Вы находитесь на странице: 1из 7

Development of hypothetical components for milk process simulation

using a commercial process simulator


Y. Zhang, Muhammad Tajammal Munir, W. Yu, B.R. Young

Industrial Information & Control Centre (I


2
C
2
), Chemical and Materials Engineering Department, The University of Auckland, New Zealand
a r t i c l e i n f o
Article history:
Received 17 December 2012
Received in revised form 24 May 2013
Accepted 19 July 2013
Available online 17 August 2013
Keywords:
Process simulation
Milk processing
Hypothetical components
a b s t r a c t
Commercial process simulators do not contain all the components required for simulation of milk
processing. The aim of this work was to create credible hypothetical components to simulate milk pro-
cessing using commercial process simulators. To create milk components in the simulator component
library, a hypothetical component database was built in a commercial process simulator with the values
and functions to estimate the necessary physical properties of the milk. As a result process simulation can
now be extended to the dairy industry.
2013 Elsevier Ltd. All rights reserved.
1. Introduction
The use of process modeling and simulation within process
engineering is well recognized as process simulators are powerful
tools for model based calculations for example in process design.
They help to predict certain process behaviours accurately enough
at low cost without putting the real process (equipment or unit
operation) at risk. Modeling and simulation play a very important
role in critical process decision making and optimization
(Rodrigues and Minceva, 2005; Theien et al., 2011). Despite some
(generally expected) differences between process simulation and
real-life operation, process simulators can provide reliable infor-
mation on process operation due to the existence of comprehen-
sive thermodynamic packages, vast component libraries and
advanced computational methods (West et al., 2008; Garca
et al., 2010).
Rigorous process simulators such as Aspen Plus/Aspen Dynam-
ics, HYSYS and VMGSim have increasingly been used in recent
years for modeling and simulation tasks in many processes from
different industries (Ruiz et al., 2010; Daz et al., 2011; Peters
et al., 2011; Zhao et al., 2011; Munir et al., 2012a,b). Process sim-
ulators can directly provide steady state models, dynamic models,
and performance analysis (e.g., Luyben, 2002).
Modeling and simulation of processes involving chemical reac-
tants or products are well developed where sufcient physical
property data and prediction models are available (Dnnebier
and Klatt, 2000; Han and Chung, 2001). Their application to food
processing has lagged behind due to the diversity of food process-
ing, the great variety of food products, and the complex physical,
chemical and biological structure of foods, which are mostly solid,
semi-solid or in some cases liquid, e.g. liquid milk. Most food mate-
rials also have highly complex compositions and their properties
change irreversibly during the process (Wang and Hirai, 2011;
Trystram, 2012). The complexities involved in the prediction of
food material properties also bring an added difculty to predict
their behaviour as a function of the operating conditions, i.e. T, P.
These food material properties cause more difculties when mod-
eling processes involving food reactants or products (Bon et al.,
2010). These difculties and complexities involved in food materi-
als are the main reasons for the absence of food components in the
component libraries of commercial process simulators. This limits
general process simulator applicability to the food industries.
Most commercial process simulators (e.g. Aspen Plus, HYSYS
and VMGSim) were developed principally for the applications in
the eld of chemical engineering. However, there are a few rare
software examples for food engineering applications.
Bon (2005) and Bon et al. (2010) presented a ProSimPlus simula-
tion ow sheet for milk pasteurisation using ideal gas and liquid
phase activity coefcient models. No validation of the milk compo-
nents thermo physical properties was presented. Halder et al.
(2011) and Abakarov and Nuez (2012) presented software examples
for food engineering applications but these are lacking the most
important process simulator capability unit operation modules.
The use of a commercial process simulator for food process stud-
ies is appropriate due to advantages derived fromthe availability of
large number of simulation modules, both for unit operations and
information management, and the ease with which it can simulate
a process. A process simulator requires food components in its
0260-8774/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jfoodeng.2013.07.033

Corresponding author. Tel.: +64 9 923 5606; fax: +64 9 373 7463.
E-mail addresses: tajammal.munir@auckland.ac.nz (M.T. Munir), b.young@
auckland.ac.nz (B.R. Young).
Journal of Food Engineering 121 (2014) 8793
Contents lists available at ScienceDirect
Journal of Food Engineering
j our nal homepage: www. el sevi er. com/ l ocat e/ j f oodeng
component library due to fact that food components are mandatory
for food process modeling and simulation. In this work, important
food components for modeling milk were built as hypothetical
components for milk process studies.
One important industrial food process is the milk processing
process, which involves the handling of uids, i.e. milk. For exam-
ple, milk (whole or concentrated) is one of the main raw materials
in milk pasteurization processes, milk powder, and cheese plants.
To develop a milk process simulation, based on a commercial pro-
cess simulator, the milk as a collection of hypothetical components
needs to be developed. For that purpose, data on the properties
(mostly physical properties) of milk (Ruiz et al., 2010; Bisig,
2011; Sindhu and Arora, 2011) is used to predict actual milk
behaviour and properties.
In this work VMGSim was selected as the commercial process
simulator for its simulation capabilities, its ability to incorporate
customized calculations using the spread sheet tool, user friendly
interface and its most recent and updated thermodynamics from
TRC/NIST for the prediction of thermodynamic data (i.e. heat
capacity and thermal conductivity) required for the hypothetical
components simulation. It is one of the latest commercial process
simulators mainly conceived for the chemical and petrochemical
industries (Daz et al., 2011; Satyro et al., 2011; Munir et al.,
2012a,b) and developed by Virtual Materials Group Inc. (VMG)
(Virtual Materials Group Inc., 2012). It is extensively used to design
a new process, troubleshoot an existing process unit or optimize
operations in a process (Saber and Shaw, 2008; Jiang et al., 2011;
Satyro et al., 2011; Motahhari et al., 2012). In addition to oil and
gas, and chemical purposes, it has also been used for biofuel pro-
cess applications (Lee et al., 2011).
The aim of this work was to develop hypothetical components
based on the process simulator to simulate actual milk. For this
purpose the data on the properties of milk needed to develop
hypothetical components were obtained from literature (Bylund,
1995; Bon et al., 2010), and the VMGThermo thermodynamic data-
base in the simulator. This would allow the simulation of milk as a
collection of new components (hypothetical components) in the
simulator, needed to simulate milk processing, and to predict milk
process behaviour closely enough to its real-life operation.
This manuscript is organized as follows. After this general intro-
duction, the materials and methods used in this work are explained
and discussed in Section 2. In Section 3 results are discussed. Final-
ly in Section 4 results are summarized, limitations are discussed,
and conclusions are made.
2. Materials and methods
2.1. Raw material (milk): composition and properties
The typical whole (13 wt.% total solids) and concentrated
(50 wt.% total solids) milk compositions considered in this work
are shown in Table 1 (Bylund, 1995; Bon et al., 2010). A material
stream was built in the simulator in order to develop a pseudo
milk mixture having hypothetical components (represented by a
superscripted asterisk,

). Depending on the component informa-
tion contained in the simulator (e.g. palmitic acid, n hexadeca-
noic acid, oleic acid, sodium chloride (NaCl) and potassium
chloride (KCl) are already present in the simulator component li-
brary) component library, each component of the milk composition
(fat, proteins, lactose and minerals) was further classied into sim-
pler components as shown in Table 2 and Fig. 1. The total solid
components include fat, proteins, lactose and minerals.
The assumptions considered in this work and the further classi-
cation of each component of the milk composition (fat, proteins,
lactose and minerals) into simplercomponents are given as
follows:
2.1.1. Assumptions
The following assumptions were adopted after the following
considerations:
(a) Fat: Milk fat is usually considered as a mix of triglyceride
esters, which are composed of various fatty-acids and glyc-
erol. As each glycerol can bind three fatty-acids and they
are not necessarily the same kind, the number of different
glycerides is extremely large. As a result, fats are usually
characterized by fatty-acids. (Bylund, 1995) In this research
we used fatty-acids instead of fatty-esters in simulating milk
due to the following considerations.
The composition of fatty esters is extremely complex,
none of which can be found in the literature or the VMG-
Sim or other commercial process simulator thermody-
namic databases.
Another compound in the VMGSim database named TRI-
GLY(C18)3 (C
54
H
105
O
6
) and belonging to the oil family
was attempted to be used to represent the total amount
of fat. However, the approach resulted in signicant dif-
ferences in density (1223 kg/m
3
vs. 1022 kg/m
3
), heat
capacity (1055 kJ/kmol K vs. 79 kJ/kmol K), thermal con-
ductivity (0.23 W/m K vs. 0.55 W/m K) and viscosity
(0.123 Pa s vs. 0.00203 Pa s) for the nal pseudo milk.
Fatty acids are the closest match to milk fat and using
these fatty acids, the pseudo milk showed a close match
of the main physical properties to actual milk.
The main drawback is that the chemical properties of
esters and acids are very different. However since this
research only focuses on the physical properties, the dif-
ferences in chemical properties can be ignored, which
need to be considered in fouling reactions and cheese
making.
The main component of milk fatty acid is palmitic acid (25
29 wt.%) and oleic acid (3040 wt.%), which can both be found in
the simulator component list. Other minor fatty acids include bu-
tyric acid, myristic acid, stearic acid, and so on. For the simplica-
tion purpose and their similar weight percentage in total fatty acid
contents, the fraction of each of these two fats was set to half
(50 wt.% palmitic acid and 50 wt.% oleic acid) and all other types
of fat components were ignored (Assumption 1).
(b) Protein: Caseins are the proteins commonly found in
mammalian milk, making up to 80% of the total proteins
in milk while whey proteins make up the rest 20%. Due
to the unavailability of public domain literature on the
physical properties of whey proteins, and considering
their relatively minor weight percentage in the total pro-
teins, the protein of milk is simplied to be casein
(Assumption 2). The average molecular weight of casein
was set to 23,000 kg/kg mole, and the density of casein
was set to be 1250 kg/m
3
(Karlsson et al., 2005; Choi
et al., 2011). A hypothetical component was created in
the simulator based on these attributes to simulate
proteins.
Table 1
Quantitative milk composition (% in mass) (Bylund, 1995; Bon et al., 2010).
Product Water Fat Proteins Lactose Minerals Total
solids
Whole milk 87.0 4.0 3.4 4.8 0.8 13.0
Concentrated
milk
50.0 16.0 13.0 18.0 3.0 50.0
88 Y. Zhang et al. / Journal of Food Engineering 121 (2014) 8793
(c) Minerals: The mineral fraction in milk, which is a small
fraction of milk (0.8 wt.%), contains cations (calcium,
magnesium, sodium and potassium) and anions (inor-
ganic phosphate, citrate and chloride) (Bylund, 1995;
Gaucheron, 2005). Due to the unavailability of calcium
salts in the VMGSim thermodynamic database, the min-
erals in raw milk were specied as 50% NaCl and 50% KCl
for simplicitys sake (Assumption 3). They can both be
found in the simulator component list.
(d) Viscosity: Since oleic acid, palmitic acid, NaCl and KCl
can be found in the component list in VMGSim, their
viscosity value cannot be manually changed. Also, the
lactose as a hypothetical compound in VMGSim is con-
sidered to have a high solubility in water, so it should
hardly affect the total liquid viscosity. Due to the insolu-
bility of proteins and the availability of changing their
viscosity value in VMGSim, they were selected as the
main contributor of pseudo milk viscosity (Assump-
tion 4).
(e) Lactose: The component list of the simulator does not
include lactose, thus another hypothetical component
was created with these attributes: molecular weight
(342.3), normal boiling point (668.9 C), density
(1525 kg/m
3
) (Herrington, 1934; Zadow, 1984), and
other properties estimated from these parameters and
the selected thermodynamic package.
2.2. Process simulation: Components and thermodynamic model
selections
Before setting up and solving the simulation cases, the pseudo
milk mixture was simulated using the assumptions, composition
and databases (databases from literature and the simulator library)
explained in Section 2.1.
The simulator library contained information for the following
components required for milk simulation: water, fat (palmitic acid
and oleic acid), and minerals (NaCl and KCl). Components not
available in the library were specied using the Hypothetical
compound manager tool. Proteins (Proteins

) and Lactose
(Lactose

) were specied in this manner. Specication of a hypo-


thetical component (Proteins

and Lactose

in this study) requires


the input of a number of properties such as normal boiling point
(NBP), density, and molecular weight, as well as the critical
properties of the substance.
The accuracyof a process simulationmainly depends onthe ther-
modynamic model used. Hence, selection of the thermodynamic
model is the crucial step in process simulation. The selection of a
property model (thermodynamic model) is based on the following
factors; type of mixture, type of molecules, operating conditions, re-
quiredproperties, level of accuracyrequired, andavailabilityof data.
Owing to the presence of both hydrocarbon-based compounds
(fat, proteins) and polar compounds such as water (87 wt.% in
whole milk and 50 wt.% in concentrated milk) in the process, the
Advanced PengRobinson (APR) equation of state model was ini-
tially selected for use as the property package for the simulation.
Other polar/activity based property packages (e.g. the Wilson
activity package) were also tried but they gave poor results show-
ing large differences in properties estimation, possibly due to the
non-polar nature of some of the compounds.
The APR model was selected because it can handle polar com-
pounds (e.g. water), has the largest applicability range of operating
conditions (T and P), has the largest binary interaction parameter
database, an extended set of interaction parameters for the pro-
cesses have strong temperature dependency, and has volume
translation to get accurate liquid phase density estimation
(Mhlbauer and Raal, 1995; Daz et al., 2011).
Bon et al. (2010) simulated the milk pasteurization process and
selected the IDEAL thermodynamic model in the process simulator
(ProSimPlus) in order to compute uid (i.e. milk) properties.
However, no validation work was reported on. The IDEAL thermo-
dynamic model in ProSimPlus is based on the GammaPhi
approach. APR in VMGSim is preferred because it is a more detailed
description and is based on a better understanding of the complex
dynamics underground the phase behaviour than IDEAL thermody-
namic model. APR is preferred over IDEAL thermodynamic model
because it estimates more accurate interaction coefcients, special
handling of water, and provides accurate volume translation cor-
rection for phase densities (ProSimPlus, 2008; Bon et al., 2010;
Jaworski and Zakrzewska, 2011).
3. Results and discussion
3.1. Simulation results and discussion
After selecting the components and the thermodynamic model,
the raw milk stream was simulated. Input data information (tem-
perature, pressure, ow and composition) of the raw milk stream
was specied to simulate the milk stream. The simulated milk
stream and the calculated physical properties (i.e. density, heat
capacity, thermal conductivity, and viscosity) using equations built
into the simulator database and property package are shown in
Table 3.
The behaviour of the simulated milk (whole and concentrated
milk) stream was compared to the real milk literature data. The
comparison of the main physical properties (i.e. density, heat
capacity, thermal conductivity, and viscosity) between the simu-
lated milk and literature data is shown in Table 3.
It should be noted that since the continuous uid (water) prop-
erties in the simulator are slightly different from actual water
properties, the milk properties were rescaled accordingly to
compare with literature data. Nevertheless, both re-scaled and
un-scaled data are presented in Table 3. All the tests in Table 3
were performed at 1 bar and 25 C.
Table 2
Composition of milk in the simulator (

pseudo milk mixture).


Product Water Fat Proteins

Lactose

Minerals Total solids


Palmitic acid Oleic acid NaCl KCl
Whole milk 87.0 2.0 2.0 3.4 4.8 0.4 0.4 13.0
Concentrated milk 50.0 8.0 8.0 13.0 18.0 1.5 1.5 50.0
Milk
Water
Fat
Minerals
Proteins*
Lactose*
Mixture of Milk
components
Components
already available
in library
Pseudo
components
Palmitic acid
Oleic acid
NaCl
KCl
Fig. 1. Composition of milk in the simulator (

= hypothetical components).
Y. Zhang et al. / Journal of Food Engineering 121 (2014) 8793 89
From Table 3 it can be observed that the raw milk liquid den-
sity, heat capacity and thermal conductivity showed very little dif-
ference between the literature and simulated values. Considering
the small differences in the raw milk density, heat capacity and
thermal conductivity between literature and simulated data, these
properties were not further considered and compared for concen-
trated milk. However, density, heat capacity and thermal conduc-
tivity can vary to some extent with varying total solids in milk.
The liquid milk viscosity showed around a 1719% difference
between the simulation results and literature data as shown in
Table 3. As a result the liquid milk viscosity was further considered
to compare simulation and literature viscosity data for raw and
concentrated milks. The milk viscosity was therefore further
optimized by regressing literature viscosity data using the Model
Regression option in the simulator.
Model regression was used to estimate the parameters required
to calculate the viscosity of the hypothetical components. The sim-
ulators model regression tool uses an optimizer to manipulate the
regressed variables and minimizes the total error of a model based
on the input data.
In this work, model regression was used to modify the viscosity
of the milk mixture stream. The viscosity of the raw milk mixture
stream was modied by manipulating the adjustable viscosity
parameters (A, B, C, and D in Eq. (1)) of the hypothetical compo-
nents. Eq. (1) can be used to regress the viscosity of any hypothet-
ical component and represents the inuence of temperature on the
viscosity of milk (viscositytemperature correlation). This is al-
ready in place in the thermo physical property database. However
it does not consider the effect of total solids on the viscosity of milk
(limitation of Eq. (1)).
lnl A B=T CT DT
2
1
where l is the viscosity (Pa s), T the Temperature (K), A, B, C, and D
is the adjustable viscosity parameters.However the viscosity of milk
is inuenced by the total solid content along with temperature
(Bakshi and Smith, 1984; McCarthy, 2002). Eq. (1) represents only
the inuence of temperature on the viscosity of milk. To re-calibrate
and to consider the combined effect of temperature and total solids
on milk viscosity, a Fernndez-Martn (1972) and Minim et al.
(2002) type viscosity model was employed as shown in Eq. (2). It in-
ter-relates viscosity, temperature and concentration (total solids). It
provides a theoretical viscosity value to compare with the simulator
viscosity results based on Eq. (1).
logg A
0
A
1
t A
2
t
2
B
0
B
1
t B
2
t
2

s
C
0
C
1
t C
2
t
2

s
2
2
where s is the total solids content (% in mass), A
i
, B
i
and C
i
are
dimensionless coefcients calculated by the least squares method,
and are given in (Fernndez-Martn, 1972).
The Fernndez-Martn (1972) and Minim et al. (2002) viscosity
model (Eq. (2)) was used to t the viscosity results calculated in the
simulator using Eq. (1) as explained in Section 3.2.
Total solids content and temperature inuence the viscosity of
milk (Bakshi and Smith, 1984; McCarthy, 2002). It is well known
that of the solids components (fat, proteins, lactose and minerals),
proteins are the main contributor to milk viscosity (Reddy and
Datta, 1994; Bienvenue et al., 2003; Herceg and Lelas, 2005; Karls-
son et al., 2005). Hence, in this work, protein was selected
(assumption 4) as the main contributor (hypothetical compound)
to milk viscosity and was the compound that the viscosity of which
was actually re-calibrated. The re-calibration steps were as
follows:
(a) Obtain whole milk viscosity data from VMGSim, from 1
to 70 C.
(b) Obtain whole milk without protein data from VMGSim,
from 1 to 70 C.
(c) Calculate the weighted natural logarithm difference of
the two sets of data.
(d) Regress the data set obtained from step c with Eq. (1).
(e) Fine-tune the adjustable viscosity parameters A, B, C, and
D in Eq. (1).
3.2. Simulation results validation
Simulation results validation was conducted to determine the
impact of different independent variables on a particular depen-
dent variable and the differences were identied between the liter-
ature and the simulated data under a given set of assumptions.
Considering the small differences in the raw milk density, heat
capacity and thermal conductivity shown between simulation
and literature, these properties were not further compared with
actual milk data. Only the milk viscosity (which showed around
a 1719% difference between simulation and literature) was
further compared with measured milk viscosity data.
In this work, the simulation of the physical properties of
density, heat capacity, and thermal conductivity (q, Cp, and k,
respectively) were validated using the thermo-physical properties
models available in Minim et al. (2002), shown as following
equations.
Cp 3744:48 1:15T 3:93E
3
T
2
; R
2
0:982 3
q 1042:01 0:37T 0:36E
3
T
2
; R
2
0:993 4
k 0:49 2:23E
3
T 1:08E
3
T
2
; R
2
0:991 5
where T is the Temperature (C), and R
2
is the squared residuals.
The simulation of the milk viscosity was validated using viscos-
ity models available in Fernndez-Martn (1972) and Minim et al.
(2002), shown as Eq. (2). Measured milk viscosity data was also
used to see actual differences in milk viscosity data between the
literature, simulated and actual data.
Table 3
Comparison of the physical properties between simulated raw milk and literature data.
Physical properties RML

RMV

Difference (%) WL

WV

RMVR

Rescaled difference (%)


q (kg/m
3
) 1030 1019 1.0 998.2 996.2 1021 1.0
Cp (kJ/kmol K) 3930 3860 2.0 4188 4220 3830 3.0
k (W/m K) 0.53 0.557 5.0 0.58 0.607 0.532 0.4
l (cp) 1.296 1.05 19.0 0.91 0.89 1.074 17.0
where RML

= Mean value of a given property of raw milk from literature.


RMV

= Mean value of a given property of raw milk from the simulator.


WL

= Mean value of a given property of water from literature.


WV

= Mean value of a given property of water from the simulator.


RMVR

= Mean value of a given property of raw milk (rescaled).


Physical properties = density (q), heat capacity (Cp), thermal conductivity (k), and viscosity (l).
90 Y. Zhang et al. / Journal of Food Engineering 121 (2014) 8793
Kessler (2002) also formulated a milk viscosity equation repre-
senting the inuence of the temperature on the viscosity of milk
which was stated to be valid for the temperature range 180 C,
but was not used in this work because it only considers the effect
of temperature on milk viscosity.
Eqs. (3)(5)represent simple linear polynomial relationships
relating the properties of milk (q, Cp, and k) to temperature (Minim
et al., 2002). Minim et al. (2002) addressed the measurement of
these properties (q, Cp, and k), developed empirical correlations
(having R
2
P0.98) for predicting these properties under different
process conditions, and concluded that water content in milk has
a large inuence on the properties while the total solids content
has the least signicant inuence as stated in Minim et al.
(2002), though not quantied.
3.3. Heat capacity, density and thermal conductivity results validation
From Table 3 it can be observed that the heat capacity (calcu-
lated at 1 bar and 25 degrees) showed a small difference (13%) be-
tween the literature and our simulated data. For the temperature
range of 170 C, Eq. (3) from Minim et al. (2002) was employed
for the validation of heat capacity simulation results and to observe
the differences between the literature model and the simulator re-
sults as shown in Fig. 2(a).
Fig. 2(a) shows the plots of the literature and the simulator
derived values for whole milk heat capacity. At lower temperatures
(130 C), the difference between the literature model and the sim-
ulator heat capacity values is larger than at higher temperatures
(3170 C). From Fig. 2(a) it can be observed that the heat capacity
(for temperature range 170 C) showed a small difference
(1.32.6%) between the literature model and the simulator data.
The results observed in Fig. 2(a) are consistent with the heat capac-
ity results in Table 1.
From Table 3 it can be observed that the density and thermal
conductivity (calculated at 1 bar and 25 degrees) showed small
differences (0.71.1%, and 0.35.1% respectively) between the lit-
erature model and the simulator data. For a temperature range of
170 C, Eqs. (4) and (5) available in Minim et al. (2002) were
employed for the validation of the density and thermal conduc-
tivity simulation results, and to observe the difference between
the literature model and simulator data as shown in Figs. 2(b)
and (c).
Figs. 2(b) and (c) show the plots of the literature and the simu-
lator based values for whole milk density and thermal conductiv-
ity, respectively.
From Fig. 2(b) it can be observed that the whole milk density
(for the temperature range of 170 C) showed a small difference
(less than 2%) between literature and simulated data and it almost
remains constant for the temperature range of 170 C. The results
observed in Fig. 2(b) are consistent with the milk density results in
Table 1.
Fig. 2(c) shows the whole milk thermal conductivity (for the
temperature of range of 170 C). A small difference (less than
0.4%) between literature model and simulator data is initially ob-
served and it further decreases with increase in temperature for
the temperature range of 170 C. At 70 C the literature model
and simulator data of thermal conductivity match within 0.1%.
From the results shown in Fig. 2(a)(c) it is clear that the heat
capacity, density and thermal conductivity values, respectively, of
whole milk calculated by the process simulator were in agreement
with the results of linear functions relating these thermo-physical
properties (heat capacity, density and thermal conductivity) to the
temperature, water and total solid contents of milk, developed by
Minim et al. (2002). Hypothetical components created in the sim-
ulator can be used to predict milk heat capacity, density and ther-
mal conductivity within 2.6%, 2%, and0.4%, respectively over the
temperature range of 170 C.
0 10 20 30 40 50 60 70
3600
3800
4000
C
p

(
K
J
/
k
g
m
o
l
e
.
K
)
0 10 20 30 40 50 60 70
800
1000
1200
R
h
o

(
K
g
/
m
3
)
0 10 20 30 40 50 60 70
0.2
0.4
0.6
0.8
Temperature (C)
Temperature (C)
Temperature (C)
k

(
W
/
m
.
K
)
Literature (Minim et al. (2002)) model based thermal conductivity
Simulator thermal conductivity
Literature (Minim et al. (2002)) model based heat capacity
Simulator heat capacity
Literature (Minim et al. (2002)) model based density
Simulator density
(a)
(b)
(c)
Fig. 2. Plots of literature and simulator values for whole milk heat capacity (a), density (b) and thermal conductivity (c).
Y. Zhang et al. / Journal of Food Engineering 121 (2014) 8793 91
3.4. Viscosity results validation
As the simulated liquid milk viscosity showed around a 1719%
difference to literature as shown in Table 3, the simulator viscosity
results were compared with literature viscosity data for the whole
and concentrated milk. Measured milk viscosity data (for whole
and concentrated milk) was also used to observe actual differences
in milk viscosity data between the literature, simulated and mea-
sured milk viscosity.
Fig. 3(a) and (b) represent the whole (13 wt.% total solids) and
concentrated (50 wt.% total solids) milk viscosities (for a tempera-
ture range of 170 C), respectively, showing the validation of sim-
ulated milk (whole and concentrated) viscosity with literature
model data and measured viscosity data. The Fernndez-Martn
(1972) and Minimet al. (2002) viscosity model (Eq. (2)) was the lit-
erature model based viscosity and Souzas (2011) work provided
measured viscosity data. Actual milk viscosity data was only avail-
able for the temperature range 2570 C. Souza (2011) robustly
measured actual viscosity of milk on a pilot scale plant (Lin et al.,
2009) over this temperature range.
Fig. 3(a) shows the plots of the differences between the litera-
ture model, simulator, and measured values of whole milk viscos-
ity. There is almost no difference between the literature model and
the simulator whole milk viscosity for a temperature range of
170 C. The actual measured milk viscosity was closer to the liter-
ature model and simulated viscosity at lower temperatures range.
A small difference (less than 0.02%) between the literature model,
simulated and actual whole milk viscosity data was observed for
the temperature range of 3070 C. A difference of 0.1% between
the literature, simulator and actual whole milk viscosity data was
observed for the temperature range of 2530 C.
Fig. 3(b) shows the plots of the literature model, simulator, and
actual values of concentrated milk viscosity. There is almost no dif-
ference between the literature model and the simulator milk vis-
cosity for a temperature range of 3570 C. The actual milk
viscosity is also closer to the literature model and simulated vis-
cosity for lower temperature ranges. A small difference (less than
0.05%) between the literature model, simulator and actual concen-
trated milk viscosity data was observed for a temperature range of
2070 C. For the temperature range of 020 C the difference
between the literature model, simulated and actual concentrated
milk viscosity data increased signicantly with a decrease in
temperature.
From the results shown in Figs. 3(a) and (b) it is clear that the
viscosity values of whole and concentrated milk, calculated by pro-
cess simulation were in agreement with the results from literature
models and actual milk viscosity data over a temperature range of
170 C for whole milk and 2070 C for concentrated milk. The
results of milk viscosity for whole and concentrated milk were
used because total solids also inuence the viscosity of milk with
temperature. Hence, hypothetical components generated in simu-
lations with viscosity re-calibrated can be used to predict milk
viscosity.
4. Summary and conclusions
In this work, hypothetical components in a process simulator
were developed for simulating pseudo milk. The properties re-
quired to simulate milk in the process simulator were obtained
from the literature and the simulators database. The simulation
results of the physical properties (heat capacity, density, thermal
conductivity and viscosity) were compared with literature models
and actual milk data. After this work, the simulation of milk as a
collection of hypothetical components in a process simulator is
possible.
The simulation results of the physical properties (heat capacity,
density, thermal conductivity and viscosity) were in agreement
with literature models and actual milk data. The hypothetical com-
ponents created in this work can be used to represent the proper-
ties of actual milk for the temperature range of 170 C for whole
milk (and 2070 C for concentrated milk). The liquid density,
heat capacity, thermal conductivity and viscosity showed very
few differences between literature models, simulated and actual
milk data over this temperature range.
In the present approach, fattyacids were used instead of fatty
esters because simulation of esters caused signicant difculties
and complexities. However, since fattyacids helped to match
the physical properties of pseudo milk with actual milk to a fairly
good extent, and milk processing does not involve any signicant
0 10 20 30 40 50 60 70
0
200
400
600
800
Temperature (C)
V
i
s
c
o
s
i
t
y

(
c
P
)
0 10 20 30 40 50 60 70
0
2
4
6
Temperature (C)
V
i
s
c
o
s
i
t
y

(
c
P
)
Literature (Minim et al. (2002)) model based viscosity
Simulated viscosity
Actual viscosity data (Souza, 2011)
Literature (Minim et al. (2002)) model based viscosity
Simulated viscosity
Actual viscosity data (Souza, 2011)
(a)
(b)
Fig. 3. Plots of literature correlations, simulator results and measured data values for whole milk (a) and concentrated milk (b) viscosities.
92 Y. Zhang et al. / Journal of Food Engineering 121 (2014) 8793
chemical reactions of these components for design purposes, we
consider the drawback as minor.
The creation of a hypothetical component database in a process
simulator allowed it to be used as a tool for milk process simula-
tion. Milk process simulation using a process simulator offers the
user the potential to analyse the process and to observe parame-
ters of interest which are difcult to study in practice.
The creation of new databases and the extension of the already
existing components in component libraries to materials like milk
and other liquid foods will expand the application of process
simulators.
References
Abakarov, A., Nuez, M., 2013. Thermal food processing optimization: algorithms
and software. Journal of Food Engineering 115 (4), 428442.
Bakshi, A.S., Smith, D.E., 1984. Effect of fat content and temperature on viscosity in
relation to pumping requirements of uid milk products. Journal of Dairy
Science 67 (6), 11571160.
Bienvenue, A., Jimnez-Flores, R., Singh, H., 2003. Rheological properties of
concentrated skim milk: importance of soluble minerals in the changes in
viscosity during storage. Journal of Dairy Science 86 (12), 38133821.
Bisig, W., 2011. Liquid Milk Products | Liquid Milk Products: Flavored Milks. In: John
(W. F. Editor-in-Chief), Encyclopedia of Dairy Sciences, second ed. Academic
Press, San Diego, pp. 301306.
Bon, J., 2005. Software for food engineering applications. In: Barbosa, G.V. (Ed.),
Food Engineering, Encyclopedia of Life Support Systems. EOLSS publishers/
UNESCO, Paris, pp. 735744.
Bon, J., Clemente, G., Vaquiro, H., Mulet, A., 2010. Simulation and optimization of
milk pasteurization processes using a general process simulator (ProSimPlus).
Computers & Chemical Engineering 34 (3), 414420.
Bylund, G., 1995. Dairy Processing Handbook. Tetra Pak Processing Systems AB,
Lund, Sweden.
Choi, J., Horne, D.S., Lucey, J.A., 2011. Determination of molecular weight of a
puried fraction of colloidal calcium phosphate derived from the casein
micelles of bovine milk. Journal of Dairy Science 94 (7), 32503261.
Daz, O.C., Modaresghazani, J., Satyro, M.A., Yarranton, H.W., 2011. Modeling the
phase behavior of heavy oil and solvent mixtures. Fluid Phase Equilibria 304 (1
2), 7485.
Dnnebier, G., Klatt, K.U., 2000. Modelling and simulation of nonlinear
chromatographic separation processes: a comparison of different modelling
approaches. Chemical Engineering Science 55 (2), 373380.
Fernndez-Martn, F., 1972. Inuence of temperature and composition on some
physical properties of milk and milk concentrates. II. Viscosity. Journal of Dairy
Research 1 (39), 7582.
Garca, M., Gonzalo, A., Snchez, J.L., Arauzo, J., Pea, J.., 2010. Prediction of
normalized biodiesel properties by simulation of multiple feedstock blends.
Bioresource Technology 101 (12), 44314439.
Gaucheron, F., 2005. The minerals of milk. Reproduction Nutrition Development 45
(4), 473483.
Halder, A., Dhall, A., Datta, A.K., Black, D.G., Davidson, P.M., Li, J., Zivanovic, S., 2011.
A user-friendly general-purpose predictive software package for food safety.
Journal of Food Engineering 104 (2), 173185.
Han, I.-S., Chung, C.-B., 2001. Dynamic modeling and simulation of a uidized
catalytic cracking process. Part I: Process modeling. Chemical Engineering
Science 56 (5), 19511971.
Herceg, Z., Lelas, V., 2005. The inuence of temperature and solid matter content on
the viscosity of whey protein concentrates and skim milk powder before and
after tribomechanical treatment. Journal of Food Engineering 66 (4), 433438.
Herrington, B.L., 1934. Some physicochemical properties of lactose: VI. The
solubility of lactose in salt solutions; the isolation of a compound of lactose and
calcium chloride. Journal of Dairy Science 17 (12), 805814.
Jaworski, Z., Zakrzewska, B., 2011. Towards multiscale modelling in product
engineering. Computers & Chemical Engineering 35 (3), 434445.
Jiang, H., Liu, Z., Yan, Q., Su, D., 2011. A simulation study of the mercury distribution
in the low temperature gas separation process. Natural Gas Industry 31 (3), 80
84.
Karlsson, A.O., Ipsen, R., Schrader, K., Ard, Y., 2005. Relationship between physical
properties of casein micelles and rheology of skim milk concentrate. Journal of
Dairy Science 88 (11), 37843797.
Kessler, H.G., 2002. Food and Bio Process Engineering-Dairytechnology. Verlag A.
Kessler, Mnchen, Germany.
Lee, S., Posarac, D., Ellis, N., 2011. Process simulation and economic analysis of
biodiesel production processes using fresh and waste vegetable oil and
supercritical methanol. Chemical Engineering Research and Design 89 (12),
26262642.
Lin, T.-I., G. De Souza and B. Young., 2009. Towards a Viscosity and Density
Correlation for Dairy Fluids - A Soft Sensor Approach. In: C. A. O. d. N. Rita Maria
de Brito Alves and B. Evaristo Chalbaud. (Eds.), Computer Aided Chemical
Engineering. Elsevier, 27, 13711376.
Luyben, W.L., 2002. Plantwide Dynamic Simulators in Chemical Processing and
Control. Marcel Dekker, Inc, New York.
McCarthy, O.J., 2002. Milk | Physical and Physico-Chemical Properties of Milk. In:
John (W. F. Editor-in-Chief), Encyclopedia of Dairy Sciences, second ed.
Academic Press, San Diego, pp. 467477.
Minim, L.A., Coimbra, J.S.R., Minim, V.P.R., Telis-Romero, J., 2002. Inuence of
temperature and water and fat contents on the thermophysical properties of
milk. Journal of Chemical & Engineering Data 47 (6), 14881491.
Motahhari, H., Satyro, M.A., Yarranton, H.W., 2012. Viscosity prediction for natural
gas processing applications. Fluid Phase Equilibria 322323, 5665.
Mhlbauer, A.L., Raal, J.D., 1995. Computation and thermodynamic interpretation of
high-pressure vapourliquid equilibriuma review. The Chemical Engineering
Journal and the Biochemical Engineering Journal 60 (13), 129.
Munir, M.T., Yu, W., Young, B.R., 2012a. Recycle effect on the relative exergy array.
Chemical Engineering Research and Design 90 (1), 110118.
Munir, M.T., Yu, W., Young, B.R., 2012b. A software algorithm/package for control
loop conguration and eco-efciency. ISA Transactions 51 (6), 827833.
Peters, L., Hussain, A., Follmann, M., Melin, T., Hgg, M.B., 2011. CO2 removal from
natural gas by employing amine absorption and membrane technologya
technical and economical analysis. Chemical Engineering Journal 172 (23),
952960.
ProSimPlus, 2008. Prosim Plus: Steady-state Simulation and Optimization of
Processes. <www.prosim.net>.
Reddy, C.S., Datta, A.K., 1994. Thermophysical properties of concentrated
reconstituted milk during processing. Journal of Food Engineering 21 (1), 3140.
Rodrigues, A.E., Minceva, M., 2005. Modelling and simulation in chemical
engineering: tools for process innovation. Computers & Chemical Engineering
29 (6), 11671183.
Ruiz, G.J., Kim, S.B., Moon, J., Zhang, L., Linninger, A.A., 2010. Design and
optimization of energy efcient complex separation networks. Computers &
Chemical Engineering 34 (9), 15561563.
Saber, N., Shaw, J.M., 2008. Rapid and robust phase behaviour stability analysis
using global optimization. Fluid Phase Equilibria 264 (12), 137146.
Satyro, M.A., Schoeggl, F., Yarranton, H.W., 2011. Temperature change from
isenthalpic expansion of aqueous triethylene glycol mixtures for natural gas
dehydration. Fluid Phase Equilibria 305 (1), 6267.
Sindhu, J.S., Arora, S., 2011. Milk | Buffalo Milk. In: John (W. F. Editor-in-Chief),
Encyclopedia of Dairy Sciences, second ed. Academic Press, San Diego, pp. 503
511.
Souza, G.D., 2011. Milk Dryer Viscosity Measurement and Advanced Control. Ph.D.
The University of Auckland, New Zealand.
Theien, M., Hai, R., Marquardt, W., 2011. A framework for work process modeling
in the chemical industries. Computers & Chemical Engineering 35 (4), 679691.
Trystram, G., 2012. Modelling of food and food processes. Journal of Food
Engineering 110 (2), 269277.
Virtual Materials Group Inc., 2012. http://www.virtualmaterials.com/vmgsim.
Wang, Z., Hirai, S., 2011. Modeling and estimation of rheological properties of food
products for manufacturing simulations. Journal of Food Engineering 102 (2),
136144.
West, A.H., Posarac, D., Ellis, N., 2008. Assessment of four biodiesel production
processes using HYSYS. Plant Bioresource Technology 99 (14), 65876601.
Zadow, J.G., 1984. Lactose: properties and uses. Journal of Dairy Science 67 (11),
26542679.
Zhao, M., Li, Y., Sun, S., 2011. Analysis and optimization of two-column cryogenic
process for argon recovery from hydrogen-depleted ammonia purge gas.
Chemical Engineering Research and Design 89 (7), 863878.
Y. Zhang et al. / Journal of Food Engineering 121 (2014) 8793 93

Вам также может понравиться