You are on page 1of 12

International Journal of Refrigeration 98 (2019) 468–479

Contents lists available at ScienceDirect

International Journal of Refrigeration


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

Development of a food temperature prediction model for real time


food quality assessment
Hajin Song a, Jiyoung Kim a,b, Byeong-Sam Kim b, Junemo Koo a,∗
a
Department of Mechanical Engineering, Kyung Hee University, Yongin, Gyeonggi-do 17104, South Korea
b
Research Group of Smart Food Distribution System, Korea Food Research Institute, Wanju-gun, Jeollabuk-do 55365, South Korea

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

Article history: Multilayer food temperature models considering the spatial temperature distribution in the food were
Received 30 May 2018 developed to predict temporal temperature changes experienced by food during distribution considering
Revised 22 November 2018
the surrounding temperature, initial food temperature, and the predetermined surface heat transfer coef-
Accepted 27 November 2018
ficient. Parameters used in the models were determined using the results of unsteady three-dimensional
Available online 30 November 2018
computational fluid dynamics simulations for the conjugated heat transfer problem between the food and
Keywords: surroundings. Models were validated by comparing model predictions with the food temperature experi-
Food temperature model mentally measured under fluctuating temperature conditions with a mean absolute deviation of less than
Freshness monitoring model 1 °C.
Real-time food quality monitoring system The developed food temperature model was integrated with a freshness monitoring model to com-
Heat transfer coefficient plete the real-time food quality monitoring system. Food quality prediction using the integrated food
Thermal network
quality monitoring system was validated for eggs and milk. Quality predictions using the food tempera-
ture model showed better accuracy than when using a single point measurement of food temperature.
© 2018 Elsevier Ltd and IIR. All rights reserved.

Développement d’un modèle prédictif de la température des aliments pour


l’évaluation en temps réel de la qualité des aliments

Mots-clés: Modèle de température des aliments; Modèle de surveillance de la fraîcheur; Système de surveillance de la qualité des aliments en temps réel; Coefficient de
transfert de chaleur; Réseau thermique

1. Introduction improving food safety and reducing FW. FQI in a system is mod-
eled as a variable that is dependent on the time history of en-
In the US, approximately 30–40% of the total amount of food vironmental factors such as temperature, relative humidity, and
fails to be consumed and is discarded every year (Buzby et al., airflow velocity. Among these factors, temperature is the most sig-
2014). Food waste (FW) occurs at both the retail and consumer nificant. Therefore, FQI is usually represented as a function of the
levels for reasons including erroneous transportation and storage temperature history only, which highlights the importance of ac-
management, improper packaging, and confusion caused by a lack curate monitoring of food temperature (Carullo et al., 2009; Je-
of clear understanding of shelf-life. FW results in economic loss dermann et al., 2009; Shih and Wang, 2016). Raab et al. (2008),
and can also have an adverse impact on the environment. Bruckner et al. (2013) and Lytou et al. (2016) developed prediction
Development of real-time food quality monitoring systems to models of aerobic microbial and Pseudomonas sp. in poultry and
keep track of the food quality index (FQI) history during distri- pork under both constant and fluctuating temperature conditions.
bution is a topic of active research with the ultimate goals of Raab et al. (2008) collected the food temperature history from the
fixed temperature sensors in the meat of large packaging unit. In

contrast, Bruckner et al. (2013) and Lytou et al. (2016) used the
Corresponding author.
chamber temperature history as the food temperature history con-
E-mail addresses: shj7034@khu.ac.kr (H. Song), jykim@kfri.re.kr (J. Kim),
bskim@kfri.re.kr (B.-S. Kim), jmkoo@khu.ac.kr (J. Koo). sidering small sample size. Gwanpua et al. (2015) presented FRIS-

https://doi.org/10.1016/j.ijrefrig.2018.11.032
0140-7007/© 2018 Elsevier Ltd and IIR. All rights reserved.
H. Song, J. Kim and B.-S. Kim et al. / International Journal of Refrigeration 98 (2019) 468–479 469

Raab et al. (2008) stressed the importance of a food temper-


Nomenclature ature monitoring technique as well as the construction of a food
quality model and infrastructure for food quality monitoring at
A Heat transfer area [m2 ] each stage of distribution. They concluded that further research is
Af Accuracy factor necessary to develop a food temperature monitoring technique and
Bf Bias factor infrastructure to complete the monitoring system despite develop-
Bi Biot number ing food quality models for pork and chicken.
Ci Heat capacitance [J °C−1 ] The rate of food temperature change can be calculated using
Cp Specific heat [J kg−1 K−1 ] the rate of convective and radiative heat transfer to the food from
FMM Food freshness monitoring model the surroundings, which are the two major heat transfer modes for
Fo Fourier number processed and packaged foods in distribution. Heat transfer coeffi-
FQI Food quality index cients are used to calculate the heat transfer rate and vary with
FQMS Real-time food quality monitoring system factors such as the temperature difference between the food and
FTM Food temperature model surroundings in the case of natural convection, fluid velocity, and
h Heat transfer coefficient [W m−2 °C−1 ] surface roughness. Denys et al. (20 03, 20 04) estimated the convec-
hconv Convective heat transfer coefficient [W m−2 °C−1 ] tive heat transfer coefficient during low-temperature pasteuriza-
h̄ Effective average heat transfer coefficient [W m−2 tion of eggs by matching the evolution of temperature distributions
°C−1 ] measured experimentally inside the eggs with that obtained from
HU Haugh unit computational fluid dynamics (CFD) simulation analysis assuming
k Thermal conductivity [W m−1 K−1 ] a heat transfer coefficient at the surface. Kondjoyan (2006) pub-
L Characteristic length of the food [m] lished a review article that detailed surface heat and mass trans-
MAPE Mean absolute percentage error fer coefficients during air chilling and the storage of food prod-
N Microbial concentration in the milk [CFU mL−1 ] ucts. They used the Nusselt number relationship with the Reynolds
Nu Nusselt number number and turbulent intensity to determine the heat transfer co-
Pr Prandtl number efficient for a single food product with a given food shape under
Q˙ conv Convective heat transfer rate [W] cross- and parallel-flow conditions. They concluded that the heat
Re Reynolds number transfer coefficient should be measured directly under real distri-
Ri Effective thermal resistance [°C W−1 ] bution conditions because it varies depending on the shape and
T Mean temperature of the food [°C] size of the food as well as flow conditions.
Ti Temperature of each layer in food [°C] Van der Sman (2003) developed models to predict the temper-
Tm Mean temperature of the food [°C] ature in high-moisture foods of regular shapes such as cylinders
Ts Food surface temperature [°C] and shell-cores involving convective and evaporative heat transfer
T∞ Surrounding temperature [°C] at the surface. They performed finite volume method-based com-
V Volume [m−3 ] putational simulations to obtain temperature distribution histories.
They observed that the mean temperature location was about
Greek letter
two thirds of the radius from the center for symmetric cases and
α Thermal diffusivity [m2 s−1 ]
developed simple differential equations to predict the evolution of
ε Emissivity
the mean and surface temperatures of foods over time using the
μ Dynamic viscosity [kg m−3 ]
predetermined thermal resistance. However, their models cannot
ρ Density [kg m−3 ]
be applied to foods with complex shapes or a non-symmetric
σ Stefan-Boltzmann constant
temperature distribution. Margeirsson et al. (2012) performed
three-dimensional (3D) heat transfer simulations of cod fillets
packaged in expanded polystyrene (EPS) boxes to investigate the
BEE tool which simulated the food quality change in cold-chain us- spatial distribution of the food temperature inside the boxes
ing the food freshness model by considering the heat transfer in under laminar natural convection conditions. Similar research was
showcase and refrigerators. It evaluated the energy consumption reported by Marai et al. (2012) for small fruits like blueberries,
as well as global warming impact of the cold chain systems based where they compared the simulated temperature distribution to
on the user inputs of cold-chain condition parameters such as per- highlight the usefulness of CFD simulations to analyze the tem-
formance index, type and configuration of refrigeration systems. perature distribution in foods. Gospavic et al. (2012) derived an
Currently, temperatures inside the delivery vehicle cargo, the analytical solution for the temperature distribution in a cuboidal
package box, warehouse, and showcases are measured and as- package of haddock fillets by separating variables for the transient
sumed to be the food temperature. However, when the outside 3D heat conduction equation under laminar natural convection
temperature changes very rapidly or the food heat capacity is large, conditions. They compared the solution with measured data and
there is a measurable discrepancy between the temperature of the CFD simulation results to validate their model. However, it is not
food and the surroundings such that accurate quality monitoring feasible to carry out CFD simulations for individual food objects in
is hindered. To address these issues, it is necessary to provide a distribution at each monitoring time point to obtain information
bridging technology capable of predicting the food temperature about the temperature distribution (Marai et al., 2012; Margeirsson
from the monitored surrounding temperature. et al., 2012). In addition, the analytical solution (Gospavić et al.,
Hoang et al. (2012) analyzed the current cold chain distribu- 2012) is limited to foods of a cuboidal shape under laminar nat-
tion process using existing information from databases and mod- ural convection conditions. It is necessary to develop a simple but
eled the evolution of food temperature and microbial growth in general model to estimate temporal changes in food quality con-
food using combined deterministic and stochastic approaches. They sidering the effects of the spatial temperature distribution in foods.
classified possible distribution environment scenarios at each stage In this study, we developed a food temperature model (FTM)
using information from existing databases, and estimated average to predict food temperature change during distribution consider-
food temperature during distribution stochastically to evaluate mi- ing the surrounding temperature, initial food temperature, and the
crobial growth in the food. predetermined surface heat transfer coefficient. The surface heat
470 H. Song, J. Kim and B.-S. Kim et al. / International Journal of Refrigeration 98 (2019) 468–479

shift duration, and average length of the long and short axes of
the egg, respectively, indicating that the temperature shift affected
the entire inside of the egg. Temperature was constantly monitored
by electronic, programmable, miniature data loggers (HL-1D/TL-
1D, Rotronic Measurement Solutions, Taiwan) in the storage cham-
bers. Data loggers were placed inside the storage chambers and
were attached to the egg surface and egg core to monitor stor-
age air, egg surface, and egg core temperatures, respectively. To
measure the egg core temperature, the tips of temperature sen-
sor were inserted into the eggs, and the gap between the temper-
ature sensor and the egg shell was filled to alleviate the impact
of the treatment. The tips of the temperature sensors were fixed
on egg shells using film tape to measure the surface temperature.
Temperature data were recorded every 10 min throughout storage.
Haugh unit was determined from egg weight and the albumen
height of a broken egg spread on a horizontal plate using Eq. (1)
(Haugh, 1937):
 
HU = 100log H + 7.51 − 1.7W0.37 (1)
Fig. 1. Schematic process diagram of the real-time food quality monitoring system.
where H is albumen height (mm) and W is the weight of the egg
when tested (g). Refer to Yimenu et al. (2017) for details of mate-
transfer coefficient was measured for each stage of circulation un- rial preparation of the eggs.
der real distribution conditions. Details of integration of this FTM Commercially packaged pasteurized milk (63 °C for 30 min) in
into a real-time food quality monitoring system (FQMS) with a pre- 200 mL and 930 mL vessels was obtained from a commercial dairy
developed food freshness monitoring model (FMM) are presented. on the day of production. Milk samples were collected three times
from April 2011 to June 2012. Wired temperature loggers (Model
2. Materials and methods TR-5i; T&D Corporation, Japan) were used to monitor the temper-
ature at the centroid of each milk package. RFID temperature log-
2.1. Real-time food quality monitoring system gers were attached to the milk package surface, and the surround-
ing temperature was monitored by temperature sensors installed
The process schematics of the real-time food quality monitoring in the refrigerator. Total aerobic counts of 10 mL samples were an-
system are shown in Fig. 1. The changes in the surrounding tem- alyzed every 2 h. Refer to Kim et al. (2016) for details of the ma-
perature are monitored by RFID temperature sensors attached to terial preparation of the milk, including details about the choice
each package unit such as a box of multiple egg cartons or a crate to use FQI for microbial counts. Samples were exposed to fluc-
containing multiples of milk packages to consider the spatial tem- tuating temperatures by incubation in a refrigerator with a pro-
perature distribution. The monitored temperature is transferred to grammable temperature history. The corresponding Fourier num-
a server computer, which calculates the food temperature and FQI bers were calculated to be 0.78 and 0.53 for 200 mL milk under
histories using FTM and FMM. Food status information is delivered type A condition and 930 mL milk respectively, where it ranged be-
to the customer on demand. tween 0.35 and 4.15 for 200 mL milk under type B condition. This
In this study, FTMs for eggs and low temperature pasteur- implies that the experiments encompassed cases of both partial
ized milk were developed and their validity and applicability were and full penetration of heat into the milk. The impact of tempera-
tested by constructing a FQMS using FMMs for eggs developed by ture shift frequency on the accuracy of FTM is discussed in a later
Yimenu et al. (2017) and for milk by Kim et al. (2016). Eggs in car- section.
tons as well as milk in bottles of different sizes were selected to Under real distribution conditions, food temperature is usually
investigate the impact of thermal capacity on the need for FTM. controlled at a constant temperature for a long period of time,
Experiments and CFD simulations were performed to determine and the impact of temporal temperature shifts decreases with the
model coefficients, including the heat transfer coefficient and ther- proper distribution process. Severe temperature abuse cases were
mal resistances. The heat transfer coefficient was estimated experi- considered to verify the performance of FTM and FQMS.
mentally using the temperature histories of food and surroundings,
while internal thermal resistances were derived from CFD analysis. 2.3. Determination of heat transfer coefficients for foods
The spatial distributions of food temperature and heat transfer co- in distribution
efficients were also analyzed using the CFD results.
The rate of convective heat transfer is proportional to the tem-
2.2. Experimental set-up perature difference. The governing law for convection is Newton’s
law of cooling, as shown in Eq. (2):
Freshly-laid, special-class, unfertilized egg samples were ob-
Q˙ conv = hconv A(Ts − T∞ ) (2)
tained directly from Ireefarm Egg Company, Seoul and storage ex-
periments were conducted at the Korea Food Research Institute in where Q˙ conv , A, hconv , Ts , and T∞ represent the rate of heat transfer
January 2015. Samples were stored in their commercial packaging due to convection, heat transfer area, convective heat transfer co-
(cartons, each containing 10 eggs) under a fluctuating tempera- efficient, food surface temperature, and surrounding temperature,
ture of 10–30 °C alternating at 24 hours. Surrounding temperatures respectively. Convective heat transfer coefficient is defined as the
were measured at multiple points by temperature sensor tips fixed rate of heat transfer per unit area for a unit temperature differ-
at the surface of the storage chamber near foods, and it was found ence. It is a complex function of fluid properties, fluid flow ve-
that the deviations were negligible between them. The Fourier locity, and the food surface. It is hard to define explicitly and is
number (Fo = α tL−2 ) for eggs was estimated to be 4.26 where usually experimentally determined, because it is difficult to repre-
α , t, and L were the thermal diffusivity of the egg, temperature sent all the complex flow and heat transfer conditions experienced
H. Song, J. Kim and B.-S. Kim et al. / International Journal of Refrigeration 98 (2019) 468–479 471

under real distribution conditions, even with computational fluid on variations according to stage of distribution. The issue of varia-
dynamics simulations (Kondjoyan, 2006). The heat transfer coef- tion according to relative position was addressed by estimating the
ficient is strongly dependent on fluid flow velocity and increases mean temperature difference between the surroundings and foods
with velocity as shown in Eqs. (3)–(5) (Cengel, 2014): at different relative positions. Furthermore, the impact of spa-
tial heat transfer coefficient variation on freshness estimation was
h·L
= Nu = C · Rem · P r n (3) analyzed by considering the standard deviations observed in the
k
experimental measurements. The mean effective heat transfer co-
ρ ·V ·L efficients should be evaluated using Eq. (9) for different stages of
Re = (4) distribution. They should be estimated using the food and sur-
μ
rounding temperature history ranges where the temperature dif-
μ · cp ference between the foods and the surroundings are significant
Pr = (5)
k and the direction of heat transfer is consistent to avoid the ef-
fect of noise signals considering the fact that the convective heat
Here h, L, k, Nu, Re, Pr, ρ , V, μ, and Cp are the heat transfer coeffi-
transfer coefficient is independent of the temperature difference in
cient, characteristic length of the food, thermal conductivity of the
the case of the forced convection mode. Table 1 shows the ma-
surrounding air, Nusselt number, Reynolds number, Prandtl num-
terial properties of foods considered in this study. The mean and
ber, air density, air velocity, air dynamic viscosity, and air specific
standard deviation values of heat transfer coefficients were 7.73 ±
heat, respectively. C, m, and n are model constants that are deter-
0.44 Wm−2 ◦ C−1 for eggs, 24.19 ± 4.88 Wm−2 ◦ C−1 for 200 mL milk
mined based on the flow regime and food shape.
under type A temperature conditions, 13.66 ± 1.64 Wm−2 ◦ C−1 for
Generally, radiative heat transfer is active in addition to convec-
200 mL milk under type B temperature conditions, and 19.48 ±
tive heat transfer for solid surfaces of different temperatures sur-
5.92 Wm−2 ◦ C−1 for 930 mL milk. Milk tended to have higher heat
rounded by fluids. Eqs. (6) and (7) can be used to calculate the
transfer coefficients for the heating phase than the cooling phase,
total heat transfer coefficient considering the additive effects of
which can be attributed to the effects of a film of condensed water
convective and radiative heat transfer, and the total heat transfer
on bottles or asymmetric fluid flow around bottles. However, the
rate for the case of the same surrounding temperatures for both
difference was statistically negligible when considering the exper-
heat transfer modes, respectively.
  imental confidence intervals. In this study, the heat transfer coef-
h = hconv + εσ (Ts + T∞ ) Ts2 + T∞
2
(6) ficients for heating (22.39 ± 3.14 Wm−2 ◦ C−1 for 930 mL milk) and
cooling (16.57 ± 6.74 Wm−2 ◦ C−1 for 930 mL milk) phases were as-
sumed to be equal. The impact of natural convection outside the
 
Q˙ = Q˙ conv + Q˙ rad = hconv A(Ts − T∞ ) + εσ Ts4 − T∞
4
= hA(Ts − T∞ ) product was analyzed to be negligible statistically (R2 = 0.07) from
the linear regression result of the heat transfer coefficients as a
(7) function of temperature difference between food surface and sur-
Here ɛ and σ are the emissivity and Stefan–Boltzmann con- roundings. The data for the analysis were selected from the mea-
stant, respectively. In this study, the two modes were not distin- sured data for 200 ml milk under type B condition where the tem-
guished, so the total heat transfer coefficient was determined and perature difference between the surface and surroundings showed
FTM was derived using the total heat transfer coefficient. The to- maximum peaks. There should be a meaningful functional rela-
tal heat transfer coefficient was estimated by applying the conser- tion between them in case of the significant impact of natural
vation of energy law to the experimentally measured food inter- convection.
nal, surface, and surrounding temperature histories; the amount of
heat transfer to a food was considered equal to the change in in- 2.4. Computational fluid dynamics set-up
ternal energy of the food, as shown in Eq. (8):
 tn Computational fluid dynamics simulations were performed to
ρV Cp (Tm,tn − Tm,t0 ) = hA(T∞ − Ts )dt (8) determine the model constants in FTM considering an egg and a
t0 single milk bottle. CFD simulations considering 3 × 3 × 3 stacked
Here ρ , V, Cp , and Tm are the density, volume, specific heat, and crates of milk consisting of 25(5 × 5) unit bottles were carried
mean temperature of the food, respectively. Tm was estimated as out to determine the spatial temperature and heat transfer co-
the arithmetic mean of the measured surface and internal tem- efficient spreads experienced during distribution. Continuity, mo-
peratures. When Tm was assumed to be the surface temperature mentum, and energy equations were solved simultaneously, and
or the core temperature, the differences in estimations of the the γ − Reθ transition model was used together with the k − ω
heat transfer coefficient were 0.4 Wm−2 ◦ C−1 for eggs and 1–4 turbulence model (Anonymous, 2017a) to consider laminar, tran-
Wm−2 ◦ C−1 for milk. These differences are statistically negligible sition, and turbulent flow regimes simultaneously for unit foods.
because they fall within the standard deviations of the estimations. The k− ∈ turbulence model (Anonymous, 2011) was used for sim-
The effective average heat transfer coefficient, h̄, was calculated ulations of stacked crates of milk to alleviate the need for a large
using Eq. (9): number of grid points. Air flow was assumed to be from the left
 tn
side with the specified inlet velocity and temperature conditions.
h̄ = ρV C p (Tm,tn − Tm,t0 )/ A(T∞ − Ts )dt (9) For example, air inlet velocity was set to 1.7 ms−1 , as at this veloc-
t0 ity, the same heat transfer coefficient value measured experimen-
The heat transfer coefficient can vary depending on the stage tally was observed in the simulation. The turbulent intensity and
of distribution as well as the position of foods, which could af- turbulent viscosity ratio were set to be 10% and 10, respectively.
fect the fluid flow and radiative heat transfer. Considering the dif- The measured temperature histories were applied as inlet bound-
ficulty in estimating heat transfer coefficients that vary with the ary conditions according to the test cases. As air flowed through
relative positions of foods, the primary objective of this study fo- the computational domain, it interacted with the milk to exchange
cuses on cases of neglecting the variations with the relative posi- heat and exited on the right side. The symmetry condition was im-
tions. Rather, we assumed that the heat transfer coefficient had a posed on the other four sides setting no normal gradient of ve-
constant value, regardless of the position of the food, and focused locity and temperature. CFD simulation was performed using the
commercial packages STAR-CCM + (Anonymous, 2017a) and Fluent
472 H. Song, J. Kim and B.-S. Kim et al. / International Journal of Refrigeration 98 (2019) 468–479

Table 1
The material properties of foods and air considered in this study.

Material Thermal conductivity, k [W m−1 K−1 ] Specific heat, Cp [J kg−1 K−1 ] Density, ρ [kg m−3 ] Dynamic viscosity, μ [kg m−3 ]

Eggs (white) (Anonymous, 1998) 0.56 3880 980 –


Eggs (yolk) (Anonymous, 1998) 0.42 2810 1020 –
Milk (Cengel, 2014; Rao et al., 2014) 0.58 3890 1030 8.89 × 10−4
Air (Cengel, 2014) 2.56 × 10−2 1007 1.18 1.85 × 10−5

(Anonymous, 2011). The impact of natural convection in milk was


evaluated using the Boussinesq model.
The computational domain was discretized with either polyhe-
dral meshes using prism layers on food surfaces to improve the
resolution of boundary layer effects or tetrahedral meshes, and the
number of elements used in the simulations was 330,0 0 0 polyhe-
dral cells for the unit milk case and 20,0 0 0,0 0 0 tetrahedral cells for
the stacked milk crates case. The mesh independence of solutions
was verified with meshes of different densities and the simulation
results showed maximum deviations in heat transfer rate less than
0.8% for milk.

3. Model development

3.1. Derivation of food temperature models for eggs and milk

The thermal resistance circuits considered for eggs and milk,


the milk circuit is shown in Fig. 2 for example. The egg circuit
was not presented for the brevity. Model equations were derived
for the thermal resistance circuit for the heat transfer phe-
nomenon between the surroundings and the food. Heat transfer
was assumed to be one-dimensional, whereas real heat transfer
phenomena are complicated and three-dimensional. Effective
values of the model constants were used to compensate for this
simplification. To use a lumped model neglecting temperature
distribution in the food, the Biot number (= hLk−1 ) should be
small, e.g. Bi < 0.1. The heat transfer coefficients required to meet
this condition were estimated to be < 1.16 Wm−2 ◦ C−1 for 200 mL
milk and < 0.69 Wm−2 ◦ C−1 for 930 mL milk, whereas the actual
values were larger than these as shown in Section 2.3. Multiple
layers with different thermal resistance and capacitance values
were added to consider possible temperature variations in foods,
which may improve the accuracy of the prediction compared to
the case of using a lumped system model or single layer model.
Eqs. (10)–(15) constitute the FTM for milk. Equations for eggs are
omitted for brevity of presentation.
Fig. 2. Diagrams illustrating (a) the configurations of three layers of milk in 200 mL
Q˙ = hA(T∞ − Ts ) (10) and 930 mL containers with the nodal temperature, (b) one-dimensional represen-
tation of the layers, and (c) the corresponding thermal resistance-capacitance net-
work.
Q˙ + (T1 − Ts )/Rs = 0 (11)

C1 T˙1 = (Ts − T1 )/Rs + (T2 − T1 )/R1 (12)


930 mL of milk were considered to take into account spatial tem-
perature distributions. Three layers were selected because this is
C2 T˙2 = (T1 − T2 )/R1 + (T3 − T2 )/R2 (13) the smallest number of layers for which a non-linear distribution
of temperature in food can be considered while minimizing com-
putational cost. The models have been developed to be used in the
C3 T˙3 = (T2 − T3 )/R2 (14) food freshness monitoring system which runs on a web server to
handle the requests of customers and distributors for each moni-
toring unit of food in distribution nation-wide. Considering the fre-
C = mC p (15)
quency of the temperature logging per a unit of food to monitor to
The series of temperatures, Ti s, in Fig. 2 represent the temper- be once in 5 minutes as well as the number of food packages to
ature of each layer in the milk to simulate the one-dimensional track simultaneously to be more than thousands and even millions,
transport of heat. T∞ is the surrounding temperature. Ri and Ci re- the computing load increases significantly with the increase of the
fer to the effective thermal resistance and capacitance between the layer numbers in FTM. It was found from this study that the accu-
nodal points, respectively, and the thermal circuits are shown in racy of the food freshness estimation using the three-layer model,
Fig. 2(c). Three layers with thermal resistance circuits for 200 and which is the simplest one allowing nonlinear temperature distribu-
H. Song, J. Kim and B.-S. Kim et al. / International Journal of Refrigeration 98 (2019) 468–479 473

quality index values to estimate egg freshness under fluctuating


temperature conditions. In this study, FMM using the HU was inte-
grated with the developed FTM to develop a real-time Food Qual-
ity Monitoring System(FQMS) for eggs. The model is shown in
Eqs. (16) and (17):
dHU
= CHU (T ) · HU 3 (16)
dt

CHU (T ) = −2.720 × 10−8 T 2 + 3.124 × 10−7 T − 1.370 × 10−6 (17)


where t, T, and CHU represent time, food temperature in Celsius,
and a coefficient reflecting the variation in HU change rate with
food temperature, respectively. To solve the differential equation in
Eq. (16), an initial value for FQI was provided as HU0 ; we used 83.4
in this study.
Fig. 3. Comparison of the temperature field in a milk bottle neglecting (left) and The model reported by Kim et al. (2016) that describes the rate
considering (right) the impact of internal natural convection.
of microbial growth in milk using the Barany and Roberts model
was adopted as the FMM for milk in the current study. The model
is shown in Eqs. (18)–(20):
tion in food, was comparable or better than the that of using the
dN Q (t )

N (t )

history of single-point food core temperature measurement. Accu-
= · μmax · 1 − · N (t ) (18)
racy improvement by putting more layers in the model was con- dt 1 + Q (t ) Nmax
sidered to be subtle due to the diffusive nature of heat conduction
in food. The three regions were assigned based on the thermal dis- dQ
= μmax · Q (t ) (19)
tribution information from CFD simulations, as shown in Fig. 3(a). dt
Due to the assumption of one-dimensional heat conduction, Ri in   
this case was not just a simple combination of food properties. In exp A1 − T B(K1 ) f or T ≤ Tc
contrast, the Ci values in the FTMs were determined explicitly us- μmax (T ) =  B2
 (20)
exp A2 − T (K )
f or T > Tc
ing Eq. (15) by considering the food properties. The number of un-
knowns for the FTM for milk was three (Rs , R1 and R2 ) whereas the Here N, N0 , and Nmax are the microbial concentration in the
other unknown, h, was determined experimentally as discussed in milk, the initial concentration, and the maximum microbial con-
Section 2.3. centration in the stationary phase, respectively. μmax represents
Fig. 4 shows the process used to derive FTMs for eggs and the maximum microbial growth rate under the given temperature
milk in 200 and 930 mL packages. The effective resistance value, conditions, represented in the form of two-zone Arrhenius equa-
Ri , was determined by adjusting this value to minimize deviations tions, with the model constants of A1 = 62.8, B1 = 18, 450, B2 =
between the right-hand side (RHS) and left-hand side (LHS) of Eqs. 11, 310, A2 = 41.4, Q0 = 1.10 × 10−3 , and Nmax = 1.31 × 107 . Here,
(10)–(14) using the temperature histories obtained from the CFD Tc = 11.8 ◦ C is the critical temperature that divides the two zones,
simulation results. For example, the LHS of Eq. (12) was calculated and needs to be converted to the Kelvin scale for use in Eq. (20).
using the mass weighted temperature of each layer, i.e., T1 and T2 , The initial microbial count in milk, N0 , was set to 38.0 CFU mL−1
and by the area weighted temperature of the bottle surface, Ts , and 36.3 CFU mL−1 for milk in 200 and 930 mL bottles, respec-
based on the CFD simulation results at each time step. By assuming tively.
the values of Rs and R1 , the RHS of Eq. (12) could be estimated at
each time step. Correct guesses of the values of the two unknowns 4. Results and discussion
Rs and R1 would lead to smaller deviation between the LHS and
RHS of the equation. The values of the unknowns were chosen to 4.1. Validation of the food temperature models for milk
minimize the sum of absolute deviations for Eqs. (10)–(14) over the
simulated period. The minimization process was performed using The determined effective values of the model constants and the
the differential evolution algorithm in the DEoptim package (Price measured heat transfer coefficients completed the FTM. The deter-
et al., 2005) as implemented in R (Anonymous, 2017b). mined model constants are shown in Table 2. When comparing the
Because FTMs are in the form of a system of ordinary differen- heat transfer resistances shown in Fig. 2(c), we noted that the re-
tial equations, food histories were obtained as solutions of the or- sistance to surface heat transfer (= h−1 A−1 , 19.0 ◦ C W−1 for eggs,
dinary differential equations. A solver for the FTMs was developed 2.0 ◦ C W−1 for 200 mL milk under type A conditions, 3.5 ◦ C W−1
using an implicit trapezoidal or Crank-Nicolson method, and food for 200 mL milk under type B conditions, and 0.9 ◦ C W−1 for
temperature histories were obtained by inputting the surrounding 930 mL milk) was greater than other thermal resistances, so that it
temperature history and initial food temperature, which was as- was the bottle neck to control the heat transfer phenomena. This
sumed to be the initial surrounding temperature in this study. The implies that accurate determination of the heat transfer coefficient
thermal capacity weighted average of the layer temperatures, i.e., is critical for accurate FTM.
T1 , T2 and T3 , was taken as the effective food temperature of FTM The impact of natural convection in milk on the temperature
to consider the temperature distribution in food. distribution in milk was investigated using CFD simulations. Re-
sults are shown in Fig. 3(a) and (b). Natural convection decreased
3.2. Freshness monitoring models for eggs and milk the temperature range in milk, reducing thermal resistance. The
impact of natural convection was considered in the current FTM.
The quality index used to estimate freshness varies depending However, this had a negligible effect on the FTM predictions, which
on the food. Weight loss, yolk index, and the Haugh unit (HU) are we attributed to the fact the impact of the heat transfer coefficient
possible candidate quality indices for eggs. Yimenu et al. (2017) de- on the prediction of FTM was greater than that of internal thermal
veloped an Food Freshness Monitoring Model(FMM) using three resistances.
474 H. Song, J. Kim and B.-S. Kim et al. / International Journal of Refrigeration 98 (2019) 468–479

Fig. 4. Flowchart of food temperature model development and its integration with a freshness monitoring model.

Table 2
Determined model constants of each food type.

Food Rs [°C W−1 ] R1 [°C W−1 ] R2 [°C W−1 ] C1 [J °C−1 ] C2 [J °C−1 ] C3 [J °C−1 ] A [m2 ]

Eggs – 1.75 16.04 57.89 132.14 16.31 6.80 × 10−3


Milk (200 mL) 0.45 0.20 1.27 348.58 353.39 79.33 2.07 × 10−2
Milk (930 mL) 0.14 0.05 0.14 1730.9 1494.5 448.8 5.50 × 10−2

Table 3
Comparison of deviations predicted by food temperature model(T3 ) and the surrounding temperature from the measured food temperature.

Food FTM predictions Surrounding temperature

mean(|T3.meas. − T3.F T M | ) [°C] max(|T3.meas. − T3.F T M | ) [°C] mean(|T3.meas. − T∞ | ) [°C] max(|T3.meas. − T∞ | ) [°C]

Eggs 0.3 1.8 1.4 16.1


Milk (200 mL): type A 0.5 2.6 1.7 12.1
Milk (200 mL): type B 0.9 4.1 2.7 17.0
Milk (930 mL) 1.0 3.6 2.4 10.1
H. Song, J. Kim and B.-S. Kim et al. / International Journal of Refrigeration 98 (2019) 468–479 475

were below 0.4 °C and around 1.0 °C on average for the FTM for
eggs and milk, respectively. The mean absolute error of FTM de-
creased for longer exposures to a constant temperature, i.e., cases
with a higher Fourier number, because the food temperature ap-
proached the surrounding temperature with time. The same result
was observed experimentally. The prediction error was the highest
for the case of 200 mL milk under type B temperature conditions,
which had the lowest Fourier number among the tested cases. In
particular, the maximum temperature differences decreased signif-
icantly using FTM, which implies improvement in the accuracy of
food freshness estimation in the case of temperature abuse un-
der real distribution conditions using FTM predictions instead of
the surrounding temperature history. Maximum deviations were as
high as 4 °C near the points of temperature shift as a result of ac-
cumulated error. However, this was for the limiting cases of tem-
perature abuse with sudden and large temperature shifts; the max-
imum deviation in real-life distribution is likely to be much lower
than this.

4.2. Spatial distribution of food temperature and heat transfer


coefficients in stacked milk

Computational fluid dynamics simulations considering 27


(3 × 3 × 3) crates of milk consisting of 25 (5 × 5) unit bottles, i.e.,
a total 675 bottles, were performed to investigate fluid flow and
heat transfer phenomena under the influence of an incoming flow
of velocity 1.7 ms−1 with a temperature shift between 5 and 20 °C
considering real-life distribution conditions.
A figure of the stacked milk bottles is shown in Fig. 6(a). Flow
velocity decreased as fluid flow passed through the stacked bottles
due to the drag force exerted by the surfaces to result in the spatial
variation in the temperature field, as shown in Fig. 6(b).
The means and standard deviations of milk temperature in each
crate were analyzed as functions of distance from the flow inlet
(ROW), location to the flow normal direction (COL), and number of
layers of milk crates (LYR) at 15 min from the start of the heating
shift. ROW had the most significant impact on milk mean temper-
ature in each crate. The mean temperature was lower by 1.2 and
2.2 °C for crates in the second and third ROWs compared to that
in the first ROW. The difference in the mean temperature was sta-
tistically negligible between the first and the third COL, and it was
lower by 0.3 °C in the second COL. The second LYR had a tempera-
ture that was lower than that of the other two LYRs by 0.1 °C. The
standard deviations of milk temperature in each crate increased
with ROW. In the third ROW, the standard deviation was higher
than in the first ROW by 0.5 °C. The effect of COL on the stan-
Fig. 5. Comparison of FTM prediction temperatures with the measured internal
dard deviation was negligible. The standard deviation increased by
food temperatures and surrounding temperatures for 200 mL milk type A, 200 mL
milk type B, and 930 mL milk. Gray lines are the average surrounding temperature 0.2 °C in the second LYR. The maximum standard deviation in a
values. Blue and red lines are average values of the measured food internal temper- crate was around 0.9 °C. This implies that there are spatial distribu-
ature and the predictions using FTMs and T3 , respectively. The red regions in the tions of temperature and heat transfer coefficients. However, these
figures represent the range of FTM predictions considering the standard deviations
distributions are strongly dependent on the layout of food in the
of the heat transfer coefficients estimations. (For interpretation of the references to
colour in this figure legend, the reader is referred to the web version of this article.) distribution and the flow field, which are likely to vary under real-
life distribution conditions, and the effects could not be considered
fully in CFD simulations. Hence, the heat transfer coefficient should
Fig. 5 compares the measured food temperature with the FTM be determined experimentally for each stage of distribution as out-
prediction results for milk under fluctuating temperature condi- lined in Section 2.3.
tions. Gray lines are the average values of the surrounding tem-
peratures measured at multiple locations during the experiments. 4.3. Application of FQMSs to eggs and milk
The average values of the measured food internal temperatures and
the predictions using FTMs are shown as blue and red lines, re- The validated FTMs were integrated with FMMs to complete
spectively. The red regions in the figures represent the range of FQMSs for eggs and milk. Food temperature histories were pre-
FTM predictions considering the standard deviations of the heat dicted by FTMs using the monitored temperature histories of the
transfer coefficient estimations. Mean and maximum absolute dif- surroundings and were fed into the FMMs to evaluate the fresh-
ferences between the measured internal food temperature and the ness variation of eggs and milk. The predicted freshness histo-
surrounding temperature as well as the FTM predictions(T3 ) are ries were compared with the measured values to validate the
shown in Table 3. Errors in the estimation of food temperature FQMSs for eggs and milk. To evaluate the improvement using the
476 H. Song, J. Kim and B.-S. Kim et al. / International Journal of Refrigeration 98 (2019) 468–479

Fig. 6. (a) Configuration of 27 (3 × 3 × 3) crates of milk consisting of 25 (5 × 5) unit bottles and (b) the temperature field of the stacked milk bottles.

Table 4 The measured HU values are represented by solid green circles. The
List of three possible combinations of food temperature scenarios and the
three prediction scenarios are distinguished by the gray dashed,
food freshness model used.
blue solid, and red solid lines. It is clear from Fig. 7(a) that the
Food temperature Food freshness model choice of egg temperature histories had no significant effect on the
Case 1 Measured surrounding temperature FMM freshness estimates.
Case 2 Measured food internal temperature FMM Fig. 7(b) and (c) compare microbial growth predictions in milk
FQMS FTM FMM in 200 mL and 930 mL packages using the three scenarios under
fluctuating temperature conditions. Green stars represent experi-
mentally obtained microbial counts. The red shaded area repre-
sents regions of possible deviation in microbial growth predictions
FTM, freshness estimations using the three combinations listed in caused by the standard deviation of the heat transfer coefficient.
Table 4 were compared for the model foods. The measured sur- The accuracy of food freshness predictions using the developed
rounding and food internal temperatures were used as the input FQMS was estimated using the bias factor (Bf ), accuracy factor (Af ),
food temperature histories of the FMMs for cases 1 and 2, respec- and mean absolute percentage error (MAPE). Table 5 compares the
tively, whereas the FTM predicted effective food temperature his- Bf , Af , and MAPE values of each temperature scenario model for
tory was used for the FQMS. the model foods. Bf represents the average logarithmic ratio of the
The predicted and measured HU histories for eggs exposed to predicted and observed values of the freshness index, as shown in
fluctuating temperature conditions shown are shown in Fig. 7(a). Eq. (21). Here n, Cpredicted , and Cobserved are the total number of ex-

Table 5
Comparison of the food freshness prediction accuracy between the three scenarios using the accuracy factor, bias factor,
and mean absolute percentage error.

Food Case 1 Case 2 FQMS

Af [-] Bf [-] MAPE [%] Af [-] Bf [-] MAPE [%] Af [-] Bf [-] MAPE [%]

Eggs 1.03 1.02 2.56 1.03 1.02 2.83 1.03 1.02 2.92
Milk (200 mL): type A 1.25 1.20 27.54 1.21 1.16 23.14 1.19 1.14 20.71
Milk (930 mL) 1.13 1.11 14.12 1.13 1.11 14.17 1.11 1.09 11.74
H. Song, J. Kim and B.-S. Kim et al. / International Journal of Refrigeration 98 (2019) 468–479 477

Fig. 7. Comparison of the predictions and measured food quality index histories with the predictions obtained using the three model scenarios in Table 4 for (a) eggs, (b)
200 mL milk (type A), and (c) 930 mL milk; solid green circles and green vertical lines in (a) represent measured Haugh units and standard deviations, respectively. The solid
green stars in (b) and (c) represent the experimentally obtained microbial counts. The upper line graphs show the temperature histories of the surroundings for each case.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

perimental samples and the predicted and observed freshness in- Bf is 1.0 when the two values are averagely identical. It is less
dex values, respectively. than unity when the model underestimates the measured data and
  
it is greater than unity for the case of overestimation. Af is an index
log C predicted / Cobserved / of the error dispersion of the model predictions and is defined in
n
B f = 10 (21)
478 H. Song, J. Kim and B.-S. Kim et al. / International Journal of Refrigeration 98 (2019) 468–479

Eq. (22). ployed as platform technologies for real-time food quality monitor-
ing systems to estimate food quality changes during distribution.
  

log Cpredicted / Cobserved


/
n
A f = 10 (22)
Acknowledgments

A model is more accurate when the value is closer to 1.0 but is This research was supported by the Main Research Program
not less than unity (Ross, 1996). MAPE is a relative error measure (E0162501) of the Korea Food Research Institute (KFRI) funded by
of prediction accuracy and is defined in Eq. (23) the Ministry of Science and ICT of South Korea.

1 Cobserved − C predicted
MAP E = × 100 (23) References
n Cobserved
Anonymous, 2017. STAR-CCM + User Guide. CD-adapco.
Anonymous, 2017. R: A Language and Environment for Statistical Computing. R Core
Case 2 showed better or compatible freshness estimation ac-
Team.
curacy than case 1, as shown in Table 5, which means that the Anonymous, 2011. FLUENT Theory Guide. ANSYS, Canonsburg, PA.
food internal temperature is a better representation of the effec- Anonymous, 1998. Thermal Properties of Foods. ASHRAE Refrigeration Handbook,
tive milk temperature for microbial growth than the surrounding Atlanta, GA.
Bruckner, S., Albrecht, A., Petersen, B., Kreyenschmidt, J., 2013. A predictive shelf
temperature. The FQI estimates of FQMS, which used FTM to pre- life model as a tool for the improvement of quality management in pork and
dict the effective food temperature, were more accurate with the poultry chains. Food Control 29, 451–460. doi:10.1016/J.FOODCONT.2012.05.048.
best Af , Bf , and MAPE values than the other two scenarios. We at- Buzby, J.C., Farah-Wells, H., Hyman, J., 2014. The estimated amount, value, and calo-
ries of postharvest food losses at the retail and consumer levels in the United
tributed this to the fact that FTM estimates the effective food tem- States. SSRN Electron. J. doi:10.2139/ssrn.2501659.
perature considering the temperature distribution in food, whereas Carullo, A., Corbellini, S., Parvis, M., Vallan, A., 2009. A wireless sensor network for
the food internal temperature is measured only at a single point in cold-chain monitoring. IEEE Trans. Instrum. Meas. 58, 1405–1411. doi:10.1109/
TIM.20 08.20 09186.
food. Cengel, Y., 2014. Heat and Mass Transfer: Fundamentals and Applications. Mc-
Graw-Hill Higher Education.
Denys, S., Pieters, J.G., Dewettinck, K., 2004. Computational fluid dynamics analysis
of combined conductive and convective heat transfer in model eggs. J. Food Eng.
63, 281–290. doi:10.1016/J.JFOODENG.20 03.06.0 02.
Denys, S., Pieters, J.G., Dewettinck, K., 2003. Combined CFD and experimental ap-
5. Conclusions proach for determination of the surface heat transfer coefficient during ther-
mal processing of eggs. J. Food Sci. 68, 943–951. doi:10.1111/j.1365-2621.2003.
In this study, food temperature models (FTMs) for eggs and tb08269.x.
Gospavić, R., Margeirsson, B., Popov, V., 2012. Mathematical model for estimation
milk were developed to predict food temperature changes dur- of the three-dimensional unsteady temperature variation in chilled packaging
ing distribution using surrounding temperature histories and units. Int. J. Refrig. 35, 1304–1317. doi:10.1016/J.IJREFRIG.2012.04.002.
initial food temperatures. Surface heat transfer coefficients were Gwanpua, S.G., Verboven, P., Leducq, D., Brown, T., Verlinden, B.E., Bekele, E.,
Aregawi, W., Evans, J., Foster, A., Duret, S., Hoang, H.M., Van Der Sluis, S.,
determined from fluctuating temperature experiments; heat trans- Wissink, E., Hendriksen, L.J.A.M., Taoukis, P., Gogou, E., Stahl, V., El Jabri, M.,
fer coefficients should be measured in the same way during each Le Page, J.F., Claussen, I., Indergård, E., Nicolai, B.M., Alvarez, G., Geeraerd, A.H.,
stage of a real-life distribution. We found that the resistance to sur- 2015. The FRISBEE tool, a software for optimizing the trade-off between food
quality, energy use, and global warming impact of cold chains. J. Food Eng. 148,
face heat transfer was much greater than the interior resistances. 2–12. doi:10.1016/j.jfoodeng.2014.06.021.
This implies that accurate determination of the heat transfer coef- Haugh, R.R., 1937. The Haugh unit for measuring egg quality. United States Egg
ficient is critical for accurate FTMs, and this should be determined Poult. Mag. 43, 522–555.
Hoang, H.M., Flick, D., Derens, E., Alvarez, G., Laguerre, O., 2012. Combined deter-
experimentally for each distribution stage. Mean absolute model ministic and stochastic approaches for modelling the evolution of food prod-
prediction errors of temperature history were about 0.3 °C and 1 °C ucts along the cold chain. Part II: a case study. Int. J. Refrig. 35, 915–926.
for eggs and milk in two different types of containers, respectively. doi:10.1016/J.IJREFRIG.2011.12.009.
Jedermann, R., Ruiz-Garcia, L., Lang, W., 2009. Spatial temperature profiling by semi-
The developed food temperature model was integrated with
passive RFID loggers for perishable food transportation. Comput. Electron. Agric.
freshness monitoring models (FMMs) to complete real-time food 65, 145–154. doi:10.1016/j.compag.20 08.08.0 06.
quality monitoring systems (FQMSs). The improvement in food Kim, B.-S., Lee, M., Kim, J.-Y., Jung, J.-Y., Koo, J., 2016. Development of a freshness-
quality prediction using the FQMSs was tested with eggs and milk, assessment model for a real-time online monitoring system of packaged com-
mercial milk in distribution. LWT - Food Sci. Technol. 68, 532–540. doi:10.1016/
where Haugh unit and microbial concentration were used as qual- j.lwt.2015.12.049.
ity indices, respectively. Haugh unit predictions of eggs using the Kondjoyan, A., 2006. A review on surface heat and mass transfer coefficients during
measured surrounding temperature, measured food internal tem- air chilling and storage of food products. Int. J. Refrig. 29, 863–875. doi:10.1016/
J.IJREFRIG.20 06.02.0 05.
perature, and effective food temperature from the FTM were sim- Lytou, A., Panagou, E.Z., Nychas, G.-J.E., 2016. Development of a predictive model
ilar. We attributed this to the low thermal capacity of eggs and for the growth kinetics of aerobic microbial population on pomegranate mari-
concluded that the surrounding temperature can be used as the nated chicken breast fillets under isothermal and dynamic temperature condi-
tions. Food Microbiol. 55, 25–31. doi:10.1016/j.fm.2015.11.009.
effective food temperature for foods with small thermal capaci- Marai, S.V., Ferrari, E., Civelli, R., 2012. Post harvest cold chain optimization of little
ties. The accuracy of milk quality predictions using the effective fruits. In: Proceedings of the COMSOL Conference. Milan.
food temperature predicted by the FTM was better than that ob- Margeirsson, B., Pálsson, H., Gospavic, R., Popov, V., Jónsson, M.Þ., Arason, S., 2012.
Numerical modeling of temperature fluctuations of chilled and superchilled cod
tained using the measured food internal temperature. While the
fillets packaged in expanded polystyrene boxes stored on pallets under dynamic
measured food internal temperature is merely a single point mea- temperature conditions. J. Food Eng. 113, 87–99. doi:10.1016/J.JFOODENG.2012.
surement in food, the effective temperature, i.e., the thermal ca- 05.017.
Price, K.V., Storn, R.M., Lampinen, J.A., 2005. Differential Evolution : A Practical Ap-
pacity weighted average temperature of the layers in the model,
proach to Global Optimization. Springer.
predicted by the FTM contains integrated information of the tem- Raab, V., Bruckner, S., Beierle, E., Kampmann, Y., Petersen, B., Kreyenschmidt, J.,
perature field in milk. Therefore, it reflects quality changes in milk 2008. Generic model for the prediction of remaining shelf life in support of
better. cold chain management in pork and poultry supply chains. J. Chain Netw. Sci.
8, 59–73. doi:10.3920/JCNS2008.x089.
We anticipate that the proposed FQMSs, by considering the spa- Rao, M., Rizvi, S., Datta, A., Ahmed, J., 2014. Engineering Properties of Foods, Fourth
tial food temperature experienced during distribution, will be em- Ed. CRC Press doi:10.1201/b16897.
H. Song, J. Kim and B.-S. Kim et al. / International Journal of Refrigeration 98 (2019) 468–479 479

Ross, T., 1996. Indices for performance evaluation of predictive models in food van der Sman, R.G.M., 2003. Simple model for estimating heat and mass transfer
microbiology. J. Appl. Bacteriol. 81, 501–508. doi:10.1111/j.1365-2672.1996. in regular-shaped foods. J. Food Eng. 60, 383–390. doi:10.1016/S0260-8774(03)
tb03539.x. 0 0 061-X.
Shih, C.-W., Wang, C.-H., 2016. Integrating wireless sensor networks with statisti- Yimenu, S.M., Kim, J.Y., Koo, J., Kim, B.S., 2017. Predictive modeling for monitor-
cal quality control to develop a cold chain system in food industries. Comput. ing egg freshness during variable temperature storage conditions. Poult. Sci. 96,
Stand. Interfaces 45, 62–78. doi:10.1016/J.CSI.2015.12.004. 2811–2819. doi:10.3382/ps/pex038.

You might also like