You are on page 1of 14

Article pubs.acs.

org/IECR

Pilot-Scale Examination of Mixing Liquid into Pulp Fiber Suspensions in the Presence of an In-Line Mechanical Mixer
Wisarn Yenjaichon,* John R. Grace, C. Jim Lim, and Chad P. J. Bennington
Department of Chemical and Biological Engineering, University of British Columbia, Vancouver, British Columbia, Canada V6T1Z3 ABSTRACT: The quality of liquid mixing into the main stream for an in-line mechanical mixer was investigated for water and pulp suspensions over a range of mass concentrations (03.0%), main-stream velocities (0.53.0 m/s), jet velocities (3.812.6 m/s), and rotational speeds (0800 rpm) based on electrical resistance tomography and a modied mixing index, derived from the coecient of variation of conductivity values. The mixing quality was worse when the jet penetrated to the far wall of the pipe for all ber mass concentrations investigated, whereas this only applied at higher mass concentrations without the impeller. For water ow, the residence time had a signicant eect on mixing at higher impeller speeds. With the impeller present, the mixing quality in pulp suspensions improved substantially and was similar to that for water when the ow approached the turbulent regime, with a considerably lower main-stream velocity required for mixing compared to a tee mixer alone. At higher mass concentrations, the energy supplied was insucient to provide the same level of turbulence as that in water, even at the highest main-stream velocity and impeller speed examined. Improved mixing with increasing impeller speed primarily occurred in the high-shear zone around the impeller, with turbulence decaying rapidly downstream, likely aided by reocculation.

1. INTRODUCTION Good mixing is essential in pulp bleaching. Poor mixing in a bleach plant causes poor chemical contacting, leading to poor product qualities such as lower pulp brightness and cleanliness. Additional bleaching chemicals are often applied to improve these qualities. However, excessive use of chemicals increases production costs and has an undesirable eect on the product strength. Mixing in bleaching processes is usually accomplished by in-line mixers ahead of bleaching towers, reactors with sucient residence time to complete the bleaching. Chemicals are injected ahead of, or inside, these static and high-shear mixers. The mixers predistribute the chemicals to achieve optimum product quality and chemical usage. In modern bleaching processes, in-line mixers have replaced continuously stirred vessels because of their increased energy dissipation rate per unit volume. Details of mixers in bleaching operations were summarized by Bennington.1,2 Static mixers are normally used in low-consistency applications (ber mass concentration, Cm, <5%), such as the rst chlorine dioxide bleaching (D0) stage, for which the suspension yield stress is relatively low. The mixing is achieved by dividing and recombining the ow or by generating turbulence. For high-shear mixers, chemicals are injected into zones of intense shear created by high rotational speed across narrow gaps through which the suspensions ow. The resulting turbulence leads to rapid mass transfer and a uniform distribution of chemicals. High-shear mixers are normally utilized in medium-consistency (8% Cm 16%) applications, such as chlorine dioxide bleaching and oxygen delignication. Experimental studies on the in-line mixing of pulp suspensions have been conducted on both laboratory and industrial scales. Most studies were carried out to evaluate the mixing performance and investigate the benets of improved mixing after the installation of high-shear mixers in bleaching processes. Past techniques include measurement of the uniformity of inert tracers,36 residual chlorine concentration,7
2012 American Chemical Society

and a radioactive tracer8 as well as alternative techniques such as temperature proling around process piping.3,9,10 Most techniques are not continuous, because pulp samples need to be taken from the process, and time-consuming because of the large number of samples at dierent locations required to give a clear picture of mixing. Temperature proling is noninvasive, continuous, and less tedious, but it quanties mixing only at the pipe periphery and not over the entire cross section, and a suitable temperature dierence between the chemical and pulp suspension must exist. Electrical resistance tomography (ERT) is a nonintrusive technique used to determine the distribution of electrical conductivity in process vessels from measurements at the vessel periphery. The ERT system injects an electrical current from a pair of neighboring electrodes and measures the voltage dierences between the remaining pairs of electrodes. This process is repeated by injecting a current for the other pairs of neighboring electrodes, and the cross-sectional distribution of the electrical conductivity is determined.11 ERT has been utilized extensively in various processes including pulp and paper systems. Recently, this technique has been used to evaluate the mixing performance of in-line mixers in pulping processes. Yenajaichon et al.12 evaluated the mixing eciency of an industrial static mixer in a chlorine dioxide bleaching stage. ERT was successful in monitoring the temporal variation of the mixing index, based on the coecient of variation (CoV), as a function of the operating conditions, including the chemical ow rate, suspension ow rate, and ber mass concentration. Kourunen et al.13 applied ERT to assess the performance of a pilot-scale medium-consistency mixer and
Received: Revised: Accepted: Published:
485

April 2, 2012 September 1, 2012 November 30, 2012 November 30, 2012
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research compared the results with those from temperature measurement. Mixing in a 10 wt % softwood pulp suspension owing at a mean velocity of 2.1 m/s was again quantied based on the CoV. ERT clearly indicated worse mixing when the tracer (cold water, air, or steam) was introduced and improved mixing when the mixer was turned on, with results similar to those from temperature proling. However, few trials were performed. Understanding in-line jet-mixing behavior is essential for investigating in-line mechanical mixers. For a turbulent jet discharging perpendicularly into a fully developed turbulent pipe ow, the ow is divided into three regions along the pipe: a ow-establishment region, a near-eld region, and a far-eld region.14,15 Jet mixing dominates the transport mechanism in the rst two regions, whereas the spread of the tracer in the far eld is governed by the velocity and turbulent diusion. A number of studies have been carried out to characterize the jet mixing of Newtonian uids downstream of T-junctions in the rst two regions for various jet-to-pipe velocity and diameter ratios. For non-Newtonian pulp ber suspensions, there have been few studies on jet mixing in pipelines. Yenjaichon et al.16 investigated the mixing of liquid tracer injected into a softwood pulp suspension ow in the jet-dominated regions for a range of conditions, with a modied mixing index based on CoV. Mixing in water at a specic jet-to-pipe diameter ratio improved with increasing jet-to-pipe velocity ratio, but was almost independent of the main-stream velocity for a specic velocity ratio. For suspension ow, however, both the main-stream velocity and jet penetration were major factors. Mixing improved with increasing main-stream velocity due to plug disintegration in the core and approached that in water for the turbulent-ow regime. For a dilute suspension, mixing improved with increasing jet velocity, whereas mixing was worse when the jet attached to the far wall of the pipe for higher suspension concentrations. The results also showed that jet mixing alone was insucient to provide good mixing for higher suspension concentrations, even at high main-stream velocities. Yenjaichon et al.17 extended the study to hardwood pulp suspensions and found that the jet-mixing behavior was similar to that for the softwood pulp. In the turbulent-ow regime, mixing for a dilute hardwood suspension was, however, better than that for water, whereas it was worse for a dilute softwood suspension, likely because of the inuences of ber-turbulence interactions in modifying turbulent structures in the bulk. The jet-to-pipe velocity ratios for optimum mixing and for a jet penetrating to the center of the pipe were also summarized. This paper applies the concepts of jet mixing to examine the mixing of pulp suspensions in the presence of an in-line mechanical mixer. One of the most popular types of in-line mechanical mixers is a rotor-stator mixer, consisting of a rotor that spins at a high speed inside a xed stator, providing short residence time and high energy dissipation rate. The design and characteristics of rotor-stator mixers were summarized by Atiemo-Obeng and Calabrese.18 Early studies on in-line rotor-stator mixers were carried out to evaluate micromixing for fast azo-coupling reactions19,20 and to determine power characteristics for foam generation.21,22 Recently, Baldyga et al.23 investigated the breakup of nanoparticle clusters in nanosuspensions by an inline Silverson rotor-stator mixer based on computational uid dynamics. The results indicated strong dependence of particle breakage on the rotor speed and suspension ow rate. Increasing the ow rate (less residence time) decreased the extent of particle breakup, especially at higher rotor speeds. The mean particle size did not decrease signicantly with increasing
486

Article

rotor speed at higher ow rate, implying less inuence of the rotor speed at shorter residence times. Baldyga et al.24 showed that the model predictions agreed well with the experimental results of particle breakage by this in-line mixer. Cooke et al.25 developed an expression for the power consumption of a Silverson high-shear in-line rotor-stator mixer to describe the impact of the rotor speed and ow rate for Newtonian and shear-thinning non-Newtonian uids. Hall et al.26 investigated droplet breakup by this in-line mixer. The emulsion drop size was found to decrease signicantly with increasing rotor speed, but was almost independent of the ow rate. Hall et al.27 showed that increasing the ow rate increased the total energy dissipation rate, with a minimal change in drop size, whereas the drop size decreased signicantly with increasing rotor energy dissipation rate, with residence times maintained similar. The mixing quality in suspensions depends on the ow regime, contact between bers, and occulation. Pulp rheology is strongly related to the suspension ow behavior. A pulp suspension behaves as a solid at stresses less than the yield stress, whereas it exhibits shear thinning for stresses exceeding the apparent yield stress. Further increasing the shear stress causes the suspension to behave as a Newtonian uid.28 Because the shear stress decreases with the radial distance from the pipe wall, a core region, where the stress is less than the yield stress, has solidlike properties and moves as a rigid plug, whereas the outer region, between the plug and pipe wall, behaves as a uid. Once the shear stress exceeds the yield stress throughout the suspension, the plug disappears and the ow becomes turbulent. The ow regimes for suspension ow in pipes are labeled plug, mixed, and turbulent.29 The suspension ows as a plug at low velocity. At higher velocity, a water layer surrounds the plug, and this layer becomes turbulent, with signicant stresses destroying the plug as the velocity increases. The plug, however, persists in this mixed-ow regime. At higher velocity, turbulent stresses disrupt the plug completely, and the suspension ow becomes turbulent. Flocculation occurs when suciently closely spaced bers collide and tangle with each other. The propensity for occulation depends on the number of bers in the volume swept by the length of a single ber called the crowding number,30 modied to a mass-based expression31

Nc

5CmL2

(1)

where Cm is the ber mass concentration, L the ber length, and the ber coarseness (ber mass per unit length). For Nc < 1, bers are free to move relative to one another, and occulation generally does not occur. At Nc = 60, bers become restrained by three-point contact and locked into a network in a bent conguration.32 At higher crowding numbers, there are continuous contacts among bers, and the ber mobility decreases signicantly. Martinez et al.33 dened a gel crowding number at Nc = 16. Below this value, the suspension behaves as essentially dilute, with bers moving freely of one another. For 16 < Nc < 60, bers interact and occulate, but the ber mobility persists. Celzard et al.34 showed that the connectivity and rigidity thresholds of ber networks correspond to Nc = 16 and 60, respectively. Suspension rheology and its impacts on the suspension behavior are summarized by Kerekes.35 The crowding number shows the level of ber contacts and probability of ber occulation in owing systems. It is, hence, a useful indicator for investigating suspension mixing.
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research

Article

2. EXPERIMENTAL DETAILS The experiments were conducted in a pilot-scale ow-loop facility, the details of which are provided in a previous paper.16 The acrylic test section, 76.2 mm in diameter and 1.88 m in length, included eight ERT sensor planes spaced at 250 mm intervals. For each of these planes, 16 circular stainless steel electrodes of 6.35 mm diameter were spaced at 22.5 intervals around the pipe periphery. The rst sensor plane was located 67 mm upstream of the side-stream injection, and the impeller was 104 mm downstream of the injection point. The mechanical mixer system consisted of a 1/2 HP motor (Leeson Electric, USA), a 1.41 N-m rotating torque sensor with an integral optical encoder (Omega Engineering, Inc., USA), a 12.7-mm-diameter stainless steel shaft, and a 63.5-mm-wide octagonal impeller with a 38.1-mm-diameter hole in the middle to prevent excessive force on the impeller when it was perpendicular to the ow. The octagon minimizes the gap between the impeller and internal wall of the pipe, hence maximizing the shear. A schematic diagram of the test section and in-line mechanical mixer is shown in Figure 1. The torque

tests showed that this did not have a signicant eect on mixing in the narrow range of temperatures investigated. To ensure fully developed ow in the main and side streams before they were mixed, the test section was 5 m downstream of the closest pipe bend, and the injection tube length-to-diameter ratio was 47, compared with an entry length of approximately 27Dj for the highest side-stream velocity investigated. The softwood pulp was Northern bleached softwood kraft with bers of arithmetic average diameter 27.6 m, coarseness 0.129 mg/m, and length-weighted average length 2.52 mm. The hardwood pulp was bleached maple kraft with bers of arithmetic average diameter 16.6 m, coarseness 0.066 mg/m, and lengthweighted average length 0.60 mm. Rheological characteristics for pulp suspensions almost identical with those examined in the present work are reported in previous work.28 An ITS P2000 ERT system (Manchester, U.K.), connected to the electrodes via 2-m-long coaxial cables, was used for data acquisition. The excitation frequency was 9.6 kHz and the injection current was 15 mA, with image reconstruction based on a linear back-projection algorithm. The data were analyzed in MATLAB 7.0 (MathWorks, Inc., Natick, MA). The mixing quality was assessed based on the modied mixing index16
M = M m 2 Ms 2 MFS
(2)

where Ms is the system mixing index in the absence of tracer (brine solution), Mm the mixing index under the given test conditions, derived from the CoV of the individual conductivity values in each of 316 pixels in each cross-sectional plane
= y
i = 1(yi y )2 n1
n

Mm =

(3)

and MFS is the mixing index for fully segregated ow at the same mean concentration, expressed in terms of the mainstream salt concentration Cp, side-stream salt concentration Cj, main-stream volumetric ow rate Qp, and side-stream volumetric ow rate Qj:
MFS = Q pQ j Q jC j + Q pCp |Cp C j|
(4)

Note that M is 1.0 (i.e., 100%) for fully segregated ow, decreases as the mixing quality improves, and reaches 0 for perfect mixing.
Figure 1. Schematic of the test section and in-line mechanical mixer (all dimensions in millimeters).

from the impeller was not measured because the frictional torque losses at the shaft seals exceeded the impeller torque and depended on the condition of the seals. The losses increased substantially as small bers entered the seal gaps and decreased signicantly when the seals became worn. Water or pulp suspension ow from the tank was mixed with a brine solution, injected at a constant ow rate with concentration from 0.9 to 2.4 g/L, at a 90 T-junction ahead of the impeller, which rotated at constant speed. The temperature of the main stream was always 20 5 C, with the temperature dierence between the main and side streams <10 C. The pulp rheology varies with the temperature, but our
487

3. RESULTS AND DISCUSSION 3.1. Eect of the Velocities and Impeller Rotational Speed on the Mixing Quality in Water. All tests were in the turbulent-ow regime for both the main stream (Rep = 38100 229000) and jet stream (Rej = 975048600). Figure 2 shows the eect of the main-stream velocity on the mixing quality with the impeller rotating at N = 400 and 800 rpm. Error bars in this and subsequent gures correspond to 90% condence intervals. At N = 400 rpm, the modied mixing index decreased, i.e., mixing quality improved, downstream and was almost independent of the main-stream velocity at the same jet-topipe velocity ratio, indicating a negligible eect of the residence time on mixing, as shown in Figure 2a. The data consistently showed the independence of the main-stream velocity at constant velocity ratios, similar to ndings for a tee mixer
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research

Article

The eect of the impeller speed on the mixing quality for water at Up = 1.0 and 3.0 m/s is presented in Figure 3. Two

Figure 2. Modied mixing index as a function of the dimensionless distance downstream, x/D, for a Newtonian uid (water) at Up = 1.0, 2.0, and 3.0 m/s and Dr = 0.05 with the impeller rotating at N = (a) 400 and (b) 800 rpm. Figure 3. Modied mixing index as a function of the dimensionless distance downstream of injection for water at Dr = 0.05, almost constant velocity ratios, and various rotation speeds at Up = (a) 1.0 and (b) 3.0 m/s.

alone.16 Figure 2b illustrates the inuence of the main-stream velocity on the quality of mixing at a higher rotational speed, 800 rpm. Improved mixing was observed with decreasing mainstream velocity, with considerably better mixing quality at Up = 1.0 m/s. The dierence was greater at higher rotational speed because the inuence of the residence time in the high-shear zone around the rotating impeller on mixing was more signicant at higher rotational speed. At a lower impeller speed (N = 400 rpm), however, the energy dissipation rate was lower, so that the residence time in the high-shear zone had an insignicant inuence on mixing. For turbulent jets in crossing pipe ow, jet mixing dominates the transport mechanism in the ow-establishment and neareld regions, whereas the velocity and turbulent diusion of the pipe ow are dominant in the far eld.14,15 The variation in the tracer concentration decays exponentially with the distance downstream in the far-eld region, solely because of turbulent mixing in the pipe ow with a negligible eect of jet mixing from the rst two regions. A plot of variance of the tracer concentration versus distance (log-normal scale) is therefore rst-order only in the far-eld region, >30 pipe diameters downstream. The experiments conducted in this study were in the jet-establishment and near-eld regions, with the furthest measurements at 22 pipe diameters, with an impeller also present downstream of injection. The plots were therefore not rst-order, likely because of the inuences of both the jet and impeller.
488

congurations of the impeller when staticparallel and perpendicular to the owwere also examined. At Up = 1.0 m/s, the mixing quality clearly improved when the impeller was added, as illustrated in Figure 3a. The stationary impeller parallel to the ow provided better mixing than the tee alone, whereas the mixing quality of water ow with the perpendicular orientation was signicantly better than that for the parallel one and was similar to that for the mechanical mixer at N = 400 rpm. For this velocity, mixing improved profoundly as the rotational speed increased. The impeller speed and conguration had less inuence on mixing at Up = 3.0 m/s, as shown in Figure 3b. At higher velocity, the uid passing through the high-shear zone around the rotating impeller likely spent less time there. An increase in the impeller speed therefore did not improve mixing signicantly, with the mixing quality similar to that of the perpendicular static mixer. At a lower main-stream velocity, however, the uid stayed longer in the high-shear zone, with faster energy dissipation with increasing rotational speed, leading to signicantly improved mixing. The results were similar to those from previous studies on an in-line rotor-stator mixer,23,24 where for a high rotor speed, particle breakup occurred mainly around the inner rotor at a low ow rate (long residence time) and around the outer rotor as well at a high
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research ow rate. For a low rotor speed, however, particle breakage occurred in the region where the outer rotor operated for both low and high ow rates. This implies that the residence time is more signicant at higher rotor speed. The results also showed that the mean particle size did not decrease signicantly with increasing rotor speed at higher ow rate, consistent with the present work, indicating less inuence of the impeller speed on mixing at a higher main-stream velocity. Figure 4 illustrates the inuence of the jet-to-pipe velocity ratio at N = 400 rpm. At this impeller speed, the residence time,

Article

Figure 4. Modied mixing index as a function of the dimensionless distance downstream of injection for various jet-to-pipe velocity ratios, R, with water at Dr = 0.05 and N = 400 rpm.

and hence the main-stream velocity, had less eect on mixing, and the mixing quality depended strongly on the jet-to-pipe velocity ratio. The mixing quality improved signicantly with increasing velocity ratio as the mixing mode changed from wall source (R < 4) to jet mixing (4 R 10) and jet impaction (R > 10). The results for the in-line mechanical mixer were similar to those for a tee mixer described previously.16 However, a signicant dierence was observed at R = 12.2 when the jet stream reached the far wall of the pipe ahead of the impeller. The mixing quality was worse than that for the jet penetrating to the center of the pipe, behavior not observed for the tee mixer alone. The mixing was more ecient when the jet penetrated to the core of the pipe, likely because of higher energy dissipation when the jet impinged on the rotating impeller. At very high velocity ratios (R > 16), energy dissipation from jet impingement on the pipe wall and rotating impeller was profound, resulting in ecient downstream mixing. Tomographic images showing jet penetration and downstream mixing appear in Figure 5. Because of misalignment between the electrode position in the ERT reconstruction process and the actual electrode position in each sensor plane, the top of the pipe in the tomographic image is rotated counterclockwise by 11 from the vertical axis, and a corresponding corrective rotation is applied to all tomographic images. From these images, the injected tracer or brine solution is represented by the high-conductivity region in red and yellow, whereas the main stream is represented by the lowconductivity blue region. The tracer was injected and reached the impeller between planes 1 and 2. The disappearance of the red and yellow colors from plane 2 to plane 8 showed improved mixing downstream. The jet penetrated to the core of the pipe at R = 6.15, reached the far wall of the pipe ahead of the
489

impeller at R = 12.2, and impinged on the far wall and recirculated around the pipe periphery ahead of the impeller at R = 24.6. 3.2. Eect of the Main-Stream Velocity on the Mixing Quality in Pulp Suspensions. Figure 6 plots the modied mixing index for various main-stream velocities for a softwood pulp suspension at Cm = 0.5%, N = 400 rpm, almost identical jet-to-pipe velocity ratios, and identical diameter ratio. Unlike mixing in water at similar velocities and rotational speed, pulp suspension mixing depended strongly on the main-stream velocity because of ow regime dierences. The mixing improved signicantly when the ow regime changed, and mixing quality approached that for water when the ow was turbulent for Up 2.0 m/s. The ow regimes for the pulp suspensions in this study were characterized previously for softwood16 and hardwood17 pulp suspensions. Without an impeller, the suspension owed as a plug at Up = 1.0 m/s and was turbulent for Up = 3.0 m/s. At Up = 2.0 m/s, the plug was disrupted, but the ow was not yet turbulent without an impeller. With the addition of the impeller, the energy supplied was sucient to provide turbulence, with the mixing quality approaching that for water. Without the impeller, the mainstream velocity required for the ow to become fully turbulent and for the mixing quality to approach that in water was as high as 4.0 m/s.16 In the high-shear zone (P2 in Figure 1) immediately downstream of the impeller, the ow was turbulent, with the mixing quality in the suspension being similar to that in water and independent of the main-stream velocity. Reocculation then likely occurred rapidly for Up = 1.0 m/s, suppressing the downstream turbulence, as shown by considerably worse downstream mixing compared to that for Up 2.0 m/s. The crowding number, Nc, of 123 from eq 1 for the 0.5% softwood pulp suspension, was higher than 60, indicating >3 contacts per ber and occulation when the shear stress was less than the suspension yield stress. When the softwood suspension ow in an empty pipe was considered at Up = 3.0 m/s, the shear stress exceeded the suspension yield stress, providing relative motion among bers and turbulent ow. At Up = 1.0 m/s, however, the ow was in the mixed ow regime, with a rigid plug in the core where the shear stress was less than the suspension yield stress.16 The rotating impeller disrupted the plug and provided turbulence in the high-shear (P2) zone, with reocculation likely occurring downstream because energy dissipation was not sustained. Figure 7 shows the inuence of the main-stream velocity on the degree of mixing of a softwood pulp suspension at Cm = 2.0% for virtually identical impeller speeds and jet-to-pipe velocity ratios and identical diameter ratio. At this concentration, the ow was essentially plug before passing through the impeller for all main-stream velocities investigated. The rotating impeller disrupted the plug, and the mixing quality improved substantially with increasing main-stream velocity compared to the tee mixer alone, where the mixing quality was poor and improved only slightly as the main-stream velocity increased.16 Enhanced mixing with increasing main-stream velocity at this concentration occurred mainly in the high-shear (P2) zone, whereas mixing improved only slightly downstream for Up 2.0 m/s compared to that for Cm = 0.5%, likely because of the reestablishment of a plug downstream of the impeller. At Up = 3.0 m/s, the quality of mixing in the pulp suspension was considerably better with the impeller than without it at the same main-stream velocity. The mixing was slightly worse than
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research

Article

Figure 5. Tomographic images for water with Dr = 0.05 and N = 400 rpm at R = (a) 6.15, (b) 12.2, and (c) 24.6. The locations of planes P1P8 are shown in Figure 1.

that for water at P2 and P3 but signicantly worse downstream, suggesting that reocculation likely occurred rapidly downstream of the impeller, with robust ber networks impeding mixing. 3.3. Eect of the Jet Velocity on the Mixing Quality in Pulp Suspensions. Figure 8 portrays the eect of the jet-topipe velocity ratio on the mixing quality for a softwood suspension at Cm = 0.5%, Up = 1.0 m/s, and N = 400 rpm. The mixing was signicantly better when the mixing mode changed from the wall source (R = 3.4) to jet mixing (R = 6.3). It was worse, however, at a higher velocity ratio when the jet reached the far wall of the pipe, whereas this behavior was not observed at the same mass concentration for a tee mixer alone.16 The
490

dierence was likely due to faster energy dissipation when the jet penetrated to the core of the pipe and impinged on the rotating impeller at R = 6.3, as discussed above for water ow. Figure 9 shows the mixing quality for a softwood suspension with Cm = 0.5%, Up = 0.5 m/s, and higher velocity ratios. Mixing improved signicantly with increasing jet velocity, likely because energy dissipation from jet impaction and the rotating impeller disintegrated the plug in the dilute pulp suspension. The energy supplied was not, however, enough to provide turbulence, even in the high-shear (P2) zone, with reocculation occurring downstream, as shown by considerably worse mixing than that for water beyond P3.
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research

Article

Figure 6. Modied mixing index as a function of the dimensionless distance downstream of injection for a softwood pulp suspension with Cm = 0.5% and water (w) at Dr = 0.05 and N = 400 rpm with various main-stream velocities and almost identical jet-to-pipe velocity ratios.

Figure 8. Modied mixing index as a function of the dimensionless distance downstream of injection for a softwood pulp suspension with Cm = 0.5%, Up = 1.0 m/s, Dr = 0.05, and various jet-to-pipe velocity ratios.

At higher concentrations (Cm 1.0%), the jet velocity had less inuence on mixing. Figure 10 plots the modied mixing index at various velocity ratios for a 2.0% softwood suspension at Up = 0.5 m/s and virtually identical impeller speeds. The degree of mixing clearly improved with the addition of the mechanical mixer. However, mixing improved only slightly with increasing jet velocity because the ber networks were still robust and energy dissipation from jet impaction was insucient to disrupt the ber networks. The mixing quality was worse when the jet reached the far wall of the pipe ahead of the impeller at R = 12.2, likely because of the robust ber networks causing the jet to adhere to the far wall downstream, behavior similar to that for a tee alone.16 Tomographic images comparing the tracer distribution for the same operating conditions are presented in Figure 11. For a jet reaching the far wall of the pipe at R = 12.2, the high-conductivity regions were concentrated at the upper part of the images and persisted as a clump downstream, as illustrated in Figure 11a. At R = 16.5, the jet impinged on the far wall and recirculated to the core of the pipe. Energy dissipation from the jet impingement and rotating impeller likely caused less nonuniformity, as shown by more spreading of the high-conductivity regions in the cross sections downstream of the P2 plane in Figure 11b. Although a higher velocity ratio of 24.1 provided slightly better mixing than R = 8.1 for a jet penetrating to the core of the pipe, this condition

Figure 9. Modied mixing index as a function of the dimensionless distance downstream for a softwood pulp suspension with Cm = 0.5% and water (w) at Up = 0.5 m/s, Dr = 0.05, and various jet-to-pipe velocity ratios.

may be undesirable from a practical point of view because impingement on the opposite pipe wall creates signicant stress there. The design of an in-line mechanical mixer should therefore be based on the velocity ratio for the jet penetrating to the axis of the pipe, as summarized by Yenjaichon et al.17

Figure 7. Modied mixing index as a function of the dimensionless distance downstream with and without the impeller for softwood pulp suspension for Cm = 2.0% and water (w) at Dr = 0.05, various main-stream velocities, and almost identical jet-to-pipe velocity ratios and rotational speeds.
491
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research

Article

Figure 10. Modied mixing index as a function of the dimensionless distance downstream for a softwood pulp suspension at Cm = 2.0%, Up = 0.5 m/s, Dr = 0.05, almost constant impeller rotation speeds, and various jet-to-pipe velocity ratios

3.4. Eect of the Impeller Rotational Speed on the Mixing Quality in Pulp Suspensions. Figure 12 shows the eect of the impeller speed on the mixing quality for a hardwood suspension with Cm = 1.0%. The mixing was poor for a tee mixer alone but improved substantially in the presence of

the impeller. At a lower main-stream velocity (Up = 0.5 m/s), a long residence time in the high-shear zone provided profoundly improved mixing with increasing impeller rotational speed, as shown in Figure 12a. Decay of turbulence likely occurred, accompanied by reocculation downstream, as is clearly shown by worse mixing downstream of P4 for a pulp suspension than for water at N = 800 rpm. The crowding number of 27.6 for the 1.0% hardwood pulp suspension exceeded 16, the gel crowding number, suggesting that bers interacted and began to occulate when the shear stress was lower than the suspension yield stress downstream of the impeller. For this velocity, the quality of mixing with the perpendicularly oriented static mixer was lower than that for all nonzero rotational speeds investigated. The suspension owed through the gaps between the impeller and inner pipe wall at relatively low velocity, and the turbulence created by the static mixer was signicantly less than that provided by the mechanical mixer. At a higher mainstream velocity (Up = 2.0 m/s), the residence time was low enough that the impeller speed had little inuence on the mixing quality. An increase in the rotational speed, however, slightly enhanced downstream mixing, as shown in Figure 12b. The mixing quality was similar for the rst three planes (P2 P4) and then became slightly better with increasing impeller speed. A comparison of suspension mixing at N = 400 and 800 rpm with water (dashed lines) at the same impeller speeds

Figure 11. Tomographic images for a softwood pulp suspension at Up = 0.5 m/s, Cm = 2.0%, and Dr = 0.05 with (a) R = 12.2 and N = 403 rpm and (b) R = 16.5 and N = 410 rpm. The locations of planes P1P8 are shown in Figure 1.
492
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research

Article

Figure 12. Modied mixing index as a function of the dimensionless distance downstream of injection for a hardwood pulp suspension with Cm = 1.0% and water (w) at Dr = 0.05 and almost constant velocity ratios with various impeller speeds at Up = (a) 0.5 and (b) 2.0 m/s.

Figure 13. Modied mixing index as a function of the dimensionless distance downstream of injection for a softwood pulp suspension with Cm = 3.0% and water (w) at Dr = 0.05 and almost identical velocity ratios with various impeller speeds at Up = (a) 1.0 and (b) 2.0 m/s.

showed that turbulence was achieved in the high-shear zone for both impeller speeds, but reocculation likely occurred faster downstream at N = 400 rpm (downstream of P5) than at N = 800 rpm (downstream of P7), because of reduced downstream turbulence. Figure 13 shows the inuence of the rotational speed on the mixing quality for a long-ber softwood suspension with a higher mass concentration of 3.0%. Mixing with the parallel static impeller was poor, not signicantly better than that for the tee mixer alone, as shown in Figure 13a. At Up = 1.0 m/s, the mixing quality improved with increasing impeller speed, predominantly occurring in the high-shear (P2) zone, with lack of downstream mixing compared to a lower-concentration, shorter-ber hardwood suspension and water, because strong ber networks reestablished immediately downstream of the impeller. The energy provided by the impeller and ow (main stream and jet) were unable to disrupt the networks in the high-shear zone, even at an impeller speed as high as 800 rpm, as shown by considerably worse mixing at P2 than for water at a lower impeller speed (N = 400 rpm). Further increasing the impeller speed likely improves mixing and provides a level of turbulence similar to that for water. However, this probably occurs only in the high-shear zone without improved downstream mixing. At this main-stream velocity, the highconcentration suspension ow, carrying high momentum, provided high energy dissipation when it impinged on the perpendicularly oriented static impeller, with the mixing quality similar to that of the mechanical mixer at N = 800 rpm. For
493

water ow at the identical main-stream velocity, however, the perpendicular conguration only matched the mixing provided by the mechanical mixer at N = 400 rpm, as shown in Figure 3a. At a higher main-stream velocity (Up = 2.0 m/s), the impeller improved mixing signicantly, as shown in Figure 13b, with slightly enhanced downstream mixing compared to a lower main-stream velocity. The impeller speed, however, had little eect on mixing, likely because of shorter residence time in the high-shear zone. 3.5. Eect of the Fiber Mass Concentration on the Mixing Quality. Figure 14 plots the modied mixing index for various ber mass concentrations of a softwood pulp suspension at Up = 1.0 m/s, N = 400 rpm, nearly identical jet-to-pipe velocity ratios, and identical diameter ratio. The mixing quality improved signicantly as the consistency decreased. The ber mass concentration strongly inuenced the mixing when turbulent ow could not be achieved. Figure 15 shows the inuence of consistency on the mixing quality for a shorter-ber hardwood suspension at a higher impeller speed, 800 rpm. The mixing quality improved profoundly as Cm decreased from 2.0 to 1.0%, at which the energy supplied was sucient to disrupt the plug and provide turbulence. The mass concentration did not, however, have a strong eect on the mixing quality for Cm 1.0% because the suspension ow was turbulent, with mixing quality similar to that for water. For Cm 1.0%, the mixing quality was similar for the rst three planes (P2P4 in Figure 1) downstream of the injection. Mixing for the hardwood pulp suspension with Cm = 1.0% was worse downstream than that for Cm = 0 and 0.5%, likely because of the reduction of turbulence accompanied by
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research

Article

Figure 14. Modied mixing index as a function of the dimensionless distance downstream for softwood pulp suspensions at Up = 1.0 m/s, R = 6, N = 400 rpm, and Dr = 0.05 for various ber mass concentrations.

Figure 15. Modied mixing index as a function of the dimensionless distance downstream for hardwood pulp suspensions at Up = 1.0 m/s, R = 6, N = 800 rpm, and Dr = 0.05 for various ber mass concentrations.

reocculation, occurring because the energy required to maintain turbulence was not sustained at the higher mass concentration. The crowding number of 13.8 for Cm = 0.5% was less than 16, indicating that bers were free to move relative to one another, so that the suspension was essentially dilute. On the other hand, the crowding number for the 1.0% hardwood pulp suspension exceeded the gel crowding number of 16. Accordingly, reocculation likely occurred only at Cm = 1.0%. The mixing quality with the mechanical mixer for a hardwood suspension ow at Cm = 0.5% was similar to that for water without bers, whereas mixing with a tee mixer and turbulent ow was better for a dilute hardwood suspension than for water.17 For tee mixing alone, mixing is achieved by turbulent shear. The shorter- and smaller-diameter bers could alter the turbulence structure in the bulk, carrying turbulent eddies with them and colliding with each other, thereby promoting turbulent dispersion. For the mechanical mixer, however, the energy dissipated near the rotating impeller probably dominated, modifying the turbulence structure in the bulk and leading to similar mixing quality for water and a dilute hardwood suspension. Tomograhic images showing turbulent and plug ow for hardwood suspensions are presented in Figure 16. For Cm =
494

0.5%, the ow was turbulent and the tracer distributed uniformly downstream of P2, as illustrated by uniform green in the cross sections, similar to that upstream of tracer injection (P1) in Figure 16a, indicating ecient mixing downstream of the impeller. Figure 16b shows the low-conductivity region in blue downstream of P2, likely a plug caused by ber reocculation at a higher ber mass concentration, 3.0%. This region occurred immediately downstream of P2, indicating rapid reocculation as the energy dissipation was not sustained. Figure 17 illustrates the inuence of the ber mass concentration on the mixing quality for a softwood suspension at a higher main-stream velocity of 3.0 m/s and virtually identical impeller speeds. For Cm 1.0%, the energy supply rate was sucient to disintegrate the ber networks and provide turbulence in the high-shear zone. Reocculation and decaying turbulence then probably occurred rapidly for Cm = 1.0%, as shown by the mixing quality being worse downstream of the P3 sensor plane compared to water and a 0.5% pulp suspension in the turbulent ow regime. At higher mass concentrations (Cm 2.0%), on the other hand, the energy supplied was insucient to provide the same level of turbulence as that for water, even in the high-shear zone immediately downstream of the impeller. Reocculation also likely occurred more rapidly than that for the 1.0% pulp suspension, with signicantly lower mixing quality downstream. 3.6. Hardwood versus Softwood Fibers. Figure 18 shows that the mixing quality for the short-ber hardwood pulp suspension was considerably better than for the softwood suspension at Cm = 1.0%, Up = 0.5 m/s, and N = 400 rpm. This was likely due to less robust ber networks. However, the energy from the impeller and the ow was not able to disintegrate the ber networks and provide turbulence in the suspension, as shown from the mixing quality being signicantly worse than that for water in turbulent ow. The inuence of the ber type on the degree of mixing for a lower mass concentration of 0.5% and a higher main-stream velocity of 2.0 m/s is presented in Figure 19. In this case, the ber properties did not have such a strong inuence on the mixing quality because the turbulence was generated in both softwood and hardwood suspensions, with the mixing quality approaching that for water. Table 1 summarizes the inuences of jet penetration, ow regime, and impeller speed on the mixing quality with an in-line mechanical mixer. For water, a jet penetrating to the center of the pipe clearly provided better mixing than the jet attaching to the pipe wall, and mixing was even better when the jet impinged on the far wall at a higher jet-to-pipe velocity ratio. A similar eect of jet penetration on mixing was observed for pulp suspensions, with considerably worse mixing when the jet reached the far wall of the pipe for the plug ow regime, likely because of the reestablishment of robust ber networks downstream of the impeller causing the jet to adhere to the wall. Jet impingement on the far wall, however, did not provide signicantly better mixing than that for jet penetration to the core of the pipe, likely because the energy supplied was insucient to disrupt the ber networks for both cases. The ow regime had a strong inuence on mixing in pulp suspensions. For similar jet penetration, mixing worsened substantially as the ow regime changed from turbulent to mixed and plug ow. For the mixed ow regime, increasing the impeller speed clearly enhanced mixing in both the high-shear (P2) zone and downstream, whereas mixing improved predominantly in the high-shear zone for the plug ow.
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research

Article

Figure 16. Tomographic images for hardwood pulp suspensions at Up = 1.0 m/s, R = 6, N = 800 rpm, and Dr = 0.05 with Cm = (a) 0.5% and (b) 3.0%. The locations of planes P1P8 are shown in Figure 1.

Figure 17. Modied mixing index as a function of the dimensionless distance downstream for softwood pulp suspensions at Up = 3.0 m/s, R = 4, Dr = 0.05, and virtually identical rotational speeds for various ber mass concentrations.

Figure 18. Modied mixing index as a function of the dimensionless distance downstream for softwood and hardwood pulp suspensions at Cm = 1.0% Up = 0.5 m/s, R = 12, and N = 400 rpm.

4. CONCLUSIONS For water in turbulent ow, the quality of mixing for a lowspeed impeller (400 rpm) was almost independent of the mainstream velocity for identical jet-to-pipe velocity and diameter ratios because the residence time in the high-shear zone had little inuence. At a higher impeller speed (800 rpm), the
495

residence time in the high-shear zone became signicant, with mixing improving considerably as the main-stream velocity decreased. Mixing improved substantially with increasing impeller speed at a low main-stream velocity (1.0 m/s), likely because of increased residence time in the high-shear zone, whereas the rotation speed had less eect at a higher mainstream velocity (3.0 m/s). At a low impeller speed (400 rpm), mixing improved substantially with increasing velocity ratio as
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research

Article

Figure 19. Modied mixing index as a function of the dimensionless distance downstream for softwood and hardwood pulp suspensions with Cm = 0.5% Up = 2.0 m/s, R = 2, and N = 400 rpm.

the mode changed from wall source to jet mixing and jet impaction. However, the mixing was worse when the jet stream reached the far wall of the pipe upstream of the impeller at a jet-to-pipe velocity ratio of 12.2. For pulp ber suspensions, mixing depended strongly on the ow regime. Mixing downstream improved profoundly when the plug was disrupted, approaching that for water when the ow became turbulent. With the addition of an impeller, the turbulent ow regime was obtained for a 0.5% softwood pulp suspension at a main-stream velocity of 2.0 m/s, whereas 4.0

m/s was required without the impeller. At a higher mass concentration (Cm = 2.0%), the ber plug was disrupted and mixing improved profoundly with increasing main-stream velocity, but downstream mixing was poor, likely because of rapid reocculation. For a dilute softwood pulp suspension (Cm = 0.5%), mixing improved with increasing jet velocity, whereas the jet velocity had little inuence at Cm = 2.0%. The mixing quality worsened when the jet penetrated to the far wall of the pipe for all mass concentrations investigated. For a dilute hardwood pulp suspension with Cm = 1.0%, mixing was improved by the addition of an impeller. A stationary impeller provided better mixing when perpendicular than when parallel to the ow. As the rotational speed increased, mixing improved signicantly, in both the high-shear zone and downstream, for a low main-stream velocity (Up = 1.0 m/s), whereas it was only slightly better at Up = 2.0 m/s, likely because of less residence time in the high-shear zone. At a higher mass concentration of 3.0% for a long-ber softwood pulp suspension, enhancement of mixing with increasing rotation speed occurred mainly in the high-shear zone, with almost negligible improvement downstream. The ber mass concentration and properties profoundly aected mixing when there was insucient supply of energy to provide turbulence. Mixing improved substantially with decreasing consistency and was better for shorter hardwood pulp bers than for softwood bers. The mass concentration and ber properties did not, however, strongly inuence the mixing quality when the ow was turbulent because the

Table 1. Summary of the Inuences of Jet Penetration, Flow Regime, and Impeller Speed on Mixing
modied mixing index, M parameter investigated jet penetration in water suspension concentration, Cm (%) 0 (water) impeller speed, N (rpm) 400 main-stream velocity, Up (m/s) 2.0, 3.0 1.0, 2.0 0.5, 1.0 0.5 jet penetration in a pulp suspension 2.0 (SW) 408 403 418 ow regime 0.5 (SW) 0.5 (SW) 0.5 (SW) 2.0 (SW) 1.0 (HW) 400 0.5 0.5 0.5 1.0 2.0 3.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 jet-to-pipe velocity ratio, R 2.06 6.15 12.2 24.6 8.14 12.2 24.1 3.99 4.13 4.13 6.14 12.5 12.5 12.0 6.35 6.14 6.12 plug mixed turbulent turbulent plug mixed plug ow regime turbulent jet penetration attaching to the near wall center reaching the far wall impinging on the far wall almost center reaching the far wall impinging on the far wall almost center almost center almost center almost center reaching the far wall reaching the far wall reaching the far wall almost center almost center almost center x/D = 2.41a 4.41 3.02 3.32 1.43 6.96 8.88 7.59 2.5 2.34 2.84 5.58 5.79 4.04 2.41 8.89 7.78 4.95 x/D = 22.1b 0.51 0.21 0.25 0.06 6.61 9.75 6.02 1.17 0.58 0.44 4.36 5.1 1.26 0.16 9.55 7.42 4.69

x/D = 12.2 0.87 0.4 0.62 0.12 6.87 9.17 6.67 1.49 0.92 0.87 4.45 5.56 1.47 0.26 9.24 7.26 4.89

impeller speed (mixed ow)

no impeller 406 802

impeller speed (plug ow)

3.0 (SW)

no impeller 406 802

High-shear zone (P2). bFurthest distance downstream (P8).


496
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research turbulence disrupted the plug, with the mixing quality approaching that for water. Unlike tee mixing, the quality of mixing with an in-line mechanical mixer for a dilute hardwood suspension ow was not signicantly better than that for water.

Article

AUTHOR INFORMATION

Corresponding Author Notes

*E-mail: wyenjaichon@chbe.ubc.ca. The authors declare no competing nancial interest. Deceased.

ACKNOWLEDGMENTS The authors thank NSERC, FPInnovations (Paprican Division), and Howe Sound Pulp and Paper Ltd. for support. We are also grateful to Drs. James Olson and Mark Martinez for providing access to the ow-loop facility in the Pulp and Paper Centre, University of British Columbia. NOMENCLATURE Cj salt concentration in the side stream for jet ow [kg/m3] Cm suspension mass concentration or consistency [%] Cp salt concentration in the main stream or pipe ow [kg/ m3] d ber diameter [m] D pipe diameter [m] Dj injection tube diameter [m] Dr jet-to-pipe diameter ratio L ber length [m] MFS mixing index for fully segregated ow dened by eq 4 [%] Mm measured mixing index dened by eq 3 [%] Ms system mixing index measured in the absence of tracer [%] M modied mixing index dened by eq 2 [%] N rotational speed [rpm] Nc crowding number dened by eq 1 Qj volumetric ow rate of the side stream [m3/s] Qp volumetric ow rate of the main stream [m3/s] n number of image pixels R jet-to-pipe mean velocity ratio, i.e., Uj/Up Rej jet Reynolds number Rep pipe Reynolds number Up main-stream or pipe velocity [m/s] Uj side-stream or jet velocity [m/s] x distance downstream of injection [m] yi local mixture electrical conductivity [mS/cm] y average electrical conductivity [mS/cm]
Greek Symbols

standard deviation of the conductivity in a tomographic image [mS/cm] ber coarseness [kg/m]
(1) Bennington, C. P. J. Mixing and Mixers. In Pulp Bleaching: Principle and Practice; Dence, C. W., Reeve, D. W., Eds.; TAPPI Press: Atlanta, GA, 1996; p 539. (2) Bennington, C. P. J. Mixing in the pulp and paper industry. In Handbook of Industrial Mixing: Science and Practice; Paul, E. L., AtiemoObeng, V. A., Kresta, S. M., Eds.; John Wiley and Sons: Hoboken, New Jersey, 2004; p 1187. (3) Torregrossa, L. O. Eect of mixing eciency on chlorine dioxide bleaching. Proceedings of TAPPI Pulping Conference, Houston, TX, 1983; TAPPI Press: Atlanta, GA, 1983; p 635.
497

REFERENCES

(4) Kolmodin, H. How to save costs by mixing chlorine dioxide and pulp homogeneously. Sven. Papperstidn. 1984, 87, 8. (5) Breed, D. B. Discovering the mechanisms of pulp mixingA pilot approach to high shear mixing. 1985 Medium Consistency Mixing Seminar, Hollywood, CA, 1985; TAPPI Press: Atlanta, GA,1985; p 33. (6) Backlund, B.; Bergnor, E.; Sandstrom, P.; Teder, A. The benefits of better mixing. Pulp Pap. Can. 1987, 88, T279. (7) Paterson, A. H. J.; Kerekes, R. J. Fundamentals of mixing in pulp suspensions: Measurement of microscale mixing in mill chlorination mixers. J. Pulp Pap. Sci. 1986, 12, J78. (8) Kouppamaki, R. The quality of mixing studied using a radiotracer technique. 1985 Medium Consistency Mixing Seminar, Hollywood, CA, 1985; TAPPI Press: Atlanta, GA, 1985; p 13. (9) Pattyson, G. W.; Kamyr, M. C. Mixer for chlorine dioxide mixing at great lakes forest products. 70th Annual Meeting Preprints, Techical Section; CPPA: Montreal, 1984; p A63. (10) Robitaille, M. A. High intensity mixing at the chlorine dioxide stage. Pulp Pap. Can. 1987, 88, T109. (11) Mann, R.; Dickin, F. J.; Wang, M.; Dyakowski, T.; Williams, R. A.; Edwards, R. B.; Forrest, A. E.; Holden, P. J. Application of Electrical Resistance Tomography to Interrogate Mixing Processes at Plant Scale. Chem. Eng. Sci. 1997, 52, 2087. (12) Yenjaichon, W.; Pageau, G.; Bhole, M.; Bennington, C. P. J.; Grace, J. R. Assessment of mixing quality for an industrial pulp mixer using electrical resistance tomography. Can. J. Chem. Eng. 2011, 89, 996. (13) Kourunen, J.; Heikkinen, L. M.; Paananen, P.; Peltonen, K.; Kay hko, J.; Vauhkonen, M. Electrical resistance tomography for evaluating a medium consistency mixer. Nord. Pulp Pap. Res. J. 2011, 26, 179. (14) Ger, A. M. Turbulent jets in crossing pipe ow. Ph.D. Thesis, University of Illinois at UrbanaChampaign, Urbana, IL, 1974. (15) Ger, A. M.; Holley, E. R. Comparison of single-point injections in pipe flow. ASCE J. Hydraul. Div. 1976, 102, 731. (16) Yenjaichon, W.; Grace, J. R.; Lim, C. J.; Bennington, C. P. J. Inline jet mixing of liquid-pulp fibre suspensions: Effect of concentration and velocities. Chem. Eng. Sci. 2012, 75, 167. (17) Yenjaichon, W.; Grace, J. R.; Lim, C. J.; Bennington, C. P. J. Inline jet mixing of liquid-pulp fiber suspensions: Effect of fiber properties, flow regime and jet penetration. AIChE J. 2012, DOI: 10.1002/aic.13913. (18) Atiemo-Obeng, V. A.; Calabrese, R. V. Rotorstator mixing devices. In Handbook of Industrial Mixing: Science and Practice; Paul, E. L., Atiemo-Obeng, V. A., Kresta, S. M., Eds.; John Wiley and Sons: Hoboken, New Jersey, 2004; p 479. (19) Bourne, J. R.; Garcia-Rosas, J. Rotor stator mixers for rapid micromixing. Chem. Eng. Res. Des. 1986, 64, 11. (20) Bourne, J. R.; Studer, M. Fast reactions in rotor stator mixers of different size. Chem. Eng. Process. 1992, 31, 285. (21) Kroezen, A. B. J.; Groot Wassink, J.; Bertlein, E. Foam generation in a rotorstator mixer. Chem. Eng. Process. 1988, 24, 145. (22) Hanselmann, W.; Windhab, E. Flow characteristics and modelling of foam generation in a continuous rotor/stator mixer. J. Food Eng. 1999, 38, 393. (23) Badyga, J.; Orciuch, W.; Makowski, L.; Malski-Brodzicki, M.; Malik, K. Break up of nanoparticle cluster in high-shear devices. Chem. Eng. Process. 2007, 46, 851. (24) Baldyga, J.; Orciuch, W.; Makowski, L.; Malik, K.; OzcanTaskin, G.; Eagles, W.; Padron, G. Dispersion of nanoparticle clusters in a rotorstator mixer. Ind. Eng. Chem. Res. 2008, 47, 3652. (25) Cooke, M.; Rodgers, T.; Kowalski, A. J. Power characterisation of an in-line Silverson high shear mixer. AIChE J. 2012, 58, 1683. (26) Hall, S.; Cooke, M.; El-Hamouz, A.; Kowalski, A. J. Droplet break-up by in-line Silverson rotorstator mixer. Chem. Eng. Sci. 2011, 662068. (27) Hall, S.; Cooke, M.; Pacek, A. W.; Kowalski, A. J.; Rothman, D. Scaling up of silverson rotorstator mixers. Can. J. Chem. Eng. 2011, 89, 1040.
dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

Industrial & Engineering Chemistry Research


(28) Derakhshandeh, B. Rheology of low to medium consistency pulp bre suspensions. Ph.D. Thesis, University of British Columbia, Vancouver, British Columbia, Canada, 2011. (29) Robertson, A. A.; Mason, S. G. The flow characteristics of dilute fiber suspensions. Tappi J. 1957, 40, 326. (30) Kerekes, R. J.; Soszynski, R. M.; Tam Doo, P. A. The occulation of pulp bres. Transactions of the 8th Fundamental Research Symposium, Oxford, U.K., 1985; p 26. (31) Kerekes, R. J.; Schell, C. J. Characterization of fibre flocculation regimes by a crowding factor. J. Pulp Pap. Sci. 1992, 18, J32. (32) Meyer, R.; Wahren, R. On the elastic properties of threedimensional fibre networks. Sven. Papperstidn. 1964, 67, 432. (33) Martinez, D. M., Buckley, K.; Jivan, S.; Lindstrom, A.; Thiruvengadaswamy, R.; Olson, J. A.; Ruth, T. J.; Kerekes, R. J. Characterizing the mobility of papermaking bres during sedimentation. Transactions of the 12th Fundamental Research Symposium, Oxford, U.K., 2001; p 225. (34) Celzard, A.; Fierro, V.; Kerekes, R. J. Flocculation of cellulose fibres: New comparison of crowding factor with percolation and effective-medium theories. Cellulose 2009, 16, 983. (35) Kerekes, R. J. Rheology of fibre suspensions in papermaking: An overview of recent research. Nord. Pulp Pap. Res. J. 2006, 21, 100.

Article

498

dx.doi.org/10.1021/ie300843z | Ind. Eng. Chem. Res. 2013, 52, 485498

You might also like