You are on page 1of 17

Applied Mathematical Modelling 31 (2007) 209225

www.elsevier.com/locate/apm

Crust density in bread baking: Mathematical modelling


and numerical solutions
D.R. Jeerson

a,1

, A.A. Lacey

a,*

, P.A. Sadd

a
b

School of Mathematical and Computer Sciences, Heriot-Watt University, Edinburgh EH14 4AS, UK
RHM Technology Ltd, The Lord Rank Centre, Lincoln Road, High Wycombe, Bucks HP12 3QR, UK
Received 1 November 2004; received in revised form 1 May 2005; accepted 30 August 2005
Available online 15 November 2005

Abstract
A model for the formation of a crust during bread baking is presented. The crust is the outermost part of the loaf
where the nal bread density is signicantly higher than in the crumb, the interior of the loaf. The model is based on a
collapse mechanism, whereby raised pressures due to thermal expansion and water evaporation squash bubbles in the
bread as the dough sets and fractures; the latter process allows vapour within bubbles to escape.
A preliminary analysis of some aspects of the model and an indication of the numerical solution are also presented.
Fuller results and their implications for bread manufacture will be given in a second paper.
2005 Elsevier Inc. All rights reserved.
Keywords: Breadcrust formation; Crust thickness; Crust density; Mathematical modelling of bread baking

1. Introduction
The texture and density of baked products such as bread and cakes, the manufacture of which is a major
commercial concern, is controlled by the way their rheology and vapour content change during the baking
process. The rising temperature causes thermal expansion of vapour and raises the saturation pressure of
water within the dough. This leads to a local expansion but at the same time forces compression and higher
densities elsewhere. The same temperature rises bring about fracture so that the bubbles, which are isolated
from each other in fresh dough, burst and the bubbly dough turns into a porous crumb or crust. The
porous structure allows a release of pressure and a consequent contraction and increased density of the material. At the same time, the dough mixture sets as gelatinisation of starch takes place. This change from a
viscous liquid to an elastic, or eectively rigid, solid results in the change in density being xed.
The rheological properties, e.g., viscosity and how this changes as setting occurs, are determined at least in
part by the quality of the our. Other key thermal properties (eective thermal conductivity, specic heat and
*

Corresponding author. Tel./fax: +44 1314513249.


E-mail address: a.a.lacey@ma.hw.ac.uk (A.A. Lacey).
Present address: Npower Limited, Oak House, 1 Bridgwater Road, Worcester WR4 9FP, UK.

0307-904X/$ - see front matter 2005 Elsevier Inc. All rights reserved.
doi:10.1016/j.apm.2005.08.017

210

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

density) are very much inuenced by the water content. Things like oven conditions, particularly its temperature, are of course also vitally important in controlling thermal eects within a baking loaf of bread. How
such features are varied inuence the interplay between pressurisation, setting and bursting and the consequent nal porosity, and density, of a loaf or cake.
This paper presents a model for the heat transport and density evolution with the particular aim of predicting the nal density variation near a loafs surface. This crust, the region where density is signicantly higher
than average, is of especial interest. With knowledge of how dierent quantities, such as water content or oven
temperature, can aect crust density and width, the bread manufacturers will be able to better adjust features
of the baking process to control the crust, making new products or the same products with dierent ours.
2. Overview of mathematical model
The process that is to be modelled mathematically is that of the movement of bread dough during baking,
which gives rise to a non-uniform density prole in baked bread. The current model incorporates a hypothesised notion of bubble collapse as the mechanism that causes this motion. This section describes this hypothesised mechanism, together with the other features incorporated into the model. Previous models of bread
baking (e.g., [1,2]) have considered the overall change in bread volume during baking. In [3] this is modelled
by considering the pressure-driven growth of a single isolated bubble in a dough. Although there is now a
fairly extensive literature on modelling of bread baking, this mostly concentrates on the crumb (for example
[46]). These papers also contain experimental measurements of temperature in bread, as, for example, does
[7]. In Section 6 we shall compare some of the temperature predictions of our model, taking an idealised spherical roll, with the experimental results of [7] (which were obtained for a more normally shaped loaf). There has
also been some work on browning and drying near the surface, [810], and a lumped model for moisture variation appears in [11]. Here we are interested in how the interaction between expanding and collapsing bubbles
produces a dense crust.
2.1. The bubble collapse mechanism
Baked bread is most dense towards its surface and in this work the word crust refers to this dense region.
On average the density decreases monotonically as the centre of the bread is approached. The process of interest is therefore one which acts most intensely at the bread surface, and whose eect tails o towards the bread
centre.2
At the start of baking the dough has a bubbly structure with an approximately uniform (macroscopic) density.3 The material surrounding the bubbles (referred to as the dough matrix) is initially liquid. During baking
the matrix changes its material properties and eventually sets. Also, as the temperature inside the dough
increases, the pressure inside the bubbles rises (due substantially to water evaporating from the matrix). This
rise in pressure will cause regions in the dough to expand, resulting in the expansion of the dough as a whole or
the contraction of the cooler inner part of the dough. This movement will be resisted by the setting of the
matrix, as well as external constraints such as a tin.
These pressure eects alone will not give rise to a higher density towards the bread surface. The increase in
density is understood to be a consequence of the fracture of bubbles. Fracturing refers to the rupture of part of
the matrix surrounding a bubble, causing the bubble to become connected to (and have the same internal pressure as) adjacent bubbles. This fracturing of the bubbly structure is important for the achievement of familiar
properties of baked bread such as its ability to soak up liquid. The cause of fracturing can be understood to be
the changing material properties of the dough, combined with the increasing pressure within bubbles. Since
the dough is heated from the outside, the bubbles at the surface will fracture rst and the pressure inside them
will become equal to the oven pressure (which will subsequently be referred to as atmospheric pressure).4 As
baking proceeds, bubbles progressively further away from the bread surface will fracture, and these will
2
3
4

For an overview of the breadmaking process and the changes undergone by bread dough during baking see [12,13,7].
This bubbly structure is formed by the mixing and proving of the dough, see [13].
When bread is baked in a microwave oven, it is not heated from the outside in this way and does not develop a crust.

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

211

become part of a network of pores that is connected to the bread surface. Consequently, upon fracture, the
pressure inside a bubble will become atmospheric. If the dough matrix surrounding the fractured bubble is
still suciently liquid, it can then squash. This can be understood as the inner, unfractured dough squashing
the fractured, liquid bubble against some outer conning barrier. This could either be the bread tin, or previously set dough.5 The amount by which a bubble collapses is thus determined by the extent to which the
surrounding matrix is set when it fractures, so the relationship between setting and fracturing is crucial. If
the matrix surrounding a bubble is completely set when a bubble fractures, the outward pushing of the inner
part of the dough will have no squashing eect. If the bubble is still partly liquid, it will be at least partly
squashed. The size of the bubble is a factor here. Upon fracture, a large bubble can potentially collapse much
more than a small bubble.
The increase in the density of baked bread towards its surface can thus be understood as follows. When
bubbles near the surface reach a temperature at which they are likely to fracture, the temperature gradient
is high, so there is cooler dough nearby which is still liquid, and this allows collapse to take place. Later
on, when bubbles nearer the centre reach a temperature at which they are likely to fracture, the temperature
gradient is lower. There is less liquid dough nearby and less collapse takes place. Rheological measurements
carried out at RHM Technology Ltd indicate that the fracture temperature is just a few degrees above the
setting temperature, and noticeably lower than 100 C. However, in Section 6 we do some simulations with
a fracture temperature above 100 C; an unrealistically thin and light crust is seen to result. (Strictly speaking,
the larger gap between setting and fracturing temperatures invalidates some of the approximations used in our
model, but the qualitative feature of an exceedingly weak crust will not be aected.)
This mechanism is incorporated into the current model in an idealised form. It is assumed that the dough
matrix sets instantaneously at a particular temperature and fractures at a particular temperature, the fracturing temperature being slightly higher than the setting temperature. When the outer edge of a bubble (the edge
nearest the bread surface) reaches the fracturing temperature, the bubble is assumed to fracture and the pressure inside the bubble becomes atmospheric. If the temperature gradient at this position inside the bread is
suciently high, a part of the matrix surrounding the bubble towards the bread centre will still be liquid.
It is assumed that this part instantaneously collapses, so that a spherical bubble becomes a spherical cap,
as represented in Fig. 1 [15].
Bubbles are taken to be spherical,6 and from Fig. 1 it is clear that the amount by which a bubble collapses is
simply determined by how the distance between the isotherms corresponding to the setting and fracturing temperatures compares with the size of the bubble. It is described later on how the mathematical model is written
down as though the bread dough were a homogeneous material without any small-scale bubbly structure. In
this homogenised model, bubble collapse is taken to occur at a temperature characteristic of the setting and
fracturing of the dough matrix, which is referred to as the collapse temperature. The amount of bubble collapse is calculated from the temperature gradient at the isotherm corresponding to the collapse temperature,
the porosity of the uncollapsed dough at that point and the distribution of bubble sizes. This calculation gives
a collapsed density.
The isotherm corresponding to the collapse temperature thus appears in the model as a moving boundary
(which we refer to as the collapse boundary), at which there is a jump in the bread density (from the uncollapsed to the collapsed density). This moves towards the centre of the bread as baking progresses. In terms
of what is taking place physically in the dough, this boundary can be imagined as a layer of collapsing bubbles
advancing towards the bread centre. At any particular time the collapsed density at this boundary is determined by the temperature gradient on the inner (uncollapsed) side of the boundary, and the other factors mentioned above. Once collapsed the dough does not move, and the nal density prole is given by the history of
the collapsed density at the collapse boundary as this boundary moves through the bread.
5

It was initially anticipated that the collapse of fractured bubbles might be due to the eect of the surface tension of the dough, but
calculations indicate that the time scale on which surface tension can have an eect is too long. Having understood the collapse of bubbles
to be the result of squashing against some obstacle, the question arises as to what this obstacle might be on an unconned dough surface at
the start of baking. Here it may be that the skin which forms on the surface of a proved dough might play an important role, but this has
not been investigated. Some experimental work on bubble fracture in bread dough is described in [14].
6
Bubbles are taken to be spherical when calculating bubble collapse, but this is not a crucial assumption.

212

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

Fig. 1. Diagram of the idealised bubble collapse mechanism.

In addition to the jump in density, other conditions which apply at the collapse boundary are continuity of
temperature and continuity of heat ow (which follows from consideration of conservation of both heat and
mass).
2.2. Inner dough movement
In the current model, dough movement is governed entirely by factors in the locality of the collapse boundary (temperature gradient, bubble size, etc.). The specics of what is taking place in the inner part of the dough
do not aect bubble collapse, which is determined entirely by the argument represented in Fig. 1 and happens
instantaneously. However, the movement of the inner part of the dough is important in so far as it aects the
thermodynamics. In the current model it is assumed that the volume of the bread is xed, and that the inner
part of the dough (the part inside the collapse boundary) expands uniformly so as to ensure conservation of
mass, given the amount of collapse at the collapse boundary.
2.3. Moisture vaporization and thermodynamics
As mentioned before, heat transfer inside the dough is modelled as though the dough were a homogeneous
material. Where moisture is present, an important process that takes place on the scale of individual bubbles is
the evaporation and condensation of water inside bubbles. Moisture evaporates at the hotter end of a bubble,
absorbing latent heat, and condenses at the cooler end, releasing latent heat. This is a mechanism for heat
transfer, which intensies at higher temperatures. In the heat transfer equations used in the model, this eect
appears as a temperature-dependent thermal conductivity.
In addition to the collapse boundary, the model incorporates another moving boundary at which water
vaporizes. This corresponds to a particular isotherm (at a temperature higher than the collapse temperature)
and its movement is governed by a Stefan condition.
The two moving boundaries (after they have appeared in the dough) divide the bread into three regions, in
which the heat-transfer equations have slightly dierent forms. The central part of the dough is moving, giving
rise to an advection term in the corresponding heat equation. In addition, thermal conductivities, specic heats
and densities change across moving boundaries.
The model thus consists of two moving boundaries coupled together with various forms of the heat equation. An appropriate law for heat transfer is applied at the bread surface. As described later on, the model is
solved in a spherically symmetric geometry. The model is summarised in Fig. 2.

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

213

Heat from oven


Vaporization boundary
Collapse boundary

Expanding
inner dough

r=rc T=Tc
r=rv T=Tv

Fig. 2. Diagram of spherically symmetric model of a loaf during baking showing moving boundaries and dough movement.

Note. Throughout this text we refer to the two sides of a moving boundary as the inside and the outside. By
the inside, we mean the side towards the bread centre, and by the outside, we mean the side towards the bread
surface.
3. The equations of the model
Here we formulate the model as a set of partial dierential equations governing heat transfer inside the
bread, coupled with a set of ordinary dierential equations governing moving boundaries (relating to bubble
collapse and to moisture vaporization) and inner dough expansion.
As explained in Section 2.1, the bread density after bubble collapse depends the distance between two isotherms (the setting and fracturing isotherms). It is clear that this will also depend on the size of the collapsing
bubbles. It is assumed that bubble diameters are initially distributed log-normally, and that as the inner part of
the dough expands, all bubbles in that part expand in proportion to one another. The ratio of the diameter of
a bubble at time t to its original diameter is denoted by a(t) (and is the same for all bubbles). At the collapse
boundary (dened in Section 2.1), there is a jump in the density. The outer density at the collapse boundary is
^t and the inner value is the central density q(t) = qm(1  /(t)), where qm is the density
the collapsed density q
of bubble-free (wet) dough and / is the porosity of the central uncollapsed part of the dough. The collapsed
density is given by
^t
q

qt
.
1  /tmt

3:1

Here m is the volume of bubble collapse per unit uncollapsed bubble volume. On the assumption that the
bubble diameters are initially distributed log-normally and expand as described above, m has the following
form:





1
ln f  m  3s2
ln f  m  s2
2 2m4s2

p
p
mt m ft
erfc
erfc
 3f e
2
s 2
s 2


ln f  m
2
p
2f3 e3m9s =2 erfc
;
3:2
s 2

214

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

where

ft dT

 !
oT 
at 
.
or rc 

Here dT denotes the specied dierence between the fracturing temperature and the setting temperature, and
the subscript means that the temperature gradient is taken on the inside of the collapse boundary (rc denotes
the collapse boundary position). The parameters m and s relate to the mean l and variance r2 of the bubble
diameters through
!
 2

l2
r
m ln p ; s2 ln 2 1 .
l
r2 l2
The expression (3.2) is derived as follows. The log-normal distribution has the probability density function
!
2
1
ln x  m
px p exp
.
2s2
xs 2p
When a bubble of diameter a(t)x (i.e., a bubble whose diameter at the start of baking was x) collapses according to the hypothesised mechanism described in Section 2.1, its uncollapsed volume v(x) = p(ax)3/6 loses a
spherical cap of volume f(dr,ax) = p((ax)3/6  axdr2/2 + dr3/3), where dr is the distance between the setting
and fracturing isotherms (see Fig. 1). The volume of collapse per unit bubble volume m is therefore given
by the ratio of the average volume lost to the average uncollapsed bubble volume, i.e.,
Z 1
Z 1
Z 1
Z 1
f dr; axpx dx
vaxpx dx
f dr=a; xpx dx
vxpx dx mdr=a.
0

0
1

This gives (3.2), in which dr has been approximated by dT oT =orjrc  . Note that dierently shaped bubbles
would give rise to a dierent function m*(f), but one with similar properties.
In the current model, the above equations (along with the equations for the temperature T) completely
determine the dough movement, the expansion of the inner part of the dough being assumed to be uniform
and related to the above so as to ensure conservation of mass. The total mass of the loaf is given by
Z R
4
~r dr;
M pr3c tqt 4p
r2 q
3
rc t
^t). Dierentiating
~r, rc(t) 6 r 6 R, is the density prole outside the collapse boundary (~
qrc t q
where q
this with respect to t and equating the result to zero gives
dq
3
drc

^
qt  qt
;
dt rc t
dt

3:3

where drc/dt denotes the velocity of the collapse boundary (discussed below). The initial value of q is the initial
density of the loaf (assumed uniform). Bubbles in the central part of the dough are assumed to grow in such a
way that the above change in density results, and this implies
da
qm at
dq

dt 3qtqt  qm dt

3:4

(this is obtained using the fact that all volume change is due to a change in bubble volume and the volume of
bubble free dough remains xed). Since a(t) is the ratio of the size of a bubble to its initial size, a(0) = 1.
In the wet part of the dough, heat transfer due to evaporation and condensation of moisture gives rise to an
eective thermal conductivity k(T) (see [4,16,12]). Experimental evidence [17] suggests that this conductivity
has the form
kT j1 j2 ej3 T .

3:5

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

215

The thermal conductivity is assumed to have this form throughout the wet part of the dough, although in
~1 , etc. In
the collapsed, wet dough it is assumed that the conductivity ~k can involve dierent constants j
the dry part of the dough (the part outside the vaporization boundary) the conductivity k is taken to be
constant.
In the central part of the dough (inside the collapse boundary) the heat equation (given a spherically symmetric geometry) takes the form


 2


oT
oT
oT
2 oT o2 T
vt; r
2 ;
qtc
kT
k 0 T
3:6
ot
or
or
r or or
where k 0 denotes the derivative of k with respect to T and c denotes the specic heat of wet dough. Up until the
point when the collapse boundary appears at the bread surface the velocity v is zero and the density q is constant and the above equation holds throughout the dough. The initial temperature of the dough is taken to be
uniform with a value characteristic of that of dough after proving. After the collapse boundary appears, the
velocity v of the central part of the dough is given by (3.11) below. In the part of the dough between the collapse boundary and the bread surface (or the vaporization boundary once it appears) the heat equation is
given by
 2


oT ~0
oT
2 oT o2 T
~
~rc
k T

q
kT
;
3:7
ot
or
r or or2
^t). Once the vaporization bound~r is the density prole outside the collapse boundary (~
where q
qrc t q
ary appears at the bread surface, the heat equation between this boundary and the bread surface is given by


oT  2 oT o2 T
k

qrc
1  mw ~
.
3:8
ot
r or or2
Here mw denotes the mass fraction of water in the wet part of the dough, and c is the specic heat of moisture
free dough, which relates to the specic heat c through
c mw cw 1  mw c;
where cw is the specic heat of water.
Conservation of heat at the moving boundary r = rc implies




oT 
drc
oT 
drc
~
^tc
.
kT c  qtc vt; rc 
kT c   q
or
or
dt
dt
rc 

rc

Here the subscripts rc and rc+ are used to denote the temperature gradients at the collapse boundary on
the inside and outside, respectively, and Tc denotes the temperature at the collapse boundary (which is a
parameter of the model). Conservation of mass at this boundary leads to


drc
drc
qt vt; rc 
.
3:9
^
qt
dt
dt
It follows that

oT 
kT c 
or

rc 

~kT c


oT 
.
or rc

3:10

Since the dough inside the collapse boundary is taken to be uniformly expanding, its velocity varies linearly
from zero at the bread centre to v(t, rc) at the collapse boundary, so from (3.9) we have


^t drc
r
q
1
; 0 6 r 6 rc t.
vt; r
3:11
rc t
qt dt
The movement of the vaporization boundary is governed by a Stefan condition:


drv ~
oT 
oT 
~rv L
mw q
kT v   k  .
or
or
dt
rv 

rv

3:12

216

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

Here L denotes the latent heat of evaporation of water and rv and Tv denote respectively the position of and
temperature at the vaporization boundary (this temperature again being a parameter of the model).
Finally, heat transfer from the oven is represented using a surface boundary condition of the form
KT

oT
hT a  T ;
or

r R;

3:13

where R is the loaf radius, h is a surface heat-transfer coecient, Ta is the ambient (oven) temperature and
K(T) is given by one of k(T), ~kT or k, depending on T (which here is the temperature at the bread surface).
A symmetry condition holds at the bread centre:
oT
0;
or

r 0.

3:14

4. Asymptotic moving boundary behaviour


Given the model parameters, and the spatial derivatives of the surface temperature up to second order at
the point in time when the moving boundaries appear at the bread surface, it is possible to calculate the leading-order behaviour of these boundaries analytically using asymptotic analysis. It is, for example, possible to
work out the nal density gradient at the bread surface in terms of these quantities.
4.1. Collapse boundary asymptotics
In making these calculations, it is convenient to work in terms of the spatial variable x = R  r, i.e., the
distance from the bread surface. The time t = 0 is taken to be the time at which the collapse boundary appears
at the bread surface. We think of the solution in terms of two functions T(t, x) and T~ t; x, where T~ satises
(3.7) and the boundary condition (3.13) and T satises (3.6) and the initial condition T(0, x) = T0(x) where T0
is the temperature prole at the time when the collapse boundary appears at the bread surface. Temperatures
T and T~ are related through (3.10) (with T replaced with T~ on the right-hand side). We denote the collapse
boundary position by s(t) = R  rc(t). For simplicity, we shift the temperature to zero at the collapse boundary, i.e.,
T t; st T~ t; st 0.

4:1

We consider the behaviour of T and T~ over a short time period after the appearance of the collapse boundary by rescaling time as t = s, 0 <   1. We consider the behaviour in a small region near the bread surface
by rescaling space to obtain balance in the heat equation: x = 1/2n. The surface boundary condition (3.13)
implies that T and T~ will be linear in x to leading order, and T and T~ are rescaled accordingly: T = 1/2h,
h. The boundary position is rescaled as s = 1/2r. The rescaled dependent variables are written as
T~ 1=2 ~
asymptotic expansions:
hs; n  h0 s; n 1=2 h1 s; n h2 s; n   
4:2
~
and similarly for h and r.
^, k and
To nd the behaviour to the orders we are concerned with here it is sucient to set the variables q, q
~k appearing in (3.6), (3.7) and (3.13) at their values at t = 0, x = 0, which we denote by q0, q
^0 , k0 and ~k 0 (that
this is justied becomes apparent once the leading-order solution has been determined). Substituting these
expressions into the dierential equations and boundary conditions gives a collection of problems for h0,
~
h0 , r0, etc. Solving these leads to
h0 s; n cn;

~
h0 s; n ~cn;

r0 s 0;

with c

h
h
T a  T c ; ~c T a  T c
~k 0
k0

4:3

and it is seen that the leading-order collapse boundary movement is not picked up so we must consider a smaller region in space. We adopt the new scaling of the spatial and dependent variables: x = y, T = u, T~ ~u,

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

217

s = v. Revised problems, given by substituting asymptotic expansions of the form (4.2) for u, ~u and v into
(3.6), lead to
u0 s; y cv0 s  y;

~
u0 s; y ~cv0 s  y;

4:4

where v0 is still to be determined. The following terms u1 and ~u1 are zero, while the next give
 


1 1
dv
2c 2
u2
c^
q0 c 0  c2 k 00 
y for large y.
2 k0
R
ds

4:5

This must match with the quadratic variation in the h1 term in the earlier asymptotic expansion temperature.
This can be found to satisfy a problem of the form:

oh1
o2 h1
oh1
n2
o2 T 
K 2 gs;
s; n 0; h1 0; n g ; with g 2  .
4:6
ox t0
os
on
2
on
x0

Then

 Z s
n2
gs ds
h1 s; n g Ks

2
0
and matching (4.5) with this leads to




dv0
1
2 g def 0
0

ck 0 k 0
s0
^0 c
R c
ds
q

and

v0 s s00 s.

Proceeding further in this manner and using



oT 
def
bt  c b00 t   
ox 

4:7

4:8

yst

^:
eventually leads to an initially linear variation of the collapsed density q
"
#
q0
q00
q / m0 f0 /00 m0
^t 
q

0 0 0 0
t 
2
1  /0 m0
1  /0 m0
1  /0 m0

4:9

Here the subscript zero indicates that the quantities are evaluated at t = 0 (the time at which the collapse
boundary appears at the bread surface). The derivative m00 of m (m is given by (3.2)) with respect to f can be
evaluated explicitly and the derivative f00 of f(t) = dT/a(t)b(t) can be found using (4.8) and (3.4). The derivatives of q and / = 1  q/qm are given by (3.3). Note that the initial value of drc/dt is s00 . The leading-order
behaviour of the collapsed density at the bread surface is then given by (4.9).
4.2. Vaporization boundary asymptotics
The asymptotic behaviour at the vaporization boundary is calculated in a similar manner to that at the collapse boundary. Here we give an outline of this calculation. The behaviour here is governed by the heat equations (3.7) and (3.8), the Stefan condition (3.12) and the boundary condition (3.13).
For these calculations t = 0 is taken to be the time at which the vaporization boundary appears at the bread
surface. The temperature inside the vaporization boundary is denoted by T~ and that outside the vaporization
boundary by T . In general, breve is used to denote quantities between the bread surface and the vaporization
 denotes the density in that region). In this section the temperature is shifted so that it is zero
boundary (e.g., q
at the vaporization boundary. As before, the equations are written in terms of the spatial coordinate x = R  r
and the vaporization boundary position is denoted by st r  rv t.
As for the collapse boundary calculations, space and time are rst rescaled so as to obtain balance in the
heat equation, and temperatures, which are linear in x to leading order, are rescaled accordingly, as is the
vaporization boundary position:
t s;

x 1=2 n;

h;
T~ 1=2 ~

T 1=2 
h;

s 1=2 r
.

218

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

The dependent variables are written as asymptotic expansions of the form


~
hs; n  ~
h0 s; n 1=2 ~
h1 s; n ~
h2 s; n    ;
etc.
Similarly to the collapse boundary calculations we nd that
~
h0 ~cn;


h0 cn;

4:10

where
~c

h
T  T v ;
~k 0 a

c

h
T  T v
k 0 a

0 0. Using (4.10) we nd from


(~k 0 is now the conductivity evaluated at Tv rather than Tc), and we nd that r
1 0. This implies that ss 1=2 r
s is of order 3/2 as  ! 0. To capture the
the Stefan condition that r
leading-order dynamics near the vaporization boundary, we look at a smaller region in space using the rescaling x = 3/2y. The dependent variables are rescaled as T~ 3=2 ~u, T 3=2 u, s 3=2 v.
We nd from the heat equation (3.8), the surface boundary condition (3.13) and the requirement that the
temperature is zero on the vaporization boundary that

v0 s  y.
u0 s; y c
The heat equation (3.7) and the Stefan condition then give that
~
v0 s  y.
u0 s; y ~c
From the heat equation (3.8) and the surface boundary condition (3.13) we nd that u1 0, and the heat equation (3.7) implies that
~
4:11
u1 s; y g1 sy g2 s.
~ with that of ~
Matching the expansion of h
u and using what has already been calculated we nd that ~h1  g1 n as
n ! 0 so ~
h1 0 at n = 0. Solving the heat equation for ~h1 obtained from (3.7) with this boundary condition
and the initial condition ~
h1 0; n ~
gn2 =2 (where ~g o2 T~ =ox2 at t = 0, x = 0), we nd that
0 s
!1
 2
~
D
s
n
n
0
~
~ 2 ~
~ @2n
~ 0 serf p A;
h1 s; n wn
exp
g=2 w
n2 2D
4:12
~ 0s
p
~ 0s
4D
2 D
where
~
~ 0 k0 ;
D
^0 c
q

~k 0 ~c2 ~c
~
w 0 .
2~k 0 R

The coecient g1 is then given by the coecient of n in the Taylor expansion of (4.12), i.e.,
s
~
~ D0 s.
g 4w
g1 s 2~
p
Using (4.11), the fact that 
u1 0 and the Stefan condition, we nd that v0 = us3/2, where
p
~ D
~0
4~
g 8w
p
u
^0 L p
3mw q
and we have that the vaporization boundary position and the temperature between the vaporization boundary
and the bread surface are given by
st  ut3=2    ;

T t; x  cut3=2  x   

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

219

5. Numerical solution
The model has been solved numerically using a method of lines approach. The heat equations presented in
the previous section were spatially discretised using usual central dierence approximations for the spatial
derivatives. The boundary conditions (3.13) and (3.14) are incorporated using ctitious points. This gives rise
to a system of ordinary dierential equations which can be solved up to the point where the collapse boundary
appears at the bread surface using a standard ODE solver such as ode15s in Matlab, and the solution at the
point when the collapse boundary appears can be captured using Matlabs event location capability.
Once the moving boundaries appear in the problem, it becomes necessary to move the spatial mesh points
with the moving boundaries in order to allow the formulation of the problem as a system of ODEs without
jumps (this would not be possible if moving boundaries were to cross mesh points). Spatial mesh points were
taken to be equally spaced between the moving boundaries and the boundaries of the domain (see Fig. 3).
We use a front tracking approach to the calculation of the movement of moving boundaries [18]. When a
moving boundary rst appears at the bread surface, it is necessary to use some additional approximation in
order to calculate its initial movement until there is a sucient gap between the moving boundary and bread
surface to allow the insertion of grid points in this gap. These starting o stages in the calculation of the solution are marked 2A and 3A in Fig. 3. The solution during these stages could be approximated using the results
of the asymptotic analysis carried out in Section 4. Here we use a dierent approach, and check that the result
agrees with the results of Section 4. To approximate the initial movement of the collapse boundary, we continue to solve the problem corresponding to the stage before the appearance of the collapse boundary (stage 1
in Fig. 3) and nd the position of the collapse boundary (which is the isotherm T = Tc) by interpolation. When
this is suciently far from the bread surface, grid points can be inserted in the gap. Values of temperature at
these grid points are set by linear interpolation in such a way that the condition (3.10) is not violated. The
values of density at these grid points (and at the bread surface) are set at the collapsed surface density, which
is given by (3.1), using the temperature gradient given by the surface boundary condition (3.13) with T = Tc.
After stage 2A, densities at grid points between the vaporization boundary and the bread surface subsequently become variables in the system of ODEs (since here the grid points are moving through a region where
~i at a particular mesh point at ri is simply taken to satisfy
the density is not constant). The density q
~i1
~ q
d~
qi
q
vi i1
;
dt
2dr

r
Bread surface

Collapse boundary

Vaporization boundary

0
1

2A

2B

3A

3B

Fig. 3. Representation of the way that the movement of spatial mesh points is coupled with the movement of the moving boundaries in the
method of lines scheme. The continuous lines represent spatial mesh points and the dashed lines represent the moving boundaries. The
vertical lines show the way that the calculation of the solution is divided into various stages.

220

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

where vi = dri/dt is the velocity of the mesh point at ri, and dr is the relevant mesh spacing. In the case of the
~i1 will be the collapsed denrst mesh point beyond the collapse boundary (i.e., towards the bread surface), q
^, which is determined by the temperature gradient inside the collapse boundary,
sity at the collapse boundary, q
whose calculation is described below.
Aside from the starting o period, the movement of the collapse boundary is also calculated using a fronttracking method. The velocity of the collapse boundary is given by
 , 
drc
oT 
oT 
;
5:1
 
ot rc  or rc 
dt
where the time derivatives on the right-hand side are taken at the collapse boundary but xed in space (i.e., not
moving with the collapse boundary). Consideration of conservation of mass shows that the time derivative on
the inside of the collapse boundary is given by
 !

 !2

oT 
oT 
2 oT 
o2 T 
0
^tc  k T c
q
kT c

.
ot rc 
or rc 
rc or rc  or2 rc 
The derivatives on the right-hand side of (5.1) must be taken either both on the inside of the boundary, or both
on the outside, and in either case the result must be the same. Together with (3.10) this implies that


oT 
oT 
kT c  ~kT c  .
5:2
ot
ot
rc 

rc

The thermodynamics at the collapse boundary are approximated by introducing two ctitious points, one
either side of the collapse boundary, which are then used in approximating the temperature gradients either
side of the boundary. This introduces two further unknowns into the problem, which are eliminated using
the discretised versions of (3.10) and (5.2). The nonlinearity resulting from the temperature-dependent thermal
conductivity means that this amounts to nding the well behaved root of a quadratic equation (i.e., the root
that does not become large for reasonable parameter values). In this way, the time and spatial derivatives at
the collapse boundary are calculated and (5.1) is incorporated into the system of ODEs. The spatial derivative
inside the collapse boundary is also used in the calculation of the collapsed density (3.1).
The movement of the vaporization boundary (determined by (3.12)) is initially approximated (in stage 3A
in Fig. 3) by xing the heat ow outside the vaporization boundary at its value at the point when the boundary
appears at the bread surface. The heat ow inside the vaporization boundary is calculated by using the temperatures at the two grid points inside the boundary to determine the temperature gradient at the boundary,
and (3.12) is incorporated into the system of ODEs. Subsequently (in stage 3B), the temperatures at the two
grid points outside the vaporization boundary are used to determine the heat ow outside the boundary (the
use of two grid points either side of a moving boundary to calculate the temperature gradients at the boundary
in a Stefan problem is again a front-tracking method as described in [18]).
Renement of the accuracy of the calculations (both by increasing the number of spatial grid points and
decreasing the time stepping error tolerances) indicated convergence of the calculated temperature and density
proles.
6. Preliminary results
Most of the graphs shown here result from numerical simulations with the parameter values7 given in Table
1. Fig. 4 shows the temperature prole near the time when the collapse boundary reaches the bread centre. The
position of the vaporization boundary (at 100 C) is clearly visible. Fig. 5 shows the evolution of temperature
at the bread surface, in which the fall in the rate of temperature growth at the onset of vaporization can be
seen. Fig. 6 shows the nal density prole. Here the density is given as a proportion of the initial density.
The drop in density resulting from the absence of water outside the vaporization boundary has been disre7

The values of the conductivity parameters have been chosen to t experimental data [17].

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

221

Table 1
The model parameter values used to produce the results shown
Initial porosity
Density of gas-free dough
Dry specic heat
Dry thermal conductivity
Convection coecient
Bread radius
Oven temperature
Collapse temperature
Vaporization temperature
Initial temperature
Mean bubble diameter
Bubble mean/standard deviation
Fracturing temperaturesetting temperature
Latent heat of vaporization
Initial mass fraction of water
Specic heat water
Conductivity parameters (see (3.5))

/0 = 0.8
qm = 1250 kg m3
c 1680 J kg1 K1
k 0:055 W m1 K1
h = 35 W m1 K1
R = 0.05 m
Ta = 200 C
Tc = 80 C
Tv = 100 C
T0 = 40 C
m = 0.001 m
br = 1/2
dT = 8 C
L = 2.272 MJ kg1 K1
mw = 0.45
cw = 4187 J kg1 K1
~1 0
j1 j
~2 0:04
j2 j
~3 0:0256
j3 j

Temperature profile after 13.5 minutes of baking


200

180

Temperature (C)

160

140

120

100

80

60

10

20

30

40

50

Distance from bread surface (mm)

Fig. 4. Temperature prole after 13.5 min of baking for the parameter values specied in Table 1.

garded. At around 3 mm from the bread surface, the density starts to drop more rapidly. This is the result of
the onset of vaporization, which slows down heat transfer to the region on the inside of the vaporization
boundary, causing the temperature there to level o more quickly. In the current model, this eect is one
of the most signicant in determining the crust prole. By carrying out simulations for various parameter values it is possible to gain an indication of the relative importance of dierent factors in determining crust properties. For example, the current model suggests that crust thickness and density depend more sensitively on the
temperature at which bubbles burst than the initial oven temperature. Figs. 79 show that the numerical
results agree well with the asymptotics from Section 4, indicating validity of the numerics. Fig. 10 shows,
for a bread loaf of radius 5 cm, both experimental8 and theoretical results for a crust-density prole. To
get the best t, a higher than normal boiling point was needed; such changes can be attributed in practice
to the presence of sugar or salt, for example, in the dough mixture.
8

Crust densities were measured by weighing a series of thin slices of bread crust (slices were taken parallel to the at crust surface) and
dividing these weights by the volume of each slice.

222

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225


Temperature evolution 0 mm from bread surface
200
180

Temperature (C)

160
140
120
100
80
60
40

6
8
Time (minutes)

10

12

14

Fig. 5. The evolution of temperature at the bread surface for the parameter values specied in Table 1.

Final density profile (excluding effect of water vaporization)


2.6

Density as a proportion of original density

2.4
2.2
2
1.8
1.6
1.4
1.2
1
0.8
0.6

10

20
30
Distance from bread surface (mm)

40

50

Fig. 6. Final density prole for parameter values specied in Table 1 (after 13.5 min of baking the collapse boundary was within 5 mm of
the bread centre).
50
45

Distance from surface (mm)

40
35
30
25
20
15
10
5
0

100

200

300

400
500
time (s)

600

700

800

900

Fig. 7. Collapse boundary position. The solid line is the numerically calculated solution and the dashed line is the asymptotic solution.

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

223

650

Density (kg/m3)

600

550

500

450

400
0

0.2

0.4
0.6
0.8
1
Distance from surface (mm)

1.2

1.4

Fig. 8. Density. The solid line is the numerically calculated solution and the dashed line is the asymptotic solution.

0.18

Distance from bread surface (mm)

0.16
0.14
0.12
0.1
0.08
0.06
0.04
0.02
0

50

100

150
200
250
t3/2 (seconds3/2)

300

350

400

Density as a proportion of central density

Fig. 9. Vaporization boundary position. The solid line is the numerically calculated solution and the dashed line is the asymptotic
solution.

500
Numerical results
Physical measurements

400

300

200

100

10

20

30

40

50

Distance from surface (mm)

Fig. 10. Comparison of the calculated nal density with experimental results.

Fig. 11 displays temperaturetime simulations at three points (centre, 2 cm from the surface, and close to
just 1 mm fromthe surface) for a 5 cm-radius spherical roll, corresponding to the measurements in [7] for a

224

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

1mm from surface


10mm from surface
50mm from surface

120

Temperature (oC)

100
80
60
40
20
0

10

12

14

16

18

20

Time (s)

Fig. 11. Temperaturetime histories near the surface (solid line), 1 cm from the surface (dots) and at the centre (dashes) for a 5 cm-radius
loaf, c.f. [7].

Final density profile (excluding effect of water vaporization)

Density as a proportion of original density

1.035
1.03
1.025
1.02
1.015
1.01
1.005
1
0.995

10

20
30
40
Distance from bread surface (mm)

50

Fig. 12. Crust-density prole taking a fracture temperature of 101 C.

5 cm bread loaf. The results agree well except that the temperature close to the surface predicted numerically
lies near 100 C longer than seen in experiments.
One numerical simulation was also carried out with the values of Tc and dT adjusted to maintain a setting
temperature of Tc  dT = 76 C but the fracture temperature, Tc + dT, raised above boiling point to 101 C. It
is seen, in Fig. 12, that the predicted crust is very much weaker than a real one, with the surface density only
3% greater than the original value, and a thickness of only about 3 mm.
7. Conclusions
We have formulated a simple model of crust formation during bread baking. The model is based on a simple idea of dough movement due to bubble collapse coupled with heat equations with a temperature-dependent thermal conductivity in the wet part of the dough, and a vaporization boundary governed by a Stefan
condition. We have successfully solved the model numerically. The numerical method converges and the
numerical results at the points where the vaporization and collapse boundaries appear at the bread surface
agree with the analytical calculated asymptotic behaviour.
The model gives an indication of the factors involved in the bread baking process that might have an important eect on crust properties. A more detailed analysis of the dependence of crust properties on the various
model parameters will appear in a later paper.

D.R. Jeerson et al. / Applied Mathematical Modelling 31 (2007) 209225

225

Acknowledgements
The rst author wishes to thank the EPSRC for support under Grant No. GR/R64254/01. The authors also
would like to thank Mr. D. Gelder and Professors D.B. Duncan, A.K. Parrott, C.P. Please and D.L. Pyle for
many useful discussion.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]

B. Zanoni, C. Peri, S. Pierucci, A study of the bread-baking process, I: A phenomenological model, J. Food Eng. 19 (1992) 389398.
B. Zanoni, C. Peri, S. Pierucci, A study of the bread-baking process, II: Mathematical modelling, J. Food Eng. 23 (1994) 321336.
C. de Cindio, S. Correra, Mathematical modelling of leavened cereal goods, J. Food Eng. 24 (1995) 379403.
U. de Vries, C. Rask, Heat and mass transfer during baking, in: P. Zeuthen (Ed.), Processing and Quality of Foods 1, Elsevier, 1990.
M. Hasatani, H. Arai, H. Harui, Y. Itaya, N. Fushida, N. Hori, Eect of drying on heat transfer of bread during baking in oven,
Drying Technol. 10 (1992) 623639.
K. Thorvaldsson, H. Janestad, A model for simultaneous heat, water and vapour diusion, J. Food Eng. 40 (1999) 167172.
P.E. Marston, T.L. Wannan, Bread baking: the transformation from dough to bread, The Bakers Digest (August) (1976) 2449.
U. Wahlby, C. Skjoldebrand, Reheating characteristics of crust formed on buns, and crust formation, J. Food Eng. 53 (2002) 177
184.
B. Zanoni, C. Peri, D. Bruno, Modelling of browning kinetics of bread crust during baking, Lebensm. Wiss. Technol. 28 (1995) 604
609.
B. Zanoni, C. Peri, S. Pierucci, A computer model of bread baking control and optimization, in: Proceedings of the 7th International
Congress on Engineering and Food (Part 2), Brighton, UK, 1997.
M. Lostie, R. Peczalski, J. Andrieu, Lumped model for sponge cake baking during the crust and crumb period, J. Food Eng. 65
(2004) 281286.
B.J. Dobraszczyk, G.M. Campbell, J. Gan, Bread: A unique food, in: A.V.D. David, B.J. Dobraszczyk (Eds.), Cereals and Cereal
Products, Aspen, 2001 (Chapter 8).
N.L. Kent, A.D. Evers, Technology of Cereals, fourth ed., Pergamon, 1994.
D.A. Hayman, K. Sipes, R.C. Hoseney, J.M. Faubion, Factors controlling gas cell failure in bread dough, Cereal Chem. 75 (1998)
585589.
D. Gelder, Private communication, 2001.
U. de Vries, P. Sluimer, A.H. Bloksma, A quantitative model for heat transport in dough and crumb during baking, in: Cereal Science
and Technology in Sweden, 1988.
P.A. Sadd, Private communication, 2003.
J. Crank, Free and Moving Boundary Problems, Oxford University Press, Oxford, 1984.

You might also like