You are on page 1of 9

Journal of Food Engineering 99 (2010) 351–359

Contents lists available at ScienceDirect

Journal of Food Engineering


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

Simulation and analysis of a sugarcane juice evaporation system


L.M.M. Jorge a,*, A.R. Righetto a, P.A. Polli a, O.A.A. Santos a, R. Maciel Filho b
a
Chemical Engineering Department, State University of Maringá (UEM), Av. Colombo, 5790, Bloco D90, Maringá CEP 87020-900, PR, Brazil
b
Chemical Engineering School, State University of Campinas (UNICAMP), CP 6066, Campinas CEP 13081-970, SP, Brazil

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

Article history: The production of sugar and alcohol is the main objective of the sugarcane processing industry. The evap-
Received 22 December 2009 oration of sugarcane juice has a high energetic cost and is usually performed in multiple-effect evapora-
Received in revised form 6 March 2010 tors. The loss of performance during operation due to fouling makes the process more complex. In this
Accepted 14 March 2010
study, modeling, simulation, validation, and analysis were performed for a sugarcane juice industrial
Available online 19 March 2010
evaporation system (IES) composed of a falling film evaporator followed by three short vertical-tube
evaporators arranged in parallel. The IES model was developed using a commercial process simulator
Keywords:
and validated with data from the plant. The IES had marked performance losses in the first 14 days of
Sugarcane juice
Evaporators
operation, mainly due to fouling in the first effect, with a 30% decrease in the evaporation rate.
Process simulation Ó 2010 Elsevier Ltd. All rights reserved.
Overall heat transfer coefficients
Fouling

1. Introduction units (Volin and Ostrovskii, 2006). Among the commercial process
simulators, HYSYS stands out due to its ease of use and quality
Sugarcane processing is one of the most important industries in databank.
the Brazilian economy, mainly due to its high efficiency and com- However, HYSYS, as well as the other primary commercial
petitiveness. There are three main types of plants: sugar mills, eth- simulators, only has models and databanks developed for the
anol distilleries, and integrated sugar and ethanol plants (Ensinas petrochemical industry (Seider et al., 1999), which requires corrob-
et al., 2009). orating or adapting for applicability in other sectors, including to
In the 2007/2008 crop, Brazil harvested 493,384,552 tons of meet the needs of the sugar–alcohol sector in Brazil, which is
sugarcane and produced 30,760,165 tons of sugar and 14,300, strongly expanding.
346 tons of alcohol, while 386,090,117 tons of sugarcane were pro- The ethanol production process in a distillery is complex and in-
cessed in 2004/2005, yielding 26,621,221 tons of sugar and volves evaporation, fermentation, and distillation steps in this or-
7,112,218 tons of alcohol (Montanini, 2008). This means an in- der. Evaporation deserves special attention, as it is the first step
crease of 27.8% in processed sugarcane, 15.6% in sugar and of the production chain and it consumes great amounts of energy
101.1% in ethyl alcohol production in three years. The larger sugar in sugarcane juice concentration before its fermentation and
and alcohol production reflected in the Brazilian exports, which in- distillation.
creased from 16,011,340 tons of sugar in 2004 to 19,618,989 tons In this research, modeling, simulation, and analysis of a full-
in 2009 (a 22.5% increase) and from 1,929,896 tons of alcohol in scale industrial evaporation system (IES) of an alcohol distillery
2004 to 4,076,446 tons in 2009 (a 111.2% increase) (CONAB, (COCAFÉ) are performed using the HYSYS process simulator. The
2009, 2005). IES was monitored over 23 days of operation, from startup to
The rapid growth experienced by the sugar–alcohol sector in cleaning shutdown. The distillery of the Cooperativa Agrícola de
Brazil necessitates the design and installation of new facilities Astorga Ltda (COCAFÉ), which is located in the northwest of Paraná
and the expansion and optimization of the existing ones. Process State, Southern Brazil, processes around 180 tons/h (50 kg/s) of
analysis using a properly validated process simulator is essential sugarcane juice, producing approximately 380 m3/day (4.398 
to reduce production costs in each of these steps as well. 103 m3/s) of hydrated ethyl alcohol.
Computational simulation is an invaluable tool in design, reduc-
ing conception time and maximizing the economy of industrial
2. Industrial evaporation system

* Corresponding author. Tel.: +55 44 3261 4747; fax: +55 44 3261 4793. After extraction, the raw sugarcane juice contains about 4% of
E-mail address: lmmj@deq.uem.br (L.M.M. Jorge). suspended solids, which are decanted with lime, yielding a

0260-8774/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jfoodeng.2010.03.017
352 L.M.M. Jorge et al. / Journal of Food Engineering 99 (2010) 351–359

Nomenclature

i
Aiext external area of the tubes of evaporator i (m2) hS enthalpy of heating steam in evaporator i (J/kg)
i
B* normalized juice concentrations (B* = B/B0) hV enthalpy of the steam generated in evaporator i (J/kg)
B0 juice concentration in the system feed stream (kg solid/ LMTDi logarithmic mean temperature difference of evaporator
kg) i (°C)
Biin juice concentration at the evaporator i inlet (kg solid/kg) Pi absolute pressure in evaporator i (Pa)
Biout juice concentration at the evaporator i outlet (kg solid/ q1in clarified juice volumetric flow rate (m3/s)
kg) Qi heating rate of evaporator i (W)
BPRi boiling point rise in the evaporator i operating pressure Rf fouling resistance (m2 °C/W)
(°C) R1f asymptotic fouling resistance (m2 °C/W)
Ci condensate mass flow rate in evaporator i (kg/s) RG overall thermal resistances (m2 °C/W)
Cpm average heat capacity (J/(kg °C)) RL juice thermal resistance (m2 °C/W)
FBLEEDING bleeding stream mass flow rate (kg/s) RV heating steam thermal resistance (m2 °C/W)
FCONC2 CONC2 stream mass flow rate (kg/s) RW tube wall thermal resistance (m2 °C/W)
FHYSYS mass flow rates estimated by the HYSYS model (kg/s) Si mass flow rate of heating steam in evaporator i (kg/s)
F iin mass flow rate of incoming juice to evaporator i (kg/s) t time (s)
F iout mass flow rate of exiting juice from evaporator i (kg/s) T iFin temperature of incoming juice to evaporator i (°C)
FModel mass flow rates estimated by the conventional model T iFout temperature of exiting juice from evaporator i (°C)
(kg/s) T ieb water boiling temperature at the operating pressure of
FV1 mass flow rate of V1 stream (kg/s) evaporator i (°C)
FV1EVAP mass flow rate of V1EVAP stream (kg/s) T iin juice temperature at the evaporator i inlet (°C)
FV2 mass flow rate of V2 stream (kg/s) T iout juice temperature at the evaporator i outlet (°C)
FVE mass flow rate of VE stream (kg/s) U0 overall heat transfer coefficient at t = 0 (W/(m2 °C))
GFF evaporation flux in the falling film evaporator UHYSYS overall heat transfer coefficients estimated by the
(kg/(m2 s)) HYSYS model (W/(m2 °C))
GR evaporation flux in the Robert-type evaporators Ui overall heat transfer coefficient of evaporator i (W/
(kg/(m2 s)) (m2 °C))
i
hC enthalpy of the condensate from the first effect (J/kg) UModel overall heat transfer coefficients estimated by the con-
i
hfgS vaporization enthalpy of the heating steam in evapora- ventional model (W/(m2 °C))
tor i (J/kg) Vi mass flow rate of steam generated by evaporator i (kg/s)
i
hfgV vaporization enthalpy of the steam generated in evapo- qCLARIF clarified juice density (kg/m3)
rator i (J/kg) s fouling time constant (s)
i
hFin enthalpy of incoming juice to evaporator i (J/kg)
i
hFout enthalpy of exiting juice from evaporator i (J/kg)

clarified juice (CLARIF) that contains 0.5% of suspended solids at 3. Materials and methods
most. The clarified juice then runs through heat exchangers,
where it is heated up to near boiling temperature, and then to 3.1. Process data collection
the evaporation system (Fig. 1). Next, the concentrated juice
(CONC2) goes to the fermentation tanks and later to the distillation During the time interval between the IES startup and cleaning
columns. shutdown, a period of 23 days of operation, operating temperature
The COCAFÉ IES consists of a falling film evaporator (FF, Fig. 2) (T), pressure (P), flow rate (q), and concentration (B) were regularly
followed by three conventional Robert-type short vertical-tube taken at the points indicated in Fig. 1 and listed in Table 1. Operat-
evaporators (R1, R2, and R3), arranged in parallel (Fig. 3). The first ing data T, P, and volumetric flow rate of clarified juice were also
effect (FF), consisting of 1942 10-m long tubes with external diam- obtained from the online supervisor system AIMAX records, while
eter of 0.0381 m, is fed with high pressurized steam, VE, from the sugar content in the liquid streams (B) was evaluated using a
steam turbines. The second effect is constituted by three evapora- refractometer by periodic laboratory analyses of samples collected
tors (R1, R2, and R3) consisting of 1274 2.315-m long tubes, with at several points (Table 1).
the same diameter. The clarified juice (CLARIF) runs through the
first effect and leaves as pre-concentrated juice (CONC1), which 3.2. Evaporation system HYSYS model
is evenly distributed to the three second effect evaporators, which
turn it into concentrated juice (CONC2). Usually only part of the Because HYSYS does not have a module specific for evaporators,
steam generated in the first effect (V1) is used to heat the second we sought to develop alternative models from the existing mod-
effect (V1EVAP). The first effect surplus steam is bled out (BLEED- ules of this process simulator. Foust et al. (1982) described an
ING) and used as a heating fluid in various sectors of the plant. evaporator as basically consisting of a heat exchanger capable of
However, at times, the demand for steam of the second effect raising the solution temperature up to the boiling point and a de-
(V1EVAP) is greater than the amount of steam generated by the vice for separating the vapor phase from the boiling liquid. Using
first effect. In such event, the steam from the first effect is comple- this concept, Chawankul et al. (2001) successfully simulated an
mented by the injection of steam from the boiler into the BLEED- agitated orange juice thin-film evaporator in laboratory scale using
ING line. The condensates from both the first effect (CondVE) and Aspen Plus. The COCAFÉ evaporation system was modeled using
the second effect (CondV1EVAP) are used to generate steam in the HYSYS (Fig. 4); each of the four evaporators was represented
the plant boiler, while the low pressure vapor produced in the sec- by the combination of a multitubular heat exchanger and a flash
ond effect (V2) is used as heating fluid in several sectors of the vessel (separator). The heat exchanger represents the evaporator
plant. calandria. The following assumptions were made for the modeling:
L.M.M. Jorge et al. / Journal of Food Engineering 99 (2010) 351–359 353

CLARIF BLEEDING
Distillery
V2
Heating

9 10 11
1

CONC1 R1 R2 R3
V1EVAP
VE 3

2
4

7 8 CONC2
6
Fermentation
12
V1

5
CondVE CondV1EVAP

FF

Fig. 1. COCAFÉ evaporation system.

single pass in the shell and in the tubes, vertical orientation, max- Polynomial fitting (R2 = 1) of data from steam tables (Perry
imum spacing between tube sheets and A–E–L classification et al., 1984) in the range from 0.4  105 to 3.4  105 Pa.
according to the Tubular Exchanger Manufacturers Association Clarified juice mass flow rate:
(Perry et al., 1984).
F 1in ¼ q1in  qCLARIF ð6Þ
3.3. Conventional modeling of the evaporation system q1in is the volumetric flow rate (m3/s) and qCLARIF is the clarified juice
density (kg/m3).
According to Prost et al. (2006), well-known mass and energy Clarified juice density (Pacheco et al., 1999):
balance equations and the heat transfer rate equation between   . 
the hot (condensing steam) and cold (juice side) streams govern
qCLARIF ¼ 1000  1 þ B1in  B1in þ 200 54; 000
  . 
the evaporator process. To assess the quality of the IES HYSYS mod-  1  0:036  T 1in  20 160  T 1in ð7Þ
el, a conventional model was developed from mass (Eqs. (1) and
(3)) and energy (Eqs. (2) and (4)) balance equations, which were Juice average heat capacities (Hugot, 1986):
simultaneously solved for each evaporator, i, with auxiliary Eqs.
Cpm ¼ 4186:8  2512:1Bm ð8Þ
(5)–(11). The overall heat transfer coefficient (U) was estimated
from Eq. (12). Cpm is the average heat capacity (J/(kg °C)) and Bm is the inlet and
  outlet stream average sugar concentration of evaporator i (kg sol-
i
V ¼ F iin 1  Biout =Biin ð1Þ ids/kg).
Boiling point rise (BPRi) in evaporator i (Traganits, 1981):
h  i ! !
i
F iin Cpm T iout  T iin þ V i hfgV Biout 0:3 þ Biout  
i
S ¼ ð2Þ BPRi ¼ i
0:22  0:0078T ieb ð9Þ
i
hfgS 0:355 1:036  Bout

T ieb is the water boiling temperature (°C) at the operating pressure


F iout ¼ F iin  V i ð3Þ of evaporator i.
Operating temperatures of evaporator i (T iout ):
i
Q i ¼ Si hfgS ð4Þ
T iout ¼ T ieb þ BPRi ð10Þ
i i
In these equations, V , S , F iin , F iout
represent the mass flow rates
Water boiling temperature:
(kg/s) of generated steam, heating steam, incoming juice, and out-
going juice, respectively, in evaporator i; Q i is the heating rate of T ieb ¼ 6:119  1021 P 4i þ 6:403  1015 P3i  2:672  109 P 2i
evaporator i (W); Biin and Biout are the juice concentrations (kg so-
þ 6:503  104 Pi þ 54:524 ð11Þ
lid/kg) in the evaporator i, T iin and T iout are the juice temperatures
(°C) in the evaporator i, hfgS and hfgV are the vaporization enthalpies Polynomial fitting (R2 = 0.99) of data from steam tables (Perry et al.,
(J/kg) of the heating steam and the steam generated of evaporator i, 1984) in the absolute pressure range from 0.4  105 to 3.4  105 Pa.
respectively. Here and elsewhere, in and out indicate the evapora- Overall heat transfer coefficients:
tor i inlet and outlet variables.
Vaporization enthalpies: Qi
Ui ¼ ð12Þ
Aiext LMTDi
hfg ¼ 4:709  1012 P3i þ 3:941  106 P2i  1:435Pi þ 2:369  106
ð5Þ U i is the overall heat transfer coefficient (W/(m2 °C)) for evaporator
i, LMTDi is the logarithmic mean temperature difference (°C), and
Pi is the absolute pressure (Pa). Aiext is the external area of the tubes (m2).
354 L.M.M. Jorge et al. / Journal of Food Engineering 99 (2010) 351–359

Fig. 3. Second effect: three parallel Robert-type evaporators.

Table 1
Process variable measurement sites.

Measurement points
Temperature (°C) 1, 2, 3, 4, 5, 9, 10, 11, 12
Pressure (Pa) 1, 2, 3, 4, 5, 9, 10, 11, 12
Concentration (°Bx) 1, 3, 6, 7, 8, 12
Flow rate (m3/h) 1, 12

Fig. 2. First effect: falling film evaporator.


coefficients (U) predicted by both models was also small (Fig. 6).
The average deviations relative to the HYSYS estimated values
The mathematical model (Eqs. (1) through (12)) was sequen- were 0.15% for the falling film evaporator and 1.5% for the Rob-
tially solved in an electronic spreadsheet from the first evaporator, ert-type evaporators. Therefore, both the conventional and HYSYS
FF (i = 1) to the outlet of the three short vertical-tube evaporators models represented the IES adequately.
that make up the second effect: R1 (i = 2), R2 (i = 3), and R3 The reliable prediction of overall heat exchange coefficients for
(i = 4). The stream mass flow rates (F) VE, V1, V1EVAP, BLEEDING, the design, analysis, simulation, and control of processes involving
V2 and CONC2 were determined from the respective stream flow evaporators still is a major challenge due to the variety of equip-
P
rates: F VE ¼ S1 , F V1 ¼ V 1 , F V1EVAP ¼ 4i¼2 Si , F BLEEDING ¼ F V1  F V1EVAP , ment and solutes and solvents involved. The range of U values
P4 i P4 i
F V2 ¼ i¼2 V out , and F CONC2 ¼ i¼2 F out . (870–1934) W/(m2 °C) found for the second effect Robert-type
evaporators (Fig. 6) is well within the range presented by McCabe
and Smith (1976) for this type of evaporator (800–2500) W/
4. Results and discussion (m2 °C), indicating satisfactory operation.
Generally, falling film evaporators are suggested for operation
4.1. Evaporation system HYSYS model evaluation with heat-sensitive materials, such as juices and milk, as they pres-
ent low residence times and high heat transfer rates and coeffi-
The IES flow rates obtained by conventional modeling and those cients (Minton, 1986; Krupiczka et al., 2002). In the range of
predicted by the HYSYS model were different by only an average of operating conditions of the first effect of the explored IES, experi-
0.11% (Fig. 5). The difference between the overall heat transfer mental results presented in literature (Prost et al., 2006; Pacheco
L.M.M. Jorge et al. / Journal of Food Engineering 99 (2010) 351–359 355

Fig. 4. Evaporation system HYSYS model.

Fig. 5. Comparison between the estimated flow rate values. Fig. 6. Overall heat transfer coefficients for the first and second effects.

ond effect feed stream temperature (1–2). All the process measure-
and Frioni, 2004; Adib et al., 2009) involving the concentration of
ments applied in the model solution were performed before the
sugar solutions in falling film evaporators point to U values in
system stream outlet (CONC2). The comparison of the HYSYS mod-
the range (2080–4000) W/(m2 °C). However, the U values pre-
el predictions with the flow rate, temperature, and concentration
sented in Fig. 6 (262–444) W/(m2 °C), are much lower than ex-
measurements of concentrated juice (CONC2) is a convenient
pected, suggesting performance loss due to incrustation in the
way to assess the quality of the model predictions.
first effect. This possibility finds support in the work of Camargo
Fig. 7 shows the estimated volumetric flow rate, soluble solid
et al. (1990) concerning a sugar industry evaporation system com-
concentration, and temperature values of the concentrated juice
posed of five evaporators in series in which the U values found for
stream (CONC2) in contrast to the measured plant values. The vol-
the first effect were far below those reported in literature. This dis-
umetric flow rate of concentrated juice presented the largest aver-
crepancy was attributed mainly to excessive incrustation in the
age deviation relative to the HYSYS model predictions (4.5%),
evaporator tubes.
followed by the soluble solid concentration (dissolved solids)
The IES HYSYS model (Fig. 4) was solved using the feed stream
(0.67%) and temperature (0.59%). These results validate the HYSYS
flow rate, pressure, and temperature (CLARIF), operating pressures,
and evaporator outlet concentrations (CONC2–1, 2, 3) and the sec- modeling of the IES.
356 L.M.M. Jorge et al. / Journal of Food Engineering 99 (2010) 351–359

Fig. 9. Generated and heating steam flow rates in both effects.


Fig. 7. HYSYS predictions compared with process measurements at the IES outlet.

4.2. Evaporation system integrated analysis efficiency loss for the falling film evaporator in the first 14 days,
possibly due to incrustation which tended to occur in the first ef-
The behavior of clarified (CLARIF), pre-concentrated (CONC1), fect and did not significantly affect the efficiency of the second
and concentrated (CONC2) juice flow rates can be seen in Fig. 8. effect.
In the first day, 195  103 kg/h of clarified juice were processed; According to Hugot (1986), incrustation in sugarcane juice
however, the flow rate gradually decreased to a minimum of evaporator tubes derive from either suspended solids or dissolved
159  103 kg/h on day 5, followed by a continuous increase to salts. The suspended solids are due to deficient previous separation
the maximum of 195  103 kg/h on day 23. Both pre-concentrated (filtration and defection) and are mainly deposited in the first ef-
(CONC1) and concentrated (CONC2) juice flow rates followed this fect, while the dissolved salts become insoluble as the juice con-
same trend, though at lower levels due to the removal of water centration increases and preferentially deposit in the subsequent
in the evaporators, which continuously increased the juice sugar effects. Thus, the deposition of solids in suspension in the first ef-
concentration. fect is believed to be the main responsible for the performance loss
The heating steam flow rates of the first and second effects (FVE of the evaporation system during operation, while incrustation on
and FV1EVAP) during the period of operation are illustrated in Fig. 9. the second effect seems negligible.
While the heating steam flow rate for the falling film evaporator During the analyzed period of operation (Fig. 10), the first effect
continuously decreased, diminishing the amount of steam gener- steam flow rate, FV1, decreased continuously, leading to a progres-
ated in the first effect (FV1), the amount of steam consumed in sive performance loss of the first effect, which resulted in a gradual
the three Robert-type evaporators (FV1EVAP) gradually increased decrease in the sugar content of the pre-concentrated juice
to compensate for the decreased efficiency of the first evaporator. (CONC1). As a result, it was necessary to increase the amount of
Fig. 9 also shows that FV2 and FV1EVAP were almost equal at the steam fed to the second effect (FV1EVAP) so that the sugar content
start of operation and remained very close up to day 23, while FV1 in the concentrated juice (CONC2) was not much lower than the
and FVE were equal at the beginning, but gradually diverted, tend- desired value of 0.18 °Bx, otherwise the subsequent fermentation
ing to stabilize after day 14. This difference indicates a pronounced process would be hindered.

Fig. 8. Juice flow rates in the system. Fig. 10. Steam flow rates of V1, V1EVAP, and BLEEDING streams.
L.M.M. Jorge et al. / Journal of Food Engineering 99 (2010) 351–359 357

Due to the decrease in FV1 (Fig. 10), the amount of surplus steam
from the evaporation system, FBLEEDING, drastically decreased after
day 5. Steam bleeding fell from 11,020 kg/h at the beginning of
operation to 9240 kg/h on day 5 and became null near day 13. At
that time, the amount of steam generated in the first effect (FV1)
was exactly equal to the amount of steam consumed in the second
effect (FV1EVAP), thus canceling the steam surplus (FBLEEDING),
according to the mass balance F V1 ¼ F V1EVAP þ F BLEEDING . From that
point on, the consumed steam flow rate of the second effect (FV1E-
VAP) was higher than the flow rate of steam generated in the first
effect (FV1), thus requiring complementation of the amount of
steam fed to the second effect with steam from the boiler. This
steam complementation was admitted by the BLEEDING stream,
and was mathematically represented by negative FBLEEDING values.
After day 13, the amount of required additional steam continu-
ously increased and tended to stabilize around 2730 kg/h after
18 days of operation.
The performance of the IES evaporators may also be assessed
from the calculation of evaporation fluxes in both effects: falling Fig. 12. Overall thermal resistances.
film evaporator (GFF = FV1/A) and Robert-type evaporators
(GR = FV1EVAP/A). As can be seen in Fig. 11, at the beginning of oper-
ation, GFF (9.62 kg/(m2 h)) was very close to GR (9.92 kg/(m2 h)). mal resistance value for the three evaporators was 7.13 
However, as time passed GFF decreased, probably due to incrusta- 104 m2 °C/W, approximately 5.8 times lower than that found for
tion in the evaporator tubes, while GR slightly increased due to the falling film evaporator at the end of operation. This can be jus-
the increasing heating steam flow rate in the second effect (FV1E- tified by the severe incrustation in the first effect.
VAP), which was required to compensate for the efficiency loss in The fouling resistance, Rf, was calculated from Eq. (14); U0 and U
the first effect (Fig. 10). After day 18, GFF remained practically con- are the overall heat transfer coefficients at startup and at time t > 0,
stant at an average value of 5.2 kg/(m2 h), 46% less than at the respectively (Bansal et al., 2008). U was determined from the
beginning, while GR increased 40% relative to the starting value HYSYS model.
and remained practically constant at around 13.87 kg/(m2 h) after
day 14.
The supposed strong influence of fouling on the first effect per- 1 1
Rf ¼  ð14Þ
formance is demonstrated by the behaviors of the overall thermal U U0
resistances (RG = 1/U) of the first and second effects during opera- Experimental Rf values obtained from Eq. (14) and presented in
tion, as shown in Fig. 12. The resistance values were calculated Fig. 13 varied from 0 m2 °C/W at startup to 22.1  104 m2 °C/W at
from U values estimated by the HYSYS model. The overall heat day 23 and were of the same order of magnitude as those found in
transfer resistance in the falling film evaporator increased 70% the literature (Perry et al., 1984; Rohsenow et al., 1998) for various
from day 1 to day 10 and had a slight further increase after that, liquid streams in different industrial processes: minimum of
tending to stabilize around 41.5  104 m2 °C/W at day 23. This 0.8  104 m2 °C/W for distilled boiler feed water and maximum
behavior was well fitted (R2 = 0.968) to the following equation: of 17.6  104 m2 °C/W for vacuum tower bottoms.
According to the fouling curve classification presented by Ban-
RG ¼ 19:19  104 þ 23:14  104 ð1  expðt=7:421ÞÞ ð13Þ sal et al. (2008), the Rf behavior observed in Fig. 13 is asymptotic,
which is typical of weak deposits. This finding confirms Hugot’s
The three Robert-type evaporators that compose the second ef- (1986) observation that incrustation in the first effect preferably
fect showed equivalent and approximately constant thermal results from the deposition of suspended solids.
behaviors throughout operation (Fig. 12). The average overall ther-

Fig. 11. Evaporation fluxes. Fig. 13. Fouling resistance curve.


358 L.M.M. Jorge et al. / Journal of Food Engineering 99 (2010) 351–359

The Kern and Seaton (1959) model, Eq. (15), is one of the first
models presented in the literature with a consistent theoretical ba-
sis. Initially, the authors used it for the analysis of oil refinery foul-
ing and found it to describe fouling behavior in a number of other
instances (Fernandez-Torres et al., 2001). The basis for this model
is the difference between the fouling removal and deposition rates,
which eventually become equal (Bansal et al., 2008). The net foul-
ing rate then becomes null.

Rf ¼ R1
f ð1  expðt=sÞÞ ð15Þ
The asymptotic fouling resistance (R1
f )
represents the maxi-
mum thermal resistance that can be produced by incrustation,
and the fouling time constant (s) represents the time required
for Rf to be 37% of the R1f value. This value is 86% when t = 2s. As
one can see in Fig. 13, the Kern and Seaton model adequately rep-
resents (R2 = 0.999) experimental Rf data obtained from Eq. (14).
The obtained R1 f value was 23.14  10
4
m2 °C/W and the fouling
time constant (s) was equal to 7.421 days (6.412  105 s). Fig. 14. Juice temperatures in the system.

Rf ¼ 23:14  104 ð1  expðt=7:421ÞÞ ð16Þ


The limiting step in the falling film evaporator is the heat trans-
fer in the incrustation of the tubes, Rf. Thus, it is expected that Rf
imposes the same behavior to the overall thermal resistance, RG.
In this way, the same functional relationship of the Kern and Sea-
ton model was admitted to represent the behavior of RG with time,
Eq. (13). When this correlation is compared with Eq. (16), it ap-
pears that both RG and Rf have exactly the same behavior through-
out the operation period with a constant difference between them
(23.14  104 m2 °C/W).
In turn, the overall thermal resistance, RG, is the sum of four ser-
ies thermal resistances (Adib et al., 2009): RG = RV + RW + Rf + RL,
where RV, RW, and RL are the thermal resistances of the heating
steam, the tube walls, and the juice, respectively. Therefore, it ap-
pears that the difference between RG and Rf represents the sum of
the three remaining thermal resistances, that is RV + RW + RL =
19.19  104 m2 °C/W.
Given the above, it seems that significant fouling in the first ef-
fect during the period of operation resulted from the deposition of
suspended solids remaining from the previous filtration and set-
tling operations. There was no evidence of significant incrustation Fig. 15. Juice concentrations in the system.
in the second effect.
Thus, it is evident that if the IES were cleaned in less than
13 days, there would be steam economy because the steam pro- would be used to raise the temperature of the clarified juice to
duced in the first effect would be more than enough to adequately the boiling point (about 117 °C).
heat the second effect and no additional steam from the boiler Due to the progressive reduction of efficiency of the first effect,
would be required. Furthermore, the first effect would generate a the juice concentration at its output (CONC1) decreased over time
significant amount of excess steam that could be reused in the fac- and imposed a juice concentration reduction at the output of the
tory, saving energy in the boiler. second effect (CONC2), as shown in Fig. 15. In this figure, B* corre-
Despite the large efficiency gain obtained in the first effect with sponds to normalized concentrations and was obtained by dividing
more frequent cleaning, a major drawback is shutdown, which B of the indicated streams by the value of B of the system feed
might be circumvented in two ways: by installing an additional stream (B0). At the beginning of operation, the difference between
evaporator that would operate in parallel with the first effect the B* values for the clarified juice fed to the first effect and for the
and/or by improving the performance of the previous suspended pre-concentrated juice at its output was 0.133, falling to 0.0664 on
solids separation steps. day 23, a decrease of 50%. In the same period, the difference be-
Besides the previously discussed factors, it is important to stress tween the normalized concentrations at the inlet and outlet of
that the usual incrustation at the clarified juice pre-heaters also the second effect Robert-type evaporators remained virtually un-
may secondarily affect the IES performance. The pre-heaters usual changed, confirming that the performance of the second effect re-
incrustation led to a continuous decrease of the clarified juice tem- mained practically constant throughout the operation period,
perature fed to the first effect (CLARIF), while the temperatures of while that of the first effect decreased sharply.
both pre-concentrated (CONC1) and concentrated (CONC2) juices
remained almost constant throughout operation (Fig. 14). The clar-
ified juice temperature has changed from 116 °C at the beginning 5. Conclusions
of the operation up to 109 °C by the end of experiment and could
exert some influence on the first effect operation and affect the The representation of evaporators by the combination of heat
IES performance since part of the first effect heating steam (FVE) exchangers with phase separator vessels was successfully used to
L.M.M. Jorge et al. / Journal of Food Engineering 99 (2010) 351–359 359

model and simulate an industrial evaporator system (IES) using the F iin Biin ¼ F iout Biout ðA:1Þ
HYSYS process simulator.
The model implemented in HYSYS, as well as using a conven- Eq. (2) was obtained by combining the energy balances in the
tional model, adequately represented the IES behavior. The main cold stream (Eq. (A.2)) and in the hot stream (Eq. (A.3)), assuming
 
i i i i i i i i
steam and juice stream flow rate values predicted with the conven- hfgV ¼ hFout  hV , hfgS ¼ hS  hC , and hFout  hFin ¼ Cpm T iFout  T iFin .
tional model showed a mean absolute difference relative to the i i i
F iin hFin þ Q i ¼ V i hV þ F iout hFout ðA:2Þ
HYSYS simulations of only 0.11%.
The COCAFÉ evaporation system showed marked performance i i
Si hS ¼ Q i þ C i hC ðA:3Þ
loss in the first 14 days of operation, mainly due to incrustation in
the first effect, which led to a 70% increase in overall heat transfer
resistance from startup to day 10, tending to an asymptote. References
Significant fouling occurred in the IES first effect during the per-
iod of operation due to deposition of suspended solids remaining Adib, T.A., Heyd, B., Vasseur, J., 2009. Experimental results and modeling of boiling
heat transfer coefficients in falling film evaporator usable for evaporator design.
from previous filtration and settling operations. There was no evi- Chem. Eng. Process. 48, 961–968.
dence of significant incrustation in the second effect. Bansal, B., Chen, X.D., Müller-Steinhagen, H., 2008. Analysis of ‘classical’ deposition
The fouling curve behavior was asymptotic and well repre- rate law for crystallisation fouling. Chem. Eng. Process. 47, 1201–1210.
Camargo, C.A., Ushida, A.H., Ribeiro, A.M.M., Souza, M.E.P., Santos, N.F., 1990.
sented by the classic Kern and Seaton model, with R1 f and s values Conservação de energia na indústria do açúcar e do álcool, first ed. Instituto de
of 23.14  104 m2 °C/W and 7.42 days, respectively. Pesquisas Tecnológicas do Estado de São Paulo (IPT), São Paulo.
The steam generation flow rate of the first effect, GFF, showed a Chawankul, N., Chuaprasert, S., Douglas, P., Luewisutthichat, W., 2001. Simulation of
an agitated thin film evaporator for concentrating orange juice using
behavior opposite to that of the thermal resistance, a 46% reduc-
AspenPlus™. J. Food Eng. 47, 247–253.
tion from day 1 to day 18, while the latter increased during the CONAB – Companhia Nacional de Abastecimento, 2005. Indicadores da
period of operation, which required heating steam complementa- Agropecuária, vol. 10, p. 18. <http://www.conab.gov.br/conabweb/
tion for the second effect (V1EVAP) with steam from the boiler index.php?PAG=212> (acessed 13.07.09).
CONAB – Companhia Nacional de Abastecimento, 2009. Indicadores da
near day 13. Agropecuária, vol. 4, p. 28. <http://www.conab.gov.br/conabweb/
The IES performance was secondarily affected by the loss of per- index.php?PAG=212> (acessed 13.07.09).
formance of the clarified juice pre-heaters due to the occurrence of Ensinas, A.V., Modesto, M., Nebra, S.A., Serra, L., 2009. Reduction of irreversibility
generation in sugar and ethanol production from sugarcane. Energy 34, 680–
usual incrustation, which led to a continuous decrease of the clar- 688.
ified juice temperature. Consequently, part of the steam fed to the Fernandez-Torres, M.J., Fitzgerald, A.M., Paterson, W.R., Wilson, D.I., 2001. A
first effect had to be used to raise the juice temperature to its boil- theoretical study of freezing fouling: limiting behaviour based on a heat and
mass transfer analysis. Chem. Eng. Process. 40, 335–344.
ing point rather than for its evaporation. Foust, A.S., Wenzel, L.A., Clump, C.W., Anderson, L.B., 1982. Princípios das Operações
The second evaporation effect did not suffer significant perfor- Unitárias, Guanabara Dois, Rio de Janeiro.
mance loss over the 23 days of operation, suggesting that it could Hugot, E., 1986. Handbook of Cane Sugar Engineering, third ed. Elsevier,
Amsterdam.
operate uninterruptedly for a longer period. This result points to Kern, D.Q., Seaton, R.A., 1959. A theoretical analysis of thermal surface fouling. Br.
the need to install an additional evaporator parallel with the first Chem. Eng. 14 (5), 258–262.
effect to allow its cleaning without process interruptions. The first Krupiczka, R., Rotkegel, A., Ziobrowski, Z., 2002. Heat transfer to evaporating liquid
films within a vertical tube. Chem. Eng. Process. 41, 23–28.
effect cleaning should be performed in less than 13 days for all the
McCabe, W.L., Smith, J.C., 1976. Unit Operations of Chemical Engineering, third ed.
heating steam required for the second effect to be produced by the McGraw-Hill, Tokyo.
first effect without the need of complementation of steam from the Minton, P.E., 1986. Handbook of Evaporation Technology. Noyes Publications, New
boiler. Jersey.
Montanini, L., 2008. Brazilian Sugar and Ethanol Guide (Anuário da Cana), Procana,
São Paulo.
Pacheco, C.R.F., Frioni, L.S.M., 2004. Experimental results for evaporation of sucrose
Acknowledgements solution using a climbing/falling film plate evaporator. J. Food Eng. 64, 471–480.
Pacheco, C.R.F., Cézar, C.A., Song, T.W., 1999. Effect of the solute concentration on
the performance of evaporators. Chem. Eng. Process. 38, 109–119.
The authors gratefully acknowledge the financial support from
Perry, R.H., Green, D.W., Maloney, J.O., 1984. Perry’s Chemical Engineers’ Handbook,
CNPq (Conselho Nacional de Desenvolvimento Científico e Tec- sixth ed. McGraw-Hill, New York.
nológico, Brazil) and the COCAFÉ’s distillery permission to obtain Prost, J.S., Gonzáles, M.T., Urbicain, M.J., 2006. Determination and correlation of heat
the operating data necessary to develop this research. transfer coefficients in a falling film evaporator. J. Food Eng. 73, 320–326.
Rohsenow, W.M., Hartnett, J.P., Cho, Y.I., 1998. Handbook of Heat Transfer, third ed.
McGraw-Hill, New York. pp. 17148–17149.
Seider, W.D., Seader, J.D., Lewin, D.R., 1999. Process Design Principles. John Wiley &
Appendix A. Development of Eqs. (1) and (3) Sons, New York.
Traganits, S.P., 1981. Designing multiple-effect evaporators in sugar factories using
HP-41C programmable calculator. La Sucrerie Belge 100, 385–394.
Eq. (1) was obtained by combining the overall mass balance in Volin, Yu.M., Ostrovskii, G.M., 2006. Three phases in the development of computer
the cold stream (juice side), Eq. (3), and the dissolved solid simulation of chemical engineering systems. Theor. Found. Chem. Eng. 40, 281–
balance: 290.

You might also like