Sunteți pe pagina 1din 10

Chemical Engineering and Processing 40 (2001) 335– 344

www.elsevier.com/locate/cep

A theoretical study of freezing fouling: limiting behaviour based


on a heat and mass transfer analysis
M.J. Fernandez-Torres a,1, A.M. Fitzgerald a, W.R. Paterson a, D.I. Wilson a,*
a
Department of Chemical Engineering, Uni6ersity of Cambridge, New Museums Site, Pembroke Street, Cambridge, CB2 3RA, UK

Abstract

An analysis is presented of freezing fouling for liquids in laminar flow through ducts. Solidification occurs on a cooled wall
while the bulk liquid remains unsaturated. The approach assumes that intrinsically rapid heterogeneous nucleation occurs at the
solid/liquid interface, so that solidification is controlled by heat and mass transfer rates rather than by the intrinsic rate of
crystallisation. A simple one-dimensional model for a single crystallising solute predicts that a range of fouling behaviours can
occur, ranging from linear fouling to asymptotic (Kern– Seaton) behaviour, depending on the operating conditions, without any
need to invoke removal effects. Maximum fouling rates can be estimated and the occurrence of quasi-asymptotic fouling can be
identified, so permitting an apt choice of parameters for the design and operation of systems subject to such fouling effects. The
model is illustrated by a case study on ‘coring’ in food fat distribution pipelines. © 2001 Elsevier Science B.V. All rights reserved.

Keywords: Fouling; Crystallisation; Freezing; Fats; Modelling

1. Introduction more, have studied the case where freezing promoted


by a subcooled wall can be usefully used to isolate a
Fouling refers to the formation of unwanted deposit section of piping for inspection or maintenance [5].
layers on heat transfer surfaces, causing increased ther- Wiegand et al., reviewed the literature on freezing in
mal resistance and other process penalties. Freezing forced convection flows inside ducts, but concentrated
fouling describes the phenomenon whereby a layer is solely on pure liquids, usually water [6]. The related
formed on a heat transfer surface which is colder than literature on wax formation (e.g. [7]) and continuous
the bulk fluid via crystallisation of dissolved species or casting of metals (e.g. [8]) features related problems,
of the solvent itself, and is therefore considered to be a complicated by aspects of mass transfer, namely con-
particular case of crystallisation fouling [1,2]. The for- centration-dependent saturation temperatures and dif-
mation of solidified layers from moving fluids is not fusive transport, including deposit ageing.
Analogous problems arise in food processing plants,
restricted to heat exchanger fouling, however; the for-
where mixtures of food fats cause deposition of semi-
mation of such layers represents a serious operating
solid layers on the walls of distribution lines; in that
problem in undersea oil pipelines when wax fractions
industry, the phenomenon is called ‘coring’. In the
deposit, reducing the duct size and increasing pressure
latter case, coring affects both the pressure drop and
drops, and has prompted extensive research in this area
the rheology of the fat suspension due to subsequent
[3,4]. Wall freezing is not always undesired; it is a
difficulties in controlling its temperature. Fouling in
fundamental part of various industrial processes, e.g.
such systems can be caused by (a) particulate fouling,
wax separation and ice cream manufacture in scraped-
where wax crystals present in the bulk liquid become
surface heat exchangers. Keary and Bowen, further-
attached to the wall; and (b) freezing fouling, where the
presence of a cool wall causes local supersaturation and
* Corresponding author. Tel.: + 44-1223-334777; fax: + 44-1223- crystallisation there. The onset of particulate fouling,
334796.
E-mail address: ian – wilson@cheng.cam.ac.uk (D.I. Wilson).
driven by crystallisation in the bulk liquid, has
1
Present address: Departamento de Ingenieria Quimica, University prompted extensive research into the thermodynamics
of Alicante, Alicante, Spain. of wax/hydrocarbon mixtures (e.g. [9]). This paper con-

0255-2701/01/$ - see front matter © 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 2 5 5 - 2 7 0 1 ( 0 1 ) 0 0 1 1 0 - 6
336 M.J. Fernandez-Torres et al. / Chemical Engineering and Processing 40 (2001) 335–344

siders the latter case, where the bulk liquid is free of This work describes an analysis of freezing fouling
crystals. In practice, both mechanisms can occur simul- based on heat and mass transfer principles alone. Crys-
taneously, or in sequence, as the bulk temperature tallisation fouling involves the diffusion of solute spe-
changes along a pipeline. Singh et al., reported that cies to the wall and the evolution (and removal) of
surface shear prevented the attachment of particulates enthalpy of crystallisation. Heat and mass transport
in their experiments, which were performed at low rates will therefore set limits on crystallisation rates and
concentrations [7]. possible fouling behaviour: crystallisation kinetics and
Experimental studies of wax solidification on heat shear removal will give rise to behaviour within these
transfer surfaces such as those reported by Bott and limits. The flow conditions in oil pipelines and fat
Gudmundsson [10] and Ghedamu et al. [11] showed distribution systems are often laminar owing to the
strongly non-linear behaviour. In these studies the de- relatively high viscosity of the bulk fluid, so that the
posit thickness, lf, expressed as a fouling resistance Rf rates of convective heat and mass transfer are relatively
via Eq. (1), exhibited decreasing rate behaviour and low. This work investigates the types of fouling be-
even asymptotic fouling behaviour Eq. (2). haviour which could be expected for such systems
1 1 lf without (i) knowledge of crystallisation kinetics —
Rf(t) = − : (1) assumed to be intrinsically rapid, or (ii) shear removal
U(t) U(0) uf
effects — assumed to be negligible, i.e. it corresponds
Rf(t) =R
f (1− exp(− t/~)) (2) to a worst case scenario. It also provides a basis for
Eq. (2) was first reported by Kern and Seaton in an assessing the true effects of these chemical and physical
analysis of refinery fouling [12], and it has been found processes. The analysis is illustrated by a case study
to describe fouling behaviour in a number of other involving a binary mixture of a palm oil fat in a
instances. The basis of the Kern– Seaton fouling model non-crystallising solvent.
is the difference between (i) a growth term and (ii) a
removal term proportional to the deposit thickness, lf,
viz. 2. Analysis
dRf 1 dlf 1
= 8 (m; −krlf) (3) Consider the laminar flow of a single-phase liquid
dt uf dt ufzf f
mixture through a duct surrounded by a fluid at an
The experiments of both groups mentioned above, ambient temperature Ta, less than the mean (mixing-
which featured Reynolds numbers in the quasi-turbu- cup) temperature of the mixture, Tb. The following
lent region (Re \5000), exhibited deposit removal. Bott analysis considers the situation at a single (axial) co-or-
and Gudmundsson [10] found that the Reynolds num- dinate; stream-wise integration as described by Ribeiro
ber had a significant effect on the asymptotic (equi- et al. [3] is not described here. The temperature profile
librium) fouling resistance, which they related to the over a cross section is shown schematically in Fig. 1,
shear stress exerted by the fluid on the deposit surface. and features a cooled wall, at temperature Tw, which
Their reported values of R f increased with wax con- may cause the process fluid in contact with it to be
centration. They discussed their results in terms of wax saturated and thus to deposit crystals. For clarity, Fig.
crystals being formed in the cool viscous sublayer, and 1 and the subsequent analysis given in this section will
a deposit being subject to removal by surface shear. be for the special case of transfer to a flat wall. The
Similar results were reported by Ghedamu et al. [11] calculations reported, however, will be for the more
Crystallisation rates, like most chemical reaction practical case of the inner wall of a pipe: the corre-
rates, are very sensitive to temperature, due both to the
temperature dependency of the reaction rate constant
and of the degree of saturation. Both of the above-men-
tioned investigations were performed under constant
temperature driving force conditions, so that the tem-
perature at the deposit surface changed (increased)
during an experiment, affecting (reducing) the solidifi-
cation rate. Under such conditions, fouling will be
auto-retarding, as discussed by Epstein [13,14], inde-
pendent of any shear removal mechanism. The effects
of shear therefore need to be decoupled from any
auto-retardation effect. Modelling of crystallisation
fouling also requires reliable crystallisation kinetics,
which are not always available, particularly for mix-
tures of waxes or fats as described above. Fig. 1. Schematic analysis of freezing fouling (near surface).
M.J. Fernandez-Torres et al. / Chemical Engineering and Processing 40 (2001) 335–344 337

sponding algebra for cylindrical geometry is sum- Finally, we assume that the solution properties, such
marised in the Appendix. Crystallisation will result in as diffusivity, are independent of composition so that
the formation of a layer of insulating solid with solid- diffusion of solute to the deposit surface can be de-
liquid surface temperature Ts; initially Tw =Ts. The scribed by the familiar ‘rich solution’ result for diffu-

 
temperature at the deposit surface can be calculated by sion through a stagnant film:
assuming that the system is in pseudo-steady state,
1− xs
using film heat transfer coefficients to describe convec- N= CTkm ln (7)
1− xb


tive heat transfer fluxes.

(Ts −Ta)
1
+Rf
n −1

=[(Tb −Ts)hi +NDHm] (4)


There are evident analogies between this analysis and
the case of condensation in the presence of inerts,
ho
where by Eq. (7) we are implicitly using the Stefan
Here, N is the molar flux of crystallising solute; for a result for diffusion of A through ‘stagnant’ B. Eqs. (4),
slab geometry, the fouling resistance Rf is given by (6) and (7) are solved simultaneously to find N, xs and
lf/uf. Eq. (4) shows that crystallisation increases the Ts for a given set of operating conditions, subject to the
amount of heat to be removed from the fluid. The following bounds:
external convective heat transfer coefficient, ho, includes (i) Composition
any resistances due to the duct wall, insulation, radia-
tion etc, i.e. it is the overall heat transfer coefficient 0 Bxs 5 xb (8)
from the inner pipe surface. (ii) Mass transport
We make the following assumptions concerning crys-
tallisation of the solute: 0B N5 Nmax (9)
1. The mixture exhibits ideal solution behaviour. The where
freezing point, Tf, of a solution containing mole frac-
Nmax − kmCT ln(1− xb): kmCTxb
 
tion x of solidifying solute is given by (9a)
DHm 1
 
1 (iii) Heat transfer
ln(x)= − (5)
R Tm Tf h
Ta + Tb i + hiRf
where Tm is the melting point of the pure solute. ho
2. The deposit consists of pure solute. In a multi- U(Rf)5 Ts 5 Tb (10)
hi
component solid system, the deposit is treated as a + hiRf + 1
ho
mixture of crystals of pure solute, as opposed to a solid
solution. This is the approach described by Firoozabadi The lower bound on Ts in Eq. (10) is obtained from
[9] for wax deposition in hydrocarbons. The presence of the heat transfer statement in Eq. (4) by setting N=0.
entrained solvent within the crystal matrix is ignored, Note that the fouling resistance used in Eq. (10) should
but could be incorporated via an appropriate void be that based on the external surface area, such that the
fraction. left hand side must be modified to include changes in
3. Homogeneous crystallisation does not occur in the surface area for curved geometries.
flowing liquid. The solidification rate can then be calculated from
4. Intrinsically rapid heterogeneous crystallisation oc- dlf N · M
curs at the wall. The achieved rate of crystallisation is = (11)
dt zf
determined by heat and mass transfer rates alone: nei-
ther crystallisation kinetics nor removal mechanisms Porous layers, and the resulting diffusion-driven age-
are included in the model. Rapid crystallisation means ing described by Singh et al. [7] are not considered here.
that negligible supersaturation is required, so that crys- The changes in time and internal energy required for
tallisation occurs if the mole fraction of solute at the the temperature and concentration profiles within the
deposit surface is equal to or greater than the mole system to adjust as the deposit grows are also neglected.
fraction which would be in equilibrium with the surface Limits of the model regime can be identified by
at temperature Ts. This mole fraction is given by Eq. inspection of Eqs. (4), (6) and (7).
(5), and we simplify the calculation by setting the solute
concentration at the surface to be in equilibrium with 2.1. Crystallisation in the bulk liquid

 
the solid at that temperature, i.e.
Crystallisation is deemed to occur in the bulk liquid
DHm 1 1 when the mean temperature reaches the saturation tem-
ln(xs )= − (6)
R Tm Ts perature for the solute concentration, T*. This could
This equation, establishing the saturation conditions, occur via homogeneous or heterogeneous nucleation
would be augmented by crystallisation rate equations if (e.g. dust, impurities), particularly for prolonged expo-
such information were available. sure at the cloud point. We therefore choose to focus
338 M.J. Fernandez-Torres et al. / Chemical Engineering and Processing 40 (2001) 335–344

boundaries need to be added to this diagram: (i) the


region where the flow can no longer be described as
laminar, at some higher value of Tb; (ii) the limit of
applicability of the mass transfer term in Eq. (7); the
predictions for higher concentrations are subject to
uncertainty, as the diffusivity used was independent of
concentration, an approximation which may not hold
at large x.

3. Case study – coring in palm oil distribution lines

Mixtures of palm oil fats are important feedstocks in


the food industry, e.g. in the generation of plasticised
fat and shortenings. Palm oil is a mixture of tri- and
Fig. 2. Explanation of fouling regime map for fixed Ta, hi and ho. di-glyceride components with a range of melting points
[15]. We here simplify the system by treating it as a
on the temperature regime where Tb \T*, with T* binary system of one solidifying solute and a hydrocar-
calculated from Eq. (6), namely
 
bon solvent of lower melting point fractions. Table 1
DHm 1 1 summarises the operating parameters for the case
ln(xb)= − (12) study. The solute parameters are those for tri-palmitin
R Tm T*
(PPP), which is the most abundant high melting point
Thus we do not consider the case of a sub-cooled component found in palm oil. The solvent and solution
solution. parameters were either taken from work in progress
[16] or estimated from correlations.
2.2. Zero solidification beha6iour We consider a typical case where the fat solution is
pumped along a pipe of di = 50 mm at constant mass
Eq. (7) indicates that xs =xb when N =0, so that flow rate w. The internal film heat transfer coefficient hi
Ts = T* (from Eq. (6)). Eq. (10) therefore allows one to is obtained from Hausen and Kays’ correlation [17]:
predict when asymptotic fouling, or no fouling, is likely
0.0668(ds/L)Red Pr
to occur. If Ts, calculated from Eq. (10) with Rf =0, is Nud = 3.66+ (13)
greater than T*, then the ‘clean’ wall is too warm for 1+0.04[(ds/L)Red Pr]2/3
solidification to occur. Heat transfer therefore sets an where L is taken to be 50 m and the duct dimension,
upper limit for the bulk temperature which will cause ds = di –2lf , is updated as fouling occurs. The film mass
freezing fouling for given Ta and solution concentration transfer coefficient, km, is obtained from the analogous
xb (and T*). mass transfer correlation:
Fouling will cause Ts to increase over time when Tb
0.0668(ds/L)RedSc
and Ta are constant, and asymptotic fouling may occur Shd = 3.66+ (14)
1+0.04[(ds/L)RedSc]2/3
when N =0 for finite Rf. Setting [ =T* allows one to
calculate the asymptotic fouling resistance Rf for a The cylindrical geometry of the system is incorpo-
given set of process conditions, and inspection of the rated into the heat transfer statement, equation (4*),
corresponding fouling layer thickness will indicate and the fouling resistance quoted is that based on the
whether this is a feasible result. An infeasible result for external surface area, Rf,o. It follows from Eq. (1) that
lf indicates that asymptotic fouling does not occur. For
    
this resistance is given by
example, a fouling resistance which gave lf \ di /2 do d d do
would be infeasible for solidification on the inside of a Rf,o = + o ln i − (15)
dshi 2uf ds dihi,c
pipe.
The next section presents the results of this analysis For convenience, the o subscript is dropped in the
for a binary food fat system. The data are presented in following discussion and presentation of results for a
two forms: as Rf – time profiles to illustrate fouling cylindrical geometry.
behaviour; and as regime maps such as the schematic For a given set of parameters (Ta, ho, xb) the limiting
shown in Fig. 2. The abscissa presents solution concen- values of Tb were calculated, and fouling profiles were
tration (xb) in terms of the saturation temperature T*, then predicted for a set of intermediate temperatures.
given by the non-linear relationship in Eq. (12). The Table 2 summarises the calculation procedure. Eqs. (4),
map shows that freezing fouling will occur in the region (6) and (7) were solved simultaneously at a given value
T* BTb B {Tb : Tw =T*}. We note that two further of lf using a modified Newton–Raphson method
M.J. Fernandez-Torres et al. / Chemical Engineering and Processing 40 (2001) 335–344 339

within Microsoft Excel, giving N, Ts and xs and there- considered to be linear if the correlation coefficient, R 2,
fore dRf/dt. The Rf – time profile was then obtained by was greater than 0.99999.
integrating Eq. (11) stepwise in Rf. The effect of chang- Fouling profiles and fouling regime maps were gener-
ing internal diameter on film transfer coefficients, and ated for different combinations of flow rates (w=0.1
the film temperature dependency of physical properties, and 0.7 kg/s) and ambient temperatures (Ta = 25, 15°C)
were incorporated in the model (Table 1). Eqs. (4), (6) over a range of concentrations (0.01B xb B 0.9) and
and (7) were solved for increasing values of lf , until N bulk temperatures Tb. The values ho ranged from 1
reached zero, or the duct dimension reduced to an W/m2 K, estimated for an insulated pipe, to 25 W/m2
unreasonable value. The latter criterion was set by K, typical of a very draughty location. Higher values of
pressure drop in the pipe, which can be readily calcu- ho would arise if the pipe were surrounded by flowing
lated assuming that the fouling layer is smooth. For cooling liquid [10,11].
laminar flow in a pipe, the ratio of pressure gradients is

 
given by:

 
dP 4. Results and discussion

 
dz w d 4i
= (16) 4.1. Fouling beha6iour
dP wclean (di −2lf)4
dz clean The effect of bulk temperature is illustrated in Fig.
A ratio of ten was used to terminate calculations. 3(a) for a typical case {xb = 0.10 (T*= 324.5 K); ho =5
The alternative scenario, of constant pressure drop and W/m2 K; w=0.7 kg/s; Ta = 25°C}. Linear fouling be-
varying mass flow rate, calculated from Eq. (16), was haviour was predicted at low Tb, i.e. no auto-retarda-
not considered here. tion, mass transfer control. The Figure shows the
The extent of auto-retardation depends on the oper- results obtained for Tb = 326 K, above which the Rf –
ating parameters, with some profiles showing only small time profile did not satisfy the linearity condition de-
reductions in fouling rate over time. In such circum- scribed above. The maximum pressure drop criterion
stances the Rf –t and lf – t profiles were almost linear, placed a limit of 1.12 cm (Rf = 0.112 W/m2 K) on the
so a criterion based on regression was used to delineate deposit thickness in this scenario. Also plotted in Fig.
linear from falling rate behaviour. The data were fitted 3(a) is the result for Tb = 327.35 K, which exhibits
to a linear regression model, i.e. Rf =a × t, and were evident auto-retardation behaviour. The figure also

Table 1
Model parameters for palm oil case study

Parameter Value Source/Comment

Tripalmitin (PPP) Tm 65.6°C


ZHm 147.7 kJ/mol Assumed constant
zM 1083 mol/m3 Assumed constant
D 1.96×10−11 m2/s (325 K) Wilke and Chang correlation (1955) [18]
2.57×10−11 m2/s (330 K)
3.28×10−11 m2/s (335 K)
uf 0.14 W/m K
RMM 807 g/mol

Solution z 773 kg/m3


CT 4500 mol/m3
vb 0.037 Pa s (325 K)
0.029 Pa s (330 K)
0.023 Pa s (335 K)
ub 0.15 W/m K
Cp 2109 J/kg K (325 K)
2121 J/kg K (330 K)
2132 J/kg K (335 K)

Geometry di 0.050 m Typical plant dimension


do 0.054 m Typical plant dimension

Variables Ta 15, 25°C


ho 1–50 W/m2 K
w 0.1–0.7 kg/s Typical plant value
340 M.J. Fernandez-Torres et al. / Chemical Engineering and Processing 40 (2001) 335–344

Table 2 Fig. 3(b) shows the corresponding results for xb =


Calculation sequence for generating fouling regime maps
0.50. The same final Rf values appear because of the
1. Select material and operating condition parameters, pressure-drop constraint used in the calculations, but
particularly Ta, ho, xb the effect of xb on the time scales is noteworthy.
Fig. 4 is a series of domain diagrams which sum-
2. Determine limiting values
marise the results for a range of concentrations and
2.1 Foulant thickness given by maximum pressure drop
criterion, lf,max fixed ho, Ta in a similar format to Fig. 2, with lines
2.2 Limiting values of Tb, given by marking the transition from linear to asymptotic foul-
(i) Tb =T* ing. The upper boundary of the asymptotic region is
(ii) Hot wall condition; Tb [ Tw = T*, lf = 0 (Eq. (10))
given by the ‘warm wall’ criterion (N=0 at Rf =0).
3. Consider intermediate values of Tb : select Tb
3.1 Check whether a true asymptote is reached The asymptotic/falling rate boundary is given by in-
Calculate Rf such that N= 0, i.e. [ (Eq. (10)) = T* spection of the calculated Rf –t behaviour: if the data
Check whether the corresponding value of lf5lf,max fitted an asymptotic model equation (Eq. (2)) with a
If satisfied, behaviour is asymptotic
regression coefficient (R 2)] 0.998, the data set was
If not satisfied, inspect behaviour as lf approaches
lf,max by 3.2 deemed to be ‘asymptotic’. The value of 0.998 was
3.2 Integrate to find Rf–t behaviour somewhat notional, but inspection of the data sets by
Discretise range of feasible foulant thicknesses: lf  eye suggested almost negligible final fouling rates. The
{0, lf,max }
boundary between the falling rate and linear fouling
For each lf, solve equations Eqs. (4), (6) and
dlf behaviours was set by regression to a linear fouling
(7) [ N [ (Eq. (11))
dt model, with a threshold regression coefficient value of
Integrate to get lf –t and thereby Rf–t profiles 0.99999.
Inspect Rf–t profiles for
(i) Linear behaviour: compare R 2 for fit to Rf
= a+b.t
(ii) Asymptotic behaviour; compare R 2 for fit to Eq.
(2)
3.3 Select next value of Tb
4. Delineate asymptotic, falling rate and linear fouling
regimes
5. Select next value of xb (and T*)

shows the result from fitting a Kern– Seaton curve


equation Eq. (2) to these data; the agreement is reason-
able (R 2 = 0.998) and might even have been considered
good if the model predictions were actually experimen-
tal data. The asymptotic behaviour is caused by the
combination of temperature conditions and the
strongly non-linear dependence of saturation on Ts.
Between 326 and 327.35 K, the fouling profiles exhib-
ited falling-rate fouling, with quasi-asymptotic be-
haviour, i.e. falling rate fouling, giving very low final
fouling rates at Tb just less than 327.35 K. Asymptotic
fouling was observed at all Tb \327.35 K, with the
value of Rf decreasing with increasing Tb until the
warm wall condition was reached, when Rf =0 (no
solidification).
Fig. 3(a) thus demonstrates that a range of fouling
behaviours can be obtained for freezing fouling from a
priori calculations without invoking any removal term;
this analysis shows that the deposition term in Eq. (3) is
strongly dependent on the extent of fouling. Introduc-
tion of a removal term incorporating shear forces Fig. 3. Fouling behaviours predicted for PPP case study with ho = 5
would result in the onset of asymptotic behaviour at a W/m2 K; Ta =25°C; w =0.7 kg/s. Fouling resistances are based on
lower value of Tb than that calculated using this external surface area, Rf,o. Linear fouling: circles, calculated points;
solid lines, regressed model. Asymptotic fouling: crosses, calculated
analysis.
points; dashed lines, regressed model.
M.J. Fernandez-Torres et al. / Chemical Engineering and Processing 40 (2001) 335–344 341

Fig. 4. Domain maps for fouling behaviours in PPP case study. Ta =298 K; w=0.7 kg/s. Dotted line, limit of asymptotic behaviour; dashed line,
transition to falling rate fouling; crossed line, limit of linear fouling behaviour; solid line, Tb =T*b (bulk crystallisation).

Fig. 4 shows that the type of fouling behaviour Also marked on Fig. 4(c) is the locus for Re (ini-
predicted is thus strongly dependent on the original tial)\ 1000, which is used here to indicate when lami-
operating conditions. Reading across the diagram, it nar flow correlations are likely to give less reliable
can be seen that at a fixed value of Tb: (i) no solidifica- descriptions of heat and mass transfer in the system. It
tion occurs for low xb (and T*); (ii) onset of fouling is can be seen that most of the initial conditions used in
observed at a critical concentration, beyond which its this work lie within the laminar regime.
regime will change from asymptotic to falling rate to Fig. 5 shows the effect of the flow rate and ambient
linear as xb increases, until (iii) the solution is saturated temperature on the domain diagram for the palm oil
and solidification will occur from the bulk fluid. The case study in Fig. 4(c). It can be seen that the width of
solidification behaviour along a pipeline can be fol- the fouling regions (expressed in terms of temperature)
lowed by reading downwards on a domain map (assum- is reduced by a drop in Ta, but increased by a reduction
ing that the effect of L on transport parameters is in mass flow rate (and hence transport rates).
small). For example, Fig. 4(c) shows that for xb corre- Such diagrams can be used to understand the pro-
sponding to 332 K, no solidification is observed until cesses occurring in a distribution line or pipeline. The
Tb is less than  343 K, when asymptotic fouling starts fat solution will initially be too warm to cause any
(at a low rate). Asymptotic behaviour will continue solidification at the wall, but downstream, after some
until ca. 337 K, when solidification (at increasing initial cooling, asymptotic fouling will arise. Even further
rate) no longer asymptotes, but continues with a downstream, assuming negligible change in solute con-
steadily falling rate. At 334 K pseudo-linear behaviour centration, linear fouling or bulk precipitation will oc-
will arise, increasing in rate until Tb =332 K, when the cur. Solidification, however, will reduce the heat loss
from the bulk solution so that at some later time, the
solution reaches saturation and bulk crystallisation is
temperature profile will be different and the fouling
likely to occur.
behaviour at a particular location is likely to have
342 M.J. Fernandez-Torres et al. / Chemical Engineering and Processing 40 (2001) 335–344

ond or higher order reaction behaviour. Such an inter-


pretation would be false, as the results were generated
entirely without reference to kinetics, being based
purely on thermodynamic and transport criteria.
The results in Fig. 6 also demonstrate how the model
can generate bounds for fouling rates (and asymptotic
levels of deposit) which could be used in estimating
worst case behaviour, or time scales likely to be en-
countered in experimental investigations of these
systems.
The greatest source of uncertainty in the model lies in
the thermodynamic relationships between equilibrium
concentrations and temperatures at the solid/liquid in-
terface, i.e. Eq. (1). Most systems of interest feature
several crystallising species, such that the thermody-
namics in solution and at the interface require careful
elucidation via experimentation in order to be incorpo-
rated into a model of this form. Furthermore, no
subcooling has been assumed to be necessary for
growth, which may not be realistic in all applications.
Comparison with experimental results is required in
order to establish the applicability of this approach; in
its current form, it does, nevertheless, provide a frame-
work to explain the different behaviours which can be
observed in these systems.

Fig. 5. Effect of operating parameters on fouling behaviour. Domain


maps for comparison with Fig. 4(c). (a) Ta = 288 K, ho = 10 W/m2 K, 5. Conclusions
w = 0.7 kg/s. (b) Ta = 298 K, ho = 10 W/m2 K, w= 0.1 kg/s.

A model is presented for solidification, on a cooled


changed. These effects shed some light on the
wall, of fat from a binary fat/solvent solution in lami-
difficulties encountered in practice in assessing coring nar flow through a pipe. The modelling is ‘parsimo-
behaviour. nious’; that is to say, in designing the model we have
The model can also provide estimates of (worst-case) deliberately excluded various complications to investi-
fouling rates. Fig. 6 shows the variation of the initial gate how wide a range of behaviour such a simple
fouling rate with PPP concentration at a constant bulk model can predict, when supplied with realistic parame-
temperature of 64.9°C. The fouling rates are presented ter values.
in the form of a dimensionless fouling Biot number, In particular, the key simplifications invoked
Bif =Rf × ho. The figure shows a non-linear increase in concern:
fouling rate with increasing concentration which, were Crystallisation kinetics: rather than assume a rate
the data experimental, might be taken to indicate sec- law, an activation energy and so forth, we have
simply assumed the intrinsic crystallisation rate to be
high.
Crystallisation location: all crystallisation is assumed
to take place on the solid surface, either, initially, on
the clean pipe wall, or later, on the exposed surface
of the deposit.
Streamwise variations: the model neglects these, ex-
cept insofar as they are a function of local variables,
e.g. local fluid radial-mean temperature, rather than
upstream variables; it is a purely one-dimensional
model which, by assumption, may be applied at any
point along the pipe wall.
Deposit properties: the solidifying component forms
Fig. 6. Variation of initial fouling rate with concentration. Tb = a pure layer of uniform thickness in the immediate
64.9°C; Ta =25°C; ho = 25 W/m2 K; w= 0.7 kg/s. neighbourhood of any point.
M.J. Fernandez-Torres et al. / Chemical Engineering and Processing 40 (2001) 335–344 343

Removal effects: these are omitted from the model. where Ao is the external pipe surface area and ho is a
Whatever the shear stress on the surface of the lumped thermal resistance based on the external diame-
deposit, the rate of removal of deposit is nil. ter. The thermal resistance of the pipe is negligible.
Unsteady state temperature and composition profiles: Rearranging,
these too are omitted; we have constructed a pseudo
steady-state model. Q
 1
+
ln(di/ds) 
= (Ts − Tw)+ (Tw − Ta)= (Ts −Ta)
The key effects incorporated are: Aoho 2yLuf
The reduction in solubility of the fat as temperature (A.2)
is reduced; The rate of heat transfer to the inner surface by
That heat and mass transfer resistances can control convection and crystallisation is given by:
the rate of solidification;

 
The effect of release of heat of crystallisation; (Ts − Ta)
Q= = [hi(Tb − Ts)+ NDHm]As
Heat transfer by conduction through the deposit; 1 ln(di/ds)
Deposit growth. +
Aoho 2yLuf
The model proves capable of predicting different (A.3)
regimes of fouling, specifically: absence of fouling, lin-
ear fouling, reducing-rate fouling and asymptotic yielding the analogous result to Eq. (4):

 
(Kern–Seaton) fouling. The predictions are presented (Ts − Ta)Ao
in two forms: as time profiles of fouling thermal resis- = [hi(Tb − Ts)+ NDHm]As (4*)
1
tance, to illustrate these different modes of behaviour; + Rd
and as regime maps. Maximum fouling rates are esti- ho
mated and conditions for the occurrence of quasi- where As, the area of the fouling layer/liquid inter-
asymptotic fouling are identified, so permitting an apt face, and ds decrease as solidification continues. This
choice of parameters for the design and operation of expression reduces to Eq. (4) for a slab geometry where
systems subject to such fouling effects. A discussion is Ao = As.
presented to demonstrate how such diagrams may be Equation 10
used to understand the processes occurring along a Eq. (10)* is obtained by setting N= 0 in equation

 
distribution line or pipeline subject to fouling. (4*), giving:
hi A
Ta + Tb + hi Rd s

 
ho Ao
Acknowledgements Ts = (10*)
hi A
+ hi Rd s + 1
MJFT wishes to acknowledge the provision of a ho Ao
fellowship from Generalitat Valenciana. A CASE
Award for AMF from the BBSRC and United Biscuits,
and discussions with Dr. Ian Smart, are also gratefully
acknowledged.
Appendix B. Nomenclature

Appendix A

The model equations are presented for a slab geome-


try; the corresponding relationships for a cylindrical Latin
configuration, as arise in the case study calculations, Ao,i surface area based on outer, inner pipe dimen-
are presented here. sion (m2)
Equation 4 CT total concentration (mol m−3)
Assuming that the system is in pseudo-steady state, di internal (clean) diameter of pipe (m)
(e.g. ignoring the time taken to establish the tempera- do external diameter of pipe (m)
ture profile through the fouling layer) the rate of heat ds Reduced diameter of pipe (di−2lf ) (m)
transfer, Q, through the core layer by conduction D Diffusivity of solute in solvent (m2 s−1)
equals that from the external surface by convection and hi Internal convective film heat transfer coeffi-
radiation: cient at time t (W m−2 K−1)
hi,c Internal (clean) convective film heat transfer


2yLuf
Q= (Ts − Tw) =Aoho(Tw −Ta) (A.1) coefficient (W m−2 K−1)
di ho External film heat transfer coefficient (W m−2
ln
ds K−1)
344 M.J. Fernandez-Torres et al. / Chemical Engineering and Processing 40 (2001) 335–344

ŠHm Molar enthalpy of melting (J mol −1


) ~ fouling time constant (s)

km Film convective mass transfer coefficient (m s


−1
)
kr Removal rate constant (kg m−1 s−1) References
M RMM of component (kg kmol−1)
m; f Mass ‘deposition’ rate (kg s−1) [1] T.R. Bott, Aspects of crystallization fouling, Exp. Thermal Fluid
L Axial length (m) Sci. 14 (1997) 356 – 360.
[2] N. Epstein, Thinking about heat transfer fouling: a 5 × 5 matrix,
N Molar flux of solute towards surface (mol
Heat Transfer Eng. 4 (1) (1981).
m−2 s−1) [3] F.S. Ribeiro, P.R. Souza Mendes, S.L. Braga, Obstruction of
Nu Nusselt number pipelines due to paraffin deposition during the flow of crude oils,
P Pressure (Pa) Intl. J. Heat Mass Transfer 40 (18) (1997) 4219 – 4328.
Q Rate of heat transfer (W) [4] G.M. Elphingstone, K.L. Greenhill, J.J.C. Hsu, Modelling of
R Gas constant (J mol−1 K−1) multiphase wax deposition, J. Energy Resources Tech. 121
R2 Correlation coefficient (dimensionless) (1999) 81 – 85.
[5] A.C. Keary, R.J. Bowen, On the prediction of local ice forma-
Re Reynolds number tion in pipes in the presence of natural convection, J. Heat
Rd Thermal resistance of deposit (m2 K W−1) Transfer 121 (4) (1999) 934 – 944.
Rf Fouling resistance (m2 K W−1) [6] B. Wiegand, J. Braun, S.O. Neumann, K.J. Rinck, Freezing in
Rf,o Fouling resistance based on external surface forced convection flows inside ducts: a review, Heat Mass Trans-
area (m2 K W−1) fer 32 (1997) 341 – 351.
R f Asymptotic fouling resistance (m2 K W−1) [7] P. Singh, V. Venkatesan, H.S. Fogler, N. Nagarajan, Formation
and aging of incipient thin film wax-oil gels, AIChEJ 46 (5)
Sc Schmidt number
(2000) 1059 – 1074.
Sh Sherwood number [8] M.J.M. Krane, F.P. Incropera, Experimental validation of con-
Tb Mean fluid temperature (K) tinuum mixture model for binary alloy solidification, J. Heat
Ta Ambient temperature (K) Transfer 119 (4) (1997) 783 – 791.
Ts Temperature at deposit–liquid interface (K) [9] A. Faroozibadi, Thermodynamics of Hydrocarbon Reservoir
Tf freezing point of solution of composition x Fluids, McGraw-Hill, New York, 1998.
[10] T.R. Bott, J.S. Gudmunsson, Deposition of paraffin wax from
(K)
kerosene in cooled heat exchanger tubes, Can. J. Chem. Eng. 55
Tm melting point of pure solute (K) (1977) 381 – 385.
Tw temperature at heat transfer surface (wall) (K) [11] M. Ghedamu, A.P. Watkinson, N. Epstein, Mitigation of wax
T* saturation temperature for the solute concen- oil buildup on cooled surfaces, in: C.B. Panchal (Ed.), Fouling
tration (K) Mitigation of Industrial Heat-Exchange Equipment, Bgell
U overall heat transfer coefficient (W m−2 K−1) House, New York, 1997, pp. 473 – 489.
t time (s) [12] D.Q. Kern, R.E. Seaton, A theoretical analysis of thermal
surface fouling, Brit. Chem. Eng. 14 (5) (1959) 258.
w mass flow rate (kg s−1) [13] N. Epstein, Fouling of Heat Exchange Surfaces, VDI-GVC,
x mole fraction of solute (mol (mol solution−1)) Dusseldorf, Germany, 1990, pp. 1.1 – 1.17.
[14] N. Epstein, Elements of particle deposition onto non-porous
Greek solid surfaces parallel to suspension flows, Exp. Thermal Fluid
lf deposit thickness (m) Sci. 14 (4) (1997) 323 – 334.
[ surface temperature estimate, Eq. (10) (K) [15] W.L. Ng, C.H. Oh, J. Am. Oil Chem. Soc. 71 (10) (1994)
uf deposit thermal conductivity (W m−1 K−1) 1135 – 1139.
um pipe wall thermal conductivity (W m−1 K−1) [16] A.M. Fitzgerald, Modelling of Palm Oil Crystallisation in Food
ub solution thermal conductivity (W m−1 K−1) Fats, 1998 CPGS dissertation,University of Cambridge (unpub-
lished data).
vb solution dynamic viscosity (Pa s) [17] W.M. Kays, citing H.Z. Hausen, Trans. ASME, 77 (9) (1943)
zf deposit density (kg m−3) 1265. Ver. Dtsch. Ing. Beih. Verfahrenstech 4 (1955) 91.
zM deposit molar density (mol m−3) [18] C.R. Wilke, P. Chang, AIChEJ 1 (1955) 264.

S-ar putea să vă placă și