in reasonably good agreement with the average value of k' - 7.35 i 0.40 l/rev (at 100 rpm) (Table 6.9) determined in this investigation. When using different values of N, the average value is k' - 17.067 i 2.64 l/rev as compared to the value of 13.8 obtained by Rao and Cooley (1984) (d/D-0.8, d/b-&.5). Differences in the magnitudes of k' are due to differences in geometry. 6.1.2.2) Estimatiop of Average Shea; Rates and Apparent Viscosity Average shear rates were determined using Eqn. (3.8). Figures 6.36 and 6.37 present the average shear rates for the paddles and the flag impeller, respectively. The effect of impeller and cup size on average shear rates are presented in Figures 6.36a and 6.36b for the paddle impellers. As it was expected, no significant effect of cup diameter is observed. Average apparent viscosities were calculated with Eqn. (3.13). Figure 6.38 presents the values of "a for two non-Newtonian fluids as a function of iav for the different paddle impellers. It may be seen that the choice of impeller will produce different values of "a' Figure 6.39 25 172 '7’ Ti ' ’7 ' T I I I '*_1 I’ I I 4 x d/D-0.327 d/b‘1.0 O d/D-0.515 20"1 c: d/D-OJOS q '7." J ‘ \\‘ 15 _ C. l ! . B 10 ‘5 .4 ‘1 l a J 5 . l J I .. 0 r r l I l ' I - I v* I ' I"' a 25 T f I f I I t F 1 T y f y . x d/b-1.8 d/D-O.327 , o d/b-1.0 204 o d/b-O.8 .. J a (Vb-0.45 q '_' V d b-0.36 0) 15d / - \ x n B 10 .. oh ‘1 x O 4 g T 5-4 a q a i J o . , . , . b 0.0 0.2 0.4 , . 0.6 r . , . r 0.8 1.0 1.2 N [rev/s] Figure 6.36: Average Mixing Shear Rate as a Function of Rotational Speed. a) Different Cups; b) Different Paddles 173 20 ‘ I ' I fir I 1 I r I I I , I Y ‘ X d/D-O.273 . o d ”3-0.429 Flog impeller ‘ a d D-0.591 16-4 / _‘ J X '7' 124 - \ 3:. I . .zi 8-4 x g 0 _‘ O 44 g _ I 3 0 T I r I 0.0 0.2 0.4 0.6. 018' 110' 1T2 1.4 1.5' N [rev/s] Figure 6.37: Average Mixing Shear Rate as a Function of Rotational Speed for the Flag Impeller 1.8 174 4s +— r . e I . r . J x d/b-1.8 o d/b-1.o n-0.352, d/D-O.32 364 o d/b-O.6 .4 A (Vb-0.45 2? 274 V d/b-O.36 O 0. 1 I—l a 18-4 ‘0 F? a" ‘0‘, x 9'4 "D O x 1 V ‘ D O X 0 # I I r l r r 10 . , I r I . l . I z 33:}: n-O.504, (VD-0,327 84 o d/b-O.6 I A d/b-O.45 I-—v - . 2’ 64 v d/b 036 C) &. x. . 44 °x' P J ‘o a x u 2-I a o x , ‘ I: o x 0 fl I I r I I r 0.0 3.2 6.4 9 5 12.3 Figure 6.38: Apparent Viscosity as a Function of Average Shear Rate 7~[1/s} Using The Paddle Impellers. a) CMC 2%; b) CMC 1% 175 55 ' T ' I I I r I * "3:3: "3:21;. I 0 our. .rn- . 44-4 :1 n-0.352. m-59.275 d/D-0.591 - I I H D 5’: 33~ .. O o. J a . h—l o 22% _ S? 4 o a J O O 114 - O + x x x 31 0 l I ' I fi fi 45 . I . I . I . r + * "-3'331- "2:212, I o n- . , m- . 36-4 a «0.352. III-59.275 d/D'O-273 - F_‘ 4 o ' - 5': 27- -I O o. I o . H . 18-4 _ F? 4 o n j | O D 94 a 1 .I 4 x x x x b O ' I ' T ' I ' 1* r*’ 0.0 3.0 6.0 9 0 12.0 15.0 Figure 6.39: Apparent Viscosity as a Function of Average Shear Rate Using The Flag Impeller; a) Small Cap; b) Large Cap 176 shows results for the flag impeller. The size of the sample container also results in different values of "a’ and the bigger the cup, the smaller the values of "a“ 6.1.3) WW Three mixer viscometry methods were reviewed and evaluated. The values of the mixer proportionality constant k' for determination of the mixing average shear rate iav evaluated using the three mixer viscometry methods have been presented in Tables 6.3 through 6.9 in the previous sections. The assumption iav - k'N [Eqn. (3.8)] is the basis for the three procedures. However, the original assumption of the work of Metzner and Otto (1957) of a constant value of k' is not always valid for variations in operating conditions (impeller rotational speed, N), fluid rheological properties and system geometry. It is important to note that the three methods predict variations in k' with the geometry of the mixing system for the ranges investigated in this study. However, the flower Curves Method shows little variation with geometry for the flag impeller (Table 6.4). When working at high mixing speeds [N equal to 100 rpm (0.167 rps)], the values of k' ob- tained with the Iorgug Cuges Method and the Slope Method show excellent agreement. Under this operating condition, the effect of the other parameters (fluid rheology and system geometry) become less significant. Thus, the assumption of a constant k' value is valid under these terms. All the three mixer viscometry techniques require the determina- tion of parameters using computer analytical techniques, i.e., small changes in experimental data result in big changes in results of regres- sion analysis. The W can be very tedious and deviation from the basic assumption for the average shear rate ("yav - k'N) may 177 occur when using different impellers. The Mixer Torgue Curves and the Slope methods are simpler since they require less data handling. Results indicate that the basic assumption of traditional mixer viscometry methods of a direct proportionality between average shear rate in the agitated fluid and the rotational speed of the impeller [Eqn. (3.8)], with k' depending only on the geometry of the impeller, may be incorrect when working at low rotational speeds. Variation in fluid rheological properties has also proven to be an important factor for determination of the average shear rate in the mixing system. 6.2) ERMI AT OLOGICAL 0 RTIES OF POWER-LAW FLUIDS USING THE ALTERNATIVE MIXER VISCOMETRY METHOD This section presents the experimental verification of the proposed procedure for determination of rheological properties of power-law fluids using a measuring system which consists of an impeller (paddle or flag) rotating in a cylindrical container. 6.2.1) Determination of the Flow Behavior IndexI n The measured torque on the impeller shaft is presented as a func- tion of the rotational speed of the impeller for each non-Newtonian fluid, with the geometry of the impeller as a parameter. Figure 6.40 shows the results for the CMC 2% solution for the different paddle/cup combinations. Figure 6.41 shows results for the flag impeller. It fol- lows from these plots that for every geometry of the impeller system the experimental points may be approximated by a straight line. All 178 1 .000 0.100 M [Ntm] 0.010 0.001 . 0.100 M [Ntm] 0.010 0.10 ' 1.00 ' ' ”10100' ' 300700 1000.00 N [rpm] Figure 6.h0: Plot of the torque on the impeller shaft M versus rotational speed of the impeller N for 2% wt% aqueous solution of CMC (Paddle Impellers) (See Table 6.9 for definition of systems) 179 1.000 . . n... : , . - ....., . . N..... . W-..“ : CMC 27: Flog 3 0.1003 -. E 3 . 1 * '1 '1 z I . . &—-l 2 0010-: -. 1 , 1 ‘ H (VD-0.591 ' ' H d/D-O.429 ' 0.001 . . ......r e. rm... .fi ....°.'," “pf???" 0.10 1.00 10.00 100.00 1000.00 N [rpm] Figure 6.41: Plot of the torque on the impeller shaft M versus rotational speed of the impeller N for 2% wt% aqueous solution of CMC (Flag Impeller) A 180 lines have similar slope, and this slope is equal to the value of the flow behavior index, n. Tables 6.11 and 6.12 present the values of the rheological parameter for all the systems. It is seen that the mag- nitudes of the flow behavior index n obtained with the impeller system for the standard fluids are in good agreement with those using the concentric cylinders viscometer when operating at the same range of shear rates (0-40 l/s). 6.2.2) Determination of Shear firregs-Shear Rate Relationships 6.2.2.1) Average Shear gate 1;; The Miring System To verify the applicability of Eqns. (4.4), (4.6) and (4.9) to approximate the values of the mixing average shear rates, the values of k' determined using the three investigated mixer viscometry tech- niques [Section (6.1)] were used (as average values) for comparison with the theoretical expressions. Initial values of the parameters of the equations ([91, a1, a2 and as) were assigned following the concentric cylinders analogy to find the values of the constant 191 and the parameters a, , a2 and as that showed best agreement with the experimental values of k'. These are the following: 131-411' 011 - a2 - 2/n and, as - n/2, with n - power-law index. Thus, Eqn. (4.5) can be written as follows, 2/n n/2 k' - 4n [ (D/d) (6.13) (d/b) (1)/d)”n - 1 J 181 Table 6.11: Values of the Flow Behavior Index of Standard Power-Law Fluids Using the Mixer System (Paddles) and a Concentric Cylinders Viscometer (Haake Rotovisko) rElow Behavior Index- n 1 2 Mixer Congenrrig Cylinders __ELQID_______§X§IEM 1 0.902 i 0.003 2 0.834 t 0.004 3 0.896 i 0.003 4 0.887 t 0.002 CMC 1% 5 0.842 t 0.002 0.829 i 0.005 6 0.808 i 0.003 7 0.807 i 0.005 8 0.857 i 0.004 9 0.869 i 0.006 10 0.857 t 0.005 n 0.856 t 0.030 avg 1 0.599 t 0.008 2 0.759 i 0.004 3 0.726 i 0.004 4 0.735 i 0.003 CMC 1.5% 5 0.718 i 0.002 0.718 i 0.005 6 0.599 t 0.007 7 0.687 i 0.005 8 0.584 i 0.006 9 0.660 i 0.005 10 0.640 i 0.005 n 0.670 i 0.060 avg 1 0.679 t 0.004 2 0.688 t 0.003 3 0.648 i 0.004 4 0.684 i 0.002 CMC 2% 5 0.676 t 0.003 0.528 i 0.003 6 0.646 i 0.001 7 0.662 i 0.002 8 0.653 t 0.003 9 0.641 i 0.001 10 0.628 i 0.001 n 0.660 i 0.020 avg 1 Brookfield Mixer. Shear rate range: 0-30 l/s [Eqns. (6.18) and (6.19)] 2 Haake MV-III (d/D-0.73). Shear rate range: 0-40 l/s 182 Table 6.12: Values of the Flow Behavior Index of Standard Power-Law Fluids Using the Mixer System (Flag) and a Concentric Cylinders Viscometer (Haake Rotovisko) Flow Behevior Index. n 1 2 ELQID SXSIEM Mixery Concentric Cylinders 1 0.933 i 0.005 CMC 1% 2 0.805 t 0.002 0.829 i 0.005 3 0.967 t 0.007 n 0.901 i 0.085 avg 1 0.811 t 0.004 CMC 1.5% 2 0.711 i 0.003 0.718 t 0.005 3 0.708 t 0.001 n 0.743 i 0.058 avg 1 0.690 i 0.003 CMC 2% 2 0.667 i 0.002 0.528 i 0.003 3 0.658 i 0.002 n 0.672 i 0.016 avg 1 Brookfield Mixer. Shear rate range: 0-30 1/s [Eqn. (6.23)] 2 Haake MV-III (d/D-0.73). Shear rate range: 0-40 1/s 183 Equatiru1 (6.13) differs from the original equation for the shear rate at the cylindrical bob, qb .. gg (1)/<1)”n (‘1-3) n (D/d)2/n - 1 or ab - am (D/dlz/n (6.14) n (D/d)2/n _ 1 with 51 - (ZEN , al - a2 - 2/n, in the value of parameter 131 , since the division by n is not present in the equation for the paddle Ln- pellers. As stated before, the effect of the power-law parameters on the value of k' is not clearly understood. For an impeller, the dependence on the value of n seems to be less significant than in the case of the concentric cylinders. The values of a1 and 012 are identical to the concentric cylinders analogy. Values of k' obtained with Eqn. (6.13) are shown in Figure 6.42a. It is interesting to note that the above expression gives reasonably good results, except for the small gap case (d/D a 0.709). (These data are represented by the crosses (+) in Figure 6.42a). It is clear that Eqn. (6.13) predicts considerably higher values of k' when the impeller- to-cup diameter ratio is small [ (d/D) 2 0.709 ]. Thus, Eqn. (6.13) is a good approximation of the data when the diameter ratio follows within the range of 0.327 S d/D 5 0.515. An equation to approximate the values of k' when (d/D) was 2 0.709 was obtained (based on the concentric 184 30.00 J 0 Wide Glap I t I i + Small 1 e 25.20- GOP _ (J 1+: . c d 2 20.404 0 T. T: . Q) * ‘I + g- 15.60-1 0 ° ° .. ‘5; 4 ° + - e :' + ‘ x 1 0.80- 3.9 Paddles -1 6.00 ' r r ‘ I a 30.00 . , 4 1o “fideIGap + Small Gap 9 25.204 (I 4.; I: g 20.40— e: .1 3. x 1 5e 60" C) \J .1 3‘ 1 0.80- Paddles 0.00 . n.1, 6.00 ' 10130 ' 15:60 ' 20140 25120 ' 30.00 k' (calculated) Figure 6.42: Experimental Versus Calculated Values of k'; a) k' defined by Eqn. (4.5); b) k' defined by Eqns. (4.5) and (6.13) for small gap (+) (Paddle Impellers) 185 cylinders analogy) by expressing the experimental values of k' by: n/2 (6.15) 2.12 1.1 _ (m [ (010) («M») (1)/d)”In - 1 J Due to the empirical nature of Eqn. (6.15) it is difficult to physically explain the difference in the value of a1 - (2-n)/n. However, this result indicates that the gap between the impeller and the fluid container (cup) is a factor to be carefully taken into account when considering the use of mixer viscometry methods. Figure 6.42b presents results for the values of k' calculated using Eqn. (6.13) for 0.327 s d/D 5 0.515 and Eqn. (6.15) for (d/D) i 0.709. The results are satisfactory for all the impellers and power-law fluids investigated. The agreement between experiment (mixer viscometry techniques) and theory [Eqn. (6.15)] was usually better than 10%. Simplified equations were attempted by using the same procedure of data fitting. The final equations obtained were as follows: n/2 k1 - a, [——f§j§i§72 ] (6.16) Simplifying, Eqn. (6.16) becomes k' - 4n (1)/b)“/2 (6.17) Figure 6.43 shows the values of k' calculated from Eqn. (6.17) versus that determined from experimental data. Results show a reasonable agreement between predicted and average values. Even though not as good 186 30.00 I I I I 25.20- a - A O «8 0 a o 1: C, 20040-1 -1 E 0 e: o o 8. ° ° 3 15.604 0 ° .- V O - -1 o :1 x. '6' 10.304 , .116 Paddles . 6.00 . , . , . r e F r 6.00 10.80 15.60 20.40 25.20 30.00 k' (calculated) Figure 6.43: Experimental Versus Calculated Values of k'; a) k' defined by Eqn. (4.7) (Paddle Impellers) 187 as Eqn. (6.13), results fall within the range of k' values obtained using mixer viscometry techniques. The maximum error was less than 20%. It would seem that for the purposes of engineering design the use of Eqn. (6.17) is reasonable. Thus, results indicate that it is possible to approximate the value of the impeller proportionality constant, k' , for a particular impeller (paddle) by using Eqn. (6.13) for large to medium gaps, and by Eqn. (6.15) for small gap (d/D 2 0.709). Only the geometry of the system and the value of the flow behavior index, n, for the investigated fluid are needed. Equations (6.13) and (6.15) [as well as Eqn. (6.17)] avoid the dangerous assumption of k' being a constant independent of the fluid properties. Also, it can.be said that k' remains constant at each value of rotational speed. In terms of the average shear rate, these equations become, 2/n 78v - 4n ‘D/g;n ] (d/b)“/2 } N (6.18.1) (D/d) - 1 0r 2/n . (D/d) n/2 1 - 20 (d/b) (6.18.2) 3" [ (1)/<1)”n - 1 l for a paddle impeller and 0.327 s d/D 5 0.515 , and zfin 18v - {4w [ (D/d) J (d/b)“/2 } N (6.19.1) (D/d)2/n - 1 188 01' (D/d)z;In (D/d)2/n - 1 “/2 (6.19.2) Iav - 20 [ J (d/b) for d/D 2 0.709 (with l 2 b 2 5 cm and 0.5 2 n 2 0.9). Also, a simplified approximation is given by n/2 +av - [4w (D/b) 1 N (6.20.1) 01' 13v - [2«o(0/b)“/2] (6.20.2) for the geometric range investigated (0.327 s d/D 5 0.709). In the case of a flag impeller (b/d-0.5), Eqn. (4.9) gave good results for the values of k' , with 51" 411 and 011 - n/2. An expression such as Eqn. (4.9) was preferred for the flag impeller because Eqn. (6.13) did not yield very good results for this type of impeller. Thus, Eqn. (4.9) becomes, k' - 4n [ (1)/d)”2 l (6.21) Equation (6.21) predicts the value of the proportionality con- stant, k', for a flag impeller rotating in a power-law fluid. Figure 6.44a shows the results obtained from Eqn. (6.21) when d- 1.5cm. Figure 6.44b shows the case for Model 3 (d - de ). It seems that the assumption 189 C o H c: o E 1: o a. x 0 V 3‘ d 1 C.‘ 25001 1 0 «OJ d c g 22.00~ , ° . e: . O o a- 18.004 0 "1 3 1 ° 0 - 0e e . x 14500-1 "‘ . Flag . 10.00 . . . . . b 10.00 14100 13100 22100 26100 30.00 k' (calculated) Figure 6.44: Experimental Versus Calculated k'; a) k' defined by Eqn. (4.7); b) R’ defined by Eqn. (4.7) (d—de ) 190 of a cylinder of equivalent diameter, de , as responsible for the shearing of the fluid when using a flag impeller tends to overestimate the values of k'. However, the values of k' fall within the range of k' determined from the traditional mixer viscometry methods with a maximum error of 15%. Again, the assumption of a constant k' value, only a function of impeller geometry, is avoided and proved not to be tnnua for the cases investigated in this study. Thus, the average shear rate in a mixing system (flag impeller) can be approximated by the following equation, ’1 - {411' [ (1)/d)“2 1} N (6.22.1) 8V or n/2 ’1 v - 201 (DA!) ] (6.22.2) a 6.2.2.2) Average Shear Strees in The Mixing System 6.2.2.2.1) Iergue hpprorimeriohe To check the applicability of the shear stress equations, Um; experimentally measured values of torque for every mixing system were compared to those calculated using the following equations: Model 1 (concentric cylinders analogy with negligible end effects): M - 2th (d/2)2 aav (4.14) 191 Model 2 (concentric cylinders analogy with end effects): a fi L+_L M- 2 I: d 3 :laav (4.21) Model 3 (Flag impeller; d- de ): 2 M - 2«b (de/2) aav (6.23 l) and d3 n ___e L-Il M 2 [ de 3 ] aav (6.23.2) where aav'- m (y'av) flwith n from mixing system or a conventional concentric cylinders viscometer, and m from concentric cylinders vis- cometer. The value of the average shear rate can be evaluated using the appropriate equations [Eqns. (6.13) and (6.15) for the paddles and (6.21) for the flag for the ranges investigated in this study].ffima average shear rate could also be determined by traditional mixer vis- cometry methods. Figure 6.45 presents the values of torque for two fluids (1% and 1.5% CMC solutions) in a system consisting of a small paddle impeller (d/b-l.8) rotating in a large cylindrical cup (d/D-0.327). It can be seen that the Model 1 (concentric cylinders analogy with negligible end effects) [Eqn. (4.14)] gives better prediction of the torque values than the Model 2 (concentric cylinders analogy with end effects) [Eqn. (4.21)]. For a more viscous fluid (2% CMC), the Model 2 gives better results for the torque values (Figure 6.46). These results indicate that 192 6111: 1' .5: (Li/6- 1.11. 670-0327) ‘ Measured Torque [Ntm]x10' Measured Torque [Nem]x10' x in f t I I I’ 0 a 0.50 0.18 0.54 0.52 0.40 Calculated Torque [Nem]x 10a Figure 6.45: Measured Versus Predicted Values of Torque on the Impeller Shaft Using Eqn. (4.14) (o) and Eqn. (4.21) (x); a) CMC 1.5%; b) CMC 1% (Paddle Impeller # l, (d/b) - 1.8) [EE - Model 2 (end effects); NEE - Model 1 (no end effects)] 193 the type of fluid being agitated is a significant factor for considera- tion when using mixer viscometry methods. When using a larger paddle impeller (d/b-0.36) in the same cup [(d/D - 0.327] (Figure 6.47), the differences between the torque \nilues predicted by the two models becomes smaller, with the assumption of negligible end effects (Model 1) showing better agreement between experiment and theory than the assumption of the presence of end effects (Model 2). Thus, when working in a large cup, the end effects can be assumed negligible. For the other mixing systems, Model 2 (presence of end effects) (Figure 6.48) predicts values in close agreement with the experimentally measured torque values. This is due to the presence of a smaller gap between the impeller and.the wall of the cup and the use of Eqn. (4.21) seems to account for any effect of the solid boundaries. Figure 6.48a indicates that Eqn. (4.14) predicts lower values of torque. Figure 6.48b indicates the applicability of Eqn. (4.21) to represent the torque on the shaft resulting from the rotation of the paddle impeller (approximated by a cylinder). Based on the results presented in Figures 6.45 through 6.48, it can be concluded that when using a mixer viscometer with a small paddle in a large cup [(d/D) 5 0.327], assumption (iv) is valid, and the average shear stress can be approximated by Eqn. (4.15), a - 3L,- (4.15) for standard power-law fluids of low to medium viscosity. For a highly viscous standard fluid (CMC 2%), the effect of the solid boundaries 194 2.00 . , . , . 1 e 1 , ., . CMC 22: (d/b=1.8, d/D=0.327) C3 3'2 1.60-1 l—! E . *1 z "" 1.20-4 0 :3 a. 3 I— 0.604 13- J 0 L. a c, 0.40-4 0) 2 ‘ )6 NE 0.00-J . , . , . , .°IEE. 0.100 0.40 0.80 1.20 1.60 2.00 Calculated Torque [Ntm]x 103 Figure 6.46: Measured Versus Predicted Values of Torque on the Impeller Shaft Using Eqn. (4.14) (o) and Eqn. (4.21) (x) CMC 2% [EE - Model 2 (end effects); NEE - Model 1 (no end effects)] 195 no 3.00 I I I I t I 3'; .l CMC 1.5% (d/b=0.36, d/D=0.327) ,_, E 2.40- I' z I—l ., 1.80-4 3 U’ -l 3 ,__ 1.20.. ‘0 .1 3’- J 3 0.60 g NE 2 0.00-l , . , E": 0.00 0 60 1.20 1.80 2.40 3.00 3 Measured Torque [Ntm]x10 x NEE‘ 6 SE 0.00 . r . , . , . , . Al 0.00 0.16 0.32 0.46 0.64 0.60 Calculated Torque [N*m]x 103 Figure 6.47: Measured Versus Predicted Values of Torque on the Impeller Shaft Using Eqn. (4.14) (o) and Eqn. (4.21) (x); a) CMC 1.5%; b) CMC 1% (Paddle Impeller # 5, (d/b) - 0.36) [EE - Model 2 (end effects); NEE - Model 1 (no end effects)] 196 5.00 . Paddles r* ' ”’r ‘2’ I . I I L_ Measured Torque [Nem]x10’ “o 3.. E 9 .2.. O 3 g g... g I! 0.00 1 1 - r r . ”E‘s" 0.00 1 .00 2.00 3.00 4.00 5.130 Calculated Torque [Nem]x 103 Figure 6.48: Measured Versus Predicted Values of Torque on the Impeller Shaft For All Systems And Fluids (Paddle Impellers); a) Eqn. (4.14) (Model 1 - NEE); b) Eqn. (4.21) (Model 2 - EE) 197 becomes significant and assumption (iv) is questionable. Thus, the average shear stress can be approximated by -1 3 a - -1§— [ -g— + -%— ] M (6.24) Equation (6.24) also applies to the other impeller/cup combina- tions for the range of viscosities of the standard fluids. The torque values calculated by equations (4.14), (4.21) and (6.23.1) were compared to those measured experimentally with the flag impeller. Figure 6.49 presents the measured torque values versus those calculated using Eqn. (4.14) (Model 1): M - 2nb (d/2)2 Gav (4.14) It is evident that this model yields values of torque considerably higher than the experimental values. Figure 6.50a presents tflue results using Eqn. (4.21) (Model 2): .1d. .0. + .1. M - 2 [ d 3 ] aav (4.21) and it yields considerably lOwer values oi'torque. Equation (6.23.1) (Model 3), 2 M - 2wb (de/2) aav (6 23 1) seems to be the one that better represents the torque in the impeller 198 3.00 I'D d CD .35, 2.40- E * d E. 1.604 0 3 .l 6' 1- 1.204 . 'u a: . 3 8 0.60- 0 2 . 0.00 0.00 I ' I ' l ' I Model 1 (No end effects) o 9 " 0 $ 1 10 e . ‘6 . ° " 66 ° Flag . ”I" ° ' 0.60 t 1.20 1.60 2.40 3.00 Calculated Torque [Nurm])<103 Figure 6.49: Measured Versus Predicted Values of Torque on the Impeller Shaft For All Systems And Fluids (Flag Impellers) Using Eqn. (4.14) (Model 1) (Negligible end effects) 199 2.00 . F . , . , ”c Model 2 (End effects) .3 1.604 E O .2... 3 1.204 6 1- 0.604 3 0.404 a 0.00 f f , e 2.00 v I V fi ' l ' I '0 . Model 3 (d I de) . 3:". 1.604 0 - E 1.204 ° ‘ *- Geno-1 " . . , 00w- - . Flag , 0-00 r. r c r c I 1 1— .— b 0.00 0.40 0.80 1 .20 1 .60 2.00 Calculated Torque [Nalsrr1]x103 Figure 6.50: Measured Versus Predicted Values of Torque on the Impeller Shaft For All Systems And Fluids (Flag Impeller); a) Eqn. (4.21) (Model 2); b) Eqn. (6.23.1) (Model 3) 200 shaft (Figure 6.50b). Thus, the average shear stress for the flag impeller may be ex- pressed by 3 '1 «d e _L i a — 2 [ (18 + ] M (6.25) av 3 Equation (6.25) implies that the effect of the end boundaries (top and bottom) is important when using a mixing system with a flag im- peller. 6.2.2.3) Elow Curves 6.2.2.3.1) Ideal Eluids Average shear stress-average shear rate curves were obtained for the standard power-law fluids using Eqns. (6.18) and (6.19) for evalua- tion of the average shear rate of paddle impellers and Eqn. (6.21) for the flag impeller. The average shear stress was determined using Eqn. (4.21) for the paddles and Eqn. (6.23.1) for the flag impellrnt. Figure 6.51 presents a typical shear stress-shear rate plot for the CMC 2% solution obtained using the mixer viscometer with a paddle impeller. Similar results were obtained with the other fluids and geometries (Figures 314 and BIS). As indicated before, the values of the flow behavior index were calculated as the slope of the log-log plot of torque, M, versus the rotational speed of the impeller, N. The values of the consistency coefficient, m, were obtained by linear regression of the power model, a - m (y'av) fl The values are shown in Tables 6.13 through 6.17 for av every system geometry and Models analized in this investigation. 201 250.00 *' I r, T’ t r ' r r d I4L61- d '4L32 . ° I” m 7 one 2:: . 200.00- .. O F; ISOeOO'l - 0L H I . O b 100.00. "' ° 1 50.00.. a Ra I 0.990 . . o n - 0.666 n . )0 m - 26.35 Pace 0000 r fi r r r r Ti 7 r f 0.00 6.00 l 2.00 1 6.00 24.00 30.00 ‘7- [1/81 Figure 6.51: Flow Curve For 2% wt% Aqueous Solution of CMC Determined Using the Mixer Viscometer with a Paddle Impeller 202 Table 6.13: Values of the Fluid Consistency Coefficient of Standard Power-Law Fluids Using the Mixer System (Paddles) [Model 1: negligible end effects; Eqn. (4.14)] and a Concentric Cylinders Viscometer (Haake Rotovisko) n Consistency Coefficient, m (Pa 3 2 1 2 FLUID SYSTEM Mixer Congentgic Cylinders 1 2.139 i 0.003 2 2.664 t 0.004 3 2.835 t 0.003 4 3.173 i 0.005 CMC 1% 5 3.557 i 0.002 2.619 t 0.003 6 2.492 i 0.003 7 2.737 i 0.004 8 2.823 i 0.002 9 2.937 i 0.001 10 2.906 t 0.003 m 2.826 i 0.378 avg 1 18.892 i 0.004 2 12.261 i 0.007 3 13.781 i 0.005 4 14.702 i 0.004 CMC 1.5% 5 15.687 i 0.003 15.620 i 0.004 6 18.892 i 0.002 7 15.299 i 0.002 8 22.967 i 0.003 9 21.619 i 0.004 10 24.742 i 0.003 m 17.877 i 4.200 avg 1 33.209 i 0.002 2 30.656 i 0.004 3 33.379 i 0.002 4 32.980 i 0.001 CMC 2% 5 35.142 i 0.003 33.170 i 0.005 6 37.737 i 0.002 7 34.793 i 0.004 8 38.149 i 0.001 9 45.239 i 0.002 10 47.738 i 0.005 m 36.900 i 5.540 avg 1 Brookfield Mixer. Shear rate range: 0-30 l/s [Eqns. (6.18), (6.19)] 2 Haake MV-III (d/D-0.73). Shear rate range: 0-40 l/s 203 Table 6.14: Values of the Fluid Consistency Coefficient of Standard Power-Law Fluids Using the Mixer System (Flag) [Model 1: negligible end effects; Eqn. (4.14)] and a Concentric Cylinders Viscometer (Haake Rotovisko) n Qongigtency Coefficient, m (Pa 5 2 1 2 ELDIQ SXSIEM Mixer Concentric Cylinders 1 0.750 1 0.007 CMC 1% 2 1.288 1 0.005 2.619 1 0.003 3 1.155 1 0.004 m 1.064 1 0.280 3 avg 1 4.152 1 0.003 CMC 1.5% 2 7.100 1 0.003 15.620 1 0.004 3 8.961 1 0.002 m - 6.738 1 2.421 avg 1 12.400 1 0.004 CMC 2% 2 15.926 1 0 002 33.170 1 0.003 3 21.318 1 0.003 m 16.548 1 4.491 avg 1 Brookfield Mixer. Shear rate range: 0-30 1/s [Eqn. (6.21)] 2 Haake MV-III (d/D-0.73). Shear rate range: 0-40 l/s 204 Table 6.15: Values of the Fluid Consistency Coefficient of Standard Power-Law Fluids Using the Mixer System (Paddles) [Model 2: concentric cylinders with end effects; Eqn. (4.21)]) and a Concentric Cylinders Viscometer (Haake Rotovisko) n angistengy CnefficientI m (Pa 3 ) 1 2 FLUID SYSTEM Mixggi Concentric Cylinders 1 1.526 t 0.003 2 2.046 t 0.003 3 2.643 t 0.002 4 2.759 t 0.005 CMC 1% 5 3.161 t 0.002 2.619 i 0.003 6 1.562 i 0.003 7 1.056 i 0.004 8 2.353 i 0.001 9 1.836 t 0.002 10 2.219 t 0.004 m 2.116 t 0.642 avg 1 7.699 i 0.004 2 9.362 t 0.007 3 11.484 i 0.005 4 12.784 i 0.004 CMC 1.5% 5 14.006 i 0.003 15.620 i 0.004 6 11.808 i 0.002 7 11.473 i 0.002 8 19.139 i 0.003 9 13.523 i 0.004 10 18.556 i 0.002 m 13.001 i 3.644 avg 1 22.115 i 0.002 2 24.334 i 0.004 3 26.354 i 0.001 4 28.348 i 0.001 CMC 2% 5 31.378 i 0.003 33.170 i 0.005 6 23.595 i 0.003 7 27.543 i 0.004 8 31.794 i 0.002 9 26.356 i 0.003 10 36.085 1 0.004 m 27.790 i 4.268 avg 1 Brookfield Mixer. Shear rate range: 0-30 l/s 2 Haake MV-III (d/D-0.73). Shear rate range: 0-40 l/s 205 Table 6.16: Values of the Fluid Consistency Coefficient of Standard Power-Law Fluids Using the Mixer System (Flag) [Model 2: concentric cylinders with end effects; Eqn. (4.21)] and a Concentric Cylinders Viscometer (Haake Rotovisko) nc e ie t Pa sn 1 2 S xe Co centric C linders 1 0.640 i 0.007 CMC 1% 2 1.104 i 0.005 2.619 t 0.003 3 1.120 f 0.004 m 0.954 t 2.720 avg 1 4.001 i 0.003 CMC 1.5% 2 6.350 1 0.003 15.620 i 0.004 3 7.574 i 0.002 m 5.975 t 1.816 avg 1 12.210 i 0.004 CMC 2% 2 13.782 i 0.002 33.170 i 0.003 3 18.324 1 0.003 m 14.772 1 3.175 avg 1 Brookfield Mixer. Shear rate range: 0-30 l/s 2 Haake MV-III (d/D-0.73). Shear rate range: 0—40 1/s 206 Table 6.17: Values of the Fluid Consistency Coefficient of Standard Power-Law Fluids Using the Mixer System (Flag) [Model 3: d - de ; Eqn. (6.23.1)] and a Concentric Cylinders Viscometer (Haake Rotovisko) enc e ient Pa sn 1 2 __ELQ.I_D__§1$_TEM—_mxer COW. 1 1.522 1 0.006 CMC 1% 2 2.648 1 0.004 2.619 1 0.003 3 2.328 1 0.004 m 2.166 1 0.580 avg 1 8.408 1 0.003 CMC 1.5% 2 14.379 1 0.003 15.620 1 0.004 3 18.417 1 0.003 m 13.735 1 5.035 avg 1 25.120 1 0.004 CMC 2% 2 32.250 1 0.002 33.170 1 0.003 3 43.280 1 0.002 m 33.550 1 9.149 avg 1 Brookfield Mixer. Shear rate range: 0-30 l/s [Eqn. (6.23.1)] 2 Haake MV-III (d/D-O.73). Shear rate range: 0-40 l/s 207 Tables 6.13 and 6.14 present the values of the fluid consistency coefficient, m, obtained by replacing the mixing system with the Model 1 [negligible end effects, Eqn. (4.14)], for the paddles and flag im- peller, respectively. Comparison with the values obtained using the concentric cylinders viscometer show that the results with the paddle impellers are satisfactory. However, the flag impeller gives con- siderably lower values of the rheological parameter, m, when using Model 1 (Table 6.14). Also, the effect of the gap seems to be more important for this impeller, since the value of the consistency coefficient ob- tained with the small gap case (System 3, d/D - 0.591) is twice as high as the value of the consistency coefficient in the large gap case (System 1, d/D - 0.273) (See Table 6.14). A small gap also gives higher" ‘malues of the rheological parameter when using a paddle impeller (Systems 9 and 10, d/D - 0.327), especially for the fluid of higher viscosity (CMC 2%) (Table 6.13). Tables 6.15 and 6.16 present the values of the consistency coeffi- cient, m, calculated by replacing the mixing system with the Model 2 [presence of end effects, Eqn. (4.21)], for the paddles and flag im- peller, respectively. No big differences with the values from Tables 6.13 and 6.14 are observed, with results from Model 2 [Eqn. (4.14)] lower than those obtained with Model 1 [Eqn. (4.21)]. The same behavior was observed in results for the flag impeller (Table 6.16). Table 6.17 presents the results obtained by replacing the flag impeller with a cylinder of dimensions as in Model 3 [d - de , Eqn. (6.23.1)]. It can be seen that this model gives considerably higher values of the fluid consistency coefficient than the previous models. In 208 comparison to the values obtained using a concentric cylinders vis- cometer, results also show better agreement. Rheogxams (shear stress-shear rate curves) were developed for all fluids and systems. Figure 6.52 shows the flow curves for the 1% CMC (Figure 6.52a) and the 2% CMC (Figure 6.52b) solutions using the mixer viscometer with the paddle impellers in a large cup [(d/D) - 0.327]. The model selected for approximation of the average shear stress was Model 2 (concentric cylinders analogy with end effect) since it gave more con- sistent values of torque (Figure 6.48b). Results were compared with those obtained with the concentric cylinders viscometer (Haake Rotovisko with the MV-III sensor which provided a range of shear rates similar to that obtained with the Brookfield viscometer and the mixer impellers) and proved to be satisfactory. Figures 6.53a and 6.53b present results for a small gap [(d/D) - 0.709]. Results indicate that the method works better when the agitated fluid is highly viscous (CMC 2%) than when agitating a low viscosity fluid (CMC 1%) when a small gap is present. Figure 6.54 presents the flow curves obtained for the flag im- peller using the Model 3 (d - de ). Results show excellent agreement with those obtained using the concentric cylinder viscometer, with better results for the more viscous fluid (CMC 2%). For comparison, Figures 6.55 and 6.56 present the flow curves obtained using Model 1 [Eqn. (4.14)] and Model 2 [Eqn. (4.21)], respectively. It is clear tfluat the Model 3 [Eqn. (6.23.1)] is still the best approximation for the mixing system when agitating standard power-law fluids. 6.2.2.3.2) Eood Erodugt Flow curves for an actual food product (Rancher's Choice Creamy 209 1000.00 . mmy . . ”.ny . .mmf y rnwnr . ,, o d/b-1.8 («VD-0.327) o d/b-0.6 e d/b-OJG 100"” + H6611. (IN-lll) , 1'— O 3'... &: . 60 .1 10.00 ° ° ' “.4! E: 'e' " 1.00 ° CMC 13 0.10y , ..yyqu . ....;IIFy .yn—Jlfv . ..ynllr . a 1000.001W t 1 fit: r r: u - rt : o d/b-1.6 (am-0.827) ' ' ""5 I o «Vb-0.6 I 1 A d/b-OJO ‘ ‘ + Pkuflut(flwhdl) . ’ ‘ .25? d 4»"’0 ‘ ._.. 100.001 " ‘ 1 a 1 23': : I: j e 1 d «0‘ . II 1 ~ . q . . . one 2:: 1 10.00““ . . . vr'I'l—‘l—I-F'I'I'W—l-‘FFl'I'Ifi—l—I-I'I'IEI'I' 0.01 0.10 1.00 10.00 100.00 1000.00 ‘7- [1/8] Figure 6.52: Comparison of the Flow Curve for a Standard Fluid Determined Using the Concentric Cylinders Viscometer with the Data Obtained Using the Mixer Viscometer with the Paddle Impellers in a Large Cup [Model 2, Eqn. (4.21)]; a) CMC 1%; b) CMC 2% 210 1 000.00 o d/b-1.6 (d/0-0.708) o d/b-l .0 + Hooke (IN-Ill) 1 00.00 H I! ,0 E. 1 ' oe .g " oe 10.00 ,4" 0 .‘ae C110 13 1 .00 I W fi I W r a lOOKLIN11--r-r1w11ur-w-1Frvunqr . . ..yuf”. ...yan—** y 1 o d/b-l .6 “ID-0.700) “a? : o d/b-1.0 1 ' + Moshe (UV-ll) 1 . «o '3' ‘ 0. a: ‘ E... 100.00-1 3116 -: 6 ' 1 b 3 . I" : a . d - 9 a . ° one 2: . 10.00-4mm . , . n. b 0.101 0.10 1 .00 10.00 1 00.00 1 000.00 1"- [Us] Figure 6.53: Comparison of the Flow Curve for a Standard Fluid Determined Using the Concentric Cylinders Viscometer with the Data Obtained Using the Mixer Viscometer with the Paddle Impellers in a Small Cup [Model 2, Eqn. (4.21)]; a) CMC 1%; b) CMC 2% 211 1000.00 I 00"!!! r U IIUItI'r I fititiq T U VIII-a! d/0-0.273 d/0-0.420 d/0-0.501 100.00 Hooke (IN-Ill) 1 0.00 Um [P0] '0' 9 a 0 3+ If e“. “x ’o “o ’0 0+ '3 o W 1.00 ° CMC 13 0.10I . . ”Hwy . ..yyy.q— . . ”Hwy . . ......, 1000.00,, . . turner , . uum . . ”may amt" I o d/D-0.273 1 . O d/D-0.429 : ‘ A d/D-0.591 " ‘ + Hooke (UV-Ill) to. " ' I ‘ . :3" . 1 "° 3 I 0.3 : b 1 ‘ +5 1 -1 . fi «1 a; " ‘ 011023 ' b 10.00-[W y . yynnf . y.r"uf . rutrm 0.131 0.10 1 .00 1 0.00 l 00.00 1 000.00 7"- [1/81 Figure 6.54: Comparison of the Flow Curve for a Standard Fluid Determined Using the Concentric Cylinders Viscometer with the Data Obtained Using the Mixer Viscometer with a Flag Impeller [Model 3, Eqn. (6.23.1)]; a) CMC 1%; b) CMC 2% 212 1000.00 . mum, y . ......y . yywm, . r ....u, o d/0-0.273 o d/0-0.420 A d/0-0.591 100'00 + Hooke (MV-lll) , 7 + ”’+ a + I: fine 4" b E‘ 10.00 ‘4‘” be ° 0 + he b be 1.00 a O CMC 1% a 0.10! y rrrva . .......r . ....yyq—. Irrrrfl“: ....yy.‘ 1000.00 . . . .....y r . . .....y y.....y ......y . o d/0-0.273 a d/DI!0.420 4. A d/0-0.501 «O + Pkuflu|(Ndefl) * 1" P; 100.00 1:. Q +’ ‘ <3 13’s a. 0. ... "0 H ‘u. 6 b 10.00 1:. C110 23b 1.00. . ...nrr—I—v-l-I-n-u“. .yrnflj:. .yywqu. run 0.01 0.10 1 .00 1 0.00 100.00 1 000.00 7?.[1/31 Figure 6.55: Comparison of the Flow Curve for a Standard Fluid Determined Using the Concentric Cylinders Viscometer with the Data Obtained Using the Mixer Viscometer with a Flag Impeller [Model 1, Eqn. (4.14)]; a) CMC 1%; b) CMC 2% 213 1000000 I I IIIIII' I I I (YTTYI I I IIIIIII I rII o d/0-0.273 n d/0-0.420 1 00.00 A d/D-0.591 + Hooke (av-111) , , E” . ‘ ... 10.00 be 5° 6 0. Que b 9 50 1.00 a 3" 0 one 17: 0.101 . 11...“, . . ...., . e. ...", . . ”3., 1000000 I I I VUI'UIr I I I IIIIII I I I III!!! I I I Irrt o d/D-0.273 a d/0-0.422 a d/0-0.591 . ' .— 100.00 1 "w“ ("V-m) .. ° ‘5. e ‘ ‘30 (3 o a 0. . ° °o 1—1 . ‘6 ‘6. 6 ., 1 10.00 o cuc 2:: 1.001 . . ......y . . r . . . 1...? 0.10 1.00 1050 106.00 100"0.00 7"- [1/81 Figure 6.56: Comparison of the Flow Curve for a Standard Fluid Determined Using the Concentric Cylinders Viscometer with the Data Obtained Using the Mixer Viscometer with a Flag Impeller [Model 2, Eqn. (4.21)]; a) CMC 1%; b) CMC 2% 214 Dressing; Kraft, INC., Glenview, Illinois) were estimated to check the applicability of the proposed procedure. Figure 6.57 presents the results using the paddle impellers [Model 2, Eqn. (4.21)] in comparison with the ones obtained with the concentric cylinders viscometer. The agreement of the data is evident for different impeller and cup sizes. Tables 6.18 and 6.19 present the values of the rheological parameters (n and m) for the salad dressing, evaluated using Model 1 [Eqn. (4.14)] and Model 2 [Eqn. (4.21)] for approximation of the average shear stress with the paddle impellers, respectively. The proposed procedure seems to be able to approximate the rheological behavior of a food product when using the mixer viscometer with the paddle impellers. The approach takes into account the variations in impeller and cup geometry. It is interesting to note that the differences due to variation in impeller are less than the observed in the developed curve flows for the standard fluids (See Figures 6.53 and 6.57, for instance). The same trend was observed with the flag impeller (Figures 6.56 and 6.58). For the flag impeller, peculiar results were obtained. Table 6.20 shows the values of the rheological parameters of the tested fluid with the three models investigated in this study. Figure 6.58 shows the flow curves obtained for the food product by approximating the flag impeller with a cylinder of diameter (18 (Model 3). Approximation of the shear stress-shear rate data with Model 3 gives higher values of the flow curve as compared with the concentric cylinders viscometer (Haake Rotovisko with MV-III). Figures 6.59 and 6.60 present results using Model 1 and 2, respectively. Results indicate that the two models show excellent agreement with the concentric cylinders data for the food Product as compared with Model 3 (Figure 6.58). It might be said that 215 1000.00 . . .....r—I—r-l-I-nwfi . o d/b-1.8 MAJ-0.327) '7 Ir o d/b-0.6 A d/b-OJd Hooks ouvuqu) '— 1 00.00 I: 0.. h-‘ a 6 11.1313" 1 0.00 1.00fi trttr‘fi'irriI—TTT It: 1 000.00 o d/b-1.8 (d/0-0.700) u d/b-1.0 1+ Hunk..fllvhdfl) 1 00.00 H ‘3 10¢ . 1: " I'v':%.'fl' 1 0.00 Salad Dreeehg, b 1.00 ' 0.01 0.10 1 .00 1 0.00 1 00.00 1 000.00 7- [V8] Figure 6.57: Comparison of the Flow Curve for Salad Dressing Determined Using the Concentric Cylinders Viscometer with the Data Obtained Using the Mixer Viscometer with the Paddle Impellers [Model 2, Eqn. (4.21)]; a) Large Cup; b) Small Cup 216 1000.00 . . . . yyyaE .yynr. 1 g o d/0-0.273 ' 1' 1 a d/D-0.420 : 1 A d/D-4L591 . , + Hooke (IN-ill) , 100.00-:1 - '3' 3 d 5, * . 0L 1 ‘* d* . 1.... . ‘ ‘ + 1 8 1 . w" - I I 1 Flag Impeller I . Salad Dressing ‘ 1 .00-W u rt rm!!— u I u I m 0.01 0.10 1 .00 10.00 1 00.00 1 000.00 '7- [1/81 Figure 6.58: Comparison of the Flow Curve for Salad Dressing Determined Using the Concentric Cylinders Viscometer with the Data Obtained Using the Mixer Viscometer with a Flag Impeller [Model 3, Eqn. (6.23.1)] 217 1 Flog Impeller . Solod Dreeehq 1.00-mi. rumM—r yuuur r..." 0.131 0.1 O 1 .00 10.00 1 00.00 1 000.00 7- [1/8] Figure 6.59: Comparison of the Flow Curve for Salad Dressing Determined Using the Concentric Cylinders Viscometer with the Data Obtained Using the Mixer Viscometer with a Flag Impeller [Model 1, Eqn. (4.14)] 1000.00 ‘ 1 00.001 - 3 <3 4 Q. .. H I e! 10.00-é 1 .00- 00‘31 +>ao, 218 . . .mq—fi-mwrr—q—u—u-nm . . . . y... d/0-0.273 ' d/0-0.420 d/0-0.501 Hooks (LN-Ill) + ... O 6 I oe’ 3'" e Flog Impeller Salad Dressing . . ..mq—. ..mIIF . ....n 0.1 0 1 .00 l 0.00 1 00.00 1000.00 7.. [1/8] Figure 6.60: Comparison of the Flow Curve for Salad Dressing Determined Using the Concentric Cylinders Viscometer with the Data Obtained Using the Mixer Viscometer with a Flag Impeller [Model 2, Eqn. (4.21)] ‘ 219 Table 6.18: Values of the Flow Behavior Index (n) and the Fluid Consistency Coefficient of Salad Dressing Using the Mixer System (Paddles) [Model 1, Eqn. (4.14)] and a Concentric Cylinders Viscometer (Haake Rotovisko) n te e i ie Pa sn 1 SYSTEM Ming;__!1§nnn§ter 1 0.248 t 0.003 29.811 1 0.003 3 0.243 1 0.001 20.186 1 0.002 5 0.236 1 0.001 30.057 1 0.005 6 0.239 1 0.002 25.475 1 0.005 8 0.227 1 0.003 19.039 1 0.003 9 0.257 t 0.002 32.007 1 0.004 10 0.258 1 0.002 27.701 1 0.003 Average 0.244 i 0.011 26.325 1 5.029 2 Haake data: n - 0.399 0.005 1 m - 15.00 1 0.003 Pa s“ 1 Brookfield Mixer. Shear rate range: 0-30 1/s [Eqns. (6.18) and (6.19)] 2 Haake MV-III (d/D-0.73). Shear rate range: 0-40 l/s 220 Table 6.19: Values of the Flow Behavior Index (n) and the Fluid Consistency Coefficient of Salad Dressing Using the Mixer System (Paddles) [Model 2, Eqn. (4.21)] and a Concentric Cylinders Viscometer (Haake Rotovisko) n s stenc Coe icient m Pa sn 1 SYSTEM MW: 1 0.248 1 0.002 19.314 1 0.003 3 0.236 t 0.003 15.484 1 0.004 5 0.243 t 0.001 18.023 1 0.002 6 0.239 1 0.002 18.793 1 0.005 8 0.227 1 0.003 15.867 1 0.002 9 0.257 1 0.002 18.313 1 0.003 10 0.258 1 0.002 18.999 1 0.004 Average 0.244 1 0.011 17.853 1 1.554 2 Haake data: n - 0.399 1 0.005 n i m - 15.00 0.003 Pa 3 1 Brookfield Mixer. Shear rate range: 0-30 1/s [Eqns. (6.18) and (6.19)] 2 Haake MV-III (d/D-0.73). Shear rate range: 0-40 1/s 221 Table 6.20: Values of the Flow Behavior Index (n) and the Fluid Consistency Coefficient of Salad Dressing Using the Mixer System (Flag) and a Concentric Cylinders Viscometer (Haake Rotovisko) n te en m Pa sn 1 __§y§1§u Mixg; Vigggneter Model System 1 1 0.257 1 0.003 14.916 1 0.004 2 0.230 1 0.002 15.030 1 0.004 3 0.235 1 0.002 16.249 1 0.003 Average 0.240 t 0.014 15.398 1 0.739 Model System 2 1 0.257 t 0.003 12.570 1 0.005 2 0.230 1 0.002 12.907 1 0.003 3 0.235 1 0.002 12.740 1 0.003 Average 0.240 1 0.014 12.739 1 0.168 Model System 3 1 0.257 1 0.003 28.691 1 0.003 2 0.230 1 0.002 29.240 1 0.002 3 0.235 1 0.002 30.690 1 0.004 Average 0.240 1 0.014 29.873 1 0.742 2 Haake data: n - 0.399 1 0.005 n m - 15.00 1 0.003 Pa 5 1 Brookfield Mixer. Shear rate range: 0-30 1/s [Eqn. (6.21)] 2 Haake MV-III (d/D-0.73). Shear rate range: 0-40 1/s 222 the area of the flag impeller of diameter d and blade height b con- tributes to the torque values when agitating the food product. In the case of a standard power-law fluid such as the CMC solutions, this is not the case, since the assumption of the total area of the flag im- peller yielded higher values of torque as compared with the geometry given by Model 3 (d - de ). These results indicate that the fluid to be investigated using mixer viscometry is of great importance when select- ing the appropriate equations to approximate the flow curves since the food product (salad dressing) shows more shear-thinning behavior (n= 0.244 1 0.011) than the standard CMC solutions (0.5 s n S 0.9). In terms of comparison, plots of the apparent viscosity, "a , versus the average shear rate, :yav , were developed for the different Models using the flag impeller and the results obtained with the con- centric cylinders viscometer. Figures 6.61 to 6.63 present the results for the three models analyzed in this investigation. The agreement of results is clear and excellent. It may be seen that the differences obtained with the Model 3 (d - de ) is not so drastic (Figure 6.63). Figure 6.64 presents the results for two different gaps. Figure 6.64a shows the results when agitating the food product with a flag impeller and a small gap is present (d/D-0.59l). Figure 6.64b shows the results for the wide gap case (d/D-0.273). It may be noted that the effect of the mixing system (in this case, the impeller-to-cup diameter ratio, d/D) is not significant. In.summany, analytical systems were developed for the mixing systems based on the original development for a power-law fluid agitated in a concentric cylinders viscometer. These systems (or models) take into account differences in impeller shape (paddle or flag) and n. [Pms] 223 40.00 . I . r I I . o d/D-O.273 . o d/D-O.429 F109 q A d/D-O.591 32°00“ + Hooks (W—lll) Model 1 (NE) "‘ 24.0046 .. 16.00~ .. O - + 4' 8.00- 00+ 4 d 3*» 4 ‘ an“ on A * 0.00 ,— rw rfi , , .— 0.00 8.00 16.00 24.00 32.00 7.. [1/8] 40.00 Figure 6.61: Apparent Viscosity as a Function of Average Shear Rate For Salad Dressing. Flag Impeller (Model 1: concentric cylinders with no end effects [NEE]). 224 40.00 . , . , . , o d/0-0.273 . u d/D-O.429 ”“9 A d/D-0.591 3100* + Hooks (NV-Ill) M°d°' 2 (EB) - :3 24.004 .. O E. 1 ‘ 5 16.004 - d g cl 8 OO- *4- an, ' " *4 a Q + an 0.00 . , . , . °° ,‘ . j . 0.00 8.00 16.00 24.00 32.00 40.00 ‘7- [1/8] Figure 6.62: Apparent Viscosity as a Function of Average Shear Rate For Salad Dressing. Flag Impeller (Model 2: concentric cylinders with end effects [EE]). n. [Pms] 225 40.00 . , . , I o d/D-0.273 . a d/D-O.429 ”“9 4» A d D-4L5 1 552-004 + Hick. (Msv—m) MM“ 3 (d'd°) -’ L o 24.00-1 .. J 16.004 i q» d Q q» a.00~ " 9' -. 4' . ++‘* ‘bA 0 * On A + 0.00 F 1 T ‘ . . ' 0.00 8.00 16.00 24.00 32.00 40.00 Figure 6.63: Apparent Viscosity as a Function of Average Shear Rate For Salad Dressing. Flag Impeller (Model 3: concentric cylinders with end effects; (1 - de ). n.- [Pets] 1). [Pas] 226 40.00 . , e r r I J o Model ; Eggs) Flog q a Model 32.00-4 A Model :5 (d-de) d/D-O.591 _ , + Hook. (av-m) (Smou GOP) A 24.00-4 0 .. .n . 16.00“ ‘ c- 0 II a d 8.00-1 1», ‘ .. D + ’ .1 g “ ... .1 0.00 r r e r r r 1 r g 40.00 . , . , . , r T , °1°331£2$ + D 0 32.00-1 A Model 3 (4.4.) d/D-O.273 _ , + Hook. (uv—m) (W140 Gap) 1 I 24.00-J, .. du d 16.00~ A .. O " 0‘ I A 8.00-1 3.,“ .. II a ’fi . ’ + ‘ 3 + 0 0.00 e r r , 1 1 1 f g 0.00 8.00 1 6.00 24.00 32.00 7.. [1/81 Figure 6.64: Apparent Viscosity as a Function of Average Shear Rate For Salad Dressing. Flag Impeller. a) Small Cap; b) Wide Gap. 227 differences in system geometry (impeller-to-cup diameter ration, d/D, and impeller blade height, b). Thus, flow curves can be directly detemw mined by only measuring the torque needed to rotate the impeller as a function of the rotational speed of the impeller. The selection of the model can be checked with data from a conventional viscometer when available. 6.3 QENEBAL REQOMMEHQAIIQNS Egg IE5 AEELICAIION QE MIXER VISCOMETRY This section presents the recommendations for a general procedure Untxing system and unknown power-law fluid). Figure 6.65 shows the flow diagram of the procedure. The procedure has been developed for use with a simple and easy to handle data collection system. However, it can be applied to different combinations of Viscometers and mixing systems. 6.3.1 Mixer System Used in This Study The procedure is as follows: 1. Maintain a constant temperature of the fluid. 2. Impeller and cup dimensions (d/D and d/b) must be known. 3. Other geometric parameters should remain constant for a set of experiments (fluid height, distance from bottom of impeller to bottom of container). 4. Check for time-dependent effects (there should be none for power law fluids). Agitate fluid for a period of 10 minutes to reach an equilibrium value of torque. 5. Select range of rotational speeds. The criterium to follow is 228 Re 5 10. Also, check for presence of surface waves or turbulence. The maximum allowable value of rotational speed, N, will be that which does not produce formation of surface waves on agitated fluid. Operating values of the rotational speed (rev/s) greater than 20 rpm are recommended. . Measure the torque in the shaft required to agitate the fluid, M, as a function of impeller rotational speed, N. . From log-log plot of M vs. N, evaluate the value of the flow behavior index, n, of the fluid. . To determine the flow curve of the fluid, a model is required. Therefore, expressions for an average mixing shear rate, fiav , and an average shear stress, a , were developed. Different models were av evaluated to determine the best model for flow curve determination. . If working with a paddle impeller [ 0.36 s (d/b) S 1.8], use Eqn. (6.19.1) or (6.19.2) if d/D 2 0.709 (small gap) and Eqn. (6.18.1) or (6.18.2) if [0.327 S (d/D) 5 0.515] (wide to medium gap). For a flag impeller [(d/b)-0.3], use Eqn. (6.21). Thus, 2/n . (D/d) n/2 7 _ 4, (d/b) N (6.18.1) av { [ (D/d)2/n _ 1 ] } or 2/n . (D/d) n/2 1 - 20 (d/b) (6.18.2) av [ (D/d)2/n _ 1 ] for a paddle impeller and 0.327 S d/D 5 0.515 , and 229 zfifi 73v - {4x [ (D/d’ J (d/b)n/2 } N (6.19.1) (D/d)2/n - 1 or Zfin . (D/d) n/2 7 - 20 (d/b) (6.19.2) 8" [ (1)/<1)”n - 1 J for d/D 2 0.709 (with 1 z b 2 S cm and 0.5 2 n 2 0.9). Flag impeller (b/d—0.5), k' - 4x [ (1)/d)“2 1 (6.21) 10. Determine the best model to approximate the average shear stress. A way to check the applicability of the shear stress equations [Model 1 (Eqn. (4.14); Model 2 (Eqn. (4.21) and Model 3 (Eqn. (6.23.1) and (6.23.2)] is to compare the values of torque calculated using these equations with the experimentally measured values of torque. The best equation will be the one that shows better agreement with the experimental data. Thus, Model 1 (concentric cylinders analogy with negligible end effects): M - 2xb (cl/2)2 aav (4.14) Model 2 (concentric cylinders analogy with end effects): 11. 12. 13. 230 __NL _‘Q_+; M 2 [ d 3 ] Gav Model 3 (Flag impeller; d- de ): 2 M - 2nb (dc/2) aav and (4.21) (6.23.1) (6.23.2) Develop plots of average shear stress versus average shear rate for the fluid (flow curves). The value of the consistency coefficient, m, can be evaluated by linear regression of the plots. Evaluate the apparent viscosity of the fluid from "a - m (iav ) n-l If available, compare results with data obtained with a traditional concentric cylinders viscometer. Make sure the range of shear rates are comparable. 6.3.2 New Mixer Viscometer System in this study is used, the following procedure is recommended: If an impeller other than a flag or a paddle impeller considered . Follow previous steps (l-7). . Evaluate k' of the particular impeller using the Slope method. The main reason for selection of this mixer viscometry method is its 231 combination (d/D.d/b, 1 I xeeo otner yariaoles " select impeller/cup <:E::> Add no . "noel“? FLAG Fino x wfin . traditional 3564/) ‘ d/b-O.3 mixer VlSC. methods ye constant (c.h, i.etcl check for 33 from find exorgsswon tor k‘ ' -. ' d tnus for mixin tine deoenden. 33v fr°r Eqn,(5.21) ‘“ the 1 ' Eqn. (6.18) cm) 0.709 svst;m agitate fluid for main. at 1 yes 123?? "add . ‘selected | iiyfrom ' .' la [Eqn.(6.19) v V calculate Gig n range of rotational speed of viscoeeter select Model (1. 2 or 3) 00 go to R.‘ 10 lower value ® yes yes I is appropri ate _ elasure torque H as a function of N find values of fluid consisten coeff..n. fro- linear regressi V free log-log plot of H vs. I. find flow behavior 1 evaluate * 7""1‘3‘ )fl-l l l ' 3.1.3:.“ Figure 6.65: Flow Diagram of General Procedure For Mixer Viscometry 232 simplicity. If differences in fluid properties are considered, the Viccccicy_mccching_mcchcc should be used. . Seek a range of variable where k' is constant (i.e., (d/D), (d/b) and N). . With the values of k' obtained in (2) find an expression for k' as a function of system geometry and fluid properties. As a start, use the ccncentric cyiigcerc cncicgy cppzoach developed in this study [Equations (6.18.1), (6.18.2), (6.19.1), (6.19.2), or (6.21)]. Different expressions may be obtained when different (d/D) ranges are considered. . From (3), find expressions for fiav' . Follow previous step (10) to find approximations of the aav of the fluid under study for the new mixing system. . Follow steps (ll-13) to develop the flow curves (shear stress-shear rate relationship). . Repeat procedure for different impeller/cup combination. 233 CHAPTER 7 SUMMARY’ARD CONCLUSIONS 1. A procedure for determination of the rheological behavior of power- law fluids using a mixer viscometer was developed. 2. Based on the concentric cylinders system analogy, an expression for the average shear rate in the mixing system can be determined. Average shear rates are a direct function of impeller to cup diameter ratios, impeller rotational speed and the value of the flow behavior index, n. 3. The size of the gap has a significant effect on the determination of expressions for the average shear rate in the mixing system. Hence, two different expressions were obtained for different d/D ranges: Eqn. (6.18) for 0.327 s d/D 5 0.515 and Eqn. (6.19) for d/D 2 0.709 when using paddle impellers. 2/n . (D/dl_ n/2 1 - a, (d/b) N (6.18) av { [ (D/d) 2/n _ 1 ] } 4 . (D/d) n/2 7 - 4, (d/b) N (6.19) a" { [(1)/d) 2/n - I] } 4. Impeller shape plays an important role in the development of equations for development of average shear stress-average shear rate 234 relationships for power—law fluids. Thus, the expression for the average shear rate for a flag impeller is different from the ones obtained for the paddle impellers [Eqns. (6.18) and (6.19)}, and is given by Eqn. (6.21). - k'N - 4n [ (1)/d)”2 iav ] N (6.21) . The analogy made with the concentric cylinders with the addition of end effects is the best model [Eqn. (4.21)] for approximation of the average shear stress of power-law fluids when agitated with a paddle impeller rotating in a cylindrical cup. Results apply for all the different impeller/cup combinations and power-law fluids tested in this study ( 0.327 s d/D 5 0.709; 0.36 S d/b S 1.8; 0.5 S n s 0.9). M - -49— {—9— + -%— ] aav (4.21) . The best model for approximating the average shear stress when agitating a standard power-law fluid with a flag impeller rotating in a cylindrical cup is given by the assumption of the flag as a cylindrical surface with two blades attached, with a total diameter equal to de , with negligible end effects [From Eqn. (6.23.1)], 2 nd e av 2 . Flow curves (average shear stress versus average shear rate) were 235 determined for power-law fluids without the need for calibration with Newtonian fluids. . The applicability of the WW procedure was experimentally verified with both ideal (standard) and actual (food product) power-law fluids. . Comparison of results obtained with the conccntcic cylinderc cncicgy procedure with those obtained using a conventional concentric cylinders viscometer show excellent agreement for the range of impeller and cup sizes investigated in this study. 10. The ccnccncgic cylinders cnciogy procedure is capable of estimating 11. 12. 13. the rheological behavior of a food product when using the mixer viscometer with the paddle impellers. The QOBCGDEELQ cylincccs cnaiogy procedure is simple and requires little data collection (torque measurements at selected values of impeller rotational speed) in comparison to the more tedious approach of established mixer viscometry methods. The developed concentric cylinder procedure has proven to yield excellent approximation of the rheological behavior of power-law fluids using a low-cost viscometer. Established mixer viscometry methods have been reviewed and evaluated for power-law (shear-thinning) fluids. 14. 15. 16. 17. 18. 236 System geometry (impeller and cup), fluid properties and operating conditions (impeller rotational speed) were varied to investigate the effect of these parameters on the value of the mixer proportionality constant, k', using traditional mixer viscometry methods. Traditional mixer viscometry methods (viscosity matching and slope) predict variations in the mixer proportionality constant, k', with geometry of the mixing system. The power curve method is very tedious and deviation from the basic assumption (48v - k'N) may occur. The mixer torque Curve and the slope methods are simpler with the slope method requiring less data handling. The common assumption of fiav - k'N, with k' a constant depending on the geometry of the impeller only, may lead to significant errors in the values of the average shear rate when measuring properties of shear-thinning fluids at low rotational speeds. The determined critical rpm range was found to be N S 20 rpm. The effect of system geometry and fluid rheological behavior becomes almost negligible at higher rotational speeds. Thus, the assumption of a constant k' is valid under these conditions (N > 20 rpm). There is a relation between the average shear rate and the type of fluid being agitated. However, the influence of the power-law rheological parameters, m and n, on mixer viscometry methods is not clearly understood. It seems that the more shear-thinning the fluid 9-" 19. 237 (lower value of n), the lower the average shear rate. However, it must be noticed that the fluid with lower n values in this study was also the more viscous fluid (larger value of m). Thus, further investigation is necessary to identify the effect of the power-law parameters in mixer viscometry methods. The shape of the impeller has an important influence on average shear rates as determined by established mixer viscometry methods since different values of k' are obtained for the two impeller shapes investigated in this study, especially at low values of rotational speed, N. ( - 238 CHAPTER 8 SUGGESTIONS FOR.PUTURB STUDY . To study the influence of the power-law parameters (n and m) on the value of the impeller proportionality constant, k'; one possibility is to vary the values of the rheological parameters and identify separate relationships for each parameter. . To test the developed procedure (concentric cylinders analogy) with different types of fluids and impeller shapes. . To test the developed procedure (concentric cylinders analogy) for a wider range of shear rates using different viscometers. . To review and evaluate the mixer viscometry methods when agitating fluids that do not obey the power-law model. The presence of a yield stress and time dependent characteristics should also be evaluated. . To extend the results from this investigation to scale-up of mixing tanks. . To evaluate shear fields in mixing systems. APPENDICES A u m u 239 APPENDIX.A DIMENSIONLESS ANALYSIS FOR MIXING VESSELS (NEWTONIAN FLUIDS) When using dimensional analysis to design an experiment, the first step is to define the pertinent quantities as shown in Table A.1. The number of required pi terms (dimensionless quantities) is defined from Buckingham's theorem as the number of pertinent quantities minus the number of dimensions. Therefore the number of pi terms required to design the experiment is 10 minus 3, which equals 7. By describing the motion of a fluid in a mixing vessel only in terms of length [LJ, time [T] and mass [M], the following set of dimen- sionless products is obtained (Rushton et al., 1950): 1 . Thus, fi1 52 53 _ N D «1 (D) - d p 000 fil 'fl2 3fi3 -1 nl-LTM-L T (M/L) D L: 51 ' 353 ‘ 1 ‘ 0 51 ‘ 1 T: ~62 - 0 M: 63 - 0 n1 - d/D = D/d By inspection, «2 - H/d n3 - b/d n‘ - L/d 240 7'5 - C/d fl1 32 53 With u - n/p, «6 (v) - d N p V then, 0 O 0 fil '32 3 BS 2 '1 L T M - L T ( M/L ) ( L /T ) L: 51 ' 3fi3 ' 2 ' 0 51 ' 2 1 o o o fi1 2 s 53 2 ' L T M - L T ( M/L ) (l/T ) L3 fl1 ' 3fl3 ' 1 ' 0 51 ' 1 T: ~82 + 2 - o p, - 2 M: fis-O 2 d N «7 - “g“ - Froude Number, Fr 51 52 53 ’1 N "a (P) ' d p P . Thus, 1 51 '5 O 0 0 2 3 33 L T M - L T ( M/L ) 2 3 ( ML /T ) 241 L1 31 ' 353 ’2 ' 0 51 ' 5 M: fls ' 1 - 0 fis - 1 5 a N ' 5 a P d N p - Power Number, P According to the Buckingham's theorem (Langhaar, 1981), the power consumption of mixer impellers is given by an.equathn10f the form: :82 193 34 :35 :86 .37 P - f [ (Re) (Fr) (D/d) (H/d) (L/d) (b/d) (c/d) ] or [Eqn. (3.91)] In the current investigation, (H/d), (L/d) and (c/d) remained constant. Therefore, Eqn. (3.91) becomes 242 Table A.l: Pertinent Quantities Involved In Fluid Agitation Processes (Newtonian Fluids) Pertinent Quantities Number Symbol Description Units Dimensions Independent Variables l d Impeller m L diameter 2 D Tank diameter m L 3 b Impeller blade m L - height ‘ 4 H Fluid depth m L 5 L Tank length m L _1 6 N Impeller rev/s T Speed 3 ,3 7 p Fluid kg/m ML Density ,1 ,1 8 n Fluid kg/ms ML T Viscosity 2 _2 9 g Gravitational m/s LT Acceleration Dependent Variable 2 3 10 P Power kgm/ 5 ML T The fundamental factors affecting the mixing,process are the configuration of the system, the behavior of the fluid and the process control variables (rotational speed). The most significant variables that can be manipulated to affect power consumption are rotational speed, N; impeller and tank geometry, (d, D, b); and fluid properties, p and 17 (Temperature dependency is built in p and 11. Therefore, tempera- ture was not included as one of the pertinent quantities for the analysis). 243 A.1 Development of Dimensionless Functions for Power-Law Fluids (Used in Slope method) When agitating non-Newtonian (power-law fluids), the viscosity n is replaced by the flow behavior index, n, and the fluid consistency coefficient, m (kg sn'Z/m). Thus, the power requirement is a function of P-f(d. N. p.m. n) 31 32 33 34 35 d N 0 - C P p m (n is dimensionless and enters in m) 2 3 31 32 33 3 34 “,2 35 0 - (ML /T ) L T (M/L ) (MT /L) L: 281 + fiz - 364 - 35 - 0 T: -3al - p. + (n-2)fl5 - 0 M: 6, + 34 + 35 - 0 Let fil - B4 - 0. Then, 35 - -l 32 ' ’3 33 ' -n-1 n _ P 1 3 m d Nn+1 Let 31 - 0 and 8‘ - 1. Then, 65 - -1 33 ' -n-2 244 32'2 APPENDIX.B EXPERIMENTAL RESULTS 3"— . '1 245 APPENDIX B EXPERIMENTAL RESULTS “1 Table 3.1: Regression Results of Po - ao(Re) (Fr)m2 (Paddle Impellers) Linear Multiple Regression.Analysis Regression Estimated Regression Estimated Standard t* Coefficient Cosfficient Error log a0 88.299 -- -- 011 -0.976 0.023 -43.53 a2 -0.023 0.023 -1.08 Analysis of Variance Sum of Degrees of Error Mean F Squares Freedom Squares Regression 48.346 2 24.173 -- Residual 0.585 312 0.022 1115.37 Totsl 481931 314 2 R = 0.988 Test of hypothesis for a2 : Cl: a2 - 0 C2: (12"0 For a level of significance of 0.05, t(0.975,w) = 1.960 * Since t - 1.08 < t(0.975,m), we accept C1 and conclude that a2 = O I 4 1 246 Table 3.2: Regression Results of Po - 190(Re)’81 (Fr)flz (Flag Impeller) Linear.Mu1tiple Regression.Ana1ysis Regression Estimated Regression Estimated Standard t* Coefficienc Qccffiicicnt Error log 60 12.567 -- -- 81 -0.983 0.045 -20.07 fig -0.056 0.045 —1.87 Analysis of variance Sum of Degrees of Error Mean F* Squsres Freedom Squares Regression 61.373 2 30.669 -- Residual 2.015 87 0.032 958.66 Totsl 63.388 90 2 R - 0.968 Test of hypothesis for 62 : C1: 62 - 0 Can/92,10 For a level of significance of 0.05, t(0.975,87) = 1.990 * Since t - 1.87 < t(0.975,87), we accept C1 and conclude that 62 = 0 247 Table B.3: Values of 81, fiz and 8, from k' - 81 (d/b)fi2 + 53 (Matching Method of Power Curves) For Power-Law Fluids 2 FLUID 51 52 33 R 1 5.029 2.618 10.159 0.985 2 5.898 1.764 7.734 1.000 yr- 3 4.137 1.459 7.179 0.971 Fluid 1: Hydroxypropyl Methylcellulose 1% (n=0.504, m=6.49 Pa sn) Fluid 2: Hydroxypropyl Methylcellulose 1.5% (n-0.374, m=28.42 Pa sn) Fluid 3: Hydroxypropyl Methylcellulose 2% (n-0.352, m-59.27 Pa 5“) 248 I I . I _ I I I I . I .0 J CMC 1.5% T: 25 C o o o I 0 144.01 0 ° "‘ '1 o d O F; 108.0" 0 -1 0- J o I H b 72.0- "‘ 4 o . 36.0.1 —1 fl 4 3 o o . f . I . r - I If I . 10.0 2 l ‘ I V r ' I T 8.0-4 ‘ H 4 o ‘ a) GJJW ‘ C, l J G 5.4 o 4.0“ O " § . O a O O 2.0-1 o - I 0 ° ° 0 o o, b CLO I r 71% 1‘ r r” ‘7 pi r 0 30 50 90 120 150 Figure 3.1: a) Shear Stress Versus Shear Rate; b) Apparent Viscosity Versus Shear Rate (Hydroxypropyl Methylcellulose 1.5%) F‘”_‘” a [P0] ’1- [PO 3]. 290.0 249 232.04 174uO-+ 116.0-1 58.0-1 0.0 T I ' I CMC 2% T= 25 C 25.0 20.0~ 15.0- 10.0-1 5.0- 0.0 BIO ' 100 Figure B.2: a) Shear Stress Versus Shear Rate; b) Apparent Viscosity Versus Shear Rate (Hydroxypropyl Methylcellulose 2%) 250 100 ' I ' T’ T‘ I I"T I ' I I I ' I I i x n=0.504 System {I 1 o n=0.374 80+ A 1130.352 (d/b‘1.8, d/D=00327) _, f— + 3 {2. 60-1 x 4 ::J J .gi 140-1 0 1 I x . 1 204 I, .. x A 4 3 2 ,a I O r r’ ' I ' ”Ti ' I ' I r* r r”r’ ' r r 60 r T ' l T ' I T I ' [ VT j r I r x n=0.504 System # 2 J o n80.374 431 1 “0.352 (d/b=1.0. d/0=0.327) q I J "a? - \\‘.36-I .2. 1 . .gi 124-1 5 a J x A '1 12‘1 a -I I , . b 0 r ‘T' f”r"r”r‘ vi I r I I r’ I I I I r 0 0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 N [rev/s] Figure B.3: Average Mixing Shear Rate as a Function of Rotational Speed of the Paddle Impellers. a) System 1; b) System 2 (Torque Curves - Matching Method) 250 ‘ I ' T ' T— T I ' 1 r 1 r T 1 l ' o n-O.374 80—1 A n=O.352 (d/b‘1.8, d/D=0.327) _ .—-. 1 { 60" x T ::J a a 404 o .I ah I I .. . I 204 . - X 5 4 3 2 a,J o . I . I . f . I . I I I . I I I I 60 T F ' I ‘ T ' I T I ' [ fl j r I r .1 x "’0'504 System {I 2 . o n=-O.374 48.1 A 0-0.352 (d/b=1.0. d/D=0.327) _I 1 1 F07 - A \ 35 J:; 4 .{L 24-1 5 .. ‘ x A '1 I 9 . b O r r T T T r I ' T ' I ' T r 00 02 04 0.6 08 1O 12 14 16 18 Figure B.3: Average Mixing Shear Rate as a Function of Rotational Speed of the Paddle Impellers. a) System 1; b) System 2 (Torque Curves - Matching Method) 251 50 I I ' I ' I 1 I ' ”T ‘j I ‘* I ' I ' ' I. " "-3.33: System # 9 3 O n- . 48.4 A n-0.352 (d/b-1.8. (II/030.709) I .— ‘ I Q 364 a .4 ::a J x I .g 24' -1 4 o 1 12 x X o . I . ~ ‘ 8 O ' r ”f ‘T’ r’ r4’r' rfi’* I ’ Ti’r’ T'*** T—“T 45 T T I I T V— T1 r r U T r r r 1 I I x "'33:: System I? 10 . 0 fl- . 364 A "-0.352 (d/b-1.0. d/D-O.709) _. I 2 4 TD... 27 -I \ l ‘ I: I . B 18% 3 a ‘?~ J . A 9-I . _ x ‘ ‘ . . b Figure 3.4: Average Mixing Shear Rate as a Function of Rotational Speed of the Paddle Impellers. a) System 9; b) System 10 (Torque Curves - Matching Method) 252 20 ' I r’ I I I T— I ' r I ‘T’ r‘ I ' r . x "-0-504 d/D-0.429 o 0-0.374 16m A III-0.352 d X I I H m o _ \ 124 . L ‘ 3 .. .5 8‘4 ; I x ‘ 44 _ " B 4 6 4 a 0 ' r’ ' r* r I I r7 r r r ”I f r* T’ r T” zotww.1..II..I-fI ‘ x 11-0.504 (”0.027 0 n-O.374 16+ 5 n-O.352 - J X F; 12- O -I \ ,. a I—l 1 8 .I ‘?~ (I z 1 O X 41 x o d J g ‘ . b O 1 I . VT I ,r, r r—er. I . I r r ' ff 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 N [1/rev] Figure 3.5: Average Mixing Shear Rate as a Function of Rotational Speed of the Flag Impeller. a) System 2; b) System 1 (Torque Curves - Matching Method) 253 50 - I I I T I T I T—I f T I r T'- — /.,-1,I n-0.504. d/Da-O.709 J x d/b-O.36 i 4o-I ,/ -—. I "a? 3.0-I ’x Q \ I," ", s; I a 4 B 20" ,”’ _‘I IN /’ . l 4 ,”’ ............... 1 10-1 ........ 2 - .I ’ ......... .I o "' """ a r r T r I r T r I I fil’ I r r r 1' 30 f r ‘11 I I j T I I— I' l 1 r I— I T I — d/b-1-8 n-O.352, d/D-O.327 x d/b-O.36 . 244 ’,,’ J ,I” 184 ’a” D 1.. [1/81 1 rA—g_J___L-~J-_._L_L-_1__ul-- A Figure 8.6: Average Shear Rate as a Function of Impeller Rotational Speed for Two Paddle Impellers (Torque Curves - Matching Method) 254 20 I. T IT I I T I x k («.5 n.0504, d/D-O.709 j o k' (50 rpm) ‘ 161 a k' (10 rpm) d/b-1'8 4: J” 4 5’3 124 '1- c1 4P .9. . ‘: é 8* I I' J I 4.4 ‘x ‘ J x“ I m D 0 T r T F T— T T r T— 15 1. l ' I ' I ' I ' x k (on) n-0.504. d/D-O.327 . o k' (50 rpm) 124 o k' (10 rpm) d/b‘1'8 " + J 7; '° ' I 9-I " o o. J. ‘ LII—l . 6% D -4 S . I. 3-4 q'x.x|- A I " ° ° a O I f r [T ' I t F t 0.0 10.0 20.0 30.0 40.0 50.0 in [1/81 Figure 3.7: Apparent Viscosity as a Function of Average Shear Rate for a Paddle Impeller (CMC 1%) (Torque Curves - Matching Method) 15 255 12% no [PO*S] X L‘(oq3 O k' (50 rpm) n '3 1" (10 rpm) I II I I T T «0.504, d/D-O.273 4 r ' l U r F X I T I n-o.504. d/D-O.591 I n— _1_- J-I_._l---...l... —_J.-__l—J .aL—_4_.J___4.--_.l.._i. ---l.-- L. _ J-- -- __J ' r’ .0 10.0 I r 20.0 30 0 'F- [1/81 50.0 Figure 8.8: Apparent Viscosity as a Function of Average Shear Rate For a Flag Impeller (CMC 1%) (Torque Curves - Matching Method) 256 l.OE-l-05 f I e I I I I I 3 d/D= 0.429 Flog Impeller 3. l.OE+O4 E 5 j 32 § m 1' 4’, 1000.0 "' 22: 4 3 A 100 rpm E + 50 10013 0 20 d U 10 3 a. o 5 1 215 10.0 I ‘ r I I l f IT I I l l.OE+05 * I I . I . I I I . d/D= 0.273 Flog Impeller j LOE+O4 '5— 3 ‘ as: 3 3 o 5 a? :3: @g '1 '53 1000.0 1.8 a? o '5 ..Es A 100 rpm .8 o E + 50 . n . 100.0 0 20 ' “a o 10 5 O 5 ‘ 10.0 I " .2'5 . t I l I , T 0.0 O 1 O 3 0.4 O 6 O 7 (l-n) 3 Figure 8.9: Plots of Dimensionless Functions [P/(md Nn+1)] Versus (l-n) For The Flag Impeller (Slope Method) 257 LOE+04 I I I I I I , I . , k‘ = 10.41 (1/rev) R3 = 0.981 ,__, 1000.0 3 I: a. ‘2 100.0 E I—l \\ o. 10.0 — d/D-0.327. d/b-0.6 1-0 I '* r** I I I I I I ' I 1 0E+04 I I I I I I I I I k‘ = 12.58 (1/rev) R3 =3 0.985 1 o — cup-0.327. d/b-1.0 - l ' I ' I I I r I I ”’fi 0.0 0.1 0.3 0.4 0.6 0.7 (l-n) 3 Figure 8.10: Plots of Dimensionless Functions [P/(md Nn+1)] Versus (l-n) For The Paddle Impellers (Slope Method) 258 100E+04 V I V r r I V T k' = 26.89 (1/rev) F?3 = 0.960 ... 1000.0 3; z a. 1: 100.0 .5. \\ o. 10.0 -— 410-04515. d/b-1.8 1-0 I I” I’ I’ I I I I I I I.oe+04 I . I . I l I k‘ = 10.00 (1/rev) R2 =3 0.960 .. 1000.0 1 3 L 2 100.0 . N- S - -_ I \\ o. 10.0 — d/D-O.327. d/b-O.45 1-0 I'* II ”1* II I I l' I I I I 0.0 0.1 0.3 0.4 0.6 (l-n) 3 Figure 8.11: Plots of Dimensionless Functions [P/(md Nn+1)] Versus (l-n) For The Paddle Impellers (Slope Method) 259 1 .OE-I-04 . I I I k' = 10.31 (1/rev) I?3 = 0.904 I. 1000.0 3 Z: a. 13 100.0 E. \\ O. 10.0 — Iva-0.515. d/b-O.6 1'0 I ' I I I f I r I I 1 OE+O4 f r I r r I I a k' = 15.99 (1/rev) F!3 = 0.960 ,_ 1000.0 3 I: a. 1: 100.0 E H \\ 0.. 10.0 1 o — ago-0.515. d/b-1.0 ' - I I I— I I I r I r I 0.1 0.3 0.4 0.6 0.7 (l-n) 3 Figure 3.12: Plots of Dimensionless Functions [P/(md Nn+1)] Versus (l-n) For The Paddle Impellers (Slope Method) 260 1.0E+04 I I I I I I I I I k‘ = 9.827 (1/rev) R3 = 0.920 1000.0 100.0 10.0 I I I I ' i O P / [mod'i-NWI’] LOE+04 I I I. I I I I I k‘ = 17.73 (1/rev) R2 =- 0.972 10.0 — d/D-O.709. (Vb-1.8 I r V I I I I' 0.0 0.1 0.3 0.4 0.6 (1 -n) 3 Figure 8.13: Plots of Dimensionless Functions [P/(md Nn+1)] Versus (l-n) For The Paddle Impellers (Slope Method) 261 1000.00 .--.....I I I ......I . ...I? one 1:: - d/D-O.327 100.00 F.. o I! an X E. 10.00 a- ° " I x- !) .x. 1.00 " ° I. x cub-0.30 . I d/b-(M 0.10 I . I ......I e. ......I I I I.-.I.? “(”3‘."I.IIII 1000.00 . I ..-...I I ...I-..I - ......II . one 1:: - d/D-O.515 100.00 :5 I. h—I 10.00 3'0 O I x- b 8.. 0 1.00 "II "' x «Vb-0.6 x I d/b-1.0 0.10 . ...II- - fl..." .-I.I.° ‘97".‘9. 0.01 0.30 1.50 Iofoo {03.00 7.. [1/81 Figure 3.14: Flow Curve For 1% wt% Aqueous Solution of CMC Determined Using the Mixer Viscometer with the Paddle Impellers 262 1000.00 I I ..III I I I . ..III I I I I. CMC 23 - d/D-0.327 100.00 x. 3' ° "" XI 0 O o- 8. O ‘-‘ X- o I x. o b IOJDO x. 01 ”I x d/b-OJG I (Vb-0.6 1.00 I WIN-II I I. .....I III... .4. .-.-..I ‘000000 I tilltvt vrvvu' v rrvvuu' r I U rrrt' CMC 23 - d/D-O.515 " file 0 '3' 100.00 a. 0 file £331 ”he: '3 ‘IO 10.00 in ’hO ' ' x d/b-(M I d/b-1.0 1.00I I . ...I-If I I . .....r ...-II?- “ébf‘fiI-Iq 0.01 O.1O 1 .00 10.00 1 00.00 7.. [1/81 Figure 8.15: Flow Curve For 2% wt% Aqueous Solution of CMC Determined Using the Mixer Viscometer with the Paddle Impellers 263 APPENDIX.C PROCEDURE FOR USE OF THE MIXER BROOKFIELD VISCOMETER FOR NEWTONIAN FLUIDS This section provides the user of Brookfield Viscometers with the information required to obtain viscosity readings with the "mixer inmellers" (flag and paddle type). This information consists of the impeller "factors" and viscosity ranges for use of the impellenfimq) combinations with a Brookfield Viscometer. These factors are equivalent to the Factor Finder supplied with the Viscometer for other spindle geometries. 6.1) Determination of Impeller Factors Brookfield Viscometers use a Factor for every spindle/speed combination the user selects. It is simply multiplied by the Viscometer reading to evaluate viscosity (in centipoise). Table 0.1 presents the values of the Factors for the flag and paddle impellers for evaluation of the viscosity of Newtonian fluids. The recommended range of viscosity for the Viscometer model is almo shown to assure proper use of the equipment. The procedure for data collection and analysis was: 1. The torque required to rotate the impeller, at each value of rotational speed for several Newtonian fluids of known viscosity at constant temperature, was measured. 264 Table C.1: Factors for Use of the Mixer Impellers with the Brookfield Viscometer Paddle Impellers FACTOR VISCOSITY RANGE (Pa 3) SYSTEM VISCOMETER MODEL Min - Max Min - Max HBTD ifiyTD, HBTD RVTD 1 45.1/N 5.7/N 4.51 - 451 0.57 - 57 2 33.8/N 4.2/N 3.38 - 338 0.42 - 42 3 22.7/N 2.8/N 2.70 - 227 0.28 - 28 4 19.0/N 2.3/N 1.90 - 190 0.23 - 23 5 15.5/N 2.0/N 1.50 - 150 0.20 - 20 6 45.0/N 5.0/N 4.50 - 450 0.50 - 50 7 29.0/N 3.3/N 2.90 - 290 0.33 - 33 8 21.0/N 2.3/N 2.10 - 210 0.23 - 23 9 34.0/N 4.1/N 3.40 - 340 0.41 - 41 10 22.5/N 2.8/N 2.25 - 225 0.28 - 28 Flag Impeller l 55.0/N 6.5/N 5.50 - 550 0.65 - 65 2 55.0/N 6.6/N 5.50 - 550 0.66 - 66 3 42.5/N 5.3/N 42.5 - 425 0.53 - 53 N in rpm Recommended RPM - 10 - 100 3 1 Pa 5 - 10 cp Viscosity ranges (in Pa 5): Maximum: Factor x 100 Minimum: Factor x 10 265 2. Plots of Newtonian viscosity, n, versus torque, M, for each value of rotational speed, N and each system geometry (impeller/cup combination) were made. Figure C.1b is a typical plot for a selected impeller/cup combination. 3. A relationship between Newtonian viscosity, 0, and torque reading, was found to be of the form: n - C1 M (C.l) where: 3 C1 - constant, s/m M - torque reading, N m 4. The proportionality constant, C1, was plotted as a function of N for each system geometry as in Figure CIla. 5. From Eqn. (0.1) and the plot of (4), the following expression for the Newtonian viscosity was obtained for the different systems (Table CIZ): B2 97 - B, (N/60) M (0.2) with N - impeller rotational speed, rpm 3 B1 - constant, l/m 32 - constant, -n Thus, the Newtonian viscosity, 0, can be calculated from the fixllow- ing equation: 0 - FACTOR * Torque (C.3) 266 3.00E+05 J -- 'Y- 4301.70;[x«(-'-0.3773)] I 4 14054-0541 - 4'. 5'" n 1 g 13054-054: - 75~ 4 I J H I I. 12054-054 - o C J I I 6.00E‘I‘04’4 ‘g "i 4 “‘ q 0.00 I "‘.""“."--,-~ -- W ------- - - I 3 0.00 0.40 0.80 1 .20 1.60 2.00 N [rev/3] 14 ' I ' I ' I I 12.. f r ,0 ..I + 1,, . 10-4 I, I J H J ll ’ d g g... d/b-0.36 I. 3:. 5 ' and P -- rpm-2.5 d 4... — rpm-5 .. -- rpm-10 ‘ 2 - rpm-20 - -- run-50 ‘ 0 I I . , I I'— m"? b 0.0000 0.0000 0.001 2 0.0010 0.0024 0.0030 Torque [Ntm] Figure C.1: a) Plot of 01 Versus Impeller Rotational Speed; b) Plot of n Versus Torque Reading (Paddle Impeller, d/b-O 45) Table C.2: Table for the fit of Eqn. 267 Paddle Ilpellers (C.2) SYSTEM 81 82 1 131.6462 -0.9541 0.999 2 97.2958 -0.9827 1.000 3 65.1144 -0.9753 1.000 4 54.3087 -0.9750 1.000 5 46.0179 -0.9773 1.000 6 139.0464 -0.8527 0.990 7 92.6439 -0.9l34 0.989 8 61.7416 -0.9388 0.995 9 99.0534 -0.9771 1.000 10 65.6778 -0.9905 1.000 Flag Lupeller 1 162.9035 -0.9816 1.000 2 158.2457 -0.9755 1.000 3 124.8972 -0.9973 1.000 268 B2 with FACTOR - 81(N/60) IFactors were converted to constants for use with the Brookfield display readings as follows: ,0 [FACTOR x spring constant (Nm) x 10 ] x N - FACTOR2 (C.3) FACTOR: with FACTOR - ——7§—- (C 4) Table C.2 shows average values for every system (at all values of N). The mixer impeller Factors for every impeller/cup combination presented in Table C.2 allow the user for direct determination of vis- cosity readings. Torque readings were converted to Viscometer display readings to facilitate the procedure. Table 0.3 shows the values of viscosity obtained with the dif- ferent impeller/cup combinations at two selected values of N (10 and 50 rpm). It may be seen that the maximum error obtained is about 10% which indicates that the prediction factors provide accurate estimation of the Newtonian viscosity. C.2) ro dure Fo D e inat o O ewton n Vi cosit The following procedure is a useful tool for analysis of vis- cosity data of food.products of unknown behavior. It is also a starting point from which more advanced techniques can be explored. The procedure for determination of Newtonian viscosity with the Mixer Brookfield Viscometer is as follows: 269 Table C.3: Values of Viscosity of a Newtonian Fluid (n - 4.84 Pa 3) Obtained with The Impeller Factors Paddle Impellers 0 (Pa 8) SYSTEM N (rpm) HBTD RVTD % Error 1 10 4.71 4.76 -2.7 «1.7 50 4.50 4.55 -7.0 -6.0 2 10 5.00 4.98 3.3 2.9 50 4.75 4.72 -1.9 -2.5 3 10 4.68 4.62 -3.3 -4.5 50 4.66 4.60 -3.7 -4.9 4 10 4.99 4.84 3.0 0.0 50 4.93 -- * 1.9 -- 5 10 4.95 5.11 2.3 5.6 50 4.84 -- 0.0 -- 6 10 5.25 4.67 8.5 -3.5 50 4.66 4.79 -3.7 -l.0 7 10 4.79 4.95 -l.0 2.3 50 4.53 4.66 -6.4 -3.7 8 10 5.29 4.99 9.3 3.1 50 4.96 -- 2.5 -- 9 10 4.85 4.67 0.2 -3.5 50 4.70 4.55 -2.3 -5.9 10 10 4.95 4.93 2.3 1.9 50 4.91 4.89 1.4 1.0 Flag Impeller 1 10 4.60 4.55 -4.9 -5.9 50 4.56 4.49 «5.7 -7.2 2 10 5.11 5.21 5.6 7.6 50 4.85 5.04 0.2 4.1 3 10 4.61 4.62 -4.7 —4.5 50 4.51 4.51 -6.8 -6.8 Reading out of range of viscometer 270 I Determine if the fluid is Newtonian. To do so, record display reading with impeller rotating at different values of rotational speed. A plot of the reading values as a function of N should give a straight line with slope equal to 1. . Select impeller/cup combination and value of rotational speed according to standard Brookfield procedure. Generally, a reading display of 10 or higher assures that proper selection of equipment has been made. . Read digital display from Viscometer. . Find FACTOR for selected system and N in Table 6.1. . To obtain viscosity readings (Pa 3), multiply Brookfield reading by the Factor at specified value of rotational speed (in rpm) from step (2). 271 REFERENCES Beavers, G.S., Yoo, J.Y. and Joseph, D.D. 1980. The free surface on a liquid between cylinders rotating at different speeds. Part III. Rheol. Acta 19, 19-31. Beckner, J.L. and Smith, JIM. 1966. Anchor-agitated systems: power input with Newtonian and pseudoplastic fluids. Trans. Instn. Chem. Engrs. 44, T224-T235. Begachev, V.I., Gurvich, A.R. and Braginski, L.N. 1980. A generalized method of power calculation in mixing highly viscous Newtonian and non- Newtonian media. Theor. Found. Chem. Engng. 14(1), 98-104. Bertrand, J. and Courdec, J.P. 1988. Numerical and experimental study of flow induced by an anchor in viscous, Newtonian and pseudoplastic fluids. Int. Chem. Engng. 28(2), 257-270. Bird, R.B., Stewart, W.E. and Lightfoot, E.D. 1960. Transport Phenomena. John Wiley & Sons Inc. Blasinski, H. and Rzyski, E. 1976. Mixing of non-Newtonian fluids with turbine, propeller and paddle agitator. Int. Chem. Engng. 16(4), 751- 754. Blasinski, HI, Kochanski, B. and Rzyski, E. 1970. The Froude number in the mixing process. Int. Chem. Engng. 10(2), 176-179. Boger, DIV. 1977/1978. A highly elastic constant-viscosity fluid. J. Non-Newt. Fluid Mech. 3, 87-91. Bohme, 0., V0 , R. and Warnecke, W. 1985. Die freie oberflache einer flussigkeit uber einer rotierenden scheibe. Rheol. Acta 24, 22-33. Bongenaar, J.J.T.M., Kossen, N.W.F., Metz, B. and Meijboom, F.W. 1973. A method for characterizing the rheological properties of viscous fermen- tation broths. Biotech. Bioengng. 15, 201-206. Bourne, J.R. 1965. Brit. J. Appl. Phys. 16, 1411. Bourne, J.R. and Butler, H. 1969. Power consumption of helical ribbon impellers in viscous liquids. Trans. Instn. Chem. Engrs. 47, T263-T270. Bourne, J.R., Buerli, M. and Renegass, W. 1981. Power and heat transfer to agitated suspensions: Use of heat flow calorimetry. Chem. Engng. Sci. 36, 782-784. Calderbank, P.H. and Moo-Young, MIB. 1959. The prediction of power consumption in the agitation of non-Newtonian fluids. Trans. Instn. Chem. Engrs. 37, 26-33. 272 Calderbank, P.H. 1958. Part I: The interfacial area in gas-liquid con- tacting with mechanical agitation. Trans. Instn. Chem. Engrs. 36, 443- 463. Calderbank, P.H. and Moo-Young, M.B. 1961. The power characteristics of agitators for the mixing of Newtonian and non-Newtonian fluids. Trans. Instn. Chem. Engrs. 39, 338-347. Castell-Perez, MIE. and Steffe, J.F. 1989. Using Mixing to evaluate rheological properties. Chapter 14 in Visggglagtig Ergperties of Solid, Elgi§_§n§_§§m1_§glig_figgg§ (Rao, M.A., ed.) Elsevier Applied Science Publishers Ltd., Barking, England. Castell-Perez, M.E., Steffe, J.P. and Morgan, R.G. 1987. Adaptation of a Brookfield (HBTD) viscometer for mixer viscometry studies. J. Texture Stud. 18, 359-365. Charles, M. 1978. Technical aspects of the rheological properties of microbial cultures. Adv. Biochem. Engng. 8, 1-61. Chavan, V.V. and Ulbrecth, J. 1973. Power correlation for off-centred helecal screw impellers in highly viscous Newtonian and non-Newtonian liquids. Trans. Instn. Chem. Engrs. 51, 349-354. Chavan, V.V. and Ulbrecht, J. 1972. Power correlation for helical ribbon impellers in inelastic non-Newtonian fluids. Chem. Engng. J. 3, 308-311. Chavan, V.V. and Mashelkar, RIA. 1980. Mixing of viscous Newtonian and non-Newtonian fluids. In Advances in Transport Ergggsses, New Delhi, 210-246. Chavan, V.V., Arumugam, M. and Ulbrecht, J. 1975. On the influence of liquid elasticity on mixing in a vessel agitated by a combined ribbon- screw impeller. AIChE J. 21(3), 613-615. Chavan, V.V., Jhaveri, A.S. and Ulbrecht, J. 1972. Power consumption for mixing of inelastic non-Newtonian fluids by helical screw agitators. Trans. Instn. Chem. Engrs. 50, 147-155. Chaplin, L. 1987. Mixing of highly viscoelastic fermentation broths. Presented at the 11th Conference on Mixing, August 1987, Henniken, NH. In Nauman, E.B., Etchells, A.W.III and Tatterson, 0.8. 1988. Mixing: The state-of-the-art. Chem. Engng. Prog. 84, 58-69. Collias, D.J. and Prud'homme, R.K. 1985. The effect of fluid elasticity on power consumption and mixing times in stirred tanks. Chem. Engng. Sci. 40(8), 1495-1505. Conti, RI, Sicardi, S. and Specchia, V. 1981. Effect of the stirrer clearance on particle suspension in agitated vessels. The Chem. Engng. J. 22, 247-249. Crochet, M.J. and Walters, K. 1983. Numerical methods in non-Newtonian fluid mechanics. Ann. Rev. Fluid Mech. 15, 241-260. 273 Deak, A., Havas, G. and Sawinsky, J. 1985. The power requirements for anchor, ribbon and helical screw agitators. Int. Chem. Engng. 25(3), 558-565. Dealy, J.M. and Vu, T.K.P. 1977/1978. The Weissenberg effect in molten polymers. J. Non-New. Fluid Mech. 3, 127-140. Ducla, J.M., Desplanches, H. and Chevalier, J.L. 1983. Effective vis- cosity of non-Newtonian fluids in a mechanically stirred tank. Chem. Engng. Comm. 21, 29-36. Edwards, M.F., Godfrey, J.C. and Kashani, M.M. 1976. Power requirement for the mixing of thixotropic liquids. J. Non-New. Fluid Mech. l, 309- 322. Eitelberg, G. 1983. Weissenberg effect and its dependence upon the experimental geometry. Rheol. Acta 22, 131-136. Ford, E.W. 1984. Rheological analysis of starch-thickened apricots using mixer viscometry techniques. M.S. Thesis, Department of Agricultural Engineering, Michigan State University, East Lansing, Michigan. Ford, EIW. and Steffe, J.F. 1986. Quantifying thixotropy in starch- thickened, strained apricots using mixer viscometry techniques. J. Texture Stud. 17, 71-85. Foresti, R. Jr. and Liu, T. 1959. How to measure power requirements for agitation of non-Newtonian liquids in the laminar region. Ind. and Eng. Chem. 51(7), 860-864. Godleski, S. and Smith, JIC. 1962. Power requirements and blend times in the agitation of pseudoplastic fluids. AIChE J. 8(5), 617-620. Green, S.J. 1953. Agitation in process design. Trans. Instn. Chem. Engrs. 31, 327-347. Griffith, D.L. and Rao, V.N.M. 1978. Flow characteristics of non- Newtonian foods utilizing a low-cost rotational viscometer. J. Food Science 43, 1876-1877. Haimoni, A. and Hannant, D.J. 1988. Developments in the shear vane test to measure the gel strength of oilwell cement slurry. Adv. Cement Res. 1(4), 221-229. Hall, K.R. and Godfrey, JIC. 1970. Power consumption by helical ribbon impellers. Trans. Instn. Chem. Engrs. 48, T201-T208. Hiraoka, SI, Yamada, I. and Mizoguchi, K. 1979. Two dimensional model analysis of flow behavior of highly viscous non-Newtonian fluid in agitated vessel with paddle impeller. J. Chem. Engng. Japan 12(1), 56- 62. Hirsekon, F.S. and Miller, SIA. 1953. Agitation of viscous solid-liquid suspensions. Chem. Engng. Prog. 49(9), 459-466. 274 Hoffman, A.H. and Gottenberg, W.G. 1973. Determination of the material functions for a simple fluid from a study of the climbing effect. Trans. Soc. Rheol. 17(3), 465-486. Holland, F.A. and Chapman, F.S. 1966. Liquid Mixing and Processing in firirrgd_1ank§. Chapman and Hall, Ltd. London Hyman, D. 1962. Mixing and agitation. Adv. Chem. Engng. 3 (Drew, Hoopes and Vermeilen, eds.) Joseph, D.D., Beavers, C.S., Cers, A., Dewald, C., Hoger, A. and Than, P.T. 1984. Climbing constants for various liquids. J. Rheol. 28(4), 325- 345. Kale, D.D., Mashelkar, R.A. and Ulbrecht, J. 1973. Drag reduction in rotational viscoelastic boundary layer flows. Nature Phys. Sci. 242, 29- 30. Keentok, Mo., Milthorpe, J.F. and O'Donovan, E. 1985. On the shearing zone around rotating vanes in plastic liquids: theory and experiment. J. Non-New. Fluid Mech. 17, 23-25. Kemblowski, Z. and Kristiansen, B. 1986. Rheometry of fermentation liquids. Biotech. Bioeng. 28, 1474-1483. Kemblowski, 2., Sek, J. and Budzynski, P. 1988. The concept of a rota- tional rheometer with helical screw impeller. Rheol. Acta 21, 82-91. Kraynik, A.M., Aubert, J.H., Chapman, R.N. and Gyure, D.C. 1984. The helical screw rheometer: A new concept in rotational rheometry. In Proceedings of the Annual Technical Conference of the Society of Plastics Engineers, 14 Fairfield Drive, PIO. Box 0403, Brookfield Center, Connecticut 06804. Krieger, I.M. and Maron, S.H. 1954. Direct determination of the flow curves of non-Newtonian fluids. III. Standardized treatment of vis- cometric data. J. Appl. Phys. 25(1), 72-75. Laughs“. ELL. 1980. W. Robert E. Krieger Publishing Company Inc., Huntington, New York. Leong, Y.K., Creasy, D.E., Boger, D.V. and Nguyen, Q.D. 1987. Rheology of brown coal-water suspensions. Rheol. Acta 26, 291-300. Mackey, K.L., Morgan, R.G. and Steffe, J.F. 1987. Effects of shear- thinning behavior on mixer viscometry techniques. J. Texture Stud. 18, 231-240. Magelli, F., Fajner, D. and Pasquali, G. 1986. Backmixing in multistage mixer columns - III. Slightly viscoelastic liquids. Chem. Engng. Sci. 41(9), 2431-2433. Magnusson, K. 1952. Effektbehouets beroende au den effektiva vis- kositeten vid omroring av strukturviskosa vatskor. IVA 23(2), 86-91. 27S Mashelkar, R.A., Kale, D.D. and Ulbrecht, J. 1975a. Rotational flows of non-Newtonian fluids. Part I. Turbulent flow of inelastic fluids around discs. Trans. Instn. Chem. Engrs. 53, 143-149. Mashelkar, R.A., Kale, D.D. and Ulbrecht, J. 1975b. Rotational flow of non-Newtonian fluids. Part II. Torque suppression with agitators. Trans. Instn. Chem. Engrs. 53, 150-153. Metzner, A.B. 1956. Non-Newtonian technology: Fluid mechanics, mixing and heat transfer. Adv. Chem. Engng. l, 77. Metzner, A.B. and Otto, RIE. 1957. Agitation of non-Newtonian fluids. AIChE J. 3(1), 3-10. Metzner, A.B. and Taylor, J.S. 1960. Flow patterns in agitated vessels. AIChE J. 6(1), 109-114. Metzner, A.B., Feehs, R.H., Lopez Ramos, H., Otto, R.E. and Tuthill, J.D. 1961. Agitation of viscous Newtonian and non-Newtonian fluids. AIChE J. 7(1), 3-9. Mitsuishi, N. and Hirai, N. 1969. Power requirements in the agitation of non-Newtonian fluids. J. Chem. Engng. Japan 2(2), 217-224. Mizushina, T., Ito, R. and Hiraoka, S. 1966. In Egg; and Mass Iransfier (in Russian), Vol 4. Nauka i Tekhnika, Minsk, 287. Nagata. S. 1975. WW3.- John Wiley & Sons. New York. Nagata, SI, Nishikawa, M., Tada, H. and Gotoh, S. 1971. Power consump- tion of mixing impellers in pseudoplastic liquids. J. Chem. Engng. Japan 4(1), 72-76. Nagata, S., Yamamoto, K., Hashimoto, K. and Naruse, Y. 1960. Studies on the flow patterns of liquids in a cylindrical mixing vessel, over a wide range of Reynolds number. Mem. Fac. Engng. Kyoto Univ. 22(1), 68-85. Nagata, S., Yokogama, T. and Yaragimoto, M. 1957. A study on the mixing of high-viscosity liquid. Chem. Engng. Japan 21, 278-286. Nguyen, Q.D. and Boger, D.V. 1983. Yield stress measurement for con- centrated suspensions. J. Rheol. 27(4), 321-349. Nguyen, Q.D. and Boger, D.V. 1985a. Direct yield stress measurement with the vane method. J. Rheol. 29(3), 335-347. Nguyen, Q.D. and Boger, D.V. 1985b. Thixotropic behavior of concentrated bauxite residue suspensions. Rheol. Acta 24, 227-237. Nguyen, Q.D. and Boger, D.V. 1987. Characterization of yield stress of fluids with concentric cylinder viscometers. Rheol. Acta 26, 508-515. 276 Nienow, AIW. and Elson, T.P. 1988. Aspects of mixing in rheologically complex fluids. Chem. Engng. Res. Des. 66, 5-15. Nienow, A.W., Wisdom, D.J., Solomon, J., Machon, V. and Vlcek, J. 1983. The effect of rheological complexities on power consumption in an aerated, agitated vessel. Chem. Engng. Comm. 19, 273-293. Oldshue, J.Y. 1983. Elgig_uixing_1§§hnglggy. McGraw-Hill Co., New York. Oliver, D.R., Nienow, A.W., Mitson, RIJ. and Terry, K. 1984. Power consumption in the mixing of boger fluids. Chem. Engng. Res. Des. 62, 123-127. Pavlushenko, 1.5. and Gluz, M.D. 1968. Shear rate gradient in a vessel with stirrer. Theor. Found. Chem. Engng. 2(1), 105-109. Pollard, J. and Kantyka, TIA. 1969. Heat transfer to agitated non- Newtonian fluids. Trans. Instn. Chem. Engrs. 47, T21-T27. ProkOpec, L. 1972. Agitated polymerization reactor for highly viscous media. Instn. Chem. Engng. 12(4), 712-719. Prud'homme, RIK. and Shaqfeh, E. 1984. Effect of elasticity on mixing torque requirements for Rushton turbine impellers. AIChE J. 30(3), 485- 486. Qiu, 0.0. and Rao, M.A. 1988. Role of pulp content and particle size in yield stress of apple sauce. J. Food Sci. 53, 1165-1170. Ranade, V.R. and Ulbrecht, J. 1977. Gas dispersion in agitated viscous inelastic and viscoelastic liquids. Proc. 2nd. European Conference on Mixing, Cambridge, F683-F100. Rao, M.A. 1975. Measurement of flow properties of food suspensions with a mixer. J. Texture Stud. 6, 533-539. Rao, M.A. 1977. Rheology of liquid foods - A review. J. Texture Stud. 8, 135-168. Rao, M.A. and Cooley, H.J. 1984. Determination of effective shear rates in rotational viscometers with complex geometries. J. Texture Stud. 15, 327-335. Reher, E. and Bohm, R. 1970. Chem. Tech. 22, 134. Cited by Rieger, F., Novak, Z. and Havelkova, ). 1988. The influence of the geometrical shape on the power requirement of ribbon impellers. Int. Chem. Engng. 28(2), 376-383. Rieger, F. and Novak, V. 1973. Power consumption of agitators in highly viscous non-Newtonian liquids. Trans. Instn. Chem. Engrs. 51, 105-111. Rieger, F. and Novak, V. 1974. Power consumption scale-up in agitating non-Newtonian fluids. Chem. Engng. Sci. 29, 2229-2234. 277 Roels, J.A., Van den Berg, J. and Voncken, R.M. 1974. The rheology of mycelial broths. Biotech. Bioengng. 16, 181-208. Rushton, J.H. and Oldshue, J.Y. 1953. Mixing - Present theory and prac- tice. Part I and II. Chem. Engng. Prog. 49(4/5), 161-168, 267-275. Rushton, J.H., Costich, E.W. and Everett, HIJ. 1950. Power characteris- tics of mixing impellers. Chem. Engng. Prog. 46(8), 395-404. Sadowska, M., Naczk, M., Sikorski, ZIE. and Ziminska, H. 1982. Effect of fish protein preparations on rheological properties of meat sausages. J. Texture Stud. 13, 371-379. Saravacos, 0.0. and Moyer, J.C. 1967. Heating rates of fruit products in an agitated kettle. Food Technol. 21(3), 372-376. Sawinsky, J., Havas, G. and Deak, A. 1976. Power requirement of anchor and helical ribbon impellers for the case of agitating Newtonian and pseudoplastic liquids. Chem. Engng. Sci. 31, 507-509. Schilo, D. 1969. Power requirement of tangential mixers in the mixing of non-Newtonian liquids. Chem.-Ing. tech. 5, 253. Sestak, J., Houska, M. and Zitny, R1 1982. Mixing of thixotropic fluids. J. Rheol. 26(5), 459-475. Sestak, J., Zitny, R. and Houska, M. 1986. Anchor-agitated systems: power input correlation for pseudoplastic and thixotropic fluids in equilibrium. AIChE J. 32(1), 155-158. Shamlou, P.A. and Edwards, MIF. 1985. Power consumption of helical ribbon mixers in viscous Newtonian and non-Newtonian fluids. Chem. Engng. Sci. 40(9), 1773-1781. Sinevic, V., Kuboi, R. and Nienow, A.W. 1986. Power numbers, Taylor numbers and Taylor vortices in viscous Newtonian and non-Newtonian fluids. Chem. Engng. Sci. 41(11), 2915-2923. Solomon, J., Elson, T.P. and Nienow, A.W. 1981. Cavern sizes in agitated fluids with a yield stress. Chem. Engng. Comm. 11, 143-164. Solomons, G.L. 1971. Fermentation equipment. Adv. Appl. Micro. 14, 231- 247. Steffe, J.F. and Ford, E.W. 1985. Rheological techniques to evaluate the shelf-stability of starch-thickened, strained apricots. J. Texture Stud. 16, 179-192. Steffe, J.F., Castell-Perez, M.F., Rose, K. and Zabik, MIE. 1989. Rapid testing method for characterizing the rheological behavior of gelatiniz- ing corn starch slurries. Cereal Chemistry 66(1), 65-68. 278 Takahashi, K., Yokota, T. and Konno, H. 1984. Power consumption of helical ribbon agitators in highly viscous pseudoplastic liquids. J. Chem. Engng. Japan 17(6), 657-659. Tamura, M.S., Shoemaker, C.F. and Henderson, J.M. 1988. Helical screw rheometer for on-line viscosity measurements. Paper # 519, presented at IFT Meeting, New Orleans, LA. Thomas, R.H. and Walters, K. 1964. The motion of an elastico-viscuos liquid due to a sphere rotating about its diameter. Quart. J. Mech. and Appl. Math. 17, 39-42. Tiu, C. and Boger, D.V. 1974. Complete rheological characterization of time-dependent food products. J. Texture Stud. 5, 329-338. Treybal, R.E. 1956. Mechanically aided liquid extraction. Adv. Chem. Engng. 1, 229-329. Ulbrecht, J. 1974. Mixing of viscoelastic fluids by mechanical agita- tion. The Chem. Engr. June, 347-367. Ulbrecht, J.J. andeau, P. 1985. Mixing of viscous non-Newtonian fluids. Chapter 4 in Mining gr Liguigs by Mecnnnignl Agitation. V01 1. (Ulbrecht, J.J. and Patterson, C.K., eds.). Gordon and Breach Science Publishers, New York. Vocadlo, J.J. and Charles, M.E. 1971. Measurement of yield stress of fluid-like viscoplastic substances. The Can. J. Chem. Engng. 49(10), 576-582. Voisey, P.W., Paton, D. and Timbers, GIE. 1977. The Ottawa starch viscometer- A new instrument for research and quality control applica- tions. Cereal Chem. 54(3), 534-557. Walker, 0.8., Ross, A.S., Wrigley, C.W. and McMaster, G.J. 1988. Accelerated starch-paste characterization with the Rapid Visco-Analyzer. Cereal Foods World 33(6), 491-494. Whalen, P.J., Anantheswaran, R.G. and Shahani, K.M. 1987. Power consump- tion of a whey-corn mash as compared to a full corn mash during liquefaction. Biotec. Letters 9(7), 523-528. White, A.M. and Brenner, E. 1934. Studies in agitation. V. The correla- tion of power data. AIChE J. 30, 585-597. White, A.M., Brenner, E., Phillips, CIA. and Morrison, M.SI 1934a. Studies in agitation. IV: Power measurements. AIChE J. 30, 570-584. White, J.L., Chankraiphon,S. and Ide, Y. 1977. Rheological behavior and flow patterns around agitators in polymer solutions. Trans. Soc. Rheol. 21(1), 1-18. Wichterle, K. and Wein, 0. 1981. Threshold of mixing of non-Newtonian liquids. Int. Chem. Engng. 21(1), 116-120. 279 Wichterle, K., Kadlec, M., Zak, L. and Mitschka, P. 1984. Shear rates on turbine impeller blades. Chem. Engng. Commun. 26, 25-32. Williams, R.W. 1979. Determination of viscometric data from the Brookfield R.V.T. viscometer. Rheol. Acta 18(3), 345-359. Wood, FIW. and Goff, T.C. 1973. The determination of the effective shear rate in the Brabender Viscograph and in other systems of complex geometry. Die Starke 25, 89-91. Yagi, H. and Yoshida, F. 1975. Gas absorption by Newtonian and non- Newtonian fluids in sparged agitated vessels. Ind. Engng. Chem. Process Des. Dev. 14(4), 488-493. Yap, C.Y., Patterson, W.I. and Carreau, P.J. 1979. Part III. Non- Newtonian fluids. AIChE J. 25(3), 516-521. Yoshimura, A.S., Prud'homme, R.K., Princen, H.M. and Kiss, AID. 1987. A comparison of techniques for measuring yield stresses. J. Rheol. 31(8), 699-710.