You are on page 1of 9

Journal of Food Engineering 115 (2013) 173–181

Contents lists available at SciVerse ScienceDirect

Journal of Food Engineering


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

Evaluation of convective heat transfer coefficient between fluids and particles


in suspension as food model systems for natural convection using two
methodologies
A. Cariño-Sarabia, J.F. Vélez-Ruiz ⇑
Chemical, Food and Environmental Engineering Department, Universidad de las Américas Puebla, Ex-Hacienda. Sta. Catarina Martir, Cholula, Puebla 72820, Mexico

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

Article history: This study was conducted to evaluate the convective heat transfer coefficient between Newtonian and
Received 21 January 2012 non-Newtonian fluids and food particles inside of a glass vessel and to assess the effect of controlled vari-
Received in revised form 4 September 2012 ables: convection phenomena, heating temperature, immersion fluid, and different food particles. The
Accepted 8 October 2012
coefficient was evaluated from the sample temperature as a function of the heating time by two
Available online 22 October 2012
approaches: a lumped parameter method and an analytical methodology. High values corresponded to
heating in Newtonian fluids (70–181, 57–248 W/m2 K), and lower corresponded in non Newtonian fluids
Keywords:
(31–169, 28–167 W/m2 K). Similarly, major heat transfer coefficient corresponded to the heating of
Convective heat transfer coefficient
Natural convection
mushrooms (32–110, 28–132 W/m2 K), followed by tomatoes (30–181, 35–183 W/m2 K) and potatoes
Newtonian and non-Newtonian fluids (31–148, 34–248 W/m2 K), respectively. Fluid properties were more important than the thermal proper-
Lumped parameter and analytical analysis ties of the particles. This coefficient was higher at 85 °C in both types of fluids. Both methods generated
Food particles satisfactory values, but the analytical approach showed more accuracy.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction Baptista et al., 1997; Palazoglu, 2006); heating/cooling of food


particles in Newtonian fluids (Dincer, 1997; Hulbert et al., 1997;
Food process operations frequently involve heating or cooling Carson et al., 2006; Laurindo et al., 2010); frying processes of food
stages between fluids and solid particles, being the heat transfer items (Costa et al., 1999; Sahin et al., 1999; Budzaki and Seruga,
by conduction and convection the dominant mechanisms. The 2005; Seruga and Budzaki, 2005), aseptic processing in continuous
increasing interest for manufacturing high quality, nutritious flow (Ramaswamy and Zareifard, 2000; Chakrabandhu and Singh,
and safe food products, as well as thermal processes modeling 2002; Palazoglu and Sandeep, 2002), industrial freezing (Amarante
and food equipment design have resulted in an imperious need and Lanoisellé, 2005), and during end-over-end agitation of canned
for evaluation of the convective heat transfer coefficient. It is nec- food items (Anantheswaran and Rao, 1985; Sablani and Ramaswamy,
essary to understand the contribution of this convective heat 1998; Varga and Oliveira, 2000; Meng and Ramaswamy, 2007;
transfer coefficient by self and to the overall heat transfer coeffi- Dwevedy and Ramaswamy, 2009), among others. Additional
cient for processes, involving solid foods within Newtonian and information on heat transfer between fluid and particulate foods is
non-Newtonian fluids (Vélez-Ruiz, 2009). The convective coeffi- required, in which its quantification has remained as a challenging
cient is commonly quantified from the temperature evolution issue of the transport phenomena in food engineering.
obtained from test materials of known shape, size and thermal Works on natural and forced convection in food particles in-
properties. clude those of Alhamdan and Sastry (1990) in which irregular
Over the past fifteen years, research efforts around the world shapes were investigated, finding hfp values between 22 and
have been conducted to study and characterize a diversity of 153 W/m2 K for cooling, and 75 and 310 W/m2 K for heating. Chan-
heating/cooling treatments in which the convective heat transfer darana et al. (1990) studied the effect of the fluid behavior at UHT
coefficient between fluids and particles has been evaluated, among conditions, they reported values of 56–90 for particles in a non-
other aspects. Including, heat exchange for non-Newtonian fluids Newtonian fluid, and 65–107 W/m2 K for particles heated in a
and irregular particles (Alhamdan and Sastry, 1990), non- Newtonian fluid. Awuah et al. (1993), evaluated the hfp in CMC
Newtonian fluids and regular items (Awuah et al., 1993, 1995; solution finding values of 76–556 W/m2 K for cylinders of carrots
and potatoes. Gadonna et al. (1996) developed predicting correla-
⇑ Corresponding author. Tel.: + 52 222 2292648; fax: +52 222 2292727. tions for hfp in forced convection, where higher values (290–
E-mail address: jorgef.velez@udlap.mx (J.F. Vélez-Ruiz). 1587 W/m2 K) were obtained as consequence of the turbulence.

0260-8774/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jfoodeng.2012.10.010
174 A. Cariño-Sarabia, J.F. Vélez-Ruiz / Journal of Food Engineering 115 (2013) 173–181

Nomenclature

a sphere radius (m) K consistency coefficient (Pasn)


a thermal diffusivity (m2/s) Lc particle characteristic length (m): radius for sphere, half
A shape dependent function of Biot number (Eq. (13)) of side for cube and ratio of volume/area for mushroom
Ap surface area of the particle (m2) mp particle mass (kg)
b thermal expansion coefficient (T1, °C1) gap apparent viscosity for non-Newtonian fluids (Pa s)
B shape dependent function of Biot number (Eq. (13)) = g viscosity for Newtonian fluids (Pa s)
dn2   n flow behavior index (dimensionless)
h L
Bi Biot number Bi ¼ fpkp c (dimensionless) Nu Nusselt number ðNu ¼
hfp D
Þ (dimensionless)
kp
C Empirical coefficient for Eqs. (17) and (18) Cpf g Cpf gap
Pr Prandtl number ðPr ¼ kf
¼ kf
Þ (dimensionless)
Cp specific heat capacity of the fluid (Cpf) or particle (Cpp)
(J/kgK)
r shear stress (Pa)
r radial position for a sphere (0 < r < a, m) equation 13
D diameter (m)
q density (kg/m3), at temperature T0 (q0) and T1 (q1), of
dn nth positive root of characteristic Eq. (13)
the fluid (qf) or particle (qp)
E Empirical coefficient for Eqs. (17) and (18)  
fH negative reciprocal slope of the straight line portion of Re Reynolds number Re ¼ Dlqv ¼ Dlqv (dimensionless)
ap
the heating curve on semi-log
 coordinates (s)
t time (s)
Fo Fourier number Fo ¼ La2t (dimensionless) T temperature (°C)
 c 
bg q2 L3c DT DT temperature difference between the heated fluid and
Gr Grashof number Gr ¼ gf 2 (dimensionless) the bulk value (°C)  
T T
T dimensionless temperature relationship Tpi T ff
g gravitational acceleration (m/s2)
Tf fluid temperature (°C)
c shear rate (s1)
Ti initial temperature (°C)
hfp surface or convective heat transfer coefficient (W/m2 K),
Tp particle temperature (°C)
also h in Eq. (13)
Xw mass fraction of water (dimensionless)
J Empirical coefficient for Eq. (18)
v velocity (m/s)
k thermal conductivity of the fluid (kf) or particle (kp)
V volume (m3) for suspended particles (Vp) and fluid (Vf)
(W/mK)
z dimension of change (m)

Related to thermal processes, several of the studies on hfp have experimental conditions; three of them that are utilized in this
been related to pasteurization and sterilization treatments. study are next.
Palazoglu and Sandeep (2002) developed a computer program to
determine the effect of hfp on microbial and nutritional responses for water : hfp ¼ 3:69Re0:434 ð1Þ
during aseptic processing. Using the most conservative value of
hfp, a high quality food product was obtained by Pannu et al. for starch solutions : hfp ¼ 2:33Re0:438 Pr0:349 ð2Þ
(2003), who carried out a study to asses the effect of flow direc-
tion and rate, food particle interaction, shape and size, as well and; for non Newtonian fluids Nu ¼ 1:41Re0:482 Pr0:355
as steam temperature on product temperature during packed
ð3Þ
bed heating of carrot and potato cubes and slices. Food center
and surface temperatures were measured by these researchers Also, the Froude number, density rate, particle concentration
under diverse process conditions reporting hfp values of 1500 and diameters relationship have been incorporated in more spe-
and 2000 W/m2 K for cubes with low Biot numbers of 14.2, and cific correlations, such those developed by Dwevedy and Ramasw-
17.8, respectively. amy (2009) for foods canning.
To quantify the convective or surface heat transfer coefficient Reviews developed by Maesmans et al. (1992) and Barigou et al.
there are different approaches, the sample temperature as a func- (1998), on the determination of fluid to particle heat transfer coeffi-
tion of time is indispensable for all of them. Some methods are very cient for food processes, mention the influence of important factors
sophisticated, employing numerical solution (Gadonna et al., 1996) in the heat transfer phenomena. These factors are: natural or forced
or finite difference technique (Palazoglu and Sandeep, 2002; convection, fluid rheology, particle size and concentration, particle
Palazoglu, 2006), other method includes the use of a time– properties and particle displacement (translational and rotational).
temperature integrator combined with mathematical modeling In the included studies of both reviews, a wide range of hfp values, 8–
(Maesmans et al., 1994). An alternative instrumental approach 7870 W/m2 K, has been reported as a function of a diversity of vari-
was utilized by Amarante and Lanoisellé (2005) to measure the ables, corresponding to Reynolds number from 0.006 to 50000, as
heat transfer coefficients in freezing equipment, in which heat the important factor; several particle shapes; different solid items;
sensors were coupled to temperature recording devices with satis- immersed in Newtonian and non Newtonian fluids.
factory results. Other approaches are practical and simple, two of Even though there are several studies on hfp evaluation and
these are the rate method and the ratio approach, both methods modeling; still there is a lack of information on it. Very little is
require accurate thermophysical properties in order to get conser- known about the hfp in food process operations in which the heat
vative hfp, and other one, is the quasi-steady method. transfer is involved, particularly the effect of particle load on the
Additionally most of the revised works, have established hfp, and there exist variations from one work to other; conse-
empirical correlations between the Nusselt number as a function quently more studies are needed. Thus, the objective of this re-
of Reynolds, Grashof, Prandtl, and other dimensional relationships search was to evaluate the convective heat transfer coefficient as
(Welty et al., 2002; Vélez-Ruiz, 2009, 2010), depending of the a function of the heating temperature (70 and 85 °C), fluid nature
A. Cariño-Sarabia, J.F. Vélez-Ruiz / Journal of Food Engineering 115 (2013) 173–181 175

(two Newtonian and two non-Newtonian fluids), and the shaped estimated at bath temperatures by using the next equations from
particle number (1, 5, and 10 items of three foods: potato cubes, different authors (Singh, 1992):
mushrooms, and tomato spheres) at natural convection, by using Specific heat (for fluids and particles):
two evaluation approaches that were compared.
Cp ¼ 0:837 þ 3:349X w ð5Þ
Thermal conductivity for fluids and, thermal conductivity for
2. Materials and methods
food particles:
2.1. Food items and test fluids kf ¼ ð326:58 þ 1:0412T  0:00337T 2 Þ

Mushrooms, potatoes and tomatoes (cherry) were selected to in- ð0:46 þ 0:54X w Þ1:73x103 ð6Þ
clude three particle shapes; they were acquired at a local supermar-
ket and stored at 5 °C. Cubes of potatoes were cut (2.0 cm side) with kp ¼ 0:148 þ 0:493X w ð7Þ
a fine knife, whereas mushrooms (0.47 cm characteristic dimen- Whereas from the density, specific heat and thermal conductiv-
sion), and tomatoes (2.5 cm diameter) were used in natural form. ity, the relation for the thermal diffusivity is:
The mass and the area of each food particle was averaged, the area
for the potato cubes was computed as 6 times of the side area, for k
a¼ ð8Þ
tomato the area was the correspondent to a sphere, and for mush- qCp
room the area was the sum of the cylindrical part plus the semi-
Thermal expansion coefficient for fluids was computed with
spherical top of the body (without the consideration of the
next relationship (Welty et al., 2002):
contact zone among the top and cylindrical parts of the mushroom).
Deionized water was taken from an institutional lab, tomato 1  ðq1 =q0 Þ
b¼ ð9Þ
puree (‘‘Del Fuerte’’ brand) was purchased from local supermarket; DT
whereas, salt (NaCl 3% w/w) and carboxymethyl cellulose (CMC Moisture content supported on water evaporation by using a
1.2% w/w) solutions were prepared by adding the corresponding vacuum oven at 70–75 °C, was obtained in according with AOAC
amounts of salt and CMC powders to water and thoroughly mixed (1995) methodology.
to break down undispersed lumps. The prepared solutions were
left one night to ensure total dispersion of the solute, finally hand
2.4. Heating experimental procedure
mixing was completed to homogenize the solution.
The four fluids were heated up to the desired temperature (70
Heat transfer experiments were carried out with four fluids
and 85 °C) in a steam kettle, taking sample for properties measure-
(two Newtonians and two non-Newtonians), three food particles
ment and transferring it to a glass vessel with a constant temper-
of different shape (potato cubes, mushrooms, and tomatoes as
ature water bath, for determination of the heat transfer coefficient.
spherical bodies), using three different particle units (1, 5 and
10), two heating temperatures (70 and 85 °C) and 1 natural con-
2.2. Determination of physical properties vection systems, generating 72 model systems or experimental
conditions; at least three experiments were done for each
Physical properties, such as density, particle dimensions and condition.
flow properties were measured by means of official or known The equipment consisted of one insulated vessel of 1 L without
methods. Density of fluids was determined with the pycnometric stirring; two or more copper-constantan thermocouples (type K),
method: a glass picnometer was used for water and sugar solution, previously calibrated, were utilized in the food items and immer-
and a Grease picnometer was utilized for CMC solution and tomato sion fluid; a data logger and a recording system were used
puree. For solid foods, a displacement method using xylene was (69200–00 Digi-Sense, USA). Electrical and recirculation (water
applied. Particle dimensions were measured with a Vernier scale. bath) mediums were utilized to control the fluid temperatures at
Mass of particles was measured by an analytical balance (Ohaus, 70 and 85 °C. Each food particle had its corresponding thermocou-
Pine Brook, NJ, USA). ple placed at the lowest point of heating that corresponded to the
Flow properties were determined instrumentally with a Brook- geometric center of the cube and sphere, and to the center of the
field Viscometer (DV-I, Brookfield Engineering Laboratories Inc., union zone between the top and base of the mushroom (it was
Middleboro, MA, USA) using a programmable sequence, with experimentally detected). The temperatures of the food center
400 mL of sample at studied temperatures. Thus a set of velocities and the fluid bulk were monitored and recorded, every 5 s, up to
(upward curve) was programmed from 0 to 100 rpm (0–36 s1) reach thermal equilibrium within the immersion fluid, when the
and the corresponding torque values were recorded. From pairs particles recorded 70 or 85 °C, correspondently.
of rotational velocity vs. torque values, and applying the manufac-
turer given constants for spindle and instrument equations, the 2.5. Data analysis
apparent shear rates (0–50 s1) and shear stresses were evaluated
by duplicate. For Newtonian fluids the viscosity was obtained, Heat transfer coefficients were determined by analyzing the
whereas for non-Newtonian fluids the flow behavior index, consis- temperature–time plots by two approaches: the so named lumped
tency coefficient and apparent viscosity were computed. From the parameter method and an analytical solution.
flow curves, the rheological parameters were fitted to the Power
Law model at both temperatures (70 and 85 °C). 2.5.1. Lumped parameter or quasi-steady state method (LP)
r ¼ K c_ n ð4Þ The evaluation of the heat transfer convective coefficients was
based on the assumption of a negligible internal resistance with
Biot number less than 0.1 (Alhamdan and Sastry, 1990; Rahman,
2.3. Thermal properties and moisture content 2009; Barigou et al., 1998; Welty et al., 2002; Vélez-Ruiz, 2010),
in which the rate of energy increase within the food item is consid-
Thermal properties (specific heat, thermal conductivity and ered equal to the rate of heat addition by convection, by mean of
thermal diffusivity) of particles and immersion fluids were the following equations:
176 A. Cariño-Sarabia, J.F. Vélez-Ruiz / Journal of Food Engineering 115 (2013) 173–181

 
mp Cp Tp  Tf following this solution, knowing fH from experimental heating data,
ln ¼ hfp t ð10Þ
Ap Ti  Tf B can be computed and substituted into the transcendental Eq. (14)
of a sphere (and similarly for the other two geometries) to deter-
or well:
mine the Biot number and hence the heat transfer coefficient.
Ap hfp The aforementioned methodology was also followed for the
ln T ¼  t ð11Þ particle cubes and mushrooms, taking the radius for tomato as
mp Cp
sphere. The half of the side for potato as cubic shape, and the vol-
From the temperature history of the particle, plots of ln ume/area relation for mushrooms, as the characteristic dimension
T ((Tp  Tf)/(Ti  Tf)) vs. time can be carried out to obtain the slope were introduced in order to simplify some of the relationships and
value; thus, the convective heat transfer coefficient may be evalu- calculation, without significant differences due to the specific form.
ated from the next equation that implies the knowledge of the
particles properties (area, mass, and specific heat):
2.6. Correlation fitting
 
mp Cp
hfp ¼ ðslopeÞ ð12Þ
Ap Finally and based on the system characteristics, fluid and parti-
cle properties and determined heat transfer coefficients, an empir-
ical correlation fitting was carried out with a computer software
2.5.2. Analytical approach (AA) (KaleidaGraph version 3.5, Synergy Software, Reading, PA, USA).
In this case, the evaluation of the convective coefficient was By using two relationships, from Awuah et al. (1993) and from
based on finite surface and internal resistances with Biot numbers Meng and Ramaswamy (2007) correspondingly.
 2

in the range 1–40. It takes the Fourier second law @T
@t
¼ a @@z2T as the
For single particles : Nu ¼ CðPr GrÞE ð17Þ
starting point, that is solved under convective boundary conditions
(T = T0 at t = 0; @T
@z
¼ 0 at the centerline of the body; k @T
@z
¼ hðT  T f Þ  J
VP
at the surface) for three geometries: infinite plate, infinite cylinder For multiple particles : Nu ¼ CðPr GrÞE ð18Þ
and sphere (Awuah et al., 1995; Welty et al., 2002; Vélez-Ruiz, Vf
2010). In which the useful equations for a sphere (Awuah et al., In which the viscosity of Newtonian liquids is substituted by the
1995; Zareifard and Ramaswamy, 1999) are the following: apparent viscosity for non-Newtonian fluids.
h i
X1 d2n þ ðBi  1Þ2 Sindn Sinðdn r=aÞ
T ¼ 2Bi  expðd2n FoÞ ð13Þ 2.7. Statistical analysis
2 2 r=a
n¼1 dn ðdn þ BiðBi  1ÞÞ
Analysis of variance and t student test were performed to
Bi ¼ 1  dCotðdÞ ð14Þ determine significance differences (p 6 0.05) of the heat transfer
coefficient magnitudes, by using the Minitab statistical software
The series solution (Eq. (13)) converges rapidly to the first term
(v.15, Minitab Inc., State College, PA, USA).
beyond a Fourier number value of 0.2. Thus, it can be generalized
for the first root and equated as:
3. Results and discussion
T ¼ A expðBFoÞ ð15Þ
or well in order to include any shape, the characteristic length (Lc) 3.1. Physical properties of food samples
was used:
! Those physical properties for the selected materials (particles
Bat and fluids) employed in this study, are included in Tables 1 and
log T ¼ log A  ð16Þ
2:303L2c 2. Some of these properties were experimentally measured or com-
puted; they are comparable with those magnitudes utilized or re-
Eq. (16) represents the straight line portion of a heating curve ob- ported by other authors. Existing discrepancies are attributed to
tained by a plot of log T vs. t, in which the slope value, also known differences of brand, food nature, variety and ripeness of the tested
 2

as heating rate index, is related to B fH ¼ 2:303LaB
c
. Therefore by food items, among others aspects of the experimentation.

Table 1
Physical properties of food particles.

Food Potato Mushroom Tomato


Property Experimentala Reported Experimentala Reported Experimentala Reported
b b
Conductivity (W/m°C) 0.52 ± 0.02 0.55 0.35 ± 0.02 0.37 0.57 ± 0.01 0.57b
Density (kg/m3) 1079 ± 28 1050c 652 ± 63 656c 1037 ± 29 1010c
Moisture (%) 81.64 ± 0.01 83.5c 94.00 ± 0.01 93.0c 92.78 ± 0.04 92.3c
Specific heat (J/kg°C) 3571 3670d 3985 3933d 3944 3928d
Thermal diffusivity (m2/s) 1.34  107 1.31  107e 1.33  107 1.40  107e 1.38  107 1.47  107e
Mass (g) 9.58 ± 0.81 – 12.0 ± 1.80 – 11.24 ± 2.33 –
Side/Vol:area/radius (m) 0.01 – 0.0047 ± 0.0004 – 0.0125 ± 0.001 –
Surface (m2) 0.0024 – 0.0041 ± 0.81 – 0.0024 + 0.00 –
Volumen (m3) 8.00  106 – 1.19 ± 0.29  105 – 1.06 ± 0.23  105 –
a
Most of data were measured or computed. Reported data from:
b
Ahmed and Rahman (2009).
C
Michailidis et al. (2009).
d
Singh et al. (2009).
e
Rahman and Al-Saidi (2009).
A. Cariño-Sarabia, J.F. Vélez-Ruiz / Journal of Food Engineering 115 (2013) 173–181 177

Table 2
Physical properties of experimental fluids.a

Food Waterb Salt solutionc CMC solutiond Tomato pureee


Property Experimental/Reported Experimental/Reported Experimental/Reported Experimental/Reported
Conductivity (W/m°C) 0.661(70) & 0.672(85) 0.652(70) & 0.672(85) 0.657(70) & 0.671(85) 0.641(70) & 0.645(85)
Density (kg/m3) 979(70) & 969(85) 1002 ± 3(70)&993 ± 5(85) 954 ± 2(70)&953 ± 4(85) 978(70) & 968(85)
Moisture (%) 100 97.0 ± 0.1 98.8 ± 0.2 87.7 ± 0.4
Specific heat (J/kg°C) 4196(70) & 4208(85) 4085(comp) & 4102(70) 4145(comp) & 4188(85) 3985(comp) & 3940(85)
Thermal diffusivitye (m2/s) 1.63(70) & 1.67  10-7(85) 4.23(70) & 4.36  10-7(85) 5.76(70) & 5.83  10-7(85) 1.49(70) & 1.54  10-7(85)
T. expansion coefficient 6.81(comp) & 6.79  104(70) 7.54  104 4.14  104 6.10  104
Consistency coefficient (Pa sn) – – 32 ± 3(70) & 29 ± 2(85) 49 ± 3(70) & 44 ± (85) g
Flow behavior index 1.0 1.0 0.81 ± 0(70) & 0.88 ± 0(85) 0.53 ± 0(70) & 0.52 ± 0(85)
Viscosityf (m Pa s) 0.411(70) & 0.344(85) 0.474(70) & 0.39(85) 14.6 ± 1.1(70) & 12.3 ± 2.4(85) 27 ± 3(70) & 22 ± 6(85) g
a
Most of data were measured or computed. Reported data from:
b
Welty et al. (2002).
c
Perry et al. (1992).
d
Alhamdan and Sastry (1990).
e
Rao and Cooley (1992), at temperatures of 70 and 85 °C.
f
Apparent viscosity for non-Newtonian fluids.
g
Viscosity in Pa s.

80

70

60
Temperature ( C)

50

40

30

20

10

0
0 100 200 300 400 500 600 700
Time (s)

Fig. 1. Temperature profile of potato cube (one particle) in water at 70 °C.

100
90
90
80
80
Temperature ( C)

70
Temperature ( C)

70
60 60
50 50
40 40
30 30
20 20
10 10
0
0 0 200 400 600 800 1000 1200
0 100 200 300 400 500 600 700 800
Time (s) Time (s)

Fig. 2. Temperature profile of mushroom (five particles) in salt solution at 85 °C. Fig. 3. Temperature profile of tomato Cherry (10 particles) in CMC solution at 85 °C.

90

3.2. Temperature profiles 80

70
Typical time–temperature profiles for different experimental
Temperature ( C)

60
systems are represented by figs. 1–4, showing the centre tempera-
50
ture of the particle as a function of time. Heating up in free convec-
tion with one particle of potato in water at 70 °C is included in fig. 40

1, taking 640 s to reach the fluid temperature. The evolution of five 30


mushrooms in salt solution is shown by fig. 2 in which a longer 20
time (695 s) was taken to reach the 85 °C, in both plots the heating 10
trend is consequence of natural convection. Whereas figs. 3 and 4,
0
are for heating in non-Newtonian fluids; in one case, tomato with 0 200 400 600 800 1000 1200 1400 1600
Time (s)
ten particles in CMC solution at 85 °C required 1085 s (Fig. 3) to
equalize the fluid temperature, whereas five units of potato in Fig. 4. Temperature profile of tomato Cherry (five particles) in tomato puree at
tomato puree at 85 °C needed 1360 s (Fig. 4), these differences in 85 °C.
178 A. Cariño-Sarabia, J.F. Vélez-Ruiz / Journal of Food Engineering 115 (2013) 173–181

natural heating rate are mainly due to the system characteristics Table 3
(food particle, immersion fluid and temperature). Heat transfer coefficients for potato, evaluated from the lumped parameter approach
and including the Biot number.
Due to the higher temperature difference, the food temperature
increased faster in a first stage, and lately in a second period in System (fluid, particles, number, temperature) h (W/m2°C) Bi
which the increasing rate diminished, as consequence of a minor Water, (1), 70 °C 123 ± 5 2.4 ± 0.1
driven force, being an asymptotic tendency. Notably and as ex- Water, (5), 70 °C 118 ± 3 2.3 ± 0.1
pected, the time required to reach a maximum temperature is low- Water, (10), 70 °C 78 ± 2 1.5 ± 0.0
Water, (1), 85 °C 145 ± 3 2.8 ± 0.0
er when the heating fluid behaves as Newtonian, due to low Water, (5), 85 °C 148 ± 8 2.9 ± 0.1
viscosity favors the movement of the fluid, in contrary to non New- Water, (10), 85 °C 93 ± 0 1.8 ± 0.0
tonian fluids. Salt sol. 3% (1), 70 °C 124 ± 8 2.4 ± 0.2
Temperature evolution in each model system exhibits the influ- Salt sol. 3% (5), 70 °C 120 ± 3 2.3 ± 0.1
Salt sol. 3% (10), 70 °C 78 ± 7 1.5 ± 0.1
ence of heating temperature, fluid nature, and number and type of
Salt sol. 3% (1), 85 °C 144 ± 7 2.8 ± 0.1
particles on the heating rates of the specific food. As expected, Salt sol. 3% (5), 85 °C 133 ± 8 2.6 ± 0.1
heating rates are higher for lower particle number (same heat rate Salt sol. 3% (10), 85 °C 89 ± 3 1.7 ± 0.0
for less particle mass), for the temperature of 85 °C (higher driven CMC 1.2% (1), 70 °C 88 ± 5 1.7 ± 0.1
force than 70 °C), for Newtonian fluids (lighter consistency than CMC 1.2% (5), 70 °C 83 ± 1 1.6 ± 0.0
CMC 1.2% (10), 70 °C 59 ± 5 1.1 ± 0.1
non Newtonians); and for potato as a consequence of its small size
CMC 1.2% (1), 85 °C 92 ± 3 1.8 ± 0.1
and mass. CMC 1.2% (5), 85 °C 90 ± 3 1.7 ± 0.1
CMC 1.2% (10), 85 °C 70 ± 4 1.3 ± 0.1
3.3. Heat transfer coefficients (hfp) Tomato puree, (1), 70 °C 56 ± 5 1.1 ± 0.1
Tomato puree, (5), 70 °C 45 ± 7 0.9 ± 0.1
Tomato puree, (10), 70 °C 32 ± 0 0.6 ± 0.0
From the experimental temperature profiles for each heating Tomato puree, (1), 85 °C 66 ± 8 1.3 ± 0.1
model system, the corresponding hfp values were computed by Tomato puree, (5), 85 °C 50 ± 3 1.0 ± 0.0
applying the lumped parameter and analytical methodologies, Tomato puree, (10), 85 °C 35 ± 1 0.7 ± 0.0
higher correlation coefficients were obtained for the analytical ap-
proach, perhaps reflecting the effect of a more accurate basis.
Figs. 5a and b show two plots of ln T and log T vs. time for mush-
Table 4
room in water for the application of both methodologies. Magni- Heat transfer coefficients for tomato, evaluated from the lumped parameter approach
tudes of the surface coefficient for lumped parameters are and including the Biot number.T1.
summarized in Tables 3–5, whereas Tables 6–8 summarize the re-
System (fluid, particles, number, temperature) h (W/m2°C) Bi
sults of the analytical method, for the three food particles.
Water, (1), 70 °C 162 ± 6 3.4 ± 0.1
Thus, taking into account the effect of the analyzed variables on
Water, (5), 70 °C 154 ± 5 3.3 ± 0.1
hfp, such as heating temperature, immersion fluid, type and shape Water, (10), 70 °C 90 ± 8 1.9 ± 0.2
of the food, as well as number of particles, some highlights may Water, (1), 85 °C 181 ± 7 3.8 ± 0.1
be mentioned next. For natural convection hfp is in the range of Water, (5), 85 °C 174 ± 2 3.7 ± 0.0
30–181 W/m2 K for the first method, being smaller than for the Water, (10), 85 °C 105 ± 4 2.2 ± 0.1
Salt sol. 3%, (1), 70 °C 157 ± 7 3.1 ± 0.1
analytical approach, 28–248 W/m2 K, and observing higher values,
Salt sol. 3% (5), 70 °C 142 ± 4 2.1 ± 0.1
Salt sol. 3% (10), 70 °C 92 ± 2 1.9 ± 0.1
(a) 5 Salt sol. 3% (1), 85 °C
Salt sol. 3% (5), 85 °C
174 ± 8
166 ± 5
3.6 ± 0.1
3.2 ± 0.1
4
Salt sol. 3% (10), 85 °C 108 ± 1 2.1 ± 0.1
3 CMC 1.2% (1), 70 °C 149 ± 4 1.1 ± 0.0
y = -0.0079x + 4.3935
R² = 0.9889
CMC 1.2% (5), 70 °C 101 ± 4 0.9 ± 0.1
2 CMC 1.2% (10), 70 °C 90.3 ± 3 0.6 ± 0.0
CMC 1.2% (1), 85 °C 168 ± 7 1.3 ± 0.1
LnT

1
CMC 1.2% (5), 85 °C 153 ± 7 1.2 ± 0.1
0 CMC 1.2% (10), 85 °C 98.9 ± 5 0.8 ± 0.1
0 100 200 300 400 500 600 700 800 900 Tomato puree, (1), 70 °C 52 ± 2 3.3 ± 0.1
-1 Tomato puree, (5), 70 °C 45 ± 4 3.0 ± 0.1
Tomato puree, (10), 70 °C 30 ± 1 2.9 ± 0.0
-2 Tomato puree, (1), 85 °C 61 ± 3 3.7 ± 0.2
Tomato puree, (5), 85 °C 57 ± 3 3.5 ± 0.1
-3
Time (s) Tomato puree, (10), 85 °C 35 ± 3 2.3 ± 0.0

(b)
0

-0.5
as expected, when the food was heated in Newtonian fluids. In gen-
-1 eral terms, hfp is highest in water, then in salt and CMC solutions
y = -0.0036x + 0.1423
R² = 0.99
and lowest in tomato puree. It is higher when one particle is im-
log T

-1.5 mersed in the heating medium than five and ten units, in that or-
der (with only four exceptions of twenty-four cases). And as
-2
expected, high convection coefficients were associated at 85 °C
-2.5
for all the particles in both types of fluids.
With respect to the immersion fluid, they exhibited the next
-3 trend: water (71–181, first method; and 57–248 W/m2 K, second
0 100 200 300 400 500 600 700 800 method) and salt solution (70–174, first method; and 63–175 W/
Time (s)
m2 K, second method) as Newtonian fluids, generating higher con-
Fig. 5. Semilog plot for mushroom (one particle) in water at 70 °C: (a) lumped vection coefficients, in comparison with CMC solution (59–168,
parameter approach, and (b) analytical method. first method; and 52–167 W/m2 K, second method) and tomato
A. Cariño-Sarabia, J.F. Vélez-Ruiz / Journal of Food Engineering 115 (2013) 173–181 179

Table 5 Table 7
Heat transfer coefficients for mushroom, evaluated from the lumped parameter Heat transfer coefficients for tomato, evaluated from the analytical approach and
approach and including the Biot number. including the Biot number.

System (fluid, particles, number, temperature) h (W/m2°C) Bi System (fluid, particles, number, temperature) h (W/m2°C) Bi
Water, (1), 70 °C 102 ± 13 1.1 ± 0.1 Water, (1), 70 °C 167 ± 5 3.5 ± 0.1
Water, (5), 70 °C 99 ± 13 1.1 ± 0.1 Water, (5), 70 °C 165 ± 4 3.5 ± 0.1
Water, (10), 70 °C 72 ± 8 0.8 ± 0.1 Water, (10), 70 °C 102 ± 5 2.2 ± 0.1
Water, (1), 85 °C 142 ± 8 1.6 ± 0.1 Water, (1), 85 °C 182 ± 9 3.9 ± 0.2
Water, (5), 85 °C 132 ± 3 1.5 ± 0.0 Water, (5), 85 °C 171 ± 5 3.6 ± 0.1
Water, (10), 85 °C 89 ± 7 1.0 ± 0.1 Water, (10), 85 °C 110 ± 7 2.3 ± 0.2
Salt sol. 3% (1), 70 °C 103 ± 12 1.1 ± 0.1 Salt sol. 3% (1), 70 °C 153 ± 6. 3.2 ± 0.1
Salt sol. 3% (5), 70 °C 99 ± 2 1.1 ± 0.0 Salt sol. 3% (5), 70 °C 146 ± 1 3.1 ± 0.0
Salt sol. 3% (10), 70 °C 70 ± 15 0.8 ± 0.2 Salt sol. 3% (10), 70 °C 90.6 ± 5 1.92 ± 0.1
Salt sol. 3% 85 °C 149 ± 5 1.7 ± 0.1 Salt sol. 3% (1), 85 °C 167 ± 3 3.5 ± 0.1
Salt sol. 3% (5), 85 °C 129 ± 6 1.5 ± 0.1 Salt sol. 3% (5), 85 °C 153 ± 10 3.2 ± 0.2
Salt sol. 3% (10), 85 °C 83 ± 6 0.9 ± 0.1 Salt sol. 3% (10), 85 °C 109 ± 4 2.3 ± 0.1
CMC 1.2% (1), 70 °C 78 ± 13 0.9 ± 0.1 CMC 1.2% (1), 70 °C 143 ± 4 3.0 ± 0.1
CMC 1.2% (5), 70 °C 82 ± 13 0.9 ± 0.1 CMC 1.2% (5), 70 °C 103 ± 1 2.2 ± 0.0
CMC 1.2% (10), 70 °C 62 ± 7 0.7 ± 0.1 CMC 1.2% (10), 70 °C 94 ± 2 2.0 ± 0.0
CMC 1.2% (1), 85 °C 110 ± 16 1.2 ± 0.2 CMC 1.2% (1), 85 °C 167 ± 9 3.5 ± 0.2
CMC1.2% (5), 85 °C 99 ± 10 1.1 ± 0.1 CMC 1.2% (5), 85 °C 159 ± 3 3.4 ± 0.1
CMC1.2% (10), 85 °C 72 ± 13 0.8 ± 0.1 CMC 1.2% (10), 85 °C 102 ± 7 2.1 ± 0.1
Tomato puree, (1), 70 °C 57 ± 11 0.6 ± 0.1 Tomato puree, (1), 70 °C 52 ± 0.1 1.1 ± 0.0
Tomato puree, (5), 70 °C 47 ± 2 0.5 ± 0.0 Tomato puree, (5), 70 °C 51 ± 0.2 1.1 ± 0.0
Tomato puree, (10), 70 °C 33 ± 14 0.4 ± 0.2 Tomato puree, (10), 70 °C 36 ± 7 0.8 ± 0.1
Tomato puree, (1), 85 °C 69 ± 16 0.8 ± 0.2 Tomato puree, (1), 85 °C 56 ± 3 1.2 ± 0.1
Tomato puree, (5), 85 °C 51 ± 6 0.6 ± 0.1 Tomato puree, (5), 85 °C 55 ± 2 1.2 ± 0.0
Tomato puree, (10), 85 °C 36 ± 13 0.4 ± 0.1 Tomato puree, (10), 85 °C 42 ± 9 0.9 ± 0.2

Table 6 Table 8
Heat transfer coefficients for potato, evaluated from the analytical approach and Heat transfer coefficients for mushroom, evaluated from the analytical approach and
including the Biot number. including the Biot number.

Sistema (fluido, número de partículas, temperatura) h (W/m2°C) Bi System (fluid, particles, number, temperature) h (W/m2°C) Bi
Water, (1), 70 °C 147 ± 7 2.8 ± 0.1 Water, (1), 70 °C 87 ± 12 1.0 ± 0.1
Water, (5), 70 °C 141 ± 9 2.7 ± 0.2 Water, (5), 70 °C 84 ± 13 0.9 ± 0.1
Water, (10), 70 °C 104 ± 2 2.0 ± 0.1 Water, (10), 70 °C 57 ± 7 0.6 ± 0.1
Water, (1), 85 °C 243 ± 9 4.7 ± 0.2 Water, (1), 85 °C 106 ± 18 1.2 ± 0.2
Water, (5), 85 °C 248 ± 8 4.8 ± 0.1 Water, (5), 85 °C 98 ± 12 1.1 ± 0.1
Water, (10), 85 °C 123 ± 8 2.4 ± 0.2 Water, (10), 85 °C 723 ± 16 0.8 ± 0.2
Salt sol. 3% (1), 70 °C 139 ± 9 3.2 ± 0.2 Salt sol. 3% (1), 70 °C 84 ± 15 0.9 ± 0.2
Salt sol. 3% (5), 70 °C 132 ± 8 3.2 ± 0.2 Salt sol. 3% (5), 70 °C 79 ± 9 0.9 ± 0.1
Salt sol. 3% (10), 70 °C 85 ± 1 1.6 ± 0.0 Salt sol. 3% (10), 70 °C 63 ± 6 0.7 ± 0.1
Salt sol. 3% (1), 85 °C 175 ± 8 3.4 ± 0.1 Salt sol. 3% (1), 85 °C 126 ± 10 1.4 ± 0.1
Salt sol. 3% (5), 85 °C 168 ± 8 3.2 ± 0.2 Salt sol. 3% (5), 85 °C 105 ± 18 1.2 ± 0.2
Salt sol. 3% (10), 85 °C 108 ± 3 2.1 ± 0.0 Salt sol. 3% (10), 85 °C 71 ± 7 0.8 ± 0.1
CMC 1.2% (1), 70 °C 102 ± 8 2.0 ± 0.2 CMC 1.2% (1), 70 °C 68 ± 12 0.8 ± 0.1
CMC 1.2% (5), 70 °C 91 ± 8 1.7 ± 0.1 CMC 1.2% (5), 70 °C 75 ± 8 0.8 ± 0.1
CMC 1.2% (10), 70 °C 73 ± 5 1.4 ± 0.1 CMC 1.2% (10), 70 °C 52 ± 7 0.6 ± 0.1
CMC 1.2% (1), 85 °C 108 ± 3 2.1 ± 0.1 CMC 1.2% (1), 85 °C 96 ± 12 1.1 ± 0.1
CMC 1.2% (5), 85 °C 96 ± 2 1.9 ± 0.0 CMC 1.2% (5), 85 °C 90 ± 5 1.0 ± 0.1
CMC 1.2% (10), 85 °C 73 ± 3 1.3 ± 0.1 CMC 1.2% (10), 85 °C 69 ± 10 0.8 ± 0.1
Tomato puree, (1), 70 °C 72 ± 4 1.4 ± 0.1 Tomato puree, (1), 70 °C 47 ± 4 0.5 ± 0.0
Tomato puree, (5), 70 °C 42 ± 9 0.8 ± 0.2 Tomato puree, (5), 70 °C 42 ± 9 0.5 ± 0.1
Tomato puree, (10), 70 °C 35 ± 2 0.7 ± 0.0 Tomato puree, (10), 70 °C 29 ± 8 0.3 ± 0.1
Tomato puree, (1), 85 °C 75 ± 0 1.4 ± 0.0 Tomato puree, (1), 85 °C 60 ± 4 0.7 ± 0.0
Tomato puree, (5), 85 °C 52 ± 0 1.0 ± 0.0 Tomato puree, (5), 85 °C 47 ± 3 0.5 ± 0.0
Tomato puree, (10), 85 °C 36 ± 0 0.7 ± 0.0 Tomato puree, (10), 85 °C 33 ± 10 0.4 ± 0.1

puree (30–69, first method; and 28–75 W/m2 K, second method). AA, being lower for non Newtonian fluids and as a function of the
Considering potato, tomato and mushroom, the associated heat food particle. Therefore, the studied heating process may be consid-
transfer coefficient had a range of 31–248, 30–181, and 28– ered, in according to Awuah et al. (1995), Welty et al. (2002),
149 W/m2 K, respectively. Finally for particle number, a decreasing Rahman (2009) and Vélez-Ruiz (2010), as a finite resistances pro-
trend was observed for the convection heat transfer coefficient as a cess (Bi: 0.1–40), and consequently the obtained results for the ana-
function of the number of particles due to a same quantity of heat lytical approach may be considered with more accuracy. For potato
is utilized for more food mass. Similar values of this convective and tomato is observed that the linear trend of the curve coincides
coefficient (50–200 W/m2 K) have been reported by several with the time in which the Fourier number is higher than 0.2 (143 s
researchers for analog heat transfer processes. for potato and 220 s for tomato particle), that is part of the funda-
In order to know the influence of the internal resistance, the mentals of the analytical approach; while for the mushroom parti-
magnitude of Biot number was evaluated from the determined cle, this Fourier number was reached after 32 s of heating and the
hfp. The range of the correspondent Biot numbers for the studied linear trend is observed after 85 s, the irregular shape of mushroom
heat transfers systems were 0.37–3.82 for LP and 0.32–4.78 for contributed to the observed difference among particles.
180 A. Cariño-Sarabia, J.F. Vélez-Ruiz / Journal of Food Engineering 115 (2013) 173–181

Our experimental Biot numbers are lower or comparable to 3.5. Empirical fittings
some reported data, Biot numbers of 8.3–16.7 were reported by
Palazoglu and Sandeep (2002) for particles of 0.5, 0.75 and 1 cm Finally, although the primary focus of this work was to evaluate
in which hfp was computed by numerical and analytical ap- the convective coefficient by two methods, an attempt was made
proaches; Maesmans et al. (1994) handled Biot numbers up to 40 to compare and fit the evaluated experimental coefficients with
as part of the applied methodology in the determination of hfp; those values obtained from correlations. Thus, in addition to corre-
and, a range of 14–17 for Biot was reported by Pannu et al. lations proposed by Awuah et al. (1993) and Alhamdan and Sastry
(2003) for the thermal processing of diced vegetables, differences (1990) among others, in which high error percentages were ob-
that basically may be attributed to the absence of forced move- tained, the best fittings from KaleidaGraph computer software are:
ment of our experiments. In according to Pannu et al. (2003) for Potato, one particle:
0.1 < Bi < 100, although the internal heat dissipation is slow, the s
ystem maintains its capacity to keep the surface temperature low- Nu ¼ 2:0515ðPr GrÞ0:0914 R2 ¼ 0:87 ð19Þ
er than the heating fluid. Potato, multiple particles:
 0:1562
Vp
3.4. Comparison of applied methods Nu ¼ 1:1075ðPr GrÞ0:0887 R2 ¼ 0:85 ð20Þ
Vf

From the aforementioned results, it can be observed that the Tomato, one particle:
Biot number was higher than 0.1, and both parameters hfp and Bi
exhibited similar trends. In which the tomato was the food particle Nu ¼ 2:8450ðPr GrÞ0:0455 R2 ¼ 0:93 ð21Þ
with highest values for the heat transfer coefficient and Biot num- Tomato, multiple particles:
ber in most of the cases, followed by the potato and mushroom
 0:0110
particles, this may be interpreted as a lowest internal resistance Vp
Nu ¼ 2:3721ðPr GrÞ0:0741 R2 ¼ 0:80 ð22Þ
of the specific food in the heating phenomena. Although there Vf
are other factors involved in the Biot number evaluation, such as
the lowest thermal conductivity and characteristic length for Mushroom, one particle:
mushroom, reducing the internal resistance of the particle, for in-
Nu ¼ 2:9411ðPr GrÞ0:0514 R2 ¼ 0:82 ð23Þ
stance; and obviously influencing the coefficient magnitudes.
On the other side, there is an important influence of the immer- Mushroom, multiple particles:
sion fluid. The linear trend of the temperature profile was lost at  0:0190
different times, for water and salt solution a time of 600 s was Vp
Nu ¼ 2:4158ðPr GrÞ0:0523 R2 ¼ 0:73 ð24Þ
coincident with this fact (Fig. 5), whereas for CMC solution and to- Vf
mato puree the elapsed times were 700 and 1500 s, respectively. At
In general, the fittings coefficients were not as high as expected
that time the particle has reached a temperature near to the fluid
(R2 < 0.93), due to the complementary effect of other studied fac-
temperature, reducing the driven force for heating.
tors of this work, such as particles shape and fluid flow properties.
It may be also observed, that there is more similarity between
It can be noticed, that correlations for one particle generated better
hfp values, obtained with both methodologies, for tomato than for
fittings (R2: 0.82–0.93) than those obtained for multiple items (R2:
potato and mushrooms particles (Tables 4 and 7). When the heat
0.73–0.85).
transfer coefficient is lower than 100 W/m2°C, the difference be-
tween methods is minor. Furthermore, for potato particles the
analytical approach generated higher values; in contrary for mush- 4. Conclusions
room particles the lumped parameter method produced higher
values. The difference between methods was quantified as 1–16% This study points out some facts related to the heat transfer
for tomato, 1.5–32% for mushroom and 1–40% for potato. From sta- coefficient evaluation; hfp is higher in Newtonian fluids than in
tistical comparison, there was not significant difference (p > 0.05) non Newtonian fluid at high temperature with natural convection;
between methods for tomato particles, in contrary to potato and the number of food particles has a decreasing effect on this convec-
mushroom particles. Therefore it may be concluded that the LP tive parameter, in which some interesting exceptions were ob-
methodology may be used for tomato spheres, whereas for potato served and related to non Newtonian behavior.
and mushroom geometries the AA should be applied to get more Both applied methods (LP and AA) may be used to evaluate the
accuracy in the convective coefficient evaluation. Additionally, all heat transfer coefficient, in which the application of the first one is
the interactions were significant (p 6 0.05) on the hfp values com- easier and faster, but when an accurate evaluation of hfp is needed,
puted by the AA method, being more important the interaction by considering the food or internal resistance, the analytical ap-
temperature  fluid type for potato; while for the mushroom, the proach should be utilized.
interaction temperature  particle number was more important, Rate of heat penetration in a system with food particles is a
that are in agreement with the results of Ramaswamy and function of several factors, rheological and thermal properties of
Zareifard (2000). the involved fluid; as well as, dimensions, number, shape and ther-
Maesmans et al. (1992) reported hfp of 239 W/m2°C for potato mal properties of the solid particles are also very important. All of
cubes (0.03 m side) in water at 75 °C and 146 W/m2°C in sugar them should be known in order to reach the objective of a thermal
solution (35%) also at 75 °C. For the same cubes in starch solution treatment.
at 144 °C, a hfp of 99 W/m2°C was computed; while for aluminiun The knowledge of physical and transport properties, such as
shaped mushroom in water at 60 °C, a hfp of 850 W/m2°C was heat capacity, thermal conductivity, consistency coefficient, and
quantified. Further, for spherical polyacetate particles, a coefficient convective heat transfer coefficient is indispensable for the accu-
of 153 W/m2°C in water at 80 °C was reported. All these values are racy of a thermal treatment of food fluids with solid particles, as
similar to our results, and the differences are mainly attributed to well as, to reach microorganisms destruction, and to avoid nutri-
variations in fluids, temperature and materials; the conductivity of tive and quality losses; therefore the research on these important
metals is very high in comparison with food items. aspects of food engineering must be continued.
A. Cariño-Sarabia, J.F. Vélez-Ruiz / Journal of Food Engineering 115 (2013) 173–181 181

References Maesmans, G.J., Hendrickx, M.E., De Cordt, S.V., Tobback, P., 1994. Feasibility of the
use of a time-temperature integrator and a mathematical model to determine
fluid-to-particle heat transfer coefficients. Food Research International 27, 39–
Ahmed, J., Rahman, M.S., 2009. Thermal conductivity data of foods. In: Rahman, M.S.
51.
(Ed.), Food Properties Handbook. CRC Press, Boca Raton, FL, USA, pp. 581–622.
Meng, Y., Ramaswamy, S., 2007. Dimensionless heat transfer correlations for high
Alhamdan, A., Sastry, S.K., 1990. Natural convection heat transfer between non-
viscosity fluid-particle mixtures in cans during end-over-end rotation. Journal
Newtonian fluids and an irregular shaped particle. Journal of Food Process
of Food Engineering 80, 528–535.
Engineering 13, 113–124.
Michailidis, A.P., Krokida, K.M., Rahman, M., 2009. Data and models of density,
Amarante, A., Lanoisellé, J.L., 2005. Heat transfer coefficients measurement in
shrinkage and porosity. In: Rahman, M.S. (Ed.), Food Properties Handbook. CRC
industrial freezing equipment by using heat flux sensors. Journal of Food
Press, Boca Raton, FL, USA, pp. 397–416.
Engineering 66, 377–386.
Palazoglu, T.K., Sandeep, K.P., 2002. Assesment of the effect of fluid-to-particle heat
Anantheswaran, R.C., Rao, M.A., 1985. Heat transfer to model non-newtonian liquid
transfer coefficient on microbial and nutrient destruction during aseptic
foods in cans during end-over-end rotation. Journal of Food Process Engineering
processing of particulate foods. Journal of Food Science 67 (9), 3359–3364.
4, 21–35.
Palazoglu, K., 2006. Influence of convective heat transfer coefficient on the heating
AOAC International, 1995. Official methods of analysis, 16th ed. AOAC
rate of materials with different thermal diffusivities. Journal of Food
International, Arlintong, VA.
Engineering 73, 290–296.
Awuah, G.B., Ramaswamy, H.S., Simpson, B.K., 1993. Surface heat transfer
Pannu, K.S., Castaigne, F., Arul, J., 2003. Thermal processing of particulte foods by
coefficients associated with heating of food particles in CMC solutions.
steam injection. Part 1. Heating rate index for diced vegetables. In: Welti-
Journal of Food Process Engineering 16, 39–57.
Chanes, J., Vélez-Ruiz, J.F., Barbosa-Cánovas, G.V. (Eds.), Transport Phenomena
Awuah, G.B., Ramaswamy, H.S., Simpson, B.K., 1995. Comparison of two methods for
in Food Processsing. CRC Press, Inc., Boca Raton, FLO, pp. 309–326.
evaluating fluid to-surface heat transfer coefficients. Food Research
Perry, H., Green, D., Maloney, J., 1992. Chemical Engineer Handbook. McGraw-Hill.
International 28 (3), 261–271.
México, In Spanish.
Baptista, P.N., Oliveira, F.A.R., Oliveira, J.C., Sastry, S.K., 1997. Dimensionless analysis
Rahman, M., Al-Saidi, G., 2009. Thermal diffusivity of foods: measurement, data and
of fluid-to-particle heat transfer coefficients. Journal of Food Engineering 31,
prediction. In: Food Properties Handbook. CRC Press, Boca Raton, FL, USA, pp.
199–218.
649–696.
Barigou, M., Mankad, S., Fryer, P.J., 1998. Heat transfer in two-phase solid-liquid
Ramaswamy, H.S., Zareifard, M.R., 2000. Evaluation of factors influencing tube-flow
food flows: A review. Transactions of IChem 76-C, 3–29.
fluid-to-particle heat transfer coefficient using a calorimetric technique. Journal
Budzaki, S., Seruga, B., 2005. Determination of heat transfer coefficient during frying
of Food Engineering 45, 127–138.
of potato dough. Journal of Food Engineering 66, 307–314.
Rao, M., Cooley, H., 1992. Rheological behavior of tomato pastes in steady and
Carson, J.M., Willix, J., North, M.F., 2006. Measurement of heat transfer coefficients
dynamic shear. Journal of Texture Studies 23, 415–425.
within convection ovens. Journal of Food Engineering 72, 293–301.
Sablani, S.S., Ramaswamy, H.S., 1998. Multi-particle mixing and its role in heat
Chakrabandhu, K., Singh, R.K., 2002. Fluid-to-particle heat transfer coefficients for
transfer during end-over-end agitation of cans. Journal of Food Engineering 38,
continuous flow of suspensions in coiled tube and straight tube with bends.
141–152.
Food Science and Technology (LWT) 35, 420–435.
Sahin, S., Sastry, S.K., Bayindirli, L., 1999. The determination of convective heat
Chandarana, D.I., Gavin, A., Wheaton, F.W., 1990. Particle/fluid interface heat
transfer coefficient during frying. Journal of Food Engineering 39, 307–311.
transfer under UHT conditions at low particle/fluid relative velocities. Journal of
Seruga, B., Budzaki, S., 2005. Determination of thermal conductivity and convective
Food Process Engineering 13, 191–206.
heat transfer coefficient during deep fat frying of ‘‘Krotula’’ dough. European
Costa, R.M., Oliveira, F.A.R., Delaney, O., Gekas, V., 1999. Analysis of the heat transfer
Food Research and Technology 221, 351–356.
coefficient during potato frying. Journal of Food Engineering 39, 293–299.
Singh, R.P., 1992. Heating and cooling processes for foods. In: Heldman, E.R., Lund,
Dincer, I., 1997. New effective Nusselt-Reynolds for food-cooling applications.
D.B. (Eds.), Handbook of Food Engineering. Marcel Dekker Inc., NY, pp. 247–276.
Journal of Food Engineering 31, 59–67.
Singh, R.P., Erdogdu, F., Rahman, M.S., 2009. Specific heat and enthalpy of foods. In:
Dwevedy, M., Ramaswamy, H., 2009. Dimensipnless correlations for convective heat
Rahman, M.S. (Ed.). Food Properties Handbook. CRC Press, Boca Raton, FL, USA.
transfer in canned particulate fluids under axial rotation processing. Journal of
pp. 517–543.
Food Process Engineering 33, 182–207.
Varga, S., Oliveira, J.C., 2000. Determination of the heat transfer coefficient between
Gadonna, J.P., Pain, J.P., Barigou, M., 1996. Determination of the convective heat
bulk medium and packed containers in a batch retort. Journal of Food
transfer coefficient between a free particle and a conveying fluid in a horizontal
Engineering 44, 191–198.
pipe. Transactions of IChemE 74-C, 27–39.
Vélez-Ruiz, J.F., 2009. Evaluation of convective heat transfer coefficient for food
Hulbert, G.J., Litchfield, J.B., Smichdt, S.J., 1997. Determination of convective heat
process operations. In: Sosa-Morales, M.E., Vélez-Ruiz, J.F. (Eds.), Food
transfer coefficients using 2D MRI temperature mapping and finite element
Processing and Engineering Topics. Ed. Nova, NY, USA, pp. 125–153.
model. Journal of Food Engineering 34, 193–201.
Vélez-Ruiz, J.F., 2010. Notes of Food Transport Phenomena. Universidad de las
Laurindo, J.B., Carciofi, B.A.M., Silva, R.R., Dannenhauer, C.E., Hense, H., 2010.
Americas Puebla. Unpublished document. Puebla, Mexico (In Spanish).
Evaluation of the effects of water agitation by air injection and water
Welty, J.R., Wicks, C.E., Wilson, R.E., 2002. Fundamentals of Momentum, Heat and
recirculation on the heat transfer coefficients in immersion cooling. Journal of
Mass Transfer. Limusa, Mexico, In Spanish.
Food Engineering 96, 59–65.
Zareifard, M.Y., Ramaswamy, H., 1999. A calorimetric approach for evaluation of
Maesmans, G.J., Hendrickx, M.E., De Cordt, S.V., Tobback, P., 1992. Fluid-to-particle
fluid-to-particle heat transfer coefficient under tube-flow conditions.
heat transfer coefficient determination of heterogeneous foods: A review.
Lebensmittel-Wissenschaft und-Technologie 32 (8), 495–502.
Journal of Food Processing and Preservation 16, 29–69.

You might also like