Sie sind auf Seite 1von 8

Journal of Food Engineering 222 (2018) 276e283

Contents lists available at ScienceDirect

Journal of Food Engineering


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

The impact of clean-in-place parameters on rinse water effectiveness


and efficiency
Mengyuan Fan, David M. Phinney, Dennis R. Heldman*
Dale A. Seiberling Food Engineering Laboratory, The Ohio State University, 2015 Fyffe Road, Columbus, OH 43210, USA

a r t i c l e i n f o a b s t r a c t

Article history: Although Cleaning-in-Place (CIP) is abundantly used throughout the food industry, it is recognized that
Received 21 April 2017 CIP operations use significant amounts of water and energy. The overall objective of this investigation
Received in revised form was to evaluate parameters needed to improve the effectiveness of water use during CIP pre-rinse. A
20 November 2017
pilot-scale CIP system was operated over a range of Reynolds number (Re) from 16,000 to 260,000, while
Accepted 21 November 2017
evaluating the effectiveness of rinse water to remove a reconstituted skim milk residue film from
Available online 22 November 2017
stainless steel pipe surfaces. Rinse water effectiveness was quantified by comparing the protein con-
centration on the pipe surface after pre-rinse to the initial level. As the Re increased, the effectiveness of
Keywords:
Clean-in-place
rinse step increased, but not in linear proportionality. The efficiency of the rinse water decreased
Rinsing significantly as the volume of rinse water increased. The results of this investigation provide the basis for
Optimization reducing water and energy requirements during CIP operations.
Water conservation © 2017 Elsevier Ltd. All rights reserved.
Milk deposit

1. Introduction fouling have evolved, including modification of surfaces, pulsed


flow, ozonated water rinse, electrolyzed oxidizing rinse water, and
Cleaning is one of the most critical stages in quality control of ultrasonic cleaning (Augustin et al., 2010; Boxler et al., 2013;
food processing operations (Kulkarni et al., 1975). Failure to prop- Christian and Fryer, 2006; Dev et al., 2014; Gillham et al., 2000;
erly clean processing equipment not only increases the processing Guzel-Seydim et al., 2000; Muthukumaran et al., 2004; Rosma-
cost due to the decreased heat transfer coefficient, but also ninho et al., 2007). All these methods require additional devices,
threatens plant operations with microbial contamination (Fryer capital investment, and in some cases; energy inputs. When
et al., 2006; Gillham et al., 1999; Kulkarni et al., 1975; Mattila compared to the aforementioned solutions, the adjustment of CIP
et al., 1990). Cleaning-In-Place (CIP) is a widely-used method of control parameters (flow characteristics, water temperature, and
automated cleaning without the need for disassembling equip- contact time) to achieve improved cleaning efficiency while
ment, while reducing labor and time for cleaning (Seiberling, 1968). reducing water consumption would seem more practical for the
However, CIP systems require significant quantities of water, industry.
detergent, and energy. In addition, in-place cleaning generates Pre-rinse, as the first step of a CIP process, is to rinse fresh water
large quantities of waste water, with the additional economic through the processing system to drain until the discharge is clear
burden on the industry, and an environmental burden on the (Seiberling, 1968). In addition to providing the initial soil removal,
community (Lyndgaard et al., 2014). Optimization of CIP operations the pre-rinse may also assist the following alkaline cleaning step by
to reduce water and energy requirements, while maintaining hy- wetting the severe deposits (Khaldi et al., 2015). The wetted deposit
giene of the CIP process is a worthwhile goal. is consequently more vulnerable to the detergent (Goode et al.,
Over the past two decades, the basic understanding of the 2013). It is known that the direct physical force provided by the
fouling and cleaning processes in dairy plants has increased fluid would help the deposit to pass an energy barrier and detach
considerably (Fryer and Robbins, 2005; Wilson, 2005; Xin et al., (Fryer et al., 2006; Grant et al., 1997; Weidemann et al., 2014). Thus,
2002). Novel approaches to enhance cleaning and reducing any parameters influencing the physical force of the fluid would
affect pre-rinse step. Reynolds number (Re) is commonly used to
determine the flow characteristics (Singh and Heldman, 2013).
Besides Re, some researchers believed other parameters to also be
* Corresponding author.
E-mail address: heldman.20@osu.edu (D.R. Heldman). important in developing the prediction models for CIP operation.

https://doi.org/10.1016/j.jfoodeng.2017.11.029
0260-8774/© 2017 Elsevier Ltd. All rights reserved.
M. Fan et al. / Journal of Food Engineering 222 (2018) 276e283 277

For example, Timperley and Smeulders (1987) concluded that flow programmed to be a single-pass step in this experiment, which
velocity had a larger impact on the cleaning rate than Re. The means the rinse water was pumped to rinse the fouled pipe sec-
correlation between wall shear stress (tu ) and the removal of the tions then drained directly without reusing. The contact time be-
foulant from the surface has been demonstrated by many re- tween the rinse water and the pipe sections with deposit films was
searchers (Sharma et al., 1992; Visser, 1970). Lelievre et al. (2002) controlled to 20 s or 60 s during this study. Due to the flow sepa-
demonstrated that both the mean wall shear stress and the asso- ration and gravity, nine velocities levels were achieved in three
ciated time-dependent fluctuation rate would impact the detach- diameter pipes (0.025, 0.035 and 0.050 m) at three system set flow
ment of the foulant under turbulent flow conditions. rates. At the same system set flow rate, the velocity of the flow in
It is important to note that the papers mentioned above were the 0.035 m (1.5 inch) pipe section is highest due to the gravity,
based on the alkaline cleaning step which follows the pre-rinse step while the velocity of the 0.025 m (1 inch) pipe section is the lowest.
in CIP operation. Therefore, results from these studies would serve The pre-rinse experiment was performed three times from which
as good benchmark for investigating the mechanism of pre-rinse three subsamples were obtained under each condition.
step, which has not been studied much by comparison. Further-
more, to our knowledge, there are no prediction models in litera- 2.3. Data treatment
ture to describe the impact of Re and its interactions with other CIP
control parameters on the effectiveness of pre-rinse. Once built, After each CIP pre-rinse, the test sections were disassembled
these models can be used to optimize the CIP pre-rinse step, and the residual deposit film was removed from each pipe section
ensuring rinse effectiveness while reducing water/energy outlay by using a sodium hydroxide solution and the protein was quantified
locating the optimum CIP parameter sets. by Quantipro assay. The effectiveness of the CIP pre-rinse was
The objectives of this investigation were: (1) to establish the determined from the percent residual film, which was calculated by
relationship between rinse water effectiveness and Reynolds comparing the protein content after pre-rinse to the protein con-
Number at various temperatures and contact times, (2) to develop a tent of the deposit film prior to rinsing (Equation (1) and Equation
relationship between the efficiency of water use and rinse water (2))
requirements, and (3) to develop recommendations on CIP pa-
rameters for maximum water effectiveness while reducing water Percent residual film (RF%) ¼ 100 [C/C0] (1)
and energy demands.
Rinse effectiveness ¼ 100 e RF% (2)
2. Materials and methods
Where C (mg cm2) is the protein concentration after pre-rinse and
2.1. Creating the deposit film C0 (mg cm2) is the initial protein concentration.
The efficiency of water consumed during the pre-rinse step was
Rinsing experiments were conducted using pipe sections with evaluated from the rinse effectiveness data. Water efficiency (WE)
known quantities of a deposit film. The deposit film was created was defined as the amount of deposit protein (kg) removed per unit
using an established protocol (Fan et al., 2015). The deposit film was volume (m3) of water, as expressed in Equation (3)
applied to pipe sections with three different diameters: 0.025 m (1
inch), 0.038 m (1.5 inch) and 0.050 m (2 inch). Seven pipe sections WE ¼ PR / WV (3)
of the same diameter were combined to create a test line. The
middle and two end sections of each test line were tested after each Where PR is the mass of protein removed per surface area (kg m2),
experiment. The deposit film contained nonfat-dry-milk (NFDM) which equals to C0 minus C. WV is the volume of rinse water spent
(U.S. Foods, Inc., Rosemont, Ill., U.S.A.) reconstituted to 20% w/w in per surface area (m3 m2).
deionized water at ambient temperature. The pipe sections were As suggested in Equation (4), the volume of water consumed
filled with the reconstituted milk and held for 5 min at room (WV) was computed from the volumetric flow rate (V) of the rinse
temperature. The reconstituted milk was then removed and the water and the contact time (t) between the rinse water and the pipe
pipe sections were allowed to drain at room temperature for section, which was either 20 or 60 s in this study.
30 min. Next, each pipe section was placed in a 75  C oven for
30 min to dry the deposit film. Three pipe sections were selected to WV ¼ (V)(t) (4)
be used in the CIP pre-rinse experiment. The remaining pipe sec-
tions were used to measure the protein content in the initial de- In order to minimize the impact of the system size on the
posit film. The protein content of the deposit film was measured by models, both WE and WV were normalized to per fouled surface
washing the pipe interior using a sodium hydroxide solution. The area. The fouled surface areas were calculated based on the inner
protein assay was completed with QuantiPro™ Bicinchoninic Acid diameter and length of the pipe sections.
(BCA) (Sigma-Aldrich Co.).
2.4. Statistical analysis
2.2. Cleaning-in-place (CIP) pre-rinse
SPSS.21 (IBM, Inc,. USA) statistics software was used for analysis
The cleaning-in-place (CIP) pre-rinse experiments were per- of all data. The data have been reported as the means ± standard
formed by mounting the fouled pipe sections on a test manifold, error. Analysis of covariance (ANCOVA) with 95% confidence in-
which is a part of a pilot-scale CIP system. The CIP system and test terval was used to evaluate the statistical difference of the intercept
manifold were described in Fan et al. (2015). Tap water e pH from and slope between the linear regressions. A one-way ANOVA with a
7.2 to 7.7 (Model HI, 2020; Hanna Instruments, Inc., Woonsocket, 95% confidence interval was used to evaluate the statistical differ-
RI., U.S.A.) e was used in this study. Tap water was recirculated ence between the means.
through the CIP system, except for pipe sections with deposit films,
until the flow rates (0.0016, 0.0028 or 0.0039 m3 s1) and tem- 3. Results and discussion
peratures (22, 45 or 67  C) stabilized. Next, the flow of rinse water
was directed to pipe sections with deposit films. Pre-rinse was The influence of three CIP parameters (temperature, velocity,
278 M. Fan et al. / Journal of Food Engineering 222 (2018) 276e283

contact time) on rinse effectiveness was evaluated. Three rinse 30 a


water temperature levels (22  C, 45  C, and 67  C) were selected to 20 Seconds
represent the range of temperatures used during CIP operations; 25 b b 60 Seconds
the ambient water temperature (22  C), 67  C - a recommended

Residual Film (RF%)


bc
rinse water temperature (between 60  C and 80  C) by researchers 20 c
(Li et al., 2013; Liu et al., 2006; Xin et al., 2002), and an intermediate
d
temperature between the two extremes (45  C). Nine rinse water 15
velocities between 0.73 and 2.48 m s1, including the recom- e e e
mended CIP operation flow velocity of 1.52 m s1, were applied in 10 f f f
f
this research. j j j j
The magnitude of the residual film on the interior surface of the 5
pipe sections after pre-rinse is expressed as the residual film per-
centage (RF%) to describe the rinse effectiveness. A lower magni- 0
tude of residual film percentage implies a better rinsing 16080 26190 62300 77160 87380 108150 176140 211060 261230
effectiveness. Reynolds Number (Re)

Fig. 2. Improvements in removal of residual film (RF%) deposits as a function of


3.1. Effect of Reynolds Number on rinse effectiveness Reynolds Number (Re) for 20 and 60-s rinse water residence times.

The analysis was conducted over a range of Reynolds Numbers


between 20,000 and 300,000. In Fig. 1, the residual film percentage significantly as the pumping rate increased (Reynolds number
(RF%) after a 20-s rinse with water at 22  C is presented as a ranges from 85,000 to 600,000). This could be attributed to the
function of Reynolds Number. A linear correlation is found between larger mechanical force and a better mixing effect provided by the
natural logarithm of residual film percentage (RF%) and Reynolds high level of turbulence (higher Re). It has been previously reported
Number (Equation (5)). that generally, a high Re implies instability of flow and significant
mixing within the region of flow (Meyer, 1971). These flow condi-
Ln[RF%] ¼ 3.23 - 1.03  105 [Re] (5) tions are expected to provide larger mechanical forces and
enhanced turbulent diffusion (Timperley and Smeulders, 1987).
Based on the linear correlation expression, the (deposit) residual Both of these mechanisms should have a positive impact on the
film removal is 81.4% at Re ¼ 20,000, and increased to 91.6% at rinse effectiveness. Larger direct mechanical forces can improve
Re ¼ 100,000. Although the rinse effectiveness improved as the rinsing by breaking down and removing the deposits during the
Reynolds Number increased (as expected), the improvement in initial phase of rinsing (Blel et al., 2009). Flow characteristic can
rinse effectiveness is not proportional to the increase in Reynolds impact the mass transfer rate of the detached deposit particles into
Number. For example, when Re increased by five times from 20,000 the cleaning fluid bulk (Gallot-Lavallee et al., 1984; Jensen et al.,
to 100,000; the deposit film removal only increased from 81.4% to 2005). A high mass transfer rate enhanced rinsing due to turbu-
91.6%. The five times higher change in Reynolds Number is achieved lent mixing and the exchange of deposit materials between surface
by elevating temperature and increasing flow velocity by increasing and bulk (Gillham et al., 1999; Weidemann et al., 2013; Xin et al.,
pump speed, both of which require a certain amount of energy 2002).
input. Fig. 2 shows the influence of Reynolds number on the RF% after
Similar to the results from this investigation, Hankinson and various rinsing times. Similar to the results presented in Fig. 1,
Carver (1968) also found that rinsing effectiveness improved higher percentages of residual film removal occurred in

4
3.5
ln (Residual Film%)

3
2.5
2
1.5
Ln[RF%]= 3.23 - 1.03 x 10-5 [Re]
1
R² = 0.77
0.5
0
0 20000 40000 60000 80000 100000 120000
Reynolds Number (Re)

Fig. 1. Influence of Reynolds Number (Re) on residual film deposit (RF%) with rinse water temperature of 22  C, and rinse water residence time of 20 s.
M. Fan et al. / Journal of Food Engineering 222 (2018) 276e283 279

Table 1 slopes of different contact times at the same rinse water temper-
ANCOVA results on Regressions of ln (residual film%) over Reynolds Number. ature were not significant. Since the slopes were not significantly
Temperature ( C) Contact time (s) Slope Intercept different, the intercepts of the regression expressions were iden-
22 20 1.03E-05a 3.23 ± 0.14a
tified as an indicator of the rinse effectiveness. Lower values of the
22 60 1.01E-05a 2.99 ± 0.21b intercept suggests the presence of less residual film on the pipe
45 20 6.53E-06b 3.10 ± 0.18b interior surface after rinsing, which meant achievement of better
45 60 6.04E-06b 2.83 ± 0.30c rinse effectiveness.
67 20 4.23E-06b 2.68 ± 0.25c
The impact of contact time on the rinsing effectiveness was
67 60 4.58E-06b 2.73 ± 0.21c
observed in Table 1. The intercept coefficient after 60 s rinsing at
Note- Values with the same superscript letter are not significantly different
22  C was significantly lower than that after 20 s of rinsing at the
(a ¼ 0.05).
same temperature. This indicates that a 60-s rinse at 22  C is
significantly more effective than the 20-s rinse at the same tem-
experiments conducted at higher Reynolds numbers. It needs to be perature. A similar trend was observed on comparison of the in-
noted that extending the contact time from 20 s to 60 s under a high tercepts after 20 and 60 s of rinsing at 45  C. However, when the
level of turbulence (high Re range) did not significantly influence rinse water temperature was increased to 67  C, the difference of
the rinse effectiveness. In another words, the effect of the Re level intercepts between the 20 and 60 s contact times was not signifi-
on rinse effectiveness seems to be the most dominant factor in the cant. These results suggest that rinse effectiveness can be improved
system. If that is the case, it would suggest that longer rinse contact by rinsing for longer periods at 22 and 45  C, but no significant
time is not needed at higher Re level. improvement is achieved when the water temperature is 67  C. It is
A one-way ANOVA with a 95% confidence interval was used to possible that 20 s of contact was sufficient to remove all the water-
analyze the difference in the means of percent residual film after 20 soluble deposit due to the higher diffusivity at elevated tempera-
and 60 s rinsing over the range of Reynolds numbers. Similar to the tures (67  C). Therefore, when the contact time was increased from
results presented in Fig. 1, higher percentages of residual film were 20 to 60 s at 67  C, the rinse effectiveness was not improved. The
removed in experiments conducted at higher Reynolds numbers. It portion of the deposit that was easily soluble in water was removed
was observed that when the Reynolds Number was equal to or after 20 s of contact with rinse water and the portion of the deposit
below 87,380, the effectiveness of rinsing was always significantly attached more tightly to the wall could not be removed even after
improved at longer contact times. Extending the contact time (from 60 s of rinsing. As previously suggested, the deposited film cannot
20 s to 60 s) at higher Re did not significantly influence the rinse be removed by rinse water only, and must be removed by the
effectiveness. This suggests Reynolds number, rather than contact cleaning cycle of the CIP operation (Goode et al., 2013).
time, was the dominant factor impacting rinse effectiveness. The impact of rinse water temperature on the rinsing effec-
Therefore it follows that longer rinse contact times are not needed tiveness is also revealed in Table 1. The slopes of regressions from
at higher Re levels. experiments when rinse water was at 22  C as compared to rinse
This result is consistent with literature. It has been previously water at 45  C were significantly different for both 20 s pre-rinse
suggested that higher turbulence plays a key role in affecting the and 60 s pre-rinse. These observations suggest that the rinse
mass transfer during the initial seconds of contact (Weidemann effectiveness was improved by increasing the rinse water temper-
et al., 2013). According to Goode et al. (2013), there is a point ature from 22 to 45  C. Otto et al. (2014) reported a similar result
during rinse where no further deposit can be removed by water with an approximately 37% increase in cleaning rate of whey pro-
alone and the remaining deposit can only be removed after tein deposit when the temperature of the cleaning fluid was
chemical agents are introduced. It was further stated that all water- increased from 5 to 40  C. These observations indicate that tem-
removable deposit is removed within a short contact time in a high perature of the rinse water impacts the diffusivity of water within
Re system due to the strong physical force introduced by the tur- the deposited film and consequently, influences the rinse water
bulence. The remaining deposit is not removable by rinse water, effectiveness. This is consistent with common observations about
which explains the observation from this study that extended water diffusion into porous media (Rastogi et al., 2002). Higher
contact times do not improve rinse effectiveness. Thus under a high temperatures are expected to enhance the diffusivity of water into a
level of rinse water turbulence, usually measured by a high level of porous media which in this case is the dairy deposit film. Other
Reynolds Number, extending the rinse water contact time is un- researchers have also reported the increase in mass transfer at the
necessary and cannot improve the rinse effectiveness significantly. interface between rinse water and deposit film at higher temper-
atures (Mercade -Prieto and Chen, 2006; Xin et al., 2002). These
studies concluded that high temperature water enhances rinsing
3.2. Combined effect of temperature and contact time on rinse through better mixing and more rapid exchange between film
effectiveness surface and water flowing over the surface (Gillham et al., 1999).
Interestingly, compared to the significant benefits of increasing
Under experimental conditions stated in materials and methods the rinse water temperature from room temperature to an inter-
section, the coefficients of natural log of residual film percentage as mediate temperature of 45  C on deposit removal, increasing the
a function of Reynolds Number were determined at various contact water temperature to the recommended CIP temperature of 67  C
times and temperatures, similar to the regression presented in did not significantly improve the slope of the regression (Table 1).
Fig. 1. These slope and intercept coefficients are reported in Table 1. In other words, the difference between regression slopes for rinse
Higher slope and/or lower intercept of these regressions generally water at 45  C and 67  C was not statistically significant. However,
implies a better rinse effectiveness. The influence of contact time the intercept for a 20-s rinse at 45  C was found to be significantly
and rinse water temperature on rinse effectiveness was evaluated higher than the following-rinsing longer at the same temperature,
by conducting “analysis of the correlation coefficients” (ANCOVA). 20 s at 67  C, and 60 s at 67  C. As mentioned previously, lower
The analysis of covariance (ANCOVA) was used to compare the values of the intercept indicate lower RF% or better rinse effec-
slopes and intercepts of each regression (a ¼ 0.05) and the results tiveness. This suggests that rinse effectiveness was improved by
were presented in Table 1. longer contact times at 45  C. Furthermore, rinse effectiveness was
The results in Table 1 indicated that the differences among not influenced by an increase in temperature from 45 to 67  C when
280 M. Fan et al. / Journal of Food Engineering 222 (2018) 276e283

3.5

3
Ln (Residual Film%)
2.5

1.5
Ln[RF%] = 3.28-0.99 [WV]
1 R² = 0.77
0.5

0
0 0.5 1 1.5
3 -2
Volume of Consumed Water/Fouled Surface Area, WV (m m )

Fig. 3. Improvements in removal of residual film deposits (RF%) with increased water use at rinse water temperature of 22  C and after 20-s rinse water residence time.

the contact time was 60 s. A possible explanation could be that 60 s effectiveness rinsing due to higher mechanical forces, along with
of rinsing is long enough to remove all the water-removable de- better mixing and transfer at the interface between the residue film
posits regardless of the mass transfer rate which is impacted by the and the rinse water.
rinse water temperature. Similar to the linear correlation obtained for 22  C and a 20-s
rinse in Fig. 3, a linear correlation was found between the natural
3.3. Effect of the used water volume on rinse effectiveness logarithm of RF% and volume of water (WV) per fouled surface at
67  C after a 60-s rinse (Eq. (7))
In this section, the correlation between percent residual film
and the volume of water used per fouled surface area was evaluated Ln [RF%] ¼ 2.66e0.3 [WV] (7)
at three different temperatures (22  C, 45  C and 67  C) and two
different contact times. The range of water volumes used per fouled Based on this regression, only 0.38 m3 of water per m2 fouled
surface area was from 0.18 to 3.54 m3 water per m2 fouled surface. surface was needed to remove 87.3% of the residue film. An addi-
The percent residual film (RF%) after the 20 s of pre-rinse at tional 3.7 m3 of water per m2 fouled surface was required to remove
22  C as a function of the volume of water used per fouled area is and additional 8.5% of residue film. This indicates that the water
presented in Fig. 3. A linear correlation was found between the cost increases by nearly 10 times for an improvement in residual
natural logarithm of the residual film (RF%) and the water volume removal from 87.3% to 95.8%. When compared to the film deposit
(WV) per fouled surface area as given by Eq. (6). removed after 20 s of rinse with water at 22  C, improved rinse
effectiveness was achieved with higher water temperatures and
Ln [RF%] ¼ 3.28e0.99 [WV] (6) longer contact times.
Similar regressions were developed for all combinations of
Based on this regression, removal of 79.6% of the film residue contact times and rinse water temperatures. Additional analysis
required only 0.27 m3 of water per m2 fouled surface. However, was accomplished by dividing the regression coefficients into two
removal of an additional 12.7% of the film residue would require groups based on contact times. An analysis of covariance (ANCOVA)
1.24 m3 of water per m2 fouled surface. These results indicate that was used to compare the slopes and intercepts from each regres-
the effectiveness of rinse water decreased as the level of RF% sion (a ¼ 0.05) and the results are presented in Table 2. The results
became smaller. It must be emphasized that relatively small vol- of the analysis indicated that the slopes of the linear regressions
umes of rinse water (0.18 m3 of water per m2 fouled surface) were
required to remove significant amounts (79.6%) of the residue film.
The decrease in effectiveness of the rinse water as larger volumes Table 2
are used suggests that reducing the duration of the rinse cycle ANCOVA results for coefficient for regressions of ln (residual film%) over volume of
would provide an opportunity for significant reduction in water consumed water per unit fouled surface after 20 and 60 s of pre-rinse.

use. Contact time (s) Temperature ( C) Slope Intercept


For a fixed CIP system, the volume of consumed water is a
20 22 0.99 3.28 ± 0.12a
dependent on the flow velocity and the contact time between the 45 0.99a 3.01 ± 0.10b
rinse water and the residue film. Larger amounts of water are used 67 0.89a 2.65 ± 0.13c
when the water velocities are higher. The water consumption 60 22 0.31a 3.09 ± 0.19a
values presented on the horizontal axis were achieved by the 45 0.32a 2.85 ± 0.14b
different flow velocities as illustrated in Fig. 3. Previous researchers 67 0.30a 2.66 ± 0.12b
(Gillham et al., 1999; Weidemann et al., 2013; Xin et al., 2002) have Note- Values with the same superscript letter are not significantly different
illustrated that higher velocities of rinse water can improve (a ¼ 0.05).
M. Fan et al. / Journal of Food Engineering 222 (2018) 276e283 281

0.7

Area, WE (kg protein removed in m-3


0.6

Water Efficiency per Fouled Surface


0.5

0.4

rinse water)
[WE] = 0.0495 [WV] -1.41
0.3 R² = 0.83

0.2

0.1

0
0 1 2 3 4
Volume of Consumed Water/Fouled Surface Area, WV (m3 m-2)

Fig. 4. The influence of rinse water volume on the efficiency of residual film removal.

were not significantly different for both contact times and tem- during the initial phases of rinsing, the amount of residue removed
peratures. For the 20-s contact time, the intercept of the linear per unit of water used is largest. Obviously, the initial contact be-
regression decreased significantly when the rinse water tempera- tween water and deposit film is when the efficiency of the rinse
ture was increased from 22 to 45  C and from 45 to 67  C. These water is highest. When the quantities of water used are greater
results suggest that an increase in rinse water temperature from 22 than 1 m3 m2, the water efficiency decreased until it reached a
to 67  C provided an improvement in the effectiveness of rinse relatively low and constant level. Even though the magnitudes of
water at the shorter contact time of 20 s. The same improvement WE continued to be positive, the small WE values may not justify
was not found for the 60 s pre-rinse. It was observed that there was the large quantities of water required (WV) as we move to the re-
no significant difference between intercepts of the linear re- gion in the right of this curve (Fig. 4). In this region, each increment
gressions when the temperature was increased from 45 to 67  C. of deposit removed required a significantly greater amount of
These results suggest that higher temperatures (67  C) do not water.
improve rinse effectiveness when the contact times are relatively These results indicate that the rinse water will continue to
long (60 s). remove film deposits from the pipe surfaces after the water effi-
Based on the results and discussions in this section, it was ciency reach a constant level. The relationship describing the water
concluded that a pre-rinse of 60 s at 67  C is excessive. At these efficiency as a function of water volume used, as given in Eq. (8),
conditions, no significant improvement occurred in deposit must be considered when attempting to optimize the entire
removal. Moreover, these conditions required high energy and time cleaning operation.
inputs. Pre-rinse for 20 s at 45  C would be an appropriate balance
1.41
between rinse effectiveness and energy and time input. A more [WE] ¼ 0.0495 [WV] (8)
specific consideration about water cost and rinse effectiveness is
discussed in next section. In order to evaluate the impact of temperature on water effi-
ciency, the relationships between water efficiency and water vol-
3.4. Changes of rinse water efficiency during the pre-rinse step ume at various temperatures were investigated. In order to change
Eq. (8) into a linear relationship, the natural logarithm was applied
In this study, water efficiency (WE) has been defined as the mass to both water efficiency and volume of water. As reported in Table 3,
of protein film removed per fouled surface area over volume of after the conversion, the natural logarithm of water efficiency
rinsing water per fouled surface area (Eq. (3)). Previous research decreased in a linear fashion (R2 ¼ 0.83) as the nature logarithm of
has indicated that the residue removal rate during the rinse step is water volume increased. The results indicate that the water effi-
not constant (Xin et al., 2002; Weidemann et al., 2013), and the ciency decreased rapidly with an increase in water volume at all
water efficiency (WE) is not expected to be constant. Based on re- temperatures investigated.
sults from this study, as well as observations from other researchers By comparing the slopes and intercepts of each linear regression
(Arnold and Maxcy, 1970; Harper, 1972), longer contact times may
not improve rinse effectiveness when the water temperatures are
high. Optimization of the water rinse step may be achieved by Table 3
knowing the relationship between water efficiency and volume of ANCOVA results on regressions between ln (water efficiency) and ln (water volume/
fouled surface area).
used water.
The results in Fig. 4 illustrate the relationship between water Temperature ( C) Slope Intercept R2
efficiency (WE) and water volume (WV) at 22  C, both WE and WV 22 1.41a 3.01 ± 0.012a 0.83
mentioned here are normalized to per unit of fouled surface area. 45 1.45b 2.94 ± 0.013b 0.83
When the volume of water used was lower, the water efficiency 67 1.45b 2.97 ± 0.014b 0.84
was very high. As the water consumption increased, the efficiency Note- Values with the same superscript letter are not significantly different
decreased rapidly. One interpretation of this relationship is that (a ¼ 0.05).
282 M. Fan et al. / Journal of Food Engineering 222 (2018) 276e283

(p ¼ 0.05) using ANCOVA, the impact of water temperature on References


water efficiency was analyzed (Table 3). The regression parameters
for 22  C rinse water were significantly lower when compared to Arnold, R., Maxcy, R., 1970. Evaluating circulation cleaning by analysis of soil
depletion from surfaces. J. Dairy Sci. 53 (11), 1540e1544.
parameters for the linear regression at 45  C and 67  C. These re- Augustin, W., Fuchs, T., Fo €ste, H., Scho €ler, M., Majschak, J.P., Scholl, S., 2010. Pulsed
sults suggest that at a lower temperature (22  C), the water effi- flow for enhanced cleaning in food processing. Food Bioprod. Process. 88 (4),
ciency is low. When the temperature was increased from 45 to 384e391.
Blel, W., Legentilhomme, P., Benezech, T., Legrand, J., Le Gentil-Lelievre, C., 2009.
67  C, the differences in regression coefficients were not significant. Application of turbulent pulsating flows to the bacterial removal during a
This suggests that an increase in rinse water temperature from 45 cleaning in place procedure. Part 2: effects on cleaning efficiency. J. Food Eng. 90
to 67  C does not improve water efficiency. The observation tends to (4), 433e440.
Boxler, C., Augustin, W., Scholl, S., 2013. Cleaning of whey protein and milk salts
confirm earlier results about the effect of rinse water temperature soiled on DLC coated surfaces at high-temperature. J. Food Eng. 114 (1), 29e38.
on the effectiveness of rinsing. Christian, G.K., Fryer, P.J., 2006. The effect of pulsing cleaning chemicals on the
cleaning of whey protein deposits. Food Bioprod. Process. 84 (4), 320e328.
Dev, S.R., Demirci, A., Graves, R.E., Puri, V.M., 2014. Optimization and modeling of an
3.5. Overall observations
electrolyzed oxidizing water based Clean-In-Place technique for farm milking
systems using a pilot-scale milking system. J. Food Eng. 135, 1e10.
Rinse effectiveness is not the only factor required during opti- Fan, M., Phinney, D.M., Heldman, D.R., 2015. Effectiveness of rinse water during in-
place cleaning of stainless steel pipe lines. J. Food Sci. 80 (7), E1490eE1497.
mization of water rinse step. Water efficiency may also be consid-
https://doi.org/10.1111/1750-3841.12914.
ered since it has a huge impact on the conservation of water for the Fryer, P.J., Robbins, P.T., 2005. Heat transfer in food processing: ensuring product
entire cleaning process. According to results of the current study, quality and safety. Appl. Therm. Eng. 25 (16), 2499e2510.
the water efficiency decreases rapidly as more water is used Fryer, P.J., Christian, G.K., Liu, W., 2006. How hygiene happens: physics and chem-
istry of cleaning. Int. J. Dairy Technol. 59 (2), 76e84.
regardless of the change in temperature. Gallot-Lavallee, T.H.I.E.R.R.Y., Lalande, M., Corrieu, G., 1984. Cleaning kinetics
A balance between rinse effectiveness and water efficiency is modeling of holding tubes fouled during milk pasteurization. J. Food Process
required for CIP optimization. A CIP rinse water regime targeting Eng. 7 (2), 123e142. https://doi.org/10.1111/j.1745-4530.1984.tb00642.x.
Gillham, C.R., Fryer, P.J., Hasting, A.P.M., Wilson, D.I., 1999. Cleaning-in-place of
higher rinse effectiveness regardless of the lower water efficiency is whey protein fouling deposits: mechanisms controlling cleaning. Food Bioprod.
not recommended. There are limited benefits from using additional Process. 77 (2), 127e136.
amounts of water during pre-rinse to remove strongly attached Gillham, C.R., Fryer, P.J., Hasting, A.P.M., Wilson, D.I., 2000. Enhanced cleaning of
whey protein soils using pulsed flows. J. Food Eng. 46 (3), 199e209.
deposits, as any residual deposit can be removed by the detergent Goode, K.R., Asteriadou, K., Robbins, P.T., Fryer, P.J., 2013. Fouling and cleaning
and acid wash steps to follow. studies in the food and beverage industry classified by cleaning type. Compr.
Rev. Food Sci. food Saf. 12 (2), 121e143. https://doi.org/10.1111/1541-
4337.12000.
4. Conclusions
Grant, C.S., Webb, G.E., Jeon, Y.W., 1997. A noninvasive study of milk cleaning pro-
cesses: calcium phosphate removal. J. food process Eng. 20 (3), 197e230.
The results of this investigation confirm the feasibility of opti- Guzel-Seydim, Z.B., Wyffels, J.T., Greene, A.K., Bodine, A.B., 2000. Removal of dairy
soil from heated stainless steel surfaces: use of ozonated water as a Prerinse1.
mizing water usage during CIP operations through the selection of
J. Dairy Sci. 83 (8), 1887e1891.
control parameters. Effectiveness of rinse water in deposit removal Hankinson, D.J., Carver, C.E., 1968. Fluid dynamic relationships involved in circu-
was 81.4% at Reynolds Numbers of 20,000, and improved to only lation Cleaning1. J. dairy Sci. 51 (11), 1761e1767. https://doi.org/10.3168/
91.6% at a Reynolds number of 100,000; an improvement of only jds.S0022-0302(68)87273-X.
Harper, W.J., 1972. Sanitation in dairy food plants. In: Gothie, R.K. (Ed.), Food Plant
10% in effectiveness. Sanitation. Avi Pub. Co., Westport, USA, pp. 130e160.
Rinse water effectiveness was not significantly improved when ne
Jensen, B.B.B., Friis, A., Be zech, T., Legentilhomme, P., Lelievre, C., 2005. Local wall
the rinse water temperature was increased from 45 to 67  C. At a shear stress variations predicted by computational fluid dynamics for hygienic
design. Food Bioprod. Process. 83 (1), 53e60. https://doi.org/10.1205/fbp.04021.
rinse water temperature of 67  C, an increase in contact time from Khaldi, M., Blanpain-Avet, P., Gue rin, R., Ronse, G., Bouvier, L., Andre , C.,
20 to 60 s did not improve the rinse water effectiveness signifi- Delaplace, G., 2015. Effect of calcium content and flow regime on whey protein
cantly. At 67  C, only 1.01 m3 of water per m2 fouled surface were fouling and cleaning in a plate heat exchanger. J. Food Eng. 147, 68e78.
Kulkarni, S.M., Maxcy, R.B., Arnold, R.G., 1975. Evaluation of soil deposition and
required to remove 89% of the residue film from the surface. The removal processes: an interpretive Review1. J. Dairy Sci. 58 (12), 1922e1936.
improvement of residual removal from 89 to 95% required 3.64 m3 vre, C., Legentilhomme, P., Gaucher, C., Legrand, J., Faille, C., Be
Lelie nezech, T., 2002.
water per m2 of fouled surface. It seems evident that significant Cleaning in place: effect of local wall shear stress variation on bacterial removal
from stainless steel equipment. Chem. Eng. Sci. 57 (8), 1287e1297. https://
reduction in rinse water use (72%) can be achieved by reducing the
doi.org/10.1016/S0009-2509(02)00019-2.
amount of residue removed during the initial rinse water by only a Li, L., Che, L., Liang, Q., Mercade -Prieto, R., Wu, X., Chen, X.D., 2013. Study on the
few percent. When the efficiency of water use is expressed as the dissolution of heat-induced ovalbumin gel in alkaline solutions relevant to the
removal of fouling deposits. J. Food Eng. 114 (4), 550e557.
quantity of residue removed as a function of water volume per unit
Liu, W., Zhang, Z., Fryer, P.J., 2006. Identification and modelling of different removal
area of fouled surface, the water efficiency decreased from modes in the cleaning of a model food deposit. Chem. Eng. Sci. 61 (22),
0.017 kg m3 at limited amounts of water to approximately 7528e7534.
0.001 kg m3 when the amounts of rinse water were very high. The Lyndgaard, C.B., Rasmussen, M.A., Engelsen, S.B., Thaysen, D., van den Berg, F., 2014.
Moving from recipe-driven to measurement-based cleaning procedures:
efficiency of rinse water use decreased rapidly with rinse time to monitoring the Cleaning-In-Place process of whey filtration units by ultraviolet
about one cubic meter of water per square meter of fouled surface. spectroscopy and chemometrics. J. Food Eng. 126, 82e88. https://doi.org/
A rinse water step at 45  C for a residence time of 20 s was more 10.1016/j.jfoodeng.2013.10.037.
Mattila, T., Manninen, M., Kyla €siurola, A.L., 1990. Effect of cleaning-in-place disin-
efficient than 65  C for 60 s. In general, the efficiency in deposit fectants on wild bacterial strains isolated from a milking line. J. dairy Res. 57
removal is not improved significantly at volumes of water used (01), 33e39. https://doi.org/10.1017/S0022029900026583.
above 1 m3 m2 of fouled surface. Mercade -Prieto, R., Chen, X.D., 2006. Dissolution of whey protein concentrate gels
in alkali. AIChE J. 52 (2), 792e803. https://doi.org/10.1002/aic.10639.
Meyer, R.E., 1971. Introduction to Mathematical Fluid Dynamics, vol. 24. Courier
Acknowledgements Corporation, North Chelmsford, 289e205.
Muthukumaran, S., Yang, K., Seuren, A., Kentish, S., Ashokkumar, M., Stevens, G.W.,
Grieser, F., 2004. The use of ultrasonic cleaning for ultrafiltration membranes in
The authors wish to acknowledge the generous donations from
the dairy industry. Sep. Purif. Technol. 39 (1), 99e107. https://doi.org/10.1016/
Dale A. Seiberling to the Food Engineering Research Laboratory at j.seppur.2003.12.013.
The Ohio State University. This work was also supported in part by Otto, C., Zahn, S., Plenker, J., Rohm, H., 2014. Application of a flow cell for the
comparative investigation of the cleaning behavior of starch and protein. J. Food
the USDA National Institute of Food and Agriculture, Hatch project
Eng. 131, 1e6.
232768.
M. Fan et al. / Journal of Food Engineering 222 (2018) 276e283 283

Rastogi, N.K., Raghavarao, K.S.M.S., Niranjan, K., Knorr, D., 2002. Recent de- 40 (1), 4e7.
velopments in osmotic dehydration: methods to enhance mass transfer. Trends Visser, J., 1970. Measurement of the force of adhesion between submicron carbon-
Food Sci. Technol. 13 (2), 48e59. black particles and a cellulose film in aqueous solution. J. Colloid Interface Sci.
Rosmaninho, R., Santos, O., Nylander, T., Paulsson, M., Beuf, M., Benezech, T., Müller- 34 (1), 26e31. https://doi.org/10.1016/0021-9797(70)90254-7.
Steinhagen, H., 2007. Modified stainless steel surfaces targeted to reduce fou- Weidemann, C., Stahl, S., Nirschl, H., 2013. Development of a qualitative test method
lingeevaluation of fouling by milk components. J. Food Eng. 80 (4), 1176e1187. for the cleanability of polymer woven filter media. Food Bioprod. Process. 91
Seiberling, D.A., 1968. Equipment and process design as related to mechanical/ (4), 515e524. https://doi.org/10.1016/j.fbp.2013.05.005.
chemical cleaning procedures. In: Chem. Eng. Prog., Symp. Ser, vol. 64, p. 86. Weidemann, C., Vogt, S., Nirschl, H., 2014. Cleaning of filter media by pulsed
Sharma, M.M., Chamoun, H., Sarma, D.S.R., Schechter, R.S., 1992. Factors controlling floweEstablishment of dimensionless operation numbers describing the
the hydrodynamic detachment of particles from surfaces. J. Colloid Interface Sci. cleaning result. J. Food Eng. 132, 29e38.
149 (1), 121e134. https://doi.org/10.1016/0021-9797(92)90398-6. Wilson, D.I., 2005. Challenges in cleaning: recent developments and future pros-
Singh, R.P., Heldman, D.R., 2013. Introduction to Food Engineering, fifth ed. Gulf pects. Heat. Transf. Eng. 26 (1), 51e59. https://doi.org/10.1080/
Professional Publishing, pp. 80e95. 01457630590890175.
Timperley, D.A., Smeulders, C.N.M., 1987. Cleaning of dairy HTST plate heat ex- €
Xin, H., Chen, X.D., Ozkan, N., 2002. Cleaning rate in the uniform cleaning stage for
changers: comparison of single-and two-stage procedures. Int. J. Dairy Technol. whey protein gel deposits. Food Bioprod. Process. 80 (4), 240e246.

Das könnte Ihnen auch gefallen