Sunteți pe pagina 1din 9

Chemical Engineering Science 66 (2011) 417–425

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Characterizing continuous powder mixing using residence time distribution


Yijie Gao 1, Aditya Vanarase 1, Fernando Muzzio, Marianthi Ierapetritou n
Department of Chemical and Biochemical Engineering, Rutgers—The State University of New Jersey, 98 Brett Road, Piscataway, NJ 08854, USA

a r t i c l e in f o abstract

Article history: Continuous powder mixing is an efficient alternative in high volume manufacturing of powder-based
Received 31 March 2010 products. A new method is communicated in this paper to determine the effects of different operating
Received in revised form conditions and mixer configurations on mixing performance. The main idea of the proposed methodology
24 October 2010
is the utilization of experimental residence time distribution (RTD) measurements to (a) determine the
Accepted 30 October 2010
Available online 4 November 2010
contributions of feeding variability, powder segregation and RTD variability on output composition
variance and (b) develop a predictive model of the output variance of a continuous mixer. The method is
Keywords: illustrated using experimental data for mixing acetaminophen and lactose in convective
Optimization continuous mixer.
Particle processing
& 2010 Elsevier Ltd. All rights reserved.
Mathematical modeling
Mixing
Relative standard deviation
Residence time distribution

1. Introduction rotary drum (Abouzeid et al., 1980, 1974) and in a convective


continuous mixer (Marikh et al., 2008, 2005). Several RTD models
Powder mixing is widely used in pharmaceutical, cement, food, were introduced characterizing the experimental RTD curves
and other industries in which product mixtures need to satisfy (Nauman, 2008). Delay and dead volume models that linked PFR
typical performance criteria. Besides traditional batch methods and CSTR units were used for rotating kilns (McTait et al., 1998),
used in the pharmaceutical industry, continuous mixing has been single screw extruders (Yeh and Jaw, 1998) and twin-screw
studied in recent years as an efficient alternative in high flux continuous mixers (Ziegler and Aguilar, 2003); the dispersion
manufacturing (Marikh et al., 2006). Studies have been mainly model based on Fokker–Planck equations with specified boundary
focused on the influence of operating conditions (feed rate, conditions were used in rotary calciners (Sudah et al., 2002),
impeller speed, etc.) and geometric designs (mixer size, impeller fluidized beds (Harris et al., 2003a, 2003b) and rotating kilns
types, etc.) on the efficiency and throughput of mixers (Pernenkil (Sherritt et al., 2003); Markov chains (Harris et al., 2002b; Marikh
and Cooney, 2006). The performance of several continuous mixers et al., 2006) and a compartment model (Portillo et al., 2008a) were
has been investigated for materials with different flow properties developed which described granular mixing using a network of
(Harwood et al., 1975). The effects of different types of stirrer on the interconnected cells based on possible flow patterns inside differ-
hold-up and quality of mixtures have been examined (Marikh et al., ent mixers.
2008). Although these studies offer a better understanding of mixer As stated above, the effects of the operating factors are
behavior, there is still a lack of predictive models of continuous complicated and thus difficult to be characterized or predicted.
mixing processes that can be used effectively for design and Operations empirically optimized in one mixer may not be directly
optimization purposes. utilized in other mixing cases. This lack of generality has limited the
Modeling of continuous mixing processes has received signifi- development of systematic design methodologies and control
cant attention in the literature mainly through applications of the strategies for continuous mixers. Moreover, although the contin-
residence time distribution (RTD) theory. These studies were able uous mixing process has been considered as the combination of
to model the dispersion of particles along the direction of powder macromixing and micromixing sub-processes (Weineköter and
delivery (Danckwerts, 1953). Efforts have been made using RTD as a Reh, 1995), few attempts have been made to analyze these process
measure for optimizing operating conditions in solid mixing in a components separately. In addition, previous studies suggest that
better mixing occurs when RTD is broader (Harris et al., 2003a;
Marikh et al., 2006; Weineköter and Reh, 1995; Williams and
n Rahman, 1972), while an inverse finding is reported based on our
Corresponding author. Tel.: + 1 732 445 2971; fax: + 1 732 445 2581.
E-mail address: marianth@soemail.rutgers.edu (M. Ierapetritou). prior work (Portillo et al., 2008b) in which we found that the most
1
These authors contributed equally to this work. dispersed RTD corresponds to the worst mixing in a convective

0009-2509/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2010.10.045
418 Y. Gao et al. / Chemical Engineering Science 66 (2011) 417–425

continuous mixer, probably due to the low number of blade passes The mass fraction of API in this case study is set as 3% for all the
experienced by some of the powder as it travels along the mixer operating conditions, which are summarized in Table 1.
axis. In fact, the main difficulty using RTD as a measure of mixing
performance is that it does not consider micromixing of the initially 2.2. Data collection and analysis of feeder fluctuations
segregated components.
In this paper, we developed a general method using the In order to investigate the influence of feed rate variability to the
experimentally measured RTD to quantitatively predict the per- output variance, the feed rate is measured using a catch-scale that
formance of continuous mixing. To distinguish between the records the weight of powder discharged by the feeder as a function
variability caused by feeder fluctuations and by the mixer itself, of time. More specifically, a container is placed on the catch-scale
the contributions to the output variance from different sources and powder is fed for about 30 min. The temporal feed rate of the
were separately considered. The rest of the paper is organized as feeder is then calculated based on the weight increase in the
follows. The experimental conditions and methods are presented in container. The power spectrum for each of the data sets was
Section 2. In Section 3, the methods of for fitting RTD parameters, computed and compared using a MATLAB program, where the
the empirical correlation between RTD and RSD and the variance variance distribution functions s(f) of the feed rates are calculated
analysis are presented. Based on the method developed in Section at different frequencies. The relationship between s(f) and the
3, RTD profiles and mixing performances of several operating overall variance of the feed rate fluctuations s2in is determined by
conditions are studied in Section 4 to test the feasibility of the
Z
developed method. Feed rate, blade configuration and rotary speed 1 fs =2
s2in ¼ sðf Þ2 df ð1Þ
are studied to quantify their effects on the optimal RTD fitting 2 0
parameters and the mixer performance. Section 5 presents con-
where fs is the frequency of feed rate measurements. Since the
clusions and discusses the applicability of the proposed method to
sampling interval is 0.1 s for all the feeders, fs ¼10 Hz. s(f) is
other configurations.
integrated between 0 and fs/2 to avoid aliasing, which may
introduce error in variance calculation. Note that s(f) represents
2. Experimental setting the contribution from different frequencies to the feed rate
variance. Details of this issue can be found in our previous work
2.1. Mixing equipment and materials (Gao et al., in press). Fig. 2 shows the variance distributions of API
and excipient feed rates at different frequencies for the two feed
RTD measurements were carried out in a GCM 250s convective rates studied. The peak of feed rate fluctuations of API shifts from
continuous mixer, provided by Gericke, shown in Fig. 1. The 0.16 to 0.24 Hz when feed rate is increased. This indicates better
impeller consists of a cylinder shaft with several blades on each feeding conditions, since fluctuations at higher frequency are more
side. The angle of the blades can be adjusted to 201 forward or easily filtered out by the mixer (Weineköter and Reh, 1995). For the
201 backward inclination, determining the blade patterns to be excipient, a sharp peak appears at 1.4 Hz at the lower feed rate,
investigated. In order to hold the fill level at the desired value, a whereas smaller peaks at higher frequencies (starts from 2.1 Hz)
weir is used to block the lower half of the outlet section. The characterize the higher feed rate. Notice that the lowest peak of
rotation speed of the impeller varies between 40 and 250 RPM. feed rate fluctuations (peak at the lowest frequency) always
Since the critical speed of fluidization is around 180 RPM for this corresponds to rotating frequency of the feeding screw. The results
mixer, both a dense-flow regime and a fluidized regime are in Table 2 show that the feed rate of 45 kg/h leads to smaller RSD of
observed in our studies. API concentration and variance components at higher frequency
Two loss-in-weight (LIW) feeders provided by Schenck-Accu- domains. Since continuous mixer always performs like a low-pass
rate are used to feed the mixer. Enough materials were pre-loaded filter through which feeding fluctuations at higher frequencies are
in the feeders so that the experiments were not interrupted or easier to be attenuated (Pernenkil and Cooney, 2006), these results
disturbed by re-filling the feeder. The overall feed rates examined indicate better feeding performance at feed rate of 45 kg/h.
include 30 and 45 kg/h. Acetaminophen (with average particle size
of 45 mm), pre-blended with 0.25% SiO2, is used as a representative 2.3. RTD and output variance measurement
active pharmaceutical engredient (API) and micro-crystalline
cellulose (Avicel PH200, 240 mm) is used as the main excipient. RTD is measured using the method originally proposed by
Danckwerts (1953). Bulk material is fed into the mixer until steady
state is reached, and tracer (10 g of APAP) is injected instanta-
neously into the inflow stream. Samples are collected at the outlet
of the mixer at different time points. The disturbance to the bulk
flow caused by the injection of the tracer is assumed to be
negligible. Pulse tests are repeated twice at each condition of
interest to examine the reproducibility of the RTD. The samples

Table 1
Operating conditions.

Operating parameters

Feed rate (kg/h) 30, 45


Impeller rotation rate (RPM) 40, 100, 160, 250

Design parameters
Blade configuration (blade angle with 1. All forward—all blades directing
the shaft and blade direction) forward, blade angle—201
2. Alternate—Alternate blades in
forward and backward directions—201
Fig. 1. Side view of the studied continuous mixing system.
Y. Gao et al. / Chemical Engineering Science 66 (2011) 417–425 419

Fig. 2. Feed rate analysis of (a) API and (b) excipient.

Table 2 before the next measurement so that the scanned surface of the
Statistics of feed rate fluctuations of API at the investigated feed rates. sample is renewed. Standard deviation, RSD and 95% confidence
interval are calculated. The results are plotted in Fig. 3, where the
Feed rate (kg/h) API%
RSD from measurements is found to be around 1% in the studied
Average value Standard deviation RSD (%) range of API%, which is insignificant compared to the output RSD
measurement (10–20%, described in results) obtained in the
30 3.00 1.09 36 continuous mixing experiments. As a result, the influence of NIR
45 3.02 0.64 21
measurement error can be neglected in the analysis of output
variance sources in this study.
collected are analyzed by NIR spectroscopy to determine the
concentration of acetaminophen (C(t)). The residence time dis-
3. Proposed methodology
tribution (E(t)) is calculated based on the collected data of C(t) using
the following equation:
3.1. General concepts and tools
CðtÞ
EðtÞ ¼ R 1 ð2Þ
0 CðtÞ dt As indicated in Section 2, the variance of the acetaminophen
concentrations measured at the mixer output is used as a measure
To determine the homogeneity of the product mixture, 20
for mixing performance. This measure typically represents two
samples are retrieved from the output stream of the continuous
sources of variability: the variability due to fluctuations in powder
mixing process for each operating condition. The concentration of
feeding rate, and the variability of composition due to the non-
acetaminophen in each sample is measured using NIR spectro-
ideality of the powder mixing process, which is dominant when
scopy, which is used to calculate to the corresponding standard
ideal feeding condition is satisfied. The output variance can thus be
deviation and the relative standard deviation (RSD) as a measure of
expressed as
mixing performance:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi s2output,solid ¼ s2output,fluid þ s2idealfeed ð4Þ
PN 2
s Standard deviation 1 ðCi CÞ
RSD ¼ ¼ s¼ ð3Þ where s2output,fluid indicates the variance contribution of unfiltered
C Average concentration N1
feeding fluctuations, and s2idealfeed the contribution due to granular
where C is the average concentration of N samples and Ci is the nature of the material (i.e., incomplete transverse mixing, finite
concentration of each sample. number of particles in the sample, agglomeration, segregation, etc.)
(Weineköter and Reh, 1995). Early studies (Beaudry, 1948;
2.4. Analysis of NIR method error Danckwerts, 1953) describing the variance component due to
feeding directly derives variance reduction ratio (VRR) from the
In order to investigate the contribution of error in the NIR (Near- convolution of RTD and feeding fluctuations. Although the compo-
infrared spectroscopy) method to the output variance, the follow- nent of ideal-feed was always ignored in these studies due to a
ing analysis is performed. Samples of 2%, 3%, 5% and 10% mass cross-sectional sampling size (Goldsmith, 1966; Williams and
fraction of API are prepared to study the measurement variability Rahman, 1972), it was investigated in a fluidized bed by
under the 3% API mass fraction condition. Each sample is well Williams and Richardson (1982), where mixing of segregating
mixed and measured ten times. The same sample is re-agitated particles was studied and compared with known or without
420 Y. Gao et al. / Chemical Engineering Science 66 (2011) 417–425

Fig. 3. Calibration test on NIR method: (a) standard deviation (STDEV) and (b) relative standard deviation (RSD) was measured at API% of 2–10% and 95% confidence intervals
were estimated.

Fig. 4. Data variability of RTD measurements are represented by the vertical difference between test points and model-fitted curve (a), while model variability is indicated in
multiple RTD measurements by the difference between single and multiple model-fitted curves (b).

feeding fluctuations. In our present work, this variance component needed to determine these components. In this study, the Taylor
of ideal-feed is further decomposed as follows: dispersion model is selected and described in the following sections.

s2idealfeed ¼ s2ITM þ s2RTDV ð5Þ


3.2. Model development
2
where the term s represents the variance component due to
ITM
incomplete transverse mixing (ITM), and the term s2RTDV represents In the original convolution formula of continuous mixing
the component due to RTD variability (RTDV). Note that the sample (Danckwerts, 1953)
Z 1
size is supposed to be much smaller than the scale of cross-
Cout,fluid ðtÞ ¼ Cin ðtyÞEðyÞ dy ð6Þ
sectional mixing to ensure that the term s2ITM can be detected. 0
While traditional variance measurements can hardly provide
Cin(t) and Cout,fluid(t) are the input and output concentrations. Due to
information about the different components of output variance, the
the variance components described in the previous section, a time-
characteristics of the RTD can reveal it. In particular, the difference
variant fluctuation component associated to E(y)in Eq. (6) gives the
between experimental and model-predicted data, or the data
following change:
variability in a single RTD measurement is caused by incomplete Z 1
transverse mixing, which can be used as an indicator of transverse
Cout,solid ðtÞ ¼ Cin ðtyÞðEðyÞ þ eE ðy,tyÞÞ dy ð7Þ
mixing (Fig. 4a). Low data variability indicates good transverse 0
mixing and homogeneous API concentration in the cross-section of where eE(y, t  y) represents the fluctuation of measured RTD
outlet stream. Similarly, the difference between single and multiple described in Fig. 4. Note that this new term sums the effects of
model-fitted RTD curves, or model variability in repeated RTD data variability and model variability. Similarly, substituting
experiments indicates variance due to RTD variability (Fig. 4b). Cin(t  y) by the sum of mean and time-dependent fluctuations
Small difference indicates good RTD reproducibility, which leads to C in þ ein ðtyÞ, results in the following form:
an insignificant output variance component s2RTDV , since little new Z 1
axial compositional heterogeneity is generated in the system Cout,solid ðtÞ ¼ Cout,fluid ðtÞ þ ðC in þ ein ðtyÞÞeE ðy,tyÞ dy ð8Þ
through superposition of RTD curves with low variability. This 0
R
issue was studied in Harris et al. (2002a) to validate a proposed where Cout,fluid(t)¼ N 0 Cin(t  y)E(y)dy. Assuming that ein(t) is much
method of RTD measurement, where the first three moments of smaller than C in :
RTD were used in comparison of repeated RTD measurements. It was Z 1
also investigated in an earlier work (Williams and Rahman, 1972) in Cout,solid ðtÞ  Cout,fluid ðtÞ þ C in eE ðy,tyÞ dy ð9Þ
0
validating the ideal-feed condition in a drum mixer. Since quantifica-
tion of both components correlates to the model fitting process of Since the variance components in the two components of Eq. (9) are
RTD, a RTD model that well represents the experimental data is independent, the output variance of the concentration equals the
Y. Gao et al. / Chemical Engineering Science 66 (2011) 417–425 421

sum of the variance components contributed from each of them. Section 3.2. In particular
Comparing Eqs. (4) and (9), it is not difficult to conclude that the
ideal-feed variance component s2idealfeed comes from the integration MSS MSS
sE ¼ R 1 ¼ ð16Þ
of RTD variability C in eE . 0 Cðt,P1 Þdt M

where C(t, P1) is the fitted curve of pulse test derived from multiple
3.3. Solution approach
RTD fitting process, and M represents the total material injected
into the system. Based on Eq. (15)
In order to estimate the time-invariant RTD component E(y)
discussed in the previous section, as well as to understand the s2E ¼ s2E,ITM þ s2E,RTDV ð17Þ
influence of operations on axial velocity and dispersion, the Taylor
dispersion model is considered as follows: where sE,ITM and sE,RTDV represent the data variability and model
@C @C 1@ C 2 variability of RTD shown in Fig. 4, respectively.
þ ¼ ð10Þ Based on Eq. (9), the variance of ideal-feed is more significant
@y @x Pe @x2
when the expectation of RTD fluctuation (sE) is larger, and when
Eq. (10) is the well-known one-dimensional Fokker–Planck equa- this fluctuation lasts longer. Based on this observation, we assume
tion that describes time evolution of particle distribution in a that the RSD, which is the standard deviation divided by the mean
continuous system (Risken, 1996). Detail of its solution in different concentration, can be expressed as follows:
initial and boundary conditions can be found in (Levenspiel and
Smith, 1957; Sherritt et al., 2003). One commonly used solution of RSDidealfeed ¼ ksE DT ð18Þ
Eq. (10), known as the Taylor dispersion model, is obtained with
open boundary conditions, which was described in Harris et al. where k is a dimensionless coefficient describing the ratio between
(2003a, 2003b) in characterizing the influence of axial gas and solid sum of RTD fluctuations and the corresponding RSD. DT represents
flux on axial solid Peclet number in fluidized bed: the time interval when the RTD data were collected. In order to
obtain reasonable DT, prolonged RTD measurements were
C0 Pe1=2 2
obtained for each operating condition so that all non-zero points
Cðx, yÞ ¼ ePeðxyÞ =4y
ð11Þ
ð4pyÞ1=2 can be recorded. Notice that to guarantee this assumed relationship
where y ¼ (t t0)/t and x ¼z/l are the dimensionless time and in Eq. (18), both RSD and sE should be determined by the same
location in continuous mixing process, where t0 represents the sampling size. Using the experimental RTD and RSD obtained under
effect of time delay in RTD measurement, t is the mean residence different operating conditions, k can be empirically estimated by
time and l is the overall length of the mixer. C0 denotes the amount applying linear regression of the pairs RSDidealfeed and sEDT. Then
of material injected in the pulse test. Pe¼vl/E is the Peclet number, the variance components due to incomplete transverse mixing and
where v and E represent the cross-section averaged axial velocity RTD variability can be derived based on the estimation of k as
and the axial dispersion coefficient of the system. When x ¼1, follows:
Eq. (11) corresponds to the RTD of the mixer.
RSDITM ¼ ksE,ITM DT ð19Þ
The parameters Pe, C0, t and t0 in Eq. (11) are determined to
minimize the mean sum of square (MSS) error between predicted
and experimentally determined values of pulse test. This para- RSDRTDV ¼ ksE,RTDV DT ð20Þ
meter set is defined as follows:
Based on the estimation of the RSD components, the corre-
P ¼ ½Pe,C0 , t,t0  ð12Þ
sponding variance components can then be conveniently derived
Fitted pulse test data can then be expressed as C(t, P). In using Eq. (3).
particular, to evaluate the best fitted curve for a single pulse test
the follow error is minimized:
X
n
½Cij Cðtij ,P1j Þ2 X
n
½Cij Cðtij ,PÞ2 3.4. Methodology
MSS2j,ITM ¼ ¼ min ð13Þ
i¼1
n i¼1
n
Based on the theoretical work developed in the previous section,
which represents the MSS error of data variability of RTD due to
the following procedure of the RTD modeling method can be
incomplete transverse mixing. The overall MSS error of repeated
established.
RTD experiments is minimized when model fitting of multiple
Step 1: Determine RTD from pulse tests, RSD at the outlet of
pulse tests are performed:
continuous mixing process (RSDout,solid) and feeding samples.
m X
X n
½Cij Cðtij ,P1 Þ2 m X
X n
½Cij Cðtij ,PÞ2 Step 2: Fit RTD using the Taylor dispersion model.
MSS2 ¼ ¼ min ð14Þ Step 3: Apply the fitted RTD curve as well as the feeding samples
j¼1i¼1
mn j¼1i¼1
mn
to calculate the RSD component of unfiltered feeding fluctuations
In this case the overall MSS represents the error of both data and RSDout,fluid.
model variability. In Eqs. (13) and (14), tij and Cij represents the pair Step 4: Determine RSDidealfeed based on RSDout,fluidand RSDout,solid.
of time and concentration of the ith point in the jth repeat of RTD Step 5: Establish the correlation between RSDidealfeed and the
pulse tests (i¼ 1,2, ..., n and j¼1,2, ..., m). P1j is the optimal para- form sEDT, which is obtained in the RTD fitting process.
meter set for the jth RTD test leading to the minimized mean sum of Step 6: Calculate RSDITM and RSDRTDV and the corresponding
square MSS2j,ITM, while P1 leads to the minimized MSS2 in the variance components, then analyze the effects of operations (feed
multiple RTD fitting process. The relationship between MSS of rate, rotary speed, blade angle, etc.) on each of them.
single and multiple pulse tests is as follows: The analysis of Step 6 can provide guidelines in order to improve
the design and performance of continuous mixing systems. In the
MSS2 ¼ MSS2ITM þMSS2RTDV ð15Þ
next section the proposed procedure was applied in a case study of
The MSS error evaluated in Eqs. (13) and (14) is then used to continuous mixing in order to investigate the influence of different
determine the expectation of RTD fluctuation eE described in operations to the mixing performance.
422 Y. Gao et al. / Chemical Engineering Science 66 (2011) 417–425

4. Experimental results and modeling hardly affected by altering feed rate. On the other hand, boarder
RTD profiles are observed for the alternating blade configuration
4.1. RTD fitting results other than for the forward blade configuration, the effect of which
is more significant at higher blade speeds.
Fig. 5 illustrates the influence of operating conditions and blade In order to analyze these effects in detail, the optimal axial
design on RTD characteristics, as obtained experimentally for the dispersion coefficient E and axial velocity v derived from P1 were
materials considered. For each operating condition listed in Table 3, investigated. As has been examined in several previous studies
pulse tests were repeated twice so that the influence of RTD (Harris et al., 2003b; Sarkar and Wassgren, 2009), particle disper-
variability could be included in the performance prediction. The sion coefficient is closely related to fill level and blade rotating
Taylor Dispersion model and the non-linear regression procedure speed. Generally, moderate fill level and high blade speed improve
described in Section 3 were applied in the RTD fitting process. The axial dispersion (Ingram et al., 2005; Laurent and Bridgwater, 2002;
curve in each plot corresponds to the best model fit using Eq. (14) Portillo et al., 2010), while an increase of feed rate or a decrease of
with optimal parameter sets listed in Table 3. It was observed that blade speed leads to higher fill level (Marikh et al., 2005). The
RTD is more sensitive to blade speed than to feed rate and blade optimal axial dispersion coefficients at different operating condi-
configuration. Lower blade speeds and lower flow rate lead to tions are shown in Fig. 6a, which indicates that increase of feed rate
larger mean residence time and wider RTD. When feed rate leads to larger dispersion coefficient at intermediate speed
increases from 30 to 45 kg/h, RTD profiles are narrower. This effect (100 RPM), but makes almost no difference when the speed is
is less significant for high speed cases in which the RTD curves are larger or smaller. At moderate blade speed, the fill level increases

Fig. 5. Experimental and fitted RTD curves in different operating conditions.

Table 3
RTD fitting results for different operating conditions.

No. Flow rate Blade configuration Blade speed Pe C0 t (s) t0 (s) E (cm2/s) v (cm/s)
(kg/h) (RPM)

1 30 Forward 40 7.40 1.14 52.17 9.91 2.37 0.53


2 30 Forward 100 9.91 0.15 44.58 0.15 2.46 0.74
3 30 Forward 160 14.77 0.18 29.17 0.00 2.53 1.13
4 30 Forward 250 3.49 0.37 6.58 0.00 47.41 5.01
5 45 Forward 40 16.59 0.15 39.74 0.43 1.63 0.82
6 45 Forward 100 5.15 0.21 24.84 6.84 6.67 1.04
7 45 Forward 160 14.03 0.19 25.97 0.03 2.99 1.27
8 45 Forward 250 3.40 0.45 5.46 1.27 47.68 4.91
9 30 Alternate 40 8.45 0.01 52.38 5.65 2.22 0.57
10 30 Alternate 100 7.33 0.17 43.83 0.55 3.35 0.74
11 30 Alternate 160 3.18 0.30 36.66 0.55 9.22 0.89
Y. Gao et al. / Chemical Engineering Science 66 (2011) 417–425 423

Fig. 6. Effects of operating conditions on (a) dispersion coefficient E and (b) axial velocity v.

significantly at larger feed rate. This behavior is not observed in Table 4


high rotation speed (250 RPM), at which fluidized particles dis- Predicted RSD values using the calculated k ¼0.424.
perse freely for both feed rates and the effect of feed rate on fill level
No. sE (1/s) DT (s) sEDT RSDreal RSDout,fluid RSDidealfeed RSDpredict
is not significant. Similar observation occurs at low blade speed
(40 RPM), which is not surprising because limited particle move- 1 0.0017 173 0.29 0.13 0.00 0.13 0.12
ment severely hinders dispersion; under such conditions the 2 0.002 118 0.24 0.10 0.00 0.10 0.10
influence of fill level is not significant. Larger dispersion coefficient 3 0.0028 78 0.22 0.10 0.00 0.10 0.09
4 0.0175 33 0.58 0.18 0.01 0.18 0.24
is also found when the alternate blade pattern is used, which
5 0.0025 123 0.31 0.10 0.00 0.10 0.13
disperses particles better than the forward blade pattern at higher 6 0.0025 104 0.26 0.07 0.00 0.07 0.11
speeds due to the significant increase of backward flux in that blade 7 0.0017 94 0.16 0.12 0.00 0.12 0.07
configuration. 8 0.009 43 0.39 0.18 0.02 0.18 0.16
The effects of operating conditions on axial velocity are shown 9 0.0015 170 0.26 0.14 0.00 0.14 0.11
10 0.0012 149 0.18 0.09 0.00 0.09 0.08
in Fig. 6b. Axial velocity was found higher at the 45 kg/h feed rate, 11 0.0016 122 0.20 0.09 0.00 0.09 0.08
and thus led to shorter mean residence time. This is due to the fact
that the increase of fill level is slower than the increase of feed rate.
This effect is offset at high speed (250 RPM) due to the much faster
fluidized dispersion. For the same reason as for the dispersion
coefficient, limited particle movement at low blade speed hinders
the change of axial velocity in different blade pattern configura-
tions. Furthermore, increased backward flux increases the fill level
in the alternate blade pattern and further increases mean residence
time. Similar effect has been reported in Harris et al. (2003a) on the
axial mixing study of fluidized beds, in which the effect of reflux
due to the existence of riser at the exit leads to increased mean
residence time and wider RTD curves.

4.2. Results and discussion

4.2.1. Output variance prediction Fig. 7. 95% confidence interval of experimental RSD are calculated and compared to
Based on Eq. (18), k was estimated to be 0.424 using 11 pairs of the predicted values.
mixing performance and RTD data corresponding to different
operating conditions. Table 4 provides the calculated intermediates Comparisons of experimental and predicted variance, calcu-
and the predicted RSD parameters. The results are plotted in Fig. 7, lated using the values listed in Table 4 for different operating
in which most predicted values of RSD are located within the 95% conditions, are shown in Fig. 8. Although some differences are
confidence interval of the experimental values. A paired t-test observed between the experimental and predicted variance, the
(a ¼0.05) was performed to determine if there is significant predicted values follow the same trends as the experimental
difference between predictions and measurements of output measurements. The main sources of variation between predicted
RSD. Significant difference is rejected if the two-tail p-value is and observed values include:
larger than a, or the t-value of the sample size N is smaller than its
critical two-tail value. The results based on the data in Table 4
shows that the difference between RSD pairs (M¼0.091%, (1) The term ein(t  y)eE(y, t  y), which in Eq. (8) was assumed to be
SD¼3.21%, N ¼11) is not significant, two-tail p-value is negligible, may contribute to output fluctuations in practice in
0.9274 a, t(10) ¼0.094 otwo-tail tcritical (2.228), providing the examined case study.
evidence that no significant difference is observed between the (2) In the RTD fitting process, the variance component due to
pairs. A 95% confidence interval about the mean difference of RSD is RTD fluctuations was estimated by two repeats of pulse
(  2.06%, 2.25%). tests. This may introduce non-representative variability on
424 Y. Gao et al. / Chemical Engineering Science 66 (2011) 417–425

Fig. 8. Output variance prediction for (a) forward blade configuration at 30 kg/h, (b) forward blade configuration at 45 kg/h, and (c) alternate blade configuration at 30 kg/h.

Fig. 9. Effects of operating conditions on the variance component due to feed rate Fig. 10. Effects of operating conditions on the variance component due to
fluctuations. incomplete transverse mixing.

the normalized mean sum of square sE and thus lead to


deviations on the predicted variance.
(3) The estimated interval DT in the RTD modeling process might
be too short, which might fail to capture fluctuations occurring
outside the recorded tail of the RTD curve.

4.2.2. Effect of operating and design parameters


The effect of feed rate, blade speed and blade angle on output
variance due to feed rate fluctuations, incomplete transverse
mixing and RTD variability are plotted in Figs. 9–11. In agreement
with the RSD results shown in Table 4, under the experimental
conditions examined here, the variance component caused by
feeder fluctuations shown in Fig. 9 is insignificant ( o0.08%) when
Fig. 11. Effects of operating conditions on the variance component due to RTD
compared with the other two components (0.2–0.8%) and is variability.
negligible ( o0.02%) in 40–160 RPM, as the mixer filters out most
input fluctuations in this range of blade speeds. At 250 RPM, the
mixer exhibits much worse filtering efficiency at 45 kg/h than at Fig. 11 shows the influence of operating conditions on the
30 kg/h, but the contribution of input fluctuations to the overall component of RTD variability. At high blade speed, short residence
variance is still low. time lead to highly variable RTD profile and contribute to larger
Ingram et al. (2005) found that the radial dispersion coefficient variance component. It is also noticed that high feed rate produces
increased at high rotating speed and moderate high fill level in large RTD variability, possibly due to large fill levels and less
rotating drums. In agreement with their previous findings, faster homogeneous dispersion of particles in the axial direction.
decay of variance component due to incomplete transverse mixing Decrease of this component in the alternate blade pattern can be
is found here at high blade speed in the studied convective mixer explained by the increase of backward flux that facilitates axial
(Fig. 10). It is also observed that at low blade speed, high feed rate mixing, thus leading to less variable RTD profiles.
facilitates transverse mixing. This is due to the significant increase Based on the analysis above, the influence of each operating
of fill level under such conditions, which leads to larger number of condition on different variance components can be summarized as
blade passes experienced by the powder and results in faster radial follows. High blade speed facilitates transverse mixing, but intro-
dispersion. Mixing is also improved in the alternate blade config- duces variability on the RTD profile and is inefficient in attenuating
uration due to the effect of reflux discussed above. Both effects are feeder fluctuations. High feed rate increases fill level at low blade
offset at high blade speed when the whole mixer is fluidized. speed, increasing the number of blade passes during the residence
Y. Gao et al. / Chemical Engineering Science 66 (2011) 417–425 425

time, which benefits transverse mixing while results in more RTD Harris, A.T., Davidson, J.F., Thorpe, R.B., 2002a. A novel method for measuring the
variability. Backward mixing produced by the alternate blade residence time distribution in short time scale particulate systems. Chemical
Engineering Journal 89 (1–3), 127–142.
pattern improves mixing due to the refluxing effect while leads Harris, A.T., Davidson, J.F., Thorpe, R.B., 2003a. The influence of the riser exit on the
to less variable RTD profiles at moderate blade speed. In summary, particle residence time distribution in a circulating fluidised bed riser. Chemical
the alternate blade pattern at moderate blade speed generates the Engineering Science 58 (16), 3669–3680.
Harris, A.T., Davidson, J.F., Thorpe, R.B., 2003b. Particle residence time distributions
most efficient continuous mixing for the materials examined here,
in circulating fluidised beds. Chemical Engineering Science 58 (11), 2181–2202.
while the effects of feed rate on transverse mixing and RTD Harris, A.T., Thorpe, R.B., Davidson, J.F., 2002b. Stochastic modelling of the particle
variability offset each other for the studied conditions, and thus residence time distribution in circulating fluidised bed risers. Chemical Engi-
is insignificant to the overall mixing performance. neering Science 57 (22–23), 4779–4796.
Harwood, C.F., et al., 1975. The performance of continuous mixers for dry powders.
Powder Technology 11 (3), 289–296.
Ingram, A., et al., 2005. Axial and radial dispersion in rolling mode rotating drums.
5. Conclusions Powder Technology 158 (1–3), 76–91.
Laurent, B.F.C., Bridgwater, J., 2002. Influence of agitator design on powder flow.
In the present study, we have developed a method based on the Chemical Engineering Science 57 (18), 3781–3793.
Levenspiel, O., Smith, W.K., 1957. Notes on the diffusion-type model for the
experimental RTD for prediction of the mixer performance in a longitudinal mixing of fluids in flow. Chemical Engineering Science 6 (4–5),
continuous mixing system. In the proposed methodology, the 227–235.
different sources of output variance are separately examined, Marikh, K., et al., 2008. Influence of stirrer type on mixture homogeneity in
which leads to better understanding of the effects of operating continuous powder mixing: a model case and a pharmaceutical case. Chemical
Engineering Research and Design 86 (9), 1027–1037.
conditions on each of the individual variance component. Based on Marikh, K., et al., 2005. Experimental study of the stirring conditions taking place in a
the established RTD–RSD correlation, the effects of two feed rates, pilot plant continuous mixer of particulate solids. Powder Technology 157 (1–3),
two blade configurations and four blade speeds are analyzed. The 138–143.
Marikh, K., et al., 2006. Flow analysis and Markov chain modelling to quantify the
results show that moderate blade speed and alternate blade
agitation effect in a continuous powder mixer. Chemical Engineering Research
configuration facilitate mixing, while change of feed rate shows and Design 84 (11), 1059–1074.
no significant influence to the overall output variance. Since the McTait, G.E., Scott, D.M., Davidson, J.F., 1998. Residence time distribution of particles
developed methodology is independent of operational conditions, in rotary kilns. In: Fan, L.-S., Knowlton, T.M. (Eds), Fluidization IX: Proceedings of
the Ninth Engineering Foundation Conference on Fluidization, Durango, Color-
sampling detection method, and mixer type, it should be applicable ado, pp. 397–404.
as a general method to investigate continuous mixing processes. Nauman, E.B., 2008. Residence time theory. Industrial & Engineering Chemistry
Moreover, the developed variance separation method has an Research 47 (10), 3752–3766.
Pernenkil, L., Cooney, C.L., 2006. A review on the continuous blending of powders.
important advantage in understanding the behavior of different
Chemical Engineering Science 61 (2), 720–742.
variance sources, and thus provides quantitative guidance in Portillo, P., Muzzio, F., Ierapetritou, M., 2008a. Using compartment modeling to
further studies in design and control of continuous powder mixing investigate mixing behavior of a continuous mixer. Journal of Pharmaceutical
processes. Innovation 3 (3), 161–174.
Portillo, P.M., Ierapetritou, M.G., Muzzio, F.J., 2008b. Characterization of continuous
convective powder mixing processes. Powder Technology 182 (3), 368–378.
Portillo, P.M., et al., 2010. Investigation of the effect of impeller rotation rate, powder
Acknowledgement flow rate, and cohesion on powder flow behavior in a continuous blender using
PEPT. Chemical Engineering Science 65 (21), 5658–5668.
Risken, H., 1996. The Fokker–Planck Equation: Methods of Solution and Applica-
This work is supported by the National Science Foundation
tions. Springer 472 pp.
Engineering Research Center on Structured Organic Particulate Sys- Sarkar, A., Wassgren, C.R., 2009. Simulation of a continuous granular mixer: effect of
tems, through Grant NSF-ECC 0540855, and by Grant NSF-0504497. operating conditions on flow and mixing. Chemical Engineering Science 64 (11),
2672–2682.
Sherritt, R.G., et al., 2003. Axial dispersion in the three-dimensional mixing of
References particles in a rotating drum reactor. Chemical Engineering Science 58 (2),
401–415.
Abouzeid, A.Z.M., Fuerstenau, D.W., Sastry, K.V., 1980. Transport behavior of Sudah, O.S., et al., 2002. Quantitative characterization of mixing processes in rotary
particulate solids in rotary drums: scale-up of residence time distribution using calciners. Powder Technology 126 (2), 166–173.
the axial dispersion model. Powder Technology 27 (2), 241–250. Weineköter, R., Reh, L., 1995. Continuous mixing of fine particles. Particle and
Abouzeid, A.Z.M., et al., 1974. The influence of operating variables on the residence Particle Systems Characterization 12 (1), 46–53.
time distribution for material transport in a continuous rotary drum. Powder Williams, J.C., Rahman, M.A., 1972. Prediction of the performance of continuous
Technology 10 (6), 273–288. mixers for particulate solids using residence time distributions: Part II.
Beaudry, J.P., 1948. Blender efficiency. Chemical Engineering 55, 112–113. Experimental. Powder Technology 5 (5), 307–316.
Danckwerts, P.V., 1953. Continuous flow systems: distribution of residence times. Williams, J.C., Richardson, R., 1982. The continuous mixing of segregating particles.
Chemical Engineering Science 2 (1), 1–13. Powder Technology 33 (1), 5–16.
Gao, Y., Muzzio, F.J., Ierapetritou, M.G. Characterization of feeder effects on Yeh, A.-I., Jaw, Y.-M., 1998. Modeling residence time distributions for single screw
continuous solid mixing using Fourier series analysis. AIChE Journal, in press extrusion process. Journal of Food Engineering 35 (2), 211–232.
(Corrected Proof: doi:10.1002/aic.12348). Ziegler, G.R., Aguilar, C.A., 2003. Residence time distribution in a co-rotating, twin-
Goldsmith, P.L., 1966. The theoretical performance of a semi-continuous blending screw continuous mixer by the step change method. Journal of Food Engineering
system. The Statistician 16 (3), 227–251. 59 (2–3), 161–167.

S-ar putea să vă placă și