Sie sind auf Seite 1von 14

A METHOD FOR DETERMINING THE CONVECTIVE HEAT

TRANSFER COEFFICIENT DURING iMMERSION FRYING


LYNN J. HUBBARD and BRIAN E. FARKAS'

Department of Food Science


North Carolina State University
Raleigh. NC 27695

Accepted for Publication January 14. 1999

ABSTRACT

m e convective heat transfer coefficient is a useful parameter in characteriz-


ing heat flow across a fluidholid intetface when the fluid flow field is complex
and solution of the coupled transport equations impractical. While convective
heat transfer comcient valuesfor many unit operations have been tabulated, the
boiling phase of immersionfrying has not been quantified. The objective of this
study was to develop a laboratory method for the measurement of the convective
heat transfer comcient during immersionfrying. The method that was developed
was applied to the immersionfrying of potato cylinders at an oil temperature of
180C. The convective heat tran@er coefficient was initially 300 W / d K , it
increased sharply to I100 W/mzK,and gradually decreased to 300 W / d K over
the duration of the process. Use of this new method will allow the study of the
effects of oil temperature, oil quality, product shapehize, andproduct quality on
heat transfer comcients.

INTRODUCTION

Immersion frying is a common unit operation used in the preparation of


foods and may be defined as the process of immersing a food product in an
edible oil or fat that is heated to a temperature above the boiling point of water.
This process creates a scenario wherein heat is transferred from the oil or fat to
the food and water is evaporated from the food. The coupled heat and mass
transfer problem makes frying one of the most difficult unit operations to
understand (Dagerskog 1977). In addition, the complex flow field around the
food material requires that boundary conditions for the thermal transport
equations utilize a convective heat transfer coefficient to characterize heat flow
across the fluidholid interface. Thus, quantification of the convective heat

'CorrespondingAuthor

Journal of Food Process Engineering 22 (1999) 201-214. All Rights Reserved.


"Copyrighi 1999 by Food & Nutrition Press, Inc.. Trumbull, Connecticut. 20 1
202 L.J. HUBBARD and B.E. FARKAS

transfer coefficient plays an integral part in understanding the complex system


of immersion Erying.
Immersion frying may be broken into four stages: (1) initial-heating, (2)
surface boiling, (3) falling rate, and (4) bubble end point (Farkas ef al. 1996).
Initial-heating is described as the initial immersion of a raw material into hot oil
and is characterized by the absence of water vaporization. During this stage heat
is transferred from the oil to the food via free convection and through the food
via conduction. Stage two, surface boiling, is characterized by the sudden loss
of free moisture at the surface, increased surface heat transfer, and inception of
crust formation. The falling rate stage parallels that of drying in which there is
continued thickening of the crust region, decreased heat transfer, and a steady
decrease in vapor mass transfer from the material. Bubble end point is
characterized by the apparent cessation of moisture loss from the food during
frying (Farkas ef al. 1996). These four stages may be further generalized as a
nonboiling phase (stages 1 and 4) or a boiling phase (stages 2 and 3).
Previous quantification of convective heat transfer coefficients (h), during
immersion frying has focused on the non-boiling phase. Miller ef al. (1994)used
lumped capacity analysis (Holman 1990). and found h, = 281 W/mzK for
soybean oil at 180C. As frying oils degrade many reactions occur that change
their physical and chemical properties. These reactions produce volatiles,
nonpolymeric polar compounds of moderate volatility, dimeric and polymeric
acids, dimeric and polymeric glycerides, and free fatty acids (Nawar 1996).
These products cause changes that include increases in viscosity and in free fatty
acid content, development of a dark color, decreases in iodine value and surface
tension, changes in refractive index, and an increased tendency to foam (Nawar
1996). These changes have a direct effect on heat transfer during the frying
process. Moreira ef al. (1995)studied the degradation of soybean oil and found
an h, of 285 W/mzK for fresh oil and 273 W/m2K for used oil. In a similar
study of soybean oil degradation at 190C. Tseng ef al. (1996)found h, to
decrease nonlinearly from 274.4 to 250.0 W/mzK as oil quality decreased.
Hallstrom (1979)reports a study measuring h, in the 250-300 W/mzK range
before vaporization begins and increasing to near lo00 W/mzK during the
boiling phase of the frying process. Procedures for this particular study are
unknown although it was likely a numerical approach.
The lack of data and a method quantifying h, during the boiling phase
necessitate the need to develop a procedure to measure h, during this phase of
immersion frying, Heat transfer rates during this phase play a critical role in
forming the characteristic sensory properties of the product. Heat transfer
induces browning, such as caramelization and Maillard reactions, that impart
integral flavor and color properties. Crust qualities formed during the boiling
phase are a function of heat transfer rates and are important to the overall
texture of the product. The objective of this research was to develop a procedure
CONVECTIVE HEAT TRANSFER DURING FRYING 203

for determination of h, during the boiling phase of immersion frying. The


procedure would then be used to quantify the convective heat transfer rate
during the immersion frying process. This paper presents an improved procedure
and utilizes the theory first presented by Farkas and Stewart (1997).

THEORY

During immersion frying of foods two regions develop within the sample,
a crust and a core (Farkas ef al. 1996). These two regions are defined by the
liquid vapor interface, x = X(t), created within the sample due to evaporation
of water (Fig. 1). Water is vaporized at a rate N, from the interface and escapes
the system at a rate N," such that:
N:=N,-e,p,U

where U = dX/dt and E@,U is the accumulation of vapor due to formation of


the porous crust matrix. As frying progresses, the surface temperature (TI,.)
of the sample approaches the temperature of the oil (T,) and the core tempera-
ture (T,)approaches the boiling point temperature (Tb) of water. Energy is
transferred to the sample from the oil by convection at the oil/sample interface
(q,) and through the crust (q,) and core (qJ by a combination of conduction and
advection (Farkas ef al. 1996). At this point the complete system of equations
describing the coupled heat and mass transfer is highly nonlinear (Farkas ef al.
1996). Solution of the inverse heat transfer problem to yield an accurate
determination of h, as a function of time is difficult if not impossible. In
addition, no experimental data exist for comparison with theoretical values. The
followingmaterial outlines the derivation used for simplification of the complex
transport problem thus allowing the development of a laboratory method for
mapping h, over time during the frying process.

x =o x = X(1) x=L
CENTERLINE INTERFACE SIIRMC'E

FIG. 1 . SCHEMATIC OF FOOD SYSTEM DURING IMMERSION FRYING


204 L.J. HUBBARD and B.E. PARKAS

Applying a macroscopic energy balance to the cross-section in Fig. 1 yields the


following:

Energy gained by sample:


~ = ~ C r , - T 1 , . Jat x = L, t>O

Energy lost by sample:


= NVL(H,-hg) at x = L, t>O (3)
%L

This may be simplified through the use of order of magnitude analysis (Whitaker
1977) of Eq. (1) to show N,> >egrU and therefore N:=N, which leads to:

at x
qN,=Ny(Hy-hg) = L, t>O (4)

Energy accumulation in the form of sensible heat occurs in the crust and core
regions. The core region equation for energy accumulation was developed by
Farkas ef al. (1996) and is given as:

ocxc X(t), t>O

and energy accumulation in the crust region (Farkas et al. 1996):

Q l=(eypyCpy+p,e,cp,,)-a-r' X(t)cxcL, t>O


at

The term for energy accumulation due to the presence of oil in the crust,
E @ ~ C ~was
. not included by Farkas et al. (1996). The assumption was made that
the mass fraction of oil in the crust during frying was very small and therefore
contributed little to the thennal accumulation term. This assumption has since
been shown to be valid (Moreira et al. 1997; Utheil and Escher 1996) with the
bulk of oil uptake occurring in the period immediately following removal of the
material from the fryer.
The macroscopic balance yields:
CONVECTIVE HEAT TRANSFER DURING FRYING 205

Input of energy - Output of energy = Accumulation of energy

or:

At this point the determination of h,, although possible, remains difficult


and inaccurate due to the presence of the two accumulation terms. These terms
contain physical and thermal property data which are difficult to accurately
quantify and, moreover, change with time and position during the frying
process. Ideally, a method for determining h, would not rely on estimat-
ing/measuring these parameters.
Simplification begins with the elimination of the core region accumulation
term (tJT"/dt). This is accomplished by creating a negligible temperature gradient
in the core region (aT"/dt =aT"/ax= 0) prior to immersion frying thus allowing
Q" = 0. A laboratory procedure of preheating the sample in saturated steam
immediately prior to frying gave the desired results. Use of this procedure
allowed Eq. (8) to be reduced to:

rearranging yields:

The system was further simplified by order of magnitude analysis (Whitaker


1977) of the two terms on the right hand side of Eq. (10) to yield qN > >
I
Q1(L-X)throughout the frying process. That is, the output of energy from the
material is much greater than the accumulation of energy in the porous matrix
206 L.J. HUBBARD and B.E. FARKAS

or crust region. The order of magnitude analysis allows Q'(L-X) to be dropped


therefore simplifying the system to yield:

Recognizing that N, = (dm/dt)(l/A) and substituting into Eq. (11) yields:

Solving for hfp and defining H, - h, = AH,, yields:

Eq. (13) was used to map h, during the boiling phase of immersion frying. The
heat of vaporization of water (AH,,) and the surface area of the sample (A)
were assumed to remain constant for the duration of the frying process. While
it is recognized that the surface area does change during the frying process, the
change is small and not easily measured due to the contour of the surface. The
heat of vaporization at atmospheric pressure was found in steam tables (Singh
and Heldman 1993) while surface area, drying rate (dddt), oil temperature
(T&, and surface temperature of the sample ("1,.3 were measured using the
methods described below.

EXPERIMENTAL METHODS

Russet potatoes purchased from a local supplier were used as the sample
food system. Each of the smaller ends of the potato were cut away and a #12
core bore used to remove a cylinder ca 1.6 cm in diameter from the center of
the potato. The diameter of the potato cylinder was measured with a micrometer
and recorded. Each cylinder was trimmed to ca 9.0 cm, the final length
measured and recorded, and the raw sample weight recorded using a laboratory
scale (Model TWOOS, Ohaus Corp., Florham Park, NJ).
Before preheating the potato sample, a 40 gauge, type T thermocouple was
placed at the sample's surface to record surface temperature during frying. The
thermocouple was threaded at an angle along the axis of the cylinder, slightly
CONVECTIVE HEAT TRANSFER DURING FRYING 207

below the surface of the potato. This angle was followed until the thermocouple
just pierced the surface of the potato. The potato cylinder was secured to a rigid
anchoring apparatus used to immerse and hold the sample under the surface of
the hot oil during frying. The potato and the anchoring apparatus were then
placed in a vegetable steamer (Flavor Scenter Handy Steamer-. Black and
Decker, Hamstead, MD) to preheat the system and thus eliminate heat transfer
in the core region during frying.
Canola oil was used as the frying medium in all tests. Four and a half liters
of oil were heated to 180C in a laboratory scale fryer (Dazey Chefs Pot@,
Dazey Corporation, New Century, KS).The fryer was placed on an analytical
scale (Metler-Toledo SB 8001, Metler-Toledo, Greifensee, Switzerland)
interfaced to a 486 personal computer. A porous rigid material was placed
between the fryer and the scale platform to thermally shield the instrumentation
of the scale. A 24 gauge, type T thermocouple was used to record oil tempera-
ture during frying. Data from the scale and two thermocouples were recorded
at 5 Hz with Advantech data acquisition software and hardware (American
Advantech, Sunnyvale, CA). The sample and holder were removed from the
steamer, placed directly in the hot oil, and fried for 15 min. This procedure was
repeated for five samples. Prior to placing the sample into the fryer, the scale
was tared thus giving the initial weight of the anchoring apparatus and potato
sample. At this point any change in mass recorded by the scale is due to
material crossing the system boundary (Fig. 2).

SCALE

FIG. 2. SCHEMATIC OF FRYER, THERMAL SHIELD, AND SCALE

Upon placing the sample into the hot oil, vaporization of water occurs. The
loss of water vapor results in a decrease of the mass of the system, as water in
the form of vapor crosses the system boundary, allowing for the on-line
measurement of change in mass over time (dm/dt) of the potato sample.
208 L.J. HUBBARD and B.E. FARKAS

RESULTS AND DISCUSSION

The surface temperature was measured using a 40 gauge, type T


thermocouple. The thermocouple was placed at the potato to oil interface by
threading the wire through the potato cylinder. Position of the thermocouples
was observed at the end of each fry to insure the final placement of the probe.
If the thennocouple tip became embedded in the crust or pushed beyond the
surface by crust formation, the data set was discounted. These errors were also
cross-referend with the raw data. A surface temperature profile that failed to
approach the oil temperature confirmed embedding of the thermocouple. On the
other hand, a surface temperature profile that reached the oil temperature
indicated the thermocouple had been pushed beyond the surface of the potato.
The accepted surface temperature profile increased from the boiling point at time
zero to an asymptote less than the oil temperature in the first 200 s of the 900
s of processing. At the end of the process a temperature differential of
approximately 1OC was observed between the oil and sample surface. Raw
surface temperature data were noisy due to turbulence created at the surface as
vapor escaped from the potato. A nonlinear least squares curve was developed
using SAS statistical software (SAS Institute Incorporated, Cary, NC) to smooth
the raw data thus eliminating the inherent noise of the system from the data set.
The fitted curve (Fig. 3) was then used in calculation of h,.

100 I
0 120 240 360 480 600 720 840 9 0

Time, s

FIG.3. SURFACE TEMPERATURE OF POTATO CYLINDER DURING FRYING AT 180C


CONVECTIVE HEAT TRANSFER DURING FRYING 209

Change in mass of the frying system was recorded using an analytical scale
and associated data acquisition system. Change in mass of the system as a
function of time was recorded at 5 Hz.As previously stated, frying of foods is
an inherently noisy system due to rapid escape of vapor from the product. This
noise was translated to measurements made from the analytical scale. A
nonlinear least squares fit was again applied using SAS statistical software to
smooth the raw data. From this smoothing procedure an equation was developed
to describe a curve representative of the raw data (Fig. 4). The equation
developed from the raw data for sample mass as a function of time was then
differentiated to yield drying rate (Fig. 5).

0 120 2 i O 3 i O 4iO 6&l 7;O 8iO 9 0


Time. s

FIG. 4. CHANGE IN MASS OF SYSTEM DURING IMMERSION FRYING OF POTATO


CYLINDERS AT 180C

Surface area of the potato sample (0.00454 m2)and heat of vaporization of


water (2,257 kJ/kg) were then used in Eq. (13) with an oil temperature of 180C.
Note that oil temperature data were collected during the frying process and did
not vary more than the accuracy of the instrumentation and was thus assumed
to be constant in all calculations. Smoothed surface temperature and drying rate
curves were incorporated in Eq. (13) to yield the mapping of h, (Fig. 6) over
the frying process.
210 L.J. HUBBARD and B.E. PARKAS

'i
0

s 0

0.0 I

0.00 I l l I I I I
0 120 240 360 480 600 120 840 9 10
Time, s

FIG. 5. DRYING RATE DURING IMMERSION FRYING OF POTATO CYLINDERS


AT 18OC

0 120 240 360 480 600 720 840 960


Time, s

FIG. 6. MAPPING OF THE CONVECTIVE HEAT TRANSFER COEFFICIENT FOR


IMMERSION FRYING OF POTATO CYLINDER AT 180C WITH 95%CONFIDENCE
INTERVALS

Confidence intervals at the 95%! level were established around the average
value of h, (Zar 1996). The maximum value of h, recorded was 1 100 f 140
CONVECTIVE HEAT TRANSFER DURING FRYING 211

W/m2K. By plotting dm/dt and surface temperature (TI,,) on the same graph,
it was found that a maximum h, occurs at the intersection of these two curves
which corresponds to the point of maximum heat flow, qrp(Fig. 7), of 133 W.
This heat flow corresponds to a maximum heat flux of over 29,000 W/m2 (Fig.
8). The unique properties of fried foods are largely due to the high heat transfer
rates which lead to formation of the crust and characteristic flavors. In mapping
the heat flux over time it is clear that heat flux is not a constant value but
instead changes dramatically as a function of time. At this point it may be
hypothesized that fried foods attain their unique properties through the
characteristic heating cycle shown in Fig. 8. The initially high rates of heat
transfer serve to rapidly form the ridged crust matrix before collapse of the
dried layer, a common phenomenon in dehydration. A decreasing, yet still high,
rate of heat transfer then allows for cooking of the core region. The maximum
drying rate (Fig. 5 ) and dip in h, both occur after approximately 60 s of frying.
This phenomenon may be explained as the transition from surface boiling to
crust inception. The falling rate stage (Fig. 5 ) is clearly seen and corresponds
to the decrease in h, (Fig. 6).

0.06 . ,180
............. ..........,.....
- 160
- 140 u
Surface Temperature
- 2o,
- 100
- 80
- 60
- 40
- 20
0 00 I I I I I
0
6 I20 240 360 480 660 720 840 960
Time, s

FIG. 7. DRYING RATE, SURFACE TEMPERATURE AND HEAT FLOW DURING


IMMERSION FRYING OF POTATO CYLINDERS AT 180C
212 L.J. HUBBARD and B.E. FARKAS

FIG. 8. HEAT FLUX AS A FUNCTION OF TIME DURING IMMERSION FRYING OF


POTATO CYLJNDERS AT 180C

CONCLUSION

The convective heat transfer coefficient was quantified and mapped over
time during the boiling phase of immersion frying with values ranging from 300
to 1100 W/mzK. The initial value of 300 W/mzK is at the onset of the boiling
phase and is within the range of values found in the literature for the nonboiling
phase of immersion frying. The heterogeneity of food products and processes
mandates the need for a range, as opposed to a specific value, to be considered
when calculating process parameters. This research is significant in that it offers
for the first time a procedure for mapping the convective heat transfer
coefficient during the boiling phase of frying as a function of time as opposed
to assigning a single value. The ability to map h, is a key tool in understanding
the process of frying and will allow for more accurate heat flux calculations,
kinetic calculations, throughput determination, theoretical process efficiency
calculations, and eventually the development of alternative frying processes.

ACKNOWLEDGMENT

This research was supported by USDA NRI Competitive Grants Program


Award #96-35500-3467.
CONVECTIVE HEAT TRANSFER DURING FRYING 213

NOMENCLATURE

surface area (mz)


heat capacity of species i (J/kg K)
heat of vaporization of water (J/kg)
volume fraction of liquid phase (-)
volume fraction of oil phase (-)
volume fraction of vapor phase (-)
volume fraction of solid phase (-)
convective heat transfer coefficient (W/m2K)
enthalpy of species i (J/kg)
thermal conductivity of species i (W/mK)
sample thickness (m)
mass (kg)
mass of species i (kg)
flux of species i (kg i/mz s)
flux of vapor across solid oil interface (kg y/m2 s)
heat flux through the crust (W/mz)
heat flux through the core (W/mz)
heat flux at the surface due to convection (W/mz)
heat flux at the surface due to loss of vapor (W/m2)
energy accumulation in the crust region (W/m3)
energy accumulation in the core region (W/m3)
density of species i (kg of i/m3 of i)
time (s)
boiling point temperature ("C)
core region temperature ("C)
oil temperature ("C)
surface temperature ("C)
temperature within the crust region ("C)
temperature within the core region ("C)
Velocity of interface X(t) (m/s)
Position of crust core interface (m)
drying rate (kg/s)
temperature gradient ("C/m)

Subscripts 1 Superscripts
4 oil phase
Y vapor phase
(r solid phase
214 L.J. HUBBARD and B.E. PARKAS

B liquid phase
bP at the boiling point of the &phase
S at the surface
0 at time = 0

REFERENCES

DAGERSKOG, M. 1977. Time-Temperature Relationships in Industrial


Cooking and Frying. Physical, Chemical and Biological Changes in Food
Caused by Thermal Processing. p 77-100.
FARKAS, B.E., SINGH, R.P. and RUMSEY, T.R. 1996. Modeling heat and
mass transfer in immersion frying. I. Model development. J. Food Eng.
29,211-226.
FARKAS, B.E. and STEWART, H.E. 1997. Direct measurement of the
convective heat transfer coefficient during immersion frying. Paper 55- 1.
Abstract. National IF", Orlando, FL.
HALLSTROM, B. 1979. Heat and Mass transfer in Industrial Cooking. Food
Process Engineering 1, 457-465.
HOLMAN, J.P. 1990. Unsteady state heat conduction. Heat Transfer 6" Ed.
pp. 131-206.
MILLER, K.S.,SINGH, R.P. and FARKAS, B.E. 1994. Viscosity and heat
transfer coefficients for canola, corn, palm, and soybean oil. J. Food
Processing and Preservation 18, 461-472.
MOREIRA, R., PALAU, J. and SUN, X. 1995. Simultaneous heat and mass
transfer during the deep fat frying of tortilla chips. J. Food Process
Engineering 18, 307-320.
MOREIRA, R.G., SUN, X.Z.and CHEN, Y.H. 1997. Factors affecting oil
uptake in tortilla chips in deep-fat frying. J. Food Eng. 31, 485-498.
NAWAR, W.W. 1996. Lipids. In Food Chemistry, 3d Ed. (0.Fennema, ed.)
pp. 225-314, Marcel Dekker, Inc., New York.
SINGH, R.P. and HELDMAN, D.R. 1993. Introduction ro Food Engineering,
2"dEd. pp. 462.
TSENG, Y.,MOREIRA, R. and SUN, X. 1996. Total frying-use time effects
on soybean-oil deterioration and on tortilla chip quality. Intl. J. Food Sci.
Te~hn01.31, 287-294.
UFHEIL, G. and ESCHER, F. 1996. Dynamics of oil uptake during deep-fat
frying of potato slices. Food Sci. Technol. -Lebensmittel-Wissenschaft&
Technologie, 29, 640-644.
WHITAKER, S. 1977. Fundamental Principles of Heat Transfer. pp. 66,
Pergamon Press, New York.
ZAR, J.H. 1996. Biostatistical Analysis. 3d Ed. pp. 101. Prentice Hall, Upper
Saddle River, NJ.

Das könnte Ihnen auch gefallen