Sie sind auf Seite 1von 14

Journal of Food Engineering 190 (2016) 80e93

Contents lists available at ScienceDirect

Journal of Food Engineering


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

Review

A review of the use and design of produce simulators for horticultural


forced-air cooling studies
Gabe P. Redding a, Angela Yang b, Young Min Shim b, Jamal Olatunji b, Andrew East b, *
a
Biological Systems Modeling Limited, Palmerston North 4442, New Zealand
b
Centre for Postharvest and Refrigeration Research, Massey University, Private Bag 11-222, Palmerston North 4442, New Zealand

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

Article history: Forced air cooling studies are hindered by the practical difficulties and cost associated with pallet scale
Received 10 December 2015 investigations. Produce simulators are a potential solution for overcoming some of these difficulties.
Received in revised form Despite their common use in precooling research, there are a lack of design guidelines for such simu-
28 April 2016
lators in the literature which in some case has led to their inappropriate use. This review examines
Accepted 19 June 2016
Available online 21 June 2016
previous use of produce simulators in produce cooling studies in the context of transport phenomena
theory. This allows constraints to be proposed which facilitate the design of produce simulators that can
replicate transport phenomena in horticultural produce. These include the requirements of geometric
Keywords:
Produce
and thermal property matching as well as the practical constraints imposed on produce simulators. It is
Simulator suggested that future computational modelling work is well suited to the establishment of produce
Cooling simulator design guidelines which incorporate complex effects such as the impact of variation in produce
Design geometry on flow. The material engineering challenges posed by produce simulators are also discussed.
Heat transfer © 2016 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
2. Simulation of heat transfer in produce during forced air pre-cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3. Geometric matching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4. Evaporative cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5. Addition of respiratory heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
6. Thermal property matching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.1. Single-domain produce simulators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.2. Multi-domain produce simulators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
6.3. Materials for thermal property matching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
7. Simplifying cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
8. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

1. Introduction post harvest due to continued physiological, biochemical, and


microbiological processes (Verboven et al., 2006; Dehghannya
The quality of horticultural produce progressively deteriorates et al., 2010). Immediate chilling of produce after harvest partially
arrests these processes and is used to maintain quality and prolong
storage (Castro et al., 2004). Such chilling is known as precooling
and its optimisation is vitally important to the food industry
* Corresponding author. (Brosnan and Sun, 2001), with up to 20% of all perishable foods
E-mail address: A.R.East@massey.ac.nz (A. East).

http://dx.doi.org/10.1016/j.jfoodeng.2016.06.014
0260-8774/© 2016 Elsevier Ltd. All rights reserved.
G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93 81

Nomenclature r radial distance from center (m)


ro radius (m)
a thermal diffusivity (m2/s) r* dimensionless distance from centre (r/ro)
Bi Biot number, hL/k T temperature (K)
c heat capacity (J/kg K) DT change in temperature associated with a period of
E activation energy of carbon dioxide respiration (J/mol) precooling (s)
V evaporative heat loss as percentage of heat load t time (s)
Fo Fourier number, at/L2 or at/ro2 Dt time period of precooling (s)
h convective heat transfer coefficient (W/m2K) q temperature ( C)
DHvap latent heat of vaporisation (J/kg) u flow velocity vector (m/s)
k thermal conductivity (W/mK) n kinematic viscosity (m2/s)
kt transpiration mass transfer coefficient (kg/kg s Pa or Y fraction of unaccomplished change in temperature,
kg/m2 s Pa) (TT∞)/(TiT∞)
kair air film mass transfer coefficient (kg/m2 s Pa) z respiratory production rate of carbon dioxide (mL/
kskin skin mass transfer coefficient (kg/m2 s Pa) kg hr)
L characteristic length, half the shortest dimension was zo respiratory pre-exponential for the production rate of
used for produce (m) carbon dioxide (mL/kg hr)
M percentage moisture loss from produce (%)
r density (kg/m3) Subscripts
P static pressure (Pa) a air
Pair partial pressure of water in air (Pa) c composite material
Psurf partial pressure of water at produce surface (Pa) f filler material
f volume fraction of filler material i initial
q heat of respiration (W/kg) m matrix material
Q heat of respiration as a percentage of heat load ∞ Ambient
R the gas constant 8.314 (J/mol K) p produce

being lost due to a lack of appropriate refrigeration (Defraeye et al., use and development of non-invasive flow measurement tech-
2015a). niques in packaging studies (Ferrua and Singh, 2008; O’Sullivan
In order to slow spoilage processes it is desirable to cool produce et al., 2014).
as rapidly as possible, provided this does not cause chilling injury to Practical considerations of expense and time result from the fact
the produce (Valente et al., 1996) and is done in a homogenous that experiments are required to be on a full pallet scale and are
fashion. Non-uniform cooling results in under or over-cooling of subject to the seasonal availability of produce. This has limited the
produce in different locations, which leads to variation in product scope of many studies (Ferrua and Singh, 2008). Experimental
quality and associated product loss (Ferrua and Singh, 2009d; replication of cooling studies is hindered by the biological nature of
Dehghannya et al., 2010). Hence, minimisation of product loss can produce (Castro et al., 2004). Degradation of produce, for example,
generally be achieved by cooling as rapidly and uniformly as can change the heat transfer properties of produce during experi-
possible. For forced air cooled systems, the rate and uniformity of mentation. Surface evaporation is a significant form of heat loss for
cooling is ultimately determined by the flowrate, distribution and some produce (Chuntranuluck et al., 1998a) and moisture loss from
temperature of the air throughout the entire packaging structure products may mean that this effect differs over time. To avoid such
(O’Sullivan et al., 2014). issues, replicate runs can be performed with fresh batches of pro-
One of the key parameters which effects airflow within horti- duce, though this is costly, time consuming and subject to any
cultural packaging is the size and location of vents. Enhancement of biological variation in parameters such as shape and ripeness
cooling via design of vents is constrained by the requirement that which will in turn lead to variability in replicate experiments.
the packaging be of sufficient strength to protect produce from These problems of experimental cost and time are magnified
damage during handling, transport and storage (Delele et al., 2008; when considering designing packaging to ensure maintenance of
Defraeye et al., 2015a). Though requirements of structural integrity temperature in transport systems such as refrigerated container
have historically been prioritised over those which promote rapid ship holds. In these cases full containers may be required to
and uniform cooling (Castro et al., 2004; Ferrua and Singh, 2009a), conduct a useful experiment (e.g. Tanner and Amos, 2003; Defraeye
ventilated package design has still been the subject of a great deal et al., 2015b), resulting in considerable cost for a single replicate.
of research as summarised in the review of Pathare et al. (2012). Hence, cost becomes a limiting constraint when studying these
This review demonstrated that there is considerable variation in systems (Tanner et al., 2002a; Ambaw et al., 2013).
the design recommendations for ventilated packages made across The difficulties associated with experimentation have led many
various studies. studies to focus on the alternative approach of developing mathe-
The lack of consensus in design recommendations is partly the matical models capable of predicting the flow fields and heat
result of the challenges associated with studies in this field which transfer within packaged commodities during forced air cooling
involve geometrically complex packaging systems and biologically (Ferrua and Singh, 2008). These mathematical approaches have
variable produce. The experimental determination of airflow been the subject of multiple reviews (Verboven et al., 2006;
around individual products is difficult to achieve without disturb- Dehghannya et al., 2010; Delele et al., 2010; Ambaw et al., 2013;
ing the packaging arrangement itself (Dehghannya et al., 2010; Zhao et al., 2016). Earlier studies such as those of Talbot (1988)
O’Sullivan et al., 2014) and these difficulties have promoted the made the simplifying assumption that the packed structure was a
82 G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93

porous medium, however this assumption has been shown to be current use in the context of design theory as it applies to the forced
invalid in many common packages where the container to produce air precooling of horticultural produce.
diameter ratio is less than 10 (Pathare et al., 2012). Continued ad-
vancements in numerical modelling techniques and computational 2. Simulation of heat transfer in produce during forced air
power have greatly reduced the need for such simplifying as- pre-cooling
sumptions and have allowed examination of transport phenomena
which can be highly localised in packaged and palletised horticul- Of the variety of different methods available for post-harvest
tural produce (Alvarez and Trystram, 1995; Verboven et al., 2006; cooling operations, forced air cooling is the most common
Dehghannya et al., 2010; Delele et al., 2010; Ambaw et al., 2013; (Ambaw et al., 2013). Fig. 2 shows a graphical representation of a
Defraeye et al., 2015a). tunnel cooler, the most common type of forced air cooling device
For the purposes of reducing the complications associated with (O’Sullivan et al., 2014). Produce is palletised and stacked in
experimentation and also for mathematical model validation, many ventilated packages and this physical structure leads to the uneven
studies have also utilised produce simulators (Minh et al., 1969; distribution of refrigeration air and subsequent cooling heteroge-
Alvarez and Trystram, 1995; Chuntranuluck et al., 1998b; Alvarez neity due to localised heat transfer (Alvarez and Trystram, 1995;
and Flick, 1999a; Alvarez and Flick, 1999b; Vissotto et al., 2000; Dehghannya et al., 2010; O’ Sullivan et al., 2014). Achieving a suf-
Alvarez et al., 2003; Castro et al., 2004; Vigneault and Castro, ficiently homogenous distribution of refrigeration air is the subject
2005; Allais et al., 2006; Delele et al., 2008; Ferrua and Singh, of a large volume of research, and can be facilitated in a variety of
2008). Produce simulators are generally employed with the ways including adjusting package stacking arrangements, baffling,
objective to sufficiently replicate the flow field and/or temperature periodic flow reversal, airflow bypass, manipulation of packaging
profiles that would be produced by real produce, often given ventilation design and the positioning of produce within packages
specified simplifying assumptions (e.g. negligible heat added by (Aswaney, 2007; Ferrua and Singh, 2009a; Dehghannya et al., 2010;
produce respiration). A produce simulator which can achieve these Pathare et al., 2012; O’ Sullivan et al., 2014).
objectives must also satisfy a number of practical requirements. In forced air cooling, heat transfer within produce is multifac-
These include being economical and convenient for use in studies eted (Fig. 3). Heat is removed from produce by conduction to the
conducted on a pallet scale across a range of produce. Such a pro- product surface followed by convective heat transfer from the
duce simulator would facilitate more rapid research progress by produce surface over which cold air is forced by a fan. Additional
overcoming the limitations associated with large scale studies heat may be lost from the surface of produce via moisture loss and
involving real horticultural produce. the associated removal of latent heat of evaporation
Produce simulators are seen to be the most appropriate way to (Chuntranuluck et al., 1998a). Finally, continued post-harvest
conduct experiments for determining the effect of different pa- respiration also adds heat to produce. Net radiation effects on
rameters on horticultural packaging design (Vigneault and Castro, packaged products are small relative to experimental error and, if
2005) as well as for validating mathematical models which are required, can be sufficiently described as pseudo convection
being increasingly used in this area of research (Pathare et al., 2012; (Alvarez and Flick, 1999b; Tanner et al., 2002b; Le Page et al., 2009).
Ambaw et al., 2013). Highly effective produce simulators would Ferrua and Singh (2009b, c) showed that transport phenomena
likely play an important role in turning optimal horticultural during forced air precooling applications can be modelled by
packaging design into a reality. decoupling the momentum transport from the transport of energy
Given the exponential increase in research interest in horticul- and mass, and that the flow field within and between packages can
tural packaging design over the past 20 years (Fig. 1), the potential be sufficiently described as the steady, laminar and incompressible
need for produce simulators has never been greater. However, flow of air for use in subsequent heat and mass transfer models
despite the use of produce simulators in horticultural studies and (Ferrua and Singh, 2009a). Under these conditions, flow within the
their potential utility to the industry, there has been very little system can be described by the following forms of the continuity
research specifically targeted at the effective design of produce and Navier-Stokes equations respectively,
simulators. This has increased the likelihood of inappropriate use
and haphazard experimental design involving produce simulators. V$u ¼ 0 (1)
The objective of this review is to highlight the challenges that need
to be addressed in produce simulator design by reviewing their VP
u$Vu ¼  þ n a V2 u (2)
ra
The thermophysical properties of both air and produce can be
considered to be constant over the duration of precooling due to
the relatively modest temperature range and small losses in
moisture (Tanner et al., 2002b; Dehghannya et al., 2010; Ferrua and
Singh, 2009a). It can also be assumed that energy transfer in air due
to the transfer of enthalpy associated with the diffusion of water
vapour is negligible (Ferrua and Singh, 2009a). Under these as-
sumptions heat transfer in the air domain is described by,
 
vTa
ðrcÞa þ u$VTa ¼ ka V2 Ta (3)
vt

and in the produce domain by,

vTp
ðrcÞp ¼ kp V2 Tp þ Q (4)
vt
Fig. 1. Web of Science™ citations for the topic keywords horticultural, produce,
packaging and design. The boundary condition at the air produce interface can be
G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93 83

Fig. 2. A tunnel cooler, the most common type of forced air cooling device.

It is implicit from Equations (1) through (5) that, for a produce


simulator to be capable of reproducing the absolute temperature
profiles of real produce, the flow, heat and mass transfer phe-
nomena must be sufficiently replicated. Specifically, the geometry,
evaporative cooling, produce respiration, and thermal properties of
each structure within the produce must be sufficiently matched.
Table 1 reports some of these properties and other relevant pa-
rameters for a range of horticultural produce.
The following sections of this review detail the use of produce
simulators in a sample of previous literature studies in relation to
the replication of these phenomena. Table 2 summarises this
comparison for the studies for which the thermophysical proper-
ties of the produce simulators employed were reported. It should
be noted that although the use of produce simulators is common in
postharvest research, these simulators are rarely the focus of such
research. Instead, they are typically employed as experimental
devices assisting in the meeting of a specific objective. For these
reasons it is not feasible to exhaustively review the previous use of
produce simulators. Rather, what is intended to be a representative
sample is discussed here.

3. Geometric matching
Fig. 3. A schematic of heat transfer mechanisms in produce.

Geometric matching refers to the degree to which a produce


described by, simulator matches the shape and size of produce. Since momentum
transport can be addressed as decoupled (Ferrua and Singh, 2009b,
    c), geometric matching alone will allow the replication of flow
kp VT ¼ h Tp  T∞ þ kt Psurf  Pair DHvap (5) fields, without the need to replicate thermal properties. Decoupling
further permits modelling procedures to be simplified and more
where the heat transfer coefficient can be described as a function of computationally feasibly by following the step by step approach of
fluid velocity or by some other appropriate correlation (Nusselt- (i) determination of air flow (ii) determination of transfer co-
Reynolds-Prandtl correlations are typically used). If the produce efficients from flow fields and (iii) determination of heat and mass
surface water vapour pressure (Psurf) is not sufficiently constant, a transfer (Le Page et al., 2009). This type of modelling approach has
suitable mass transfer model which appropriately describes water also been utilised in studies examining the air cooling of meat
transport through the produce domain as a function of time will products (Hu and Sun, 2000, 2001; Mirade et al., 2002; Trujillo and
need to be incorporated. Alternatively, it may be sufficient to Pham. 2006; Pham et al., 2009).
incorporate the effect of evaporation on heat transfer by removing Attempts to utilise produce simulators which deliberately
the 2nd term on right hand side of Equation (5) and incorporating replicate the shape of specific produce are sparse. Valente et al.
evaporative effects into the heat transfer coefficient (Alvarez and (1996) used avocado shaped produce simulators in their study of
Flick, 1999b). heat transfer in avocados. All of the studies in Table 2 used idealised
84
Table 1
Thermophysical parameters for various produce.

k (W/m K) a (x 106 m2/s) r (kg/m3) c (J/kgK) rc L (103 m) Bi q (103 W/kg) z0d (mL/kg hr) Ed (kJ/mol) Source(s)
(103 J/m3 K)

Apples 0.45a 0.14 840 3810 3200 19e44 0.46e4.11c 20e61 (at 10  C) 4.96  1013 65.7 Sweat (1974), ASHRAE (2006), Tabatabaeefar and
Rajabipour (2005)
Blueberries 0.54 0.14e0.15a 990 3720e4040 3682e4000 2e8 0.04e0.62c 101e183 (at 15  C) 9.53  1017 92.9 Marai et al. (2012), MacGregor (2005), ASHRAE (2006),
Oliveira et al. (2012)
Cherries 0.51e0.53a 0.13 1050 3730e3850 3917e4043 15 0.31e1.24c 81e148 (at 15  C) 4.96  1015 75.3 Milosevic and Milosevic (2012), ASHRAE (2006)
Kiwifruit 0.40e0.45a 0.11 940e1040 3893 3659e4049 21e26 0.51e2.73c 39 (at 10  C) 2.70  1011 58.7 Chen et al. (1998), ASHRAE (2006), Razavi and
BahramParvar (2007), Oliveira et al. (2012)
Melon 0.57 0.15a 930 4030 3747 64b 1.24e4.72c 22e46 (at 10  C) 1.42  1010 50.5 Eifert et al. (2006), Sweat (1974), ASHRAE (2006),
Oliveira et al. (2012)
Peaches 0.45a 0.12 930 3910 3636 23e44b 0.56e4.11c 47 (at 10  C) 1.12  1017 87.6 Sweat (1974), ASHRAE (2006), Alibeyglu et al. (2013)
Potatoes 0.50e0.51a 0.13 1040e1070 3670 3817e3927 28b 0.60e2.35c 20e62 (at 10  C) 4.54  108 41.6 Maurer and Eaton (1971), ASHRAE (2006)
Watere 0.14a

G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93


0.582 1000 4198 4198 n/a n/a n/a n/a n/a Incropera and Dewitt (2006)
Range 0.40e57 0.11e0.15 840e1070 3670e4040 3200e4049 2e64 0.04e4.72 20e183 n/a n/a n/a
a
Calculated from Equation (2) (k ¼ arc).
b
Calculated from reported surface area by assuming spherical. Values of thermal diffusivity and thermal conductivity are intended to represent the bulk portion of the produce.
c
Bi values calculated across a convective heat transfer coefficient range of 11e42 W/m2K which is the range reported by ASHRAE (2006) for a variety of produce being air cooled. The characteristic length used is half of the
shortest dimension.
d
Values taken from Exama et al. (1993) except for kiwifruit which was calculated from the work of Heyes et al. (2010).
e
The thermophysical properties of water at ~7  C are provided on Table 1 for comparison.

Table 2
Previous use of produce simulators, in the literature.

k (W/m K) a (106 m2/s) rc (103 J/m3 K) L (103 m) Bi Size Respiration Evaporation Application Source
matched simulation simulation

Acrylic spheres 0.21a 0.12 1731 38 2.00e7.60b n/a No No Determination of heat transfer Minh et al. (1969)
correlations for forced air cooling design
3% carraghenan gel spheres 0.86a 0.21 4153 38 0.49e1.86b n/a No Yes Model validation and development Alvarez and Trystram (1995)
of control design strategy for
produce cooling
Tylose filled aluminium cylinders 0.50 0.13a 3890 36e71 0.79e5.96b n/a No Yes Validation of chilling time prediction Chuntranuluck et al. (1998b)
model
Phenolic vinyl spheres 0.68 0.37a 1857 26 0.42e1.60b n/a No No Determination of cooling rate of Vigneault and Castro (2005)
produce for indirect airflow assessment
Celluloid spheres filled 0.52 0.12a 4153 19 0.40e1.53b n/a No No Validation of a mathematical Allais et al. (2006)
with 3% carraghenan gel model of mist chilling of foods
a
Calculated from Equation (2) (k ¼ arc).
b
Bi calculated across the same range of h described in Table 1. The characteristic length used is half of the shortest dimension.
G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93 85

spherical or cylindrical geometries, and were intended to be (2003) to examine flow heterogeneity and its relationship to heat
applicable to produce which sufficiently approximates these transfer coefficients. Although neither of these studies used mate-
shapes, and/or used for the purposes of model validation. rials that are thermally matched to produce, the results would be
Chuntranuluck et al., 1998b found little difference in heat transfer expected to apply to geometrically similar produce and are an
observed in their produce simulators as a result of variation in appropriate use of produce simulators which are not thermally
shape among the three idealised geometries investigated (spheres, matched.
cylinders and slabs). The experiments conducted by Chuntranuluck Since flow fields can be difficult to measure in stacked produce,
et al., 1998b involved placing produce simulators in a wind tunnel Vigneault and Castro (2005) devised a methodology in which the
which allowed homogenous flow to be delivered, and suggests in average approach velocity around a produce simulator could be
such cases it may be sufficient for produce simulators to approxi- calculated from the rate at which the simulator cooled. The produce
mate the shape of real produce, however this is unlikely to be the simulators employed were solid phenolic vinyl spheres which had
case in complex packaging structures. similar size and thermal conductivity to that of produce but were
The lack of available data means that the impact of variation in relatively poorly matched for thermal diffusivity and volumetric
produce geometry on the flow fields and subsequent heat transfer heat capacity. The authors suggested that the size and thermo-
in produce may be a valuable area of future investigation. The in- physical properties of a given produce would need to be matched
tricacies of geometric variation in real produce dictate that such before it could be used to deduce airflow within a group of produce
data may be most readily tractable in silico using modern compu- or other produce simulators. However, theory suggests that a
tational modelling techniques. Indeed, the determination of design geometrically matched produce simulator could have any known
guidelines for produce simulators which consider the aspects of thermal properties and be used to deduce airflow provided internal
geometry, evaporation, respiration and thermal matching is well temperature and air temperature local to the produce simulator
suited to the domain of computational modelling. It should be were known/measured. In theory, this provides a method by which
noted that the accurate replication and mathematical description of flow field data can be collected without impacting on the flow field
produce geometry is not necessarily a simple task and remains an itself.
active area of research (e.g. Olatunji et al., 2015; Rogge et al., 2015).
Although produce simulators which replicate produce shape are 4. Evaporative cooling
rare, most reported use of produce simulators attempts to
approximate produce size. This is evident in a comparison of As shown in the produce surface heat transfer boundary con-
Tables 1 and 2 which shows the characteristic lengths of produce dition (Equation (5)), the heat lost from produce by evaporation is
simulators to be comparable with those of real produce. Although proportional to the driving force between the vapour pressure of
the matching of characteristic length will likely allow approxima- water in the air stream and that at the produce surface. These will
tion of internal heat transfer, the failure to match geometry will be determined by the temperature and relative humidity of the air
result in different flow conditions which will result in some degree stream and the water activity of the produce surface respectively. In
of failure to replicate external convective heat transfer. Further- addition, the rate of water loss at the surface of produce is related to
more, for studies involving complex packaging structures, objects the resistance provided by the skin to transpiration and hence is
of similar characteristic length but different shape may not be able dependent on the type and state of the product (Maguire et al.,
to be packed in the same arrangement or bulk volumes, resulting in 2001).
variation in both flow and thermal mass within a specified volume. Of the studies summarised in Table 2 only Alvarez and Trystram
The simplicity afforded by replicating flow using thermally un- (1995) and Chuntranuluck et al., 1998b attempted to simulate the
matched produce simulators has seen them used in this capacity to effects of evaporative cooling in their produce simulators. The
validate CFD models of flow through ventilated horticultural produce simulators employed by Alvarez and Trystram (1995) were
packages (Ferrua and Singh, 2008; Delele et al., 2008) and for the composed of 3% carrageenan gum, the exposed surface of which
determination of heat transfer correlations and heat transfer co- allowed evaporation from a fully wetted surface to be simulated.
efficient determination. Minh et al. (1969) evaluated Nusselt- Chuntranuluck et al., 1998b simulated evaporative effects by
Reynolds-Prandtl heat transfer correlations of bulk potatoes for wrapping produce simulators in a layer of cloth which was soaked
use in the design of precooling systems. As a comparative standard, in a wetting agent. A fully wetted surface was maintained during
correlations were also evaluated for acrylic spheres which were the course of experimental trials by pumping a gentle flow of
dispersed throughout the bed of potatoes. Significantly different wetting agent over the surface of simulators. Different surface
transfer coefficients and heat transfer correlations were found be- water activities were achieved by using various salt solutions as
tween spheres and potatoes which were positioned in similar lo- wetting agents. Difficulties associated with crystallisation of these
cations in the airflow field. Although the thermophysical properties salt solutions were reported by the authors. The complexity of this
of the acrylic spheres (Table 2) were not well matched to produce, system for the replication of evaporative effects would be chal-
differences in heat transfer correlations are not expected to be due lenging to implement on a large scale. A single immersion of a cloth
to this, but to geometric differences (shape, dimension, roughness) wrapped produce simulator in wetting agent at the beginning of a
between potatoes and spheres. This is because for geometrically trial may be a viable alternative to the replication evaporative ef-
similar objects such correlations are only dependant on the ther- fects during the course of produce precooling.
mophysical and flow properties of the fluid (Alvarez et al., 2003). Le Page et al. (2009) used wetted plaster casts in their work
This was demonstrated by Valente et al. (1996) who measured developing CFD models of stacked food products. These casts
similar heat transfer coefficients for both real avocados and avo- maintained a constant surface water activity of 1.0 and were used to
cado shaped aluminium objects with substantially different ther- provide a model of water transport through food and subsequent
mal properties to produce. surface evaporation. Although the thermal properties of the casts
Similar experiments were carried out by Alvarez and Flick were not detailed, similar systems may be useful as produce
(1999b) to establish heat transfer correlations for approximately simulators.
spherical produce stacked in bins. PVC balls were used as produce The studies of Alvarez and Trystram (1995), Chuntranuluck et al.,
simulators along with aluminium balls and cylinders equipped 1998b and Le Page et al. (2009) all simulated evaporative effects of a
with thermocouples. PVC balls were also employed by Alvarez et al. fully wetted surface. In reality, all produce will vary somewhere
86 G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93

between the two extremes of a continuously fully wetted surface constant over the period of precooling.
and one where there is no surface evaporation and hence such ASHRAE (2006) reports values of the transpiration mass transfer
simulation is likely to result in greater evaporative losses than that coefficient (kt) for a variety of produce per unit mass. For example,
which would be observed for real produce. For some produce, apples are reported as having kt values ranging from 35 to
evaporation may initially be more significant at the beginning of 58  1012 kg/kg s Pa. It should be noted that kt is also usefully
the pre cooling process but become less significant as the surface expressed per unit area for some purposes (see Equation (5) for
moisture loss outstrips mass transport of water to the produce example). Using an average apple mass of 0.165 kg (Tabatabaeefar
surface. This effect has been demonstrated by Ansari et al. (1984) and Rajabipour, 2005) which would have a surface area of
for the cooling of apples and potatoes. In addition, as the product approximately 0.0169 m2 (Clayton et al., 1995), the range listed for
temperature decreases, the partial pressure difference between the apples above can be expressed per unit area as, 34 to 57  1011 kg/
product and airstream also reduces rapidly, meaning that the m2 s Pa. This range is included by the range reported by Maguire
importance of evaporative losses becomes less as cooling pro- et al. (2001) of 1e261  1011 kg/m2 s Pa based on a literature
gresses. Evaporative heat losses are also less likely to be important survey of values reported for 24 different varieties of apple. The
in products that have natural water loss resistance (Chuntranuluck width of this range highlights that even within the same type of
et al., 1998a). For example, Chuntranuluck et al. (1998c) showed produce there is large variability and that values of kt are specific to
that evaporative heat losses resulted in peeled carrots cooling the circumstances under which they were measured and should be
significantly faster than unpeeled carrots whose modest skin used with caution between systems (Maguire et al., 2001). There-
resistance was sufficient for them to behave essentially as a non- fore, such values should only be considered appropriate for per-
evaporative surface. forming preliminary design calculations. For this reason, the
A humidified air stream also reduces the driving force for transpiration mass transfer coefficient is more typically estimated
evaporation and hence reduces evaporative heat losses. Horticul- by assuming it to consist of an air film mass transfer coefficient and
tural cooling and storage processes are typically carried out at a produce skin mass transfer coefficient,
relative humidity levels of >85% (Tanner et al., 2002c). Although
these humid environments minimise moisture losses from pro- 1
kt ¼ 1
(8)
duce, the high latent heat of vaporisation of water dictates that very kair
þ k1
skin
little moisture loss is required to make a significant contribution to
precooling heat losses. Ferrua and Singh (2009c) reported that Modelling kt in this way allows the effects of air flow and skin
moisture losses of approximately 0.5% from strawberries contrib- resistance to be separately accounted for when estimating kt, and
uted 15e26% of the total heat lost during a 2 h precooling period. various methods of estimation including calculation from the heat
For small fractional losses in moisture and assuming the latent transfer coefficient using the classical Lewis relation (Lewis, 1922)
heat of vaporisation of water is constant over the temperature drop are discussed elsewhere for produce systems by Becker et al.
associated with precooling, the percentage of heat removed from (1996), Valente et al. (1996), Chuntranuluck et al. (1998a),
produce can be expressed as a percentage of the heat load as, ASHRAE (2006), Ferrua and Singh (2009a), Le Page et al. (2009),
and Dehghannya et al. (2010).
M DHvap It should also be noted that there is potential for moisture loss to
V¼ (6) alter the thermal properties of produce during precooling. How-
cp DT
ever, given the typically small moisture losses associated with
Over the duration of precooling typically used for common precooling, these properties are usually assumed constant and
agricultural products the loss in mass is less than 0.1% (Tanner et al., unaffected by moisture loss (Tanner et al., 2002b; Ferrua and Singh,
2002a). At a level of 0.1% moisture loss (M) Equation (6) can be used 2009c). Furthermore, moisture losses of a magnitude that would
to estimate the contribution of evaporative heat losses as approx- significantly affect thermal properties would likely result in prod-
imately 2e7%. The lower and upper end of this range were calcu- uct loss through, for instance, shrivel development (Maguire et al.,
lated at DHvap ¼ 2400  103 J/kg, cp ¼ 4040 J/kg K, DT ¼ 30  C, and 2001; Aswaney, 2007) or firmness degradation (Paniagua et al.,
DHvap ¼ 2500  103 J/kg, cp ¼ 3670 J/kg K, DT ¼ 10  C respectively, 2013).
and provide a low and high estimate of variability in these pa- Ultimately, whether or not evaporation plays a significant role
rameters. The range in heat capacity comes from the observed will be determined by the particular precooling system. However, it
range of produce reported in Table 1. The range of 2e7% shows that is likely that for many precooling processes the effects of evapo-
it is unlikely for typical precooling that evaporation represents a rative cooling can be omitted from produce simulators without a
significant portion of the heat removed, however, the significance major effect on the observed heat transfer. For many applications
of evaporative heat losses increases rapidly as moisture loss in- the effect of evaporation on temperature profiles and/or the time
creases, with a percentage evaporative contribution ranging from required to achieve a certain degree of cooling is of most impor-
10 to 34% at 0.5% moisture loss. This range encompasses the 26% tance and may be small in some cases even where the contribution
contribution reported by Ferrua and Singh (2009c) at a similar relative to the overall heat load seems significant. Once again,
moisture loss level. design guidelines in this regard will be very useful and can likely be
The percentage moisture loss from the product can be difficult established using computational modelling.
to determine experimentally as it can be confounded by water
absorption by corrugated fibreboard packaging and surface 5. Addition of respiratory heat
condensation once cold. Alternatively, it can be estimated from the
rate of moisture loss taken over the duration of precooling as, None of the studies reported in Table 2 attempted to incorporate
  the addition of heat of respiration into the produce simulators
M ¼ 100kt Psurf  Pair Dt (7) employed. Indeed, the assumption that heat added via produce
respiration is a negligible portion of the heat load is commonly
Estimation of the percentage evaporative contribution via made (Tanner et al., 2002a; Ferrua and Singh, 2009a; Dehghannya
Equations (6) and (7) will produce a conservatively high estimate of et al., 2010). Tanner et al. (2002a) suggest that in a typical pre-
the upper bound of this value since moisture loss is assumed cooling process the heat added by respiration is ~0.5% of the heat
G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93 87

load. 
Assuming the respiratory heat generation is constant, the fz0 exp E
RT
approximate heat added by respiration over the period of pre- q¼ (11)
cooling can also be expressed as a percentage of the heat load, T

where f is a conversion factor equal to 1.59 (J hr K/s mL). Values for


the pre-exponential factor and activation energies can be readily
calculated from measured carbon dioxide production rates at
100qDt various temperatures. Gross et al. (2004) provides a source of such
Q¼ (9)
cp DT measured data for fruit and vegetables. Using data from a variety of
literature sources, Exama et al. (1993) calculated the pre-
Table 1 gives values of the heat of respiration for various pro-
exponential factor and activation energies for different produce,
duce at temperatures approximately mid-range to the typical pre-
some of which are displayed in Table 1.
cooling process. Using these values, a minimum and maximum
Other approaches to calculating the rate of respiratory heat
range for Q can be calculated as varying from 0.03 to 22%. Where
generation include use of the empirical relationship reported by
the lower and upper values were calculated at q ¼ 20  103 W/kg,
Gaffney et al. (1985),
Dt ¼ 0.5 h, cp ¼ 4040 J/kg K, DT ¼ 30  C, and q ¼ 183  103 W/kg,
Dt ¼ 12 h, cp ¼ 3670 J/kg K, DT ¼ 10  C respectively. The range in
q ¼ a(T255.35)b (12)
heat capacity again comes from the observed range of produce
reported in Table 1.
This relationship has a high correlation coefficient and so is
The calculated range is very wide and whilst this range includes
widely used to calculate the heat of respiration of food products
the value of 0.5% suggested by Tanner et al. (2002a) as being typical
(Tanner et al., 2002b). The product specific constants, a and b, can
of produce cooling, it also extends well beyond this. The width of
be determined experimentally or estimated from existing data. For
this range is due almost entirely to variation in the duration of
example, Tanner et al. (2002b) obtained these values via regression
precooling (the upper value of this range becomes 0.90% if 0.5 h is
of data obtained from Hardenburg et al. (1986).
used instead of 12 h), showing that the assumption of a negligible
contribution of respiratory heat made by most studies is much
more likely to be valid over shorter durations of precooling, but 6. Thermal property matching
likely to be untrue for lengthy precooling processes. This will be
more likely in areas of maldistribution of flow which results in slow 6.1. Single-domain produce simulators
cooling. In these cases, the contribution of evaporative cooling
would also be more significant. Single domain produce simulators are constructed of one ma-
However, as with evaporative losses, because the interest in terial, or if constructed of multiple materials, only one part has
produce simulators is primarily to replicate temperature profiles significant resistance to heat transfer and thermal mass. These
during precooling, the effect of respiration on temperature needs to simulators can be employed to replicate heat transfer in produce
be quantified and design guidelines established. Using an analytical whose heat transfer properties are sufficiently homogenous
treatment, Awbery (1927) showed that an apple placed in a throughout.
refrigerator had only 0.02  C excess temperature at its centre as a In order to replicate internal heat transfer, a thermally matched
result of respiration after 2 h of cooling. Similar research utilising produce simulator will sufficiently match the thermal conductivity,
computational models should be able to produce appropriate thermal diffusivity and volumetric heat capacity of the produce
design guidelines demonstrating the impact of product respiration being simulated. These thermal properties are related by,
on temperature at times scales of interest to produce precooling.
Because evaporation and respiration represent heat loss and k ¼ a(rc) (13)
addition respectively, it is possible that their net effect may result in
an overall negligible effect on heat transfer in some applications. and hence matching any two of these properties will also match the
The constant values of q used in calculating the range above will third.
also tend to provide a conservative overestimate of the contribution The range reported in Table 1 shows that the bulk thermal
of respiration, since respiratory heat generation reduces with properties of produce vary little across different types of produce.
temperature. This effect can be incorporated into dynamic models The mid-range values of k ¼ 0.49 W/mK, a ¼ 0.13  106 m2/s and
using Arrhenius type temperature dependence which studies have (rc) ¼ 3625 kJ/m3 K are considered here to be representative of
shown describes the relationship between produce respiration and typical produce.
temperature well (Bhande et al., 2008; Benítez et al., 2012; A number of studies have attempted to use single domain pro-
Waghmare et al., 2013, Azevedo et al., 2015), where the produc- duce simulators which sufficiently replicate the thermal properties
tion of carbon dioxide can be described in the form, of produce. Alvarez and Trystram (1995) used spheres composed of
3% carrageenan gum solution to simulate approximately spherical
produce cooling in bins. Gel spheres were prepared by injecting the
solution into aluminium moulds which were subsequently cooled
 in ice water for gel formation and returned to ambient conditions
E
z ¼ z0 exp (10) before experimental runs. The heat transfer resistance and thermal
RT
mass of the aluminium mould can be considered negligible and
Carbon dioxide production rates can be converted to rates of hence these produce simulators are considered single-domain. The
respiratory heat generation based on the amount of energy thermal diffusivity of the gel was about 1.5 times greater than that
released during respiration. Using a value of 470 kJ/mol of respi- of typical produce, whilst the volumetric heat capacity was very
ratory carbon dioxide produced (Gross et al., 2004), the ideal gas similar to that of produce (Table 2). Since these produce simulators
law, and Equation (10), the rate of respiratory heat generation can were spherical and do not incorporate respiration or evaporation,
be calculated at a given temperature as, the cooling profile can be analytically represented by (Incropera
88 G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93

and Dewitt, 2006), gel spheres. These were chosen to give similar aerodynamic
behaviour. Although they would achieve this, their different ther-
X
∞   1   mal properties will affect heat transfer within the system such that
sin dn r *
2
Y¼ Cn exp dn Fo *
(14) the temperature profiles would be expected to be different to what
n¼1
d n r
would have been observed had all spheres been composed of gel.
where, The magnitude of any such discrepancies is a worthy subject of
further investigation and if not too large may permit large scale
4½sinðdn Þ  dn cosðdn Þ experiments with a few thermally matched produce simulators
Cn ¼ (15) amongst a bulk of aerodynamic placeholder simulators.
2dn  sinð2dn Þ
Chuntranuluck et al., 1998b devised an extremely comprehen-
And the values of dn are the positive roots of the equation, sive produce simulator for use in validating their chilling time
1dncot(dn) ¼ Bi of which the 1st root is sufficient to describe heat prediction model. Intended to be a general food analogue, simu-
transfer at longer time intervals (Fo > 0.2). lators were constructed of various size aluminium cylinders filled
Using Equation (14) it can be calculated that, at a convective with Tylose. Aluminium was chosen to confine Tylose to the desired
heat transfer coefficient value of 27 W/m2K (mid-range of the shape without introducing significant heat transfer resistance or
values reported by ASHRAE, 2006 for cooling of a wide variety of thermal capacity. Table 2 indicates that the Tylose simulator of
produce), the spheres used by Alvarez and Trystram (1995) would Chuntranuluck et al., 1998b is very closely thermally matched to
take 90 min to cool to 87.5% (7/8th) of the overall temperature typical produce. This suggests that Tylose based simulators with
change at the centre when cooled from 25 to 0  C. If the spheres housings which match the shape and size of specified produce
were thermally matched to typical produce (k ¼ 0.49 W/m K, would represent a very good produce simulator from a technical
(rc) ¼ 3625 KJ/m3K) this time would be 104 min. Hence the pro- perspective, provided the resistance to transfer of the housing was
duce simulator predicts cooling to be approximately 14 min too fast negligible or well matched to the produce. The practicality of such a
as a result of thermal mismatching (this discrepancy will vary at produce simulator would be impacted by its stability. For example,
different convective heat transfer coefficients). Whether or not this if frequent cleaning and refilling were required, regular use on a
is significant will depend on the intended application, which in the large scale may not be feasible. In addition, by their nature any
case Alvarez and Trystram (1995) was to demonstrate a design water based produce simulator is a candidate for shelf stability is-
strategy for controlling forced air cooling of produce. Because the sues associated with microbial growth.
produce simulators were not thermally matched, absolute tem-
peratures and cooling rates should at least be extrapolated to other 6.2. Multi-domain produce simulators
systems with this in mind.
Table 2 shows the values reported by Alvarez and Trystram Accurate representation of some produce may require it to be
(1995) and Allais et al. (2006) for the thermal conductivity and split into multiple subdomains, such as skin, flesh and core. This
diffusivity of 3% carraghenan gel differ, indicating that these values will be required for any region in which the thermophysical prop-
may need to be more accurately quantified. The value reported by erties differ significantly enough from adjacent regions. If such
Allais et al. (2006) of 0.52 W/m K is much closer to that of typical subdomain differentiation is required, then a separate set of
produce than the value of 0.86 W/m K reported by Alvarez and transfer equations can be written for each subdomain (Equation (4)
Trystram (1995). If this value is instead used in the calculation of with new subdomain properties plus internal boundary conditions)
the time to cool to the centre to 87.5% of the overall temperature which in turn dictates that a produce simulator will need to
change when cooled from 25 to 0  C, this time increases to 109 min replicate the geometry and thermophysical properties of each
and the cooling is now 5 min too slow as a result of thermal mis- subdomain if the internal temperature profiles of produce are to be
matching. Reducing the thermal conductivity even further so that it replicated.
matches typical produce at 0.49 W/m K actually makes the match It appears that multi domain produce simulators in the litera-
worse by increasing the cooling time to 112 min. This is because the ture have not necessarily been deliberate attempts to mimic
reduction in thermal conductivity acts to slow heat transfer, which, structural components of produce. Rather they have been restricted
in this case is the same effect as increasing the thermal mass of the to plastic housings employed as a convenient container for another
produce simulator which is larger than that of typical produce. In water based material. Whether deliberate or not, such produce
the case of the simulators used by Alvarez and Trystram (1995), a simulators may be a convenient way of replicating heat transfer in
produce simulator with thermal conductivity of 0.58 W/mK leads produce where the properties of the skin differ significantly from
to the time required to achieve 87.5% temperature change the bulk produce.
becoming equal to that of typical produce. If the time required to Allais et al. (2006) used simulators composed of 3% carraghenan
achieve a specified degree of temperature change at a specific gel housed in celluloid spheres to validate their mathematical
location is a suitable measure of cooling (as opposed to matching model of stacked spheres during mist chilling (a variation of forced
the entire temperature profile), then this raises the possibility of a air cooling where the air stream contains suspended water drop-
produce simulator with thermal conductivity and volumetric heat lets). Although the carraghenan gel provides a reasonable thermal
capacity combinations which are either both higher or both lower match with produce, unlike aluminium, the resistance and thermal
than produce. However, such a simulator would only achieve this mass of the housing is unlikely to be negligible in this case. The
match at a single flow condition. In the example above, the produce effectiveness of the produce simulator will depend on whether the
simulator with thermal conductivity of 0.58 W/m K is matched at housing sufficiently matches the bulk produce, or the skin of the
h ¼ 27 W/m K and becomes 3 min too slow at h ¼ 20 W/m K and produce if this is sufficiently different from the bulk. Sufficient
4 min too slow at h ¼ 15 W/m K. The sensitivity of the time to replication of the thermal properties of the skin of produce may
achieve a certain fraction of temperature change to changes in have played a role in the work of Vissotto et al. (2000) who
thermal properties is further discussed in Section 7 below. employed PVC balls filled with 3% carrageenan gum as produce
It is also important to note that although the study of Alvarez simulators and found that the temperature profiles produced dur-
and Trystram (1995) employed gel spheres, the majority of ing precooling compared well to those of oranges. This is despite
spheres in the study were air filled PVC balls of the same size as the some degree of thermal mismatching in carrageenan gum based
G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93 89

systems as discussed above. In this case the PVC housings may have 6.3. Materials for thermal property matching
improved the match by simulating the skin of oranges which have a
low thermal conductivity of 0.18 W/m K (ASHRAE, 2006), which is The studies discussed here which achieved reasonable thermal
very similar to that of PVC (Table 3). Vissotto et al. (2000) also matches between bulk produce properties and simulator used
attempted to produce set gel based simulators from carrageenan water based materials (Alvarez and Trystram, 1995; Chuntranuluck
gum without PVC housings, but experienced issues associated with et al., 1998b; Castro et al., 2004; Allais et al., 2006). This is because
movement after stacking, evaporation and fungal growth. full thermal matching is constrained by the requirement of
Castro et al. (2004) used a similar experimental setup to Alvarez matching volumetric heat capacity. The high water content of
and Trystram (1995) with the objective of examining which pa- produce dictates that its volumetric heat capacity is similar to that
rameters effect the cooling rate and uniformity of stacked horti- of water. As very few materials have volumetric heat capacities as
cultural produce. 28 PVC balls filled with 3% agar-agar solution high or higher than water, produce simulators have naturally been
were distributed throughout a stacking of 224 hollow PVC balls. water based.
Whilst the thermal properties of 3% agar are similar to those of The difficulties associated with the preparation and stability of
water (Denys et al., 2000) and hence produce (see Table 1), the water based produce simulators, make produce simulators con-
authors concluded that the presence of air filled balls would affect structed of alternative materials a desirable proposition. However,
heat transfer and dictate that their findings should be used care- it is unlikely that a pre-existing material will be initially well
fully. Practical difficulties were also encountered, with the authors matched to the desired produce. Composite materials offer a po-
attributing significant experimental errors to difficulties in uni- tential way to construct a desired simulator material and overcome
formly filling PVC balls with agar solution and recommended that this issue.
the accuracy and reproducibility of such studies could be improved A simple two component composite material is constructed by
through the design of more stable produce simulators. combining a base matrix material with a filler material. The volu-
Due to the potential increased complexity of creating multi metric heat capacity of a two component composite can be
domain produce simulators, it should be verified that the ther- described by the law of mixtures,
mophysical properties are sufficiently variable to justify this. In
many cases, despite stark differences in the visual or structural ðrcÞc ¼ fðrcÞf þ ð1  fÞðrcÞm (16)
properties of regions of produce, the thermal properties do not
differ drastically. An example comes from a comparison of the pulp The law of mixtures is generally insufficient to predict the
and stone of avocado. These structures differ in composition and thermal conductivity of composite materials. Many empirical and
also have obvious visual and physical differences. However, data experimental models describing the effective thermal conductivity
reported by Valente et al. (1996) shows the thermal conductivity of of composite systems have been reported in the literature. The
the pulp and stone to be 0.43 and 0.45 W/m K respectively, with applicability and effectiveness of such models have been sum-
volumetric heat capacitates of 3380 kJ/m3 K and 3240 kJ/m3 K marised by Progelhof et al. (1976) and Kumlutas et al. (2003). The
respectively. Hence, despite stark differences these two structures field continues to develop and utilise modern modelling techniques
appear to be closely thermally matched. The difference in assuming such as neural networks (Bhoopal et al., 2012).
a central stone within such an object had similar properties to the An example of such relationships is the classical model of
pulp may be negligible for the purposes of most applications. Maxwell (Maxwell, 1873),
Another example comes from the thermal diffusivity of kiwifruit
flesh and core which Chen et al. (1998) found to be similar at 2  3
kf þ 2km þ 2f kf  km
1.14  107 and 0.92  107 m2/s respectively. The water contents kc ¼ km 4   5 (17)
of these structures are also similar at approximately 85% and 75% kf þ 2km  f kf  km
respectively (MacCrae et al., 1989), and hence their volumetric heat
capacities would also be expected to be similar. This would suggest Composite design can be achieved by solution of these equa-
an overall thermal similarity, and whether or not these similarities tions. At any given filler fraction, for a produce simulator of speci-
are sufficient to justify a single domain produce simulator would be fied desired thermal properties and a base matrix material of
worth considering before pursuing a more complex alternative. known thermal properties, Equation (16) can be used to solve for
Indeed, reasonable thermal matching would be expected wherever the required volumetric heat capacity of a filler material and
different structures have similar water contents which will dictate Equation (17) for the required thermal conductivity of the filler (in
to a large degree the thermophysical properties. The thermophys- the case of Maxwell’s equation this requires an iterative solution).
ical properties of water at approximately 7  C are provided in In practice, a thermal conductivity model that is appropriate to
Table 1 for comparison, and show the expected similarity between the particular design scenario being investigated should be sought
water and produce. and its limits of applicability should be understood. For example,
Whether single or multi-domain, the sensitivity of produce Maxwell’s model is able to predict the effective thermal conduc-
simulators to thermal mismatching in pre-cooling is an important tivity of composites well at lower filler concentrations (Ebadi-
topic for future research which, once again will be well suited to Dehaghani and Nazempour, 2012; Kumlutas et al., 2003) and as-
computational modelling. These aspects are discussed further in sumes that the filler particles are randomly distributed and non-
Section 7 when various simplifying assumptions are permitted. interacting homogenous spheres in a homogeneous medium

Table 3
Thermophysical properties of common thermoplastics.

Type of thermoplastic k (W/m K) r (kg/m3) c (J/kg K) rc (103 J/m3 K) Source

High density polyethylene (HDPE) 0.543 (at 36  C) 950 1800e2700 1710e2565 Kumlutas et al. (2003)
Low density polyethylene (LDPE) 0.31 (at 25  C) 918 1900e2300 1744e2111 Luyt et al. (2006)
Polypropylene (PP) 0.11 (at 25  C) 908 1920 1743 Ebadi-Dehaghani and Nazempour (2012)
Polyvinyl Chloride (PVC) 0.19 (at 25  C) 1400 840e1170 1176e1638 Ebadi-Dehaghani and Nazempour (2012)
90 G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93

(Kumlutas et al., 2003). Fig. 4 shows the sensitivity of the duration required to achieve
Thermoplastics are an attractive proposition for use as a com- 87.5% centre cooling of typical produce to variation in some pa-
posite matrix material due to their ability to be moulded into shape rameters when cooled from 25 to 0  C. Calculations were carried
and recast into different shapes if required. Furthermore, the out using Equation (14) with parameters values held constant at
modification of thermoplastic properties through the creation of h ¼ 27 W/m2 K, k ¼ 0.49 W/m K, (rc) ¼ 3625 kJ/m3 K, and Bi ¼ 1.82,
composites is common. This includes research into the modifica- except when they were varied as the dependant variable.
tion of thermal conductivity, especially by combining with metals Although the absolute magnitude of these predictions are
(Mamunya et al., 2002; Kuriger and Alam, 2002; Kumlutas et al., technically applicable to spherical geometry, the observed re-
2003; Weidenfeller et al., 2004; Boudenne et al., 2005; Luyt et al., lationships offer useful information for produce simulator design.
2006). Little or no research has investigated modification of the Fig. 4a shows the cooling time is highly sensitive to variation in the
volumetric heat capacity via the creation of thermoplastic convective heat transfer coefficient, increasingly so as flow condi-
composites. tions stagnate at low values. This serves to highlight the importance
The thermal conductivity of thermoplastics is similar to that of of replicating flow through geometric matching in produce simu-
produce, while the volumetric heat capacity is typically less than lator design, but also the importance of the challenge posed in
half (Table 3). According to the design equations above, a produce ventilated packaging design trying to achieve homogenous flow to
simulator built on a thermoplastic matrix would require a filler all parts of the package in order to achieve cooling uniformity.
material with thermal mass higher than water but with comparable Fig. 4b shows cooling time rapidly becomes highly sensitive to
thermal conductivity. The rarity of this type of material may limit variation in thermal conductivity at values less than that of typical
the feasibility of thermoplastic based produce simulators. Ther- produce (~0.5 W/m K), but becomes relatively insensitive as ther-
moplastic produce simulators may be better suited to Biot matched mal conductivity increases beyond this level. The opposing effects
simulators, as discussed in Section 7. of thermal conductivity and volumetric heat capacity are clearly
Other potential matrix materials for the creation of composite evident through comparison of Fig. 4b and c, with increasing
produce simulators are those with high volumetric heat capacity. thermal conductivity reducing cooling time and increasing volu-
These include stainless steel and ceramic materials. Some ceramic metric heat capacity increasing it. These figures do suggest that
materials have volumetric heat capacities comparable to and there is some scope for thermal mismatching with small deviations
greater than water/produce (Warren, 1992) and thermal conduc- in thermal conductivity and volumetric heat capacity in the vicinity
tivity is typically much higher than that of water based materials of produce not leading to large deviations in cooling times. The
(Carter and Norton, 2013). The same is true of stainless steel, hence linear sensitivity to volumetric heat capacity is described by,
both ceramics and stainless steel could in theory be used to create a
composite produce simulator by adding a low thermal mass, low FoðrcÞro2
thermal conductivity material. The greater the thermal conduc- t¼ (18)
k
tivity of the initial matrix above that of produce, the greater the
fraction of low thermal conductivity filler required to reduce the where t is the time required to achieve a specified degree of
thermal conductivity to the levels of produce. The more filler completion (Y) and Fo is the dimensionless time at the same degree
required, the closer its volumetric heat capacity will have to be to of completion (Fo is constant when (rc) is varied). Hence for any
that of produce. Hence, in these cases a low thermal conductivity system where the time required to achieve a certain degree of
material with thermal mass comparable to water/produce may be completion is known (either by prediction or experimentally), the
required, and once again the rarity of this type of material may be time required to achieve the same degree of completion as a result
problematic. of variation in volumetric heat capacity can be calculated as,
However, even if ideal filler materials do not exist, such com-
posites may be worth investigating, especially if the sensitivity of ðrcÞ2
produce simulator temperature replication permits a certain de- t2 ¼ t1 (19)
ðrcÞ1
gree of property mismatching. Complex composites composed of
more than two materials may also provide potential solutions. This relationship can be utilised to quickly determine the effects
It is finally interesting to note the potential of using produce of mismatching of the volumetric heat capacity of a produce
itself as a produce simulator material. Potatoes, for example have simulator on the cooling time.
similar thermal properties to other produce (Table 1) and might be Under the assumptions of negligible evaporation, respiration
useful to simulate different produce which is less available and/or and a single homogenous domain, further simplification of produce
more expensive. simulator requirements may be permitted at extremes of the ratio
of internal to external heat transfer resistances in horticultural
7. Simplifying cases produce. This ratio is described by the Biot number,

Effective and convenient produce simulators which replicate the Bi ¼ hL/k (20)
effects of evaporation, respiration and include multiple domains of
varying thermal properties will likely be difficult to achieve from a and can be used to divide the heat transfer processes involved into
practical perspective. The expense of such simulators may also limit three categories (Vigneault and Castro, 2005), Bi ≪ 1, Bi [1 and
their application, especially for research on a pallet scale. Bi z 1.
The complexity of produce simulator design reduces consider- When Bi ≪ 1, convection limits heat transfer. In this case,
ably if evaporation and respiration can be assumed negligible, and assuming geometry and airflow are mimicked by the produce
the produce thermal properties can be considered sufficiently ho- simulator, a temperature profile which matches that of produce can
mogenous to justify a single-domain produce simulator. Although be achieved using any thermal conductivity, provided the volu-
further investigation is required, the information discussed in this metric heat capacity is matched. Hence, only a partial thermal
review suggests that these assumptions may be valid for some match is required.
produce. This section considers the implications for simulator When Bi [ 1 conduction limits heat transfer. In this case a
design when these assumptions can be considered valid. temperature profile which matches that of produce can be achieved
G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93 91

Fig. 4. Sensitivity of the duration required to achieve a 87.5% cooling of typical produce to variation in some parameters. Calculations made utilising Equation (14). Parameters held
constant at h ¼ 27 W/m2K, k ¼ 0.49 W/mK, (rc) ¼ 3625 KJ/m3K, and Bi ¼ 1.82 when not the dependant variable and cooling is from 25 to 0  C. Black circles indicate the location of
typical produce.

by only matching thermal diffusivity, provided the Bi numbers are Y ¼ f(dimensionless position, Fo, Bi) (21)
equal. However, if flow and geometry are also to be matched, then
Bi equality can only be achieved through matched thermal con- where Y is the fraction of unaccomplished change in temperature,
ductivity. Hence, in order to match the thermal diffusivity, the
volumetric heat capacity would also need to be matched. The net T  T∞
Y¼ (22)
result is that the requirements of matching flow and geometry force Ti  T∞
this scenario back to one where full property matching is required,
which offers no simplification for the purposes of produce simu-
lator design.
Using Equation (20), Fig. 5 plots the Bi number for each produce
item in Table 1 as a function of convective heat transfer coefficient.
Where a range of L and/or k have been given in Table 1, mid-range
values have been used in the calculations of Fig. 5. The Bi lines of all
produce in Fig. 5 cross the line Bi ¼ 1 (dotted line in Fig. 5) during
the convective heat transfer coefficient range of 11e42 W/m K re-
ported by ASHRAE (2006) for a variety of produce being cooled by
forced air chilling, which demonstrates that the case of Bi z 1 is
likely to apply to typical produce systems. Table 1 also gives a Bi
range for each individual produce item calculated across the heat
transfer coefficient range of 11e42. For blueberries this range is
from 0.04 to 0.62 (low range value is calculated at low h, L and high
k and vice versa for the high range value, thus creating a conser-
vatively wide range). This range also falls into the category of Bi z 1,
except for blueberries at the extreme low end of the range.
Therefore, the simplifying case of Bi ≪ 1 will not be valid to apply in
produce simulator systems. Rather, the dimensionless temperature Fig. 5. Biot numbers of various produce as a function of convective heat transfer co-
will be a function of dimensionless position, dimensionless time efficient. Values for L and k are taken from Table 1. Where a range is given in Table 1 the
and Biot number. That is, mid range value has been used. 1The lines for potatoes and kiwifruit are visually
indistinguishable. The dotted line is that of Bi ¼ 1.
92 G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93

and dimensionless time is defined by the Fourier number as, Acknowledgements

∝t This work is an output of the Ministry of Business, Employment


Fo ¼ (23)
L2 and Innovation funded Fibreboard Packaging Design research
project (MAUX 1302).
In this case, if a system composed of produce simulators is of the
same scale and geometrically similar, then for a given flow field it
will produce the same dimensionless temperature profile as the References
system it is simulating provided the thermal conductivities are
matched. This is because the Biot numbers will be matched. This Alibeyglu, S.Z., Ghaffari, H., Alipasandi, A., 2013. Estimating surface area of three
allows the possibility of utilising produce simulators which only peach varieties using allometric relationships obtain from project area. Int. J.
Agron. Plant Prod. 4 (8), 1978e1984.
match the thermal conductivity of the produce being simulated to Allais, I., Alvarez, G., Flick, D., 2006. Modelling cooling kinetics of a stack of spheres
carry out experiments, with post processing reconstruction of the during mist chilling. J. Food Eng. 72 (2), 197e209.
absolute temperature profile. Table 3 suggests high density poly- Alvarez, G., Flick, D., 1999a. Analysis of heterogeneous cooling of agricultural
products inside bins Part I: aerodynamic study. J. Food Eng. 39 (3), 227e237.
ethylene (HDPE) would be well suited to this purpose given its Alvarez, G., Flick, D., 1999b. Analysis of heterogeneous cooling of agricultural
thermal conductivity is similar to produce and it can be moulded products inside bins Part II: thermal study. J. Food Eng. 39 (3), 239e245.
into any shape desired. Fig. 4d also shows the sensitivity of the Alvarez, G., Trystram, G., 1995. Design of a new strategy for the control of the
refrigeration process: fruit and vegetables conditioned in a pallet. Food Control.
dimensionless time required to achieve 87.5% temperature change 6 (6), 347e355.
to variation in the Biot number. At Biot numbers greater than ~2 Alvarez, G., Bournet, P., Flick, D., 2003. Two-dimensional simulation of turbulent
this dimensionless time is relatively insensitive to variations in the flow and transfer through stacked spheres. Int. J. Heat Mass Transf. 46,
2459e2469.
Biot number, indicating that there may be some scope for Biot
Ambaw, A., Delele, M.A., Defraeye, T., Ho, Q.T., Opara, L.U., Nicolaï, B.M., Verboven, P.,
mismatching when operating at Biot numbers around and beyond 2013. The use of CFD to characterize and design post-harvest storage facilities:
this level. past, present and future. Comput. Electron. Agric. 93, 184e194.
Furthermore, for these systems, heat transfer theory can be used Ansari, F.A., Charan, V., Varma, H.K., 1984. Heat and mass transfer analysis in air-
cooling of spherical food products. Int. J. Refrig. 7 (3), 194e197.
to show that, given geometric similitude, two systems of different ASHRAE, 2006. Thermal properties of foods. In: ASHRAE Handbook e Refrigeration
characteristic length can also produce the same dimensionless (SI). American Society of Heating, Refrigerating and Air Conditioning Engineers.
temperature profiles. For example, this will be achieved in a given Aswaney, M., 2007. Forced-air precooling of fruits and vegetables. Air Cond. Refrig. J.
57e62. January-March.
system if characteristic length is scaled by a factor of s, fluid velocity Awbery, J.H., 1927. The flow of heat in a body generating heat. Lond. Edinb. Dublin
is scaled by a factor of 1/s and the thermal conductivity is matched. Philosophical Mag. J. Sci. 4 (22), 629e638.
This allows for the possibility of carrying out experiments on a Azevedo, S., Cunha, L.M., Fonseca, S.C., 2015. Modelling the influence of time and
temperature on the respiration rate of fresh oyster mushrooms. Food Sci.
smaller scale and only partially thermally matching produce, which Technol. Int. 21 (8), 593e603.
may offer further practical advantages. Becker, B.R., Misra, A., Fricke, B.A., 1996. Bulk refrigeration of fruits and vegetables,
Part I: theoretical considerations of heat and mass transfer. Int. J. HVAC R Res. 2
(2), 122e134.
Benítez, S., Chiumenti, M., Sepulcre, F., Achaerandio, I., Pujola , M., 2012. Modeling
the effect of storage temperature on the respiration rate and texture of fresh cut
8. Conclusions pineapple. J. Food Eng. 113 (4), 527e533.
Bhande, S.D., Ravindra, M.R., Goswami, T.K., 2008. Respiration rate of banana fruit
Produce simulators have the potential to accelerate research under aerobic conditions at different storage temperatures. J. Food Eng. 87 (1),
116e123.
progress, especially in the areas of postharvest cooling and tem- Bhoopal, R.S., Sharma, P.K., Kumar, S., Pandey, A., Beniwai, R.S., Singh, R., 2012.
perature control operations. However, a survey of the literature has Prediction of effective thermal conductivity of polymer composites using an
shown there is a lack of appropriate design guidelines for produce artificial neural network approach. Special Top. Rev. Porous Media 3 (2),
115e123.
simulators, which in some cases has resulted in inappropriate use.
Boudenne, A., Ibos, L., Fois, M., Majeste, J.C., Gehin, E., 2005. Electrical and thermal
Consideration of transport theory as it applies to forced air behaviour of polypropylene filled with copper particles. Compos. Part A 36 (1),
cooling allows design constraints to be established for produce 1545e1554.
Brosnan, T., Sun, D., 2001. Precooling techniques and applications for horticultural
simulators. Geometric matching allows the replication of flow
products e a review. Int. J. Refrig. 24 (2), 154e170.
fields. Combined with thermal property matching this allows the Carter, C.B., Norton, M.G., 2013. Conducting charge or not. In: Ceramic Composites:
replication of heat transfer in systems in which evaporative and Science and Engineering. Springer, New York.
respiratory heat losses are negligible. Future work examining the Castro, L.R., Vigneault, C., Cortex, L.A.B., 2004. Container opening design for horti-
cultural produce cooling efficiency. J. Food, Agric. Environ. 2 (1), 135e140.
extent to which geometric and thermal parameters can be Chen, X.D., McLellan, D.N., Rahman, M.S., 1998. Thermal diffusivity of kiwifruit skin,
matched/mismatched and still achieve suitable replication of the flesh and core measured by a modified fitch method. Int. J. Food Prop. 1 (2),
desired transport phenomena is recommended, as is further 113e119.
Chuntranuluck, S., Wells, C.M., Cleland, A.C., 1998a. Prediction of chilling times of
research examining the scenarios under which the simplifying as- foods in situations where evaporative cooling is significant e Part 1. Method
sumptions of negligible evaporation, respiration and produce ho- development. J. Food Eng. 37, 111e125.
mogeneity are considered appropriate. These issues can likely be Chuntranuluck, S., Wells, C.M., Cleland, A.C., 1998b. Prediction of chilling times of
foods in situations where evaporative cooling is significant e Part 2. Experi-
addressed with the assistance of modern computational modelling mental testing. J. Food Eng. 37, 127e141.
which allows aspects like the effect of variation in the geometry of Chuntranuluck, S., Wells, C.M., Cleland, A.C., 1998c. Prediction of chilling times of
produce to be established in silico, where such information might foods in situations where evaporative cooling is significant e Part 3. Applica-
tions. J. Food Eng. 37, 143e157.
be otherwise difficult to obtain experimentally. Clayton, M., Amos, N.D., Banks, N.H., Morton, R.H., 1995. Estimation of apple fruit
Practical constraints including the ability to be produced on a surface area. N. Z. J. Crop Hortic. Sci. 23 (3), 345e349.
large scale at low cost, convenient experimental setup, long life and Defraeye, T., Cronje, P., Berry, T., Opara, U., East, A., Hertog, M., et al., 2015a. Towards
integrated performance evaluation of future packaging for fresh produce in the
reusability are also important to produce simulator systems. If
cold chain. Trends Food Sci. Technol. 44, 201e225.
multiple types of different produce are to be simulated the ability to Defraeye, T., Verboven, P., Opara, U.L., Nicolai, B., Cronje, P., 2015b. Feasibility of
be reshaped and manipulate thermophysical properties may also ambient loading of citrus fruit into refrigerated containers for cooling during
be desired. Meeting these requirements represents a materials marine transport. Biosyst. Eng. 134, 20e30.
Dehghannya, J., Ngadi, M., Vigneault, C., 2010. Mathematical modelling procedures
engineering challenge to which composite materials are one po- for airflow, heat and mass transfer during forced cooling of produce: a review.
tential solution. Food Eng. Rev. 2, 227e243.
G.P. Redding et al. / Journal of Food Engineering 190 (2016) 80e93 93

Delele, M.A., Tijskens, E., Atalay, Y.T., Ho, Q.T., Ramon, H., Nicolai, B.M., et al., 2008. J. 48, 82e87.
Combined discrete element and CFD modelling of airflow through random Maxwell, J.C., 1873. A Treatise on Electricity and Magnetism. Clarendon Press,
stacking of horticultural products in vented boxes. J. Food Eng. 89, 33e41. Oxford.
Delele, M.A., Verboven, P., Ho, Q.T., Nicolaï, B.M., 2010. Advances in mathematical Milosevic, T., Milosevic, N., 2012. Fruit quality attributes of sour cherry cultivars.
modelling of postharvest refrigeration processes. Stewart Postharvest Rev. 6 (2), ISRN Agron. http://dx.doi.org/10.5402/2012/593981.
1e8. Minh, T.V., Perry, J.S., Bennett, A.H., 1969. Forced-air precooling of white potatoes in
Denys, S., Ludikhuyze, L.R., Van Loey, A.M., Hendrickx, M.E., 2000. Modeling bulk. ASHRAE Trans. 75 (2), 143e152.
conductive heat transfer and process uniformity during batch high-pressure Mirade, P.S., Kondjoyan, A., Daudin, J.D., 2002. Three-dimensional CFD calculations
processing of foods. Biotechnol. Prog. 16 (1), 92e101. for designing large food chillers. Comput. Electron. Agric. 34, 67e88.
Ebadi-Dehaghani, H., Nazempour, M., 2012. Thermal conductivity of nanoparticle Olatunji, J., Love, R.J., Shim, Y.M., Ferrua, M.J., East, A.R., 2015. Numerical determi-
filled polymers. In: Hashim, A. (Ed.), Smart Nanoparticles Technology. Intech. nation of kiwifruit shape, volume and surface area using ellipsoid approxima-
Eifert, J.D., Sanglay, G.C., Lee, D., Sumner, S.S., Pierson, M.D., 2006. Prediction of raw tion and outside diameter function. In: Proceedings of the International
produce surface area from weight measurement. J. Food Eng. 74 (4), 552e556. Congress of Engineering and Food, Quebec City, Canada, June 14th- 18th.
Exama, A., Arul, J., Lencki, R.W., Lee, L.Z., Toupin, C., 1993. Suitability of plastic films Oliveira, J.M., Lessio, B.C., Morgante, C.M., Santos, M.M., Augosto, P.E.D., 2012. Spe-
for modified atmosphere packaging of fruits and vegetables. J. Food Sci. 58 (6), cific heats of tropical fruits: caja, cashew apple, cocoa, kiwi, pitanga, soursop
1365e1370. fruit and yellow melon. Int. Food Res. J. 19 (3), 811e814.
Ferrua, M.J., Singh, R.P., 2008. A noninstrusive flow measurement technique to O’Sullivan, J., Ferrua, M., Love, R., Verboven, P., Nicolai, B., East, A., 2014. Airflow
validate the simulated laminar fluid flow in a packed container with vented measurement techniques for the improvement of forced-air cooling, refriger-
walls. Int. J. Refrig. 31, 242e255. ation and drying operations. J. Food Eng. 143, 90e101.
Ferrua, M.J., Singh, R.P., 2009a. Modeling the forced-air cooling process of fresh Paniagua, A.C., East, A.R., Hindmarsh, J.P., Heyes, J.A., 2013. Moisture loss is the
strawberry packages, Part I: numerical model. Int. J. Refrig. 32, 335e348. major cause of firmness change during postharvest storage of blueberry.
Ferrua, M.J., Singh, R.P., 2009b. Modeling the forced-air cooling process of fresh Postharvest Biol. Technol. 79, 13e19.
strawberry packages, Part II: experimental validation of the flow model. Int. J. Pathare, P.B., Opara, U.L., Vigneault, C., Delele, M.A., Al-Said, F.A., 2012. Design of
Refrig. 32, 349e358. packaging vents for cooling fresh horticultural produce. Food Bioprocess
Ferrua, M.J., Singh, R.P., 2009c. Modeling the forced-air cooling process of fresh Technol. 5, 2031e2045.
strawberry packages, Part III: experimental validation of the energy model. Int. Pham, Q.T., Trujillo, F.J., McPhail, N., 2009. Finite element model for beef chilling
J. Refrig. 32, 359e368. using CFD-generated heat transfer coefficients. Int. J. Refrig. 32, 102e113.
Ferrua, M.J., Singh, R.P., 2009d. Design guidelines for the forced-air cooling process Progelhof, R.C., Thrine, J.L., Ruetsch, R.R., 1976. Methods for predicting the thermal
of strawberries. Int. J. Refrig. 32, 1932e1943. conductivity of composite systems: a review. Polym. Eng. Sci. 16 (9), 615e625.
Gaffney, J.J., Baird, C.D., Chau, K.V., 1985. Influence of airflow rate, respiration, Razavi, S.M.A., BahramParvar, M., 2007. Some physical and mechanical properties of
evaporative cooling, and other factors affecting weight loss calculations for kiwifruit. Int. J. Food Eng. 3 (6) http://dx.doi.org/10.2202/1556-3758.1276.
fruits and vegetables. ASHRAE Trans. 91 (1), 690e707. Rogge, S., Defraeye, T., Herremans, E., Verboven, P., Nicolai, B.M., 2015. A 3D contour
Gross, K.C., Wang, C.Y., Saltveit, M., 2004. The Commercial Storage of Fruits, Vege- based geometrical model generator for complex-shaped horticultural products.
tables, and Florist and Nursery Stocks. United States Department of Agriculture. J. Food Eng. 157, 24e32.
Handbook 66, Washington DC. Sweat, V.E., 1974. Experimental values of thermal conductivity of selected fruits and
Hardenburg, R.E., Watada, A.E., Wang, C.Y., 1986. The Commercial Storage of Fruits, vegetables. J. Food Sci. 39, 1080e1083.
Vegetables, and Florist and Nursery Stocks. United States Department of Agri- Tabatabaeefar, A., Rajabipour, A., 2005. Modeling the mass of apples by geometrical
culture. Handbook 66, Washington DC. attributes. Sci. Hortic. 105, 373e382.
Heyes, J.A., Tanner, D.J., East, A.R., 2010. Kiwifruit respiration rates, storage tem- Talbot, M.T., 1988. An approach to better design of pressure-cooled produce con-
peratures and harvest maturity. Proc. Int. Symp. Postharvest Pacifica 2009 tainers. Proc. Fla. State Hortic. Soc. 101, 165e175.
(880), 167e173. Tanner, D.J., Cleland, A.C., Opara, L.U., Robertson, T.R., 2002a. A generalised math-
Hu, Z., Sun, D.W., 2000. CFD simulation of heat and moisture transfer for predicting ematical modelling methodology for design of horticultural food packages
cooling rate and weight loss of cooked ham during air-blast chilling process. exposed to refrigerated conditions: part 1, formulation. Int. J. Refrig. 25, 33e42.
J. Food Eng. 46, 189e198. Tanner, D.J., Cleland, A.C., Opara, L.U., Robertson, T.R., 2002b. A generalised math-
Hu, Z., Sun, D.W., 2001. Effect of fluctuation in inlet airflow temperature on CFD ematical modelling methodology for design of horticultural food packages
simulation of air blast chilling process. J. Food Eng. 48, 311e316. exposed to refrigerated conditions: part 2, heat transfer modelling and testing.
Incropera, F.P., Dewitt, D.P., 2006. Transient conduction. In: Fundamentals of Heat Int. J. Refrig. 25, 43e53.
and Mass Transfer, fourth ed. John Wiley & Sons, New York. Tanner, D.J., Cleland, A.C., Opara, L.U., Robertson, T.R., 2002c. A generalised math-
Kumlutas, D., Tavman, I.H., Coban, M.T., 2003. Thermal conductivity of particle filled ematical modelling methodology for design of horticultural food packages
poluethylene composite materials. Compos. Sci. Technol. 63 (1), 113e117. exposed to refrigerated conditions: part 3, mass transfer modelling and testing.
Kuriger, R.J., Alam, M.K., 2002. Thermal conductivity of thermoplastic composites Int. J. Refrig. 25, 54e65.
with submicrometer carbon fibers. Experimental heat transfer. A J. Therm. Tanner, D.J., Amos, N.D., 2003. Temperature variability during shipment of fresh
Energy Generation, Transp. Storage Convers. 15 (1), 19e30. produce. Acta Hortic. 599, 193e203.
Le Page, J., Chevarin, C., Kondjoyan, A., Daudin, J., Mirade, P., 2009. Development of Trujillo, F.J., Pham, Q.T., 2006. A computational fluid dynamic model of the heat and
an approximate empirical-CFD model estimating coupled heat and water moisture transfer during beef chilling. Int. J. Refrig. 29, 998e1009.
transfers of stacked food products placed in airflow. J. Food Eng. 92, 208e216. Valente, M., Chamberel, A., Cordonnier, J., Pumborios, M., 1996. Finite element
Lewis, W.K., 1922. The evaporation of a liquid into a gas. Trans. Am. Soc. Mech. Eng. modelling of heat transfer in avocados. Int. Agrophys. 10, 123e129.
14 (7), 445e446. Verboven, P., Flick, D., Nicolai, B.M., Alvarez, G., 2006. Modelling transport phe-
Luyt, A.S., Molefi, J.A., Krump, H., 2006. Thermal, mechanical and electrical prop- nomena in refrigerated food bulks, packages and stacks: basics and advances.
erties of copper powder filled low-density and linear low-density polyethylene Int. J. Refrig. 29, 985e997.
composites. Polym. Degrad. Stab. 91 (7), 1629e1636. Vigneault, C., Castro, L.R., 2005. Produce-simulator property evaluation for indirect
MacCrae, E.A., Bowen, J.H., Stec, M.G.H., 1989. Maturation of kiwifruit (Actinidia airflow distribution measurement through horticultural crop package. J. Food,
deliciosa cv Hayward) from two orchards: differences in composition of the Agric. Environ. 3 (2), 67e72.
tissue zones. J. Sci. Food Agric. 47 (4), 401e416. Vissotto, F.Z., Kieckbusch, T.G., Neves Filho, L.C., 2000. Pre-cooling of model prod-
MacGregor, W., 2005. Effects of air velocity, air temperature, and berry diameter on ucts with forced air. Braz. J. Food Technol. 3, 1e10.
wild blueberry drying. Dry. Technol. 23, 387e396. Waghmare, R.B., Mahajan, P.V., Annapure, U.S., 2013. Modelling the effect of time
Maguire, K.M., Banks, N.H., Opara, L.U., 2001. Factors affecting weight loss of apples. and temperature on respiration rate of selected fresh-cut produce. Postharvest
Hortic. Rev. 25, 197e234. Biol. Technol. 80, 25e30.
Mamunya, Y.P., Davydenko, V.V., Pissis, P., Lebedev, E.V., 2002. Electrical and ther- Warren, R., 1992. Ceramic-Matrix Composites. Blackie, Glasgow and London.
mal conductivity of polymers filled with metal powders. Eur. Polym. J. 38, Weidenfeller, B., Hofer, M., Schilling, F.R., 2004. Thermal conductivity, thermal
1887e1897. diffusivity and specific heat capacity of particle filled polypropylene. Compos.
Marai, S., Ferrar, E., Civelli, R., 2012. Postharvest cold chain optimisation of little Part A Appl. Sci. Manuf. 35 (1), 423e429.
fruits. In: Proceedings of the 2012 COMSOL Conference, Milan. Zhao, C.J., Han, J.W., Yang, X.T., Qian, J.P., Fan, B.L., 2016. A review of computational
Maurer, A.R., Eaton, G.W., 1971. Calculation of potato tuber surface area. Am. Potato fluid dynamics for forced-air cooling process. Appl. Energy 168, 314e331.

Das könnte Ihnen auch gefallen