Anda di halaman 1dari 47

Accepted Manuscript

Mathematical modelling of the performance parameters of a new decanter cen-


trifuge generation

Alessandro Leone, Roberto Romaniello, Riccardo Zagaria, Antonia Tamborrino

PII: S0260-8774(15)00222-8
DOI: http://dx.doi.org/10.1016/j.jfoodeng.2015.05.011
Reference: JFOE 8168

To appear in: Journal of Food Engineering

Received Date: 7 February 2015


Revised Date: 29 April 2015
Accepted Date: 10 May 2015

Please cite this article as: Leone, A., Romaniello, R., Zagaria, R., Tamborrino, A., Mathematical modelling of the
performance parameters of a new decanter centrifuge generation, Journal of Food Engineering (2015), doi: http://
dx.doi.org/10.1016/j.jfoodeng.2015.05.011

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Mathematical modelling of the performance parameters of a new decanter centrifuge

generation.

Alessandro Leonea, Roberto Romaniello a*, Riccardo Zagariaa, Antonia Tamborrino b.

a
Department of the Science of Agriculture, Food and Environment, University of Foggia, Via

Napoli, 25 - 71122 Foggia, ITALY.


b
Department of Agricultural and Environmental Science, University of Bari Aldo Moro,

Via Amendola 165/A 70126 Bari, ITALY.

*Corresponding author: Roberto Romaniello

Tel.: +39 0881 589120

E-mail address: roberto.romaniello@unifg.it

Keywords: decanter centrifuge; olive oil, centrifugal separation.

Abbreviations

Bowl rotation speed

Screw rotation speed

n Conveyor-bowl differential speed

cyl Length of the bowl cylindrical section

con Length of the bowl conical section

rb Bowl radius

rs Solis discharge radius

ro Olive oil discharge radius


rwp Wastewater/pt discharge radius

Cone half-angle

Mp Olive paste flow mass rate (kg h-1);

Mo Olive oil flow mass rate (kg h-1);

Mw Process water flow mass rate (kg h-1);

Mwp Wastewater/pt flow mass rate (kg s-1);

Mh Husk flow mass rate (kg s-1);

Oh Residual of olive oil in husk (%);

Op Residual of olive oil in pt (%);

Ow Residual of olive oil in the wastewater (%);

w.m. Wet matter;

d.m. Dry matter;

Oo Olive oil in olives (%)

Hh Water content in husk (%)

Hp Water content in pt (%)

Hw Water content in wastewater (%)

s.s Soft solid

r.s. Rigid solid

f Cavity pump motor frequency (Hz)

E Extraction efficiency (%);

SSE Sum of squared errors of prediction (SSE)

R2 Coefficient of determination

R2adj Adjusted R2

MPE Mean percentage error

MBE Mean bias error


RMSE Root mean square error

EF Modelling efficiency

Pexp,i Experimental parameter

Ppre,i Predicted parameter

Pexp,ave Experimental average parameter

2cal Chi-square calculated

2tab Chi-square distribution table

N Number of data points

H0 Null hypothesis

Abstract

A new decanter centrifuge, called Eureka, of the pt generation was studied in two different

configurations (with and without water added) to evaluate its performance parameters.

Mathematical models were developed to predict the extraction efficiency and the oil content

in the husk, wastewater and pt as a function of the olive paste mass flow rate. The

suitability of the mathematical models of the performance parameters was evaluated using

various statistical parameters, such as the mean percentage error, the mean bias error, the root

mean square error, the modelling efficiency and the chi-square test. The models developed

showed excellent generalization capabilities.

The decanter was shown to be able to process olive paste with a high extraction efficiency

that, without water added, could reach 92.5%; sufficiently dry solids and olive oil clarified by

light solids were produced in both of the configurations studied. The decanter was also

demonstrated to be able to switch from one configuration to the other without stopping

operation.
1. Introduction

A decanter machine is a horizontal centrifuge that operates continuously to separate solid and

liquid fractions using centrifugal acceleration. The principal inner elements of a decanter

machine include a screw-scroll conveyor mounted concentrically within a bowl. These two

components allow for the mechanical separation of solids that are forced outside the bowl

from liquids (Anlauf, 2007; Letki, 2007; Records and Sutherland, 2001).

The fluid is pumped through a tube placed inside the bowl, which has an internal radius rb

and rotates at an angular velocity . Inside the bowl, the fluid reaches a high centrifugal

acceleration 2r that causes the solids to adhere to the bowl wall. The centrifugal acceleration

intensity varies with the radius and angular velocity of the bowl (Anlauf, 2007; Bell et al.,

2014; Leung, 2007; Records and Sutherland, 2001). The solids that settle around the internal

wall of the bowl are transported outside it by a scroll conveyor, which rotates at a different

angular velocity n respect to the bowl. The conveyors rotation permits the solids transfer

from the cylindrical bowl (i.e., pool section) to the wet and dry conical beach zones via the

solids discharge holes. (Records and Sutherland, 2001; Reif et al., 1990). The liquids flow in

the opposite direction to the solids towards the weir at the end of the cylindrical bowl. The

weir is placed at an optimum distance from the rotation centre (i.e., liquid level radius).

(Berk, 2013; Corner-Walker and Records, 2000; Madsen, 1989). For the fluid to be separated

into its two immiscible liquids constituents with different specific gravities and a solid with a

higher specific gravity than that of either liquid, the decanter is modified to have two liquid

weirs at two different radii: one close to the rotation centre for the lighter liquid and one

farther from the rotation centre for the heavier liquid (Bell et al., 2014, Letki and Corner-

Walker, 2003; Leung, 2007). To avoid cross contamination of the two liquids discharged

from the bowl, floodgates are placed on the external side of the bowl in the casing of the

machine.
The high versatility of decanter centrifuges has led to their widespread use in various

industrial applications. In the olive oil extraction industry, the decanter is the most important

mechanical innovation for separating the husk from the liquid phases (Altieri et al., 2013;

Tamborrino et al., 2015). It is used in a 2- or 3-phase configuration in conjunction with the

recent innovation of olive paste conditioning (Ayr et al., 2015; Catania et al., 2013; Leone et

al., 2014a; Leone et al., 2014b; Tamborrino et al., 2014a; Tamborrino et al., 2014b).

In the 3-phase decanter configuration (Fig. 1a), the olive paste is often conveyed to the

conical side of the bowl into the centrifugal extractor by a variable flow-rate cavity pump.

Inside the bowl, the two liquids (i.e., oil and vegetable water) are discharged through two

different liquid weirs placed on the same side of the bowl (i.e., the cylindrical side), whereas

the solids are discharged through husk holes placed on the opposite side of the bowl (i.e., the

conical side).

Three-phase sedimentation is possible due to the addition of warm water (at the same

temperature of the olive paste) to the paste inlet at a variable percentage of 10-30%. The

addition of water creates a significant flow of wastewater, that requires careful disposal along,

and husks at 50-55% humidity (Dermeche et al., 2013; Pastore et al., 2014; Salvador et al.,

2003; Tamborrino et al., 2015).

In the 2-phase configuration, the decanter (Fig. 1b) is pre-set with only two outlets: one for

the oil and the other for the wastewater and husk together. Generally in this configuration, the

olive paste is fed from the cylindrical side of the bowl. Two-phase separation can be

performed without the addition of process water. In this process, the final products are olive

oil and husks at 65-70% humidity (Alburquerque et al., 2004; Baccioni and Peri, 2014;

Dermeche et al., 2013; Roig et al., 2006). An important advantages of this type of decanter is

the absence of wastewater, the absence of the process water added and the different olive oil

quality compared the 3-phase separation; however, the husk is significantly moister than the
3-phase decanter husk, which represents a significant disadvantage in terms of disposal costs

and natural resources. Often in order to reduce the humidity, the wet husk is processed again

in a 3-phase decanter, until to obtain husk with low humidity and wastewater. The latest

generation of decanter features innovative geometry, with a longer cylindrical portion of the

bowl, shorter conical sections compared to than those in the traditional geometry, and a

special baffle that ensures greater product performance (Altieri, 2010; Amirante et al., 2010;

Catalano et al., 2003).

Today, decanters used in the olive oil extraction process have specific configurations for 2- or

3-phase operation, and their configuration is not interconvertible from one to the other. Some

olive oil machines manufacturers produce decanters that can operate in 2- or 3-phase

configurations, depending on what adjustments are made. In this case, however, switching

between 2- and 3-phase operation requires stopping the decanter, removing its protective

casing and manually closing the heavy-liquid weirs. This manual operation is difficult during

regular plant management.

This study was designed to model the process parameters of a new type of decanter (Eureka,

Barracane s.r.l., Modugno, BA, Italy), specifically, those of the recent generation of decanter

called pt. To date, only three papers have been published on this topic, those by Altieri et

al. (2013), Tamborrino et al. (2015), and Caponio et al. (2014). The authors previously

investigated the performances of two different types of decanter: Megala 450 and Megala

650, Barracane s.r.l., Modugno, BA, Italy. The new model examined in this study, called

Eureka, utilizes mechanical changes to perform two different 3-phase processes: one with

water addition (3-phase-WA), which is similar to the 3-phase decanter previously described,

and one innovative configuration without water addition (3-phase-WW). This decanter is new

and innovative for two different reasons: (1) it uses an innovative 3-phase-WW configuration,

separately discharging olive oil, husks rich in heavy solids at 50-55% humidity and a new by-
product called pt, which contains wastewater enriched by soft solids without pit

fragments with semi-solid consistency; and (2) the new decanter is the only olive oil decanter

that can switch from one configuration to another without stopping the machine. The latter

feature allows for the modulation of by-product characteristics according to the requirements

for disposal or opportunities for their recovery. The pt could represent a by-product having

cheap cost if coupled with improving technologies for the production of biogas or for human

or animal feeding. In addition the absence of the wastewater will avoid its disposal costs.

The purpose of this study was to evaluate the performance of the new decanter, to define the

best process parameters and to characterize the composition of the husk and wastewater/pt

discharged by changing certain process parameters.

Finally, modelling of the decanter performance parameters was made. The modelling will

allow for predictions of changes in the decanter's extraction efficiency and in the composition

of the husk and wastewater/pt as a function of the set olive paste mass flow rate (Mp) and

n.

2. Materials and methods

2.1. Experimental plan

All tests were performed in the 2014 crop season in continuous mode without stopping the

decanter using a homogeneous batch of olive fruits of the cultivar Coratina (Olea europaea

L.) with a maturity index of 1.4 (IOOC, 2001). Different mass flow rates and different n

values were used. Additionally, three different wastewater/pt discharge radius rwp were set.

The olive oil discharge radius was the same in all tests and was equal to 274 mm, which

represents the distance from the oil discharge weir to the bowl rotation centre. Two test

conditions corresponding to the two different decanter settings (3-phase-WA and 3-phase-

WW) were compared. For each test condition, seven different mass flow rates performed in
triplicate were considered. The minimum and maximum mass flow rates were chosen near the

minimum and maximum flow rate acceptable for the decanter used.

The n were chosen on the basis of mass flow rate values used for the experimental tests,

taking also in account the decanters project specifications whose associate a specific n

value for each mass flow rate value. For 3-phase-WA condition, warm process water (14.4%

of olive paste mass) was added. Table 1 describes the experimental plan.

To calculate the mass flow rates in the input and output components of the decanter, olive

paste, olive oil and wastewater/pt mass flow rates were measured experimentally during

every test and the mass flow balance was calculated.

According to the experimental plan, mathematical models were defined for the prediction of

extraction efficiency (E), residual of olive oil in the husk (Oh), residual of olive oil in pt

(Op) and residual of olive oil in wastewater (Ow), based on Mp. The definition of the

mathematical models involved two steps: training and testing. For training in each decanter

configuration, a set of the 21 experimental parameters mentioned above (i.e., seven different

Mp values replicated three times) were experimentally determined and used for the definition

of the equations that best fit the parameter values versus Mp. The models determined were

validated during testing using a set of 24 parameters for each decanter configuration

considered (i.e., twelve different Mp values replicated twice). Using the mathematical models

determined in the training step, the parameters E, Oh and Ow for the 3-phase-WA

configuration, and E, Oh and Op, for the 3-phase-WW configuration were calculated and

compared to the corresponding experimental values to determine the models goodness of fit.

In the training step, the duration of each test was approximately 60 minutes, the first 10

minutes of which was used for decanter stabilization. For the validation step, the duration of

each test was approximately 30 minutes, the first 5 minutes of which was used for decanter

stabilization. In all tests, the olive paste was conditioned in a group of open malaxer machine
connected in series, for 60 minutes at 27 C. In order to have the same malaxation time for

each mass flow rate chosen, the number or the fill volume of the malaxers used for each test

was varied.

2.2. Industrial olive oil plant

The olive oil extraction plant was equipped with an olive cleaning system consisting of a leaf

removing machine (model Tornado, Clemente and C. Snc, Altamura, Ba) and a washing

machine (model Ocean 3, Clemente and C. Snc, Altamura, Ba). The washed olives were

conveyed to a hammer crusher (model Frangolea, Barracane s.r.l., Modugno, BA, Italy), and

then to a group of five malaxer machines (model Gramola 3000, Barracane s.r.l., Modugno,

BA, Italy) connected in series and having a capacity of 2200 kg each one. After malaxation,

the olive pastes were conveyed to the decanter by a cavity pump. For the solid-liquid and

liquid-liquid separation, respectively, a new model of decanter (model Eureka, Barracane

s.r.l., Modugno, BA, Italy) and vertical plate centrifuge (mod. Matic 3000, Barracane s.r.l.,

Modugno - BA Italy) were used.

2.3. Horizontal centrifugal decanter

The new model of decanter centrifuge, Eureka, is able to operate in the two different

configurations (3-phase-WA and 3-phase-WW) without stopping the machine.

In the 3-phase-WA configuration (Fig. 2a), the decanter discharges olive oil, husks (i.e., rigid

and soft solids) at 50-55% humidity and wastewater, whereas in the innovative 3-phase-WW

configuration (Fig. 2b), the decanter discharges olive oil, husks (rich in rigid solids) at 50-

55% humidity and pt consisting of wastewater and soft olive solids. The pt has a semi-

solid consistency.
The decanter is also able to operate in a third configuration. By combining the pt and husks,

it is possible to obtain a by-product equal to that obtained in the 2-phase configuration

without adding water.

A schematic representation of the Eureka decanter centrifuge model with dimensions and

operating parameters is shown in Fig. 3; the corresponding details are listed in Table 2.

The bowl and screw were connected to two different electric motors with 45 and 30 kW of

electric power, powered by an electro-mechanical system consisting of two inverters and a

planetary gearbox with a transmission ratio equal to 1:59.

The decanter was equipped with an automatic system controlled by an operator, who could

adjust the n value. By controlling the inverters of the two electric motors of the bowl and

screw, the system regulates the rotation speeds until they reach an established n value.

The olive oil discharge radius (ro) and waste water/pt discharge radius (rwp) can be

regulated manually without stopping the machine by two different adjusting screws that act

on two mechanical systems that define the position of the exit holes of the wastewater/pt

and oil, changing their distance from the bowl rotation centre.

2.4. Mass flow rate determination

The experimental tests were conducted using five different mass flow rate values. To rapidly

set the mass flow rate required in a specific test condition, the authors have determined the

correlation existing between the cavity pump motor frequency and the mass flow rate. For

this purpose, an electronic welting system was used. In the experimental tests, seven mass

flow rates were used, which produced seven set frequencies. For each seven frequencies, the

exact mass flow rate was calculated by measuring the time required to welt 500 kg of olive

paste. The measurement was replicated 4 times.


The mass flow rate of the olive oil and wastewater/pt were determined using the same

welting system; otherwise, the husk mass flow rate was determined indirectly based on the

abovementioned parameters. The process water added was measured using a digital flow

meter.

2.5. Mass balance

A mass balance of the inputs and outputs was performed to calculate the decanter extraction

efficiency and the weight distribution between the output phases.

The mass balance was calculated by considering the following parameters: Mp, the olive

paste flow mass rate (kg h-1); Mo, the olive oil flow mass rate (kg h-1); Mw, the process water

flow mass rate (kg h-1); Mwp, the wastewater/pt flow mass rate (kg h-1); and Mh, the husk

flow mass rate (kg h-1). Mh was calculated based on the other terms using the following

equations depending on the test conditions:

Mh = Mp + Mw Mo Mwp (1)

for the 3-phase-WA test condition; and

Mh = Mp Mo Mwp (2)

for the 3-phase-WW test condition.

E value was calculated using the following equation:

Mo
E 100
Oo
Mp
100

where Oo represents the amount of olive oil contained in the olives.

2.6. Sampling
For each test, olives were sampled and stored at +1 C until analysis. The olive oil, husk,

wastewater and pt were sampled at regular time intervals of 5 minutes and stored at +1 C

until analysis.

2.7. Oil content in the olives, husk, pt and wastewater

The total oil content was determined using 30 g of olive paste, husk, pt and using 50 g of

wastewater, for each test condition considered. All samples were previously dehydrated until

they reach a constant weight. Each sample was extracted with hexane in an automatic

extractor (Randall 148, Velp Scientifica, Milano, Italy) according to the analytical technique

described by Cherubini et al. (2009). The sample was initially subjected to an immersion

phase at 139 C for 60 min; the porous container of the sample was immersed directly in the

boiling solvent. The sample was then subjected to washing at 139 C for 40 min; the sample

container was then removed from the solvent and reflux washed. The final part of the process

was solvent recovery, which was conducted at 139 C over a period of 30 min. The results

were expressed as a percentage of oil on wet and dry matter.

2.8. Composition of olive, husk, pt and wastewater

An Mp value of 5640 kg h-1 was used to determine the composition of the olive, husk, pt

and wastewater by sampling them from the decanter during the experimental test. The process

was replicated three times.

To determine the water and oil contents, the procedure described in Chapter 2.7 was used.

To determine the amount of rigid solids present, 200 g aliquots of olive paste, pt and

wastewater were weighed and placed into Imhoff cones to separate the soft solids from the

rigid solids (i.e., pit fragments). The obtained pits were washed with hexane to remove oil

residues and then dried in a ventilated oven at 105 C until a constant weight was reached.
The mass of soft solids was obtained by subtracting the masses of oil, water and rigid solids

from the sample weight (Leone et al., 2105).

2.9. Statistical analysis

All experimental data were analysed using an analysis of variance (ANOVA) and a post-hoc

Kurskal Wallis test via the Statistical toolbox of Matlab (Mathworks, Inc., Natik, MA,

USA).

The performances of the proposed mathematical models were evaluated using the following

statistical parameters:

- mean percentage error (MPE);

- mean bias error (MBE);

- root mean square error (RMSE);

- modelling efficiency (EF);

- chi-square (2).

These statistical analysis allowed for the detection of differences between the experimental

data (Pexp,i) and the model estimates (Ppre,i). These parameters can be calculated as follows:

N
1 Pexp,i Ppre,i
MPE 100
N i 1 Pexp,i

N
1
MBE Ppre,i Pexp,i
N i 1

N
1 2
RMSE Pexp,i Ppre,i
N i 1

N N
2 2
Pexp,i Pexp,ave Ppre,i Pexp,i
EF i 1
N
i 1
2
Pexp,i Pexp,ave
i 1
2
2
N
Pexp,i Ppre,i
i 1 Ppre,i

For the chi-square test (2), the significance level was set to p<0.05, the number of degrees of

freedom was N-1 and the null hypothesis of independence (H0) was that there are no

significant differences among the performance expected and the observed parameters

investigated.

3. Results and discussions

The experimental plan investigated variations in Mp and n to test the decanter behaviour and

to evaluate the decanters performance parameters. The values of Mp were set based on the

frequency (f) variation in the cavity pump motor used to feed the decanter, whose values are

reported in table 3. The correlation between f and the Mp measured experimentally resulted

high, with an R2 value equal to 0.999.

3.1. Modelling of decanter performance parameters

The performance results obtained are summarized in Table 3 and Figures 4-9 shows the

corresponding trends. Figures 4, 6 and 8 show the trained prediction models for E, Oh, and

Ow versus Mp for the 3-phase-WA configuration, respectively; figures 5, 7 and 9 show the

trained prediction models for E, Oh, and Op versus Mp for the 3-phase-WW configuration,

respectively. The model that best fit the experimental data was the second-order polynomial

model. The goodness of fit results are shown in Table 4. All models worked well on the

training dataset. In fact, the Radj values were near 90%, with the unique exception for 3-phase-

WA condition showing and Radj of 0.78%, and the RMSE was lower than 0.02 for all

parameters; the SSE values were also low. Higher SSE values were recorded near 1.38 x 10-3,

representing a low error of estimate.


Table 5 shows the statistical parameters used to evaluate the performance of the

mathematically trained model in the validation step. For each parameter considered, the

predicted and experimental data series were subjected to the 2 test to reject the null

hypothesis. All data series showed a 2 value significantly lower than those in the 2 table

(2tab = 35.17), considering that the number of degrees of freedom was equal to 23 and that

p<0.05. Thus, the data series were not significantly different, and good correlations were

observed between the predicted and observed parameter values. High values of EF were

recorded ranged 89-98%. Additionally, the MPE and MBE values resulted low, confirming

the good generalization of the mathematical models for predicting the considered parameters

based on Mp. Figs. 10 and 11 show the predicted versus the observed values of E for the 3-

phase-WA and 3-phase-WW conditions, respectively; the R2 values were equal to 0.89 (R2adj

= 0.88) and 0.91 (R2adj = 0.89), respectively, indicating good correlation. This finding was

also confirmed by the low RMSE value, equal to 0.45 and 1.01, respectively.

3.2. Discussion of decanter performance parameters

By analysing the data gathered in Table 3, it is possible to conclude that in the 3-phase-WA

configuration up to an Mp near 8000 kg h-1, there were no significant differences in E. In the

3-phase-WW configuration up to an Mp near 7000 kg h-1, there were no significant

differences in E. For values of Mp greater than 7000 kg h-1, E decreased significantly.

In the 3-phase-WA configuration, the efficiency remained stable for a range of flow rates

greater than that of the 3-phase-WW configuration. In addition, comparing the efficiency

trends of the two configurations decanters (Fig. 12), the values of the 3-phase-WW

configuration were significantly higher than those of the 3-phase-WA configuration up to an

Mp near 7500 kg h-1; from 7500 kg h-1 to 9000 kg h-1, significant differences were not
observed, and above 9000 kg h-1, the value of E for the 3-phase-WA configuration was

observed to be significantly higher than the value of E for the 3-phase-WW configuration.

Similar efficiency values were obtained in a previous experiment conducted by Tamborrino et

al. in 2015 on a Megala 650 decanter, which is part of the pt generation of decanters but

required the addition of 7-9% water added into the olive pastes. In addition, as reported in a

previous study conducted by Altieri et al. in 2013, efficiency could be increased by

optimizing the regulation of n and the radius of the liquid weirs.

The results of E are also confirmed by the values of oil content in the husk and

wastewater/pt for the two different configurations, reported in Table 3. The oil content in

the husk increased significantly when either low or high Mp were used but decreased

significantly when intermediate Mp values were used. For both configurations, a significant

increase in oil content in the wastewater and pt between low and high Mp was observed.

Fig. 13 shows that the use of water in the 3-phase-WA configuration enhanced the separation

of olive oil from the wastewater compared to that from pt.

The opposite situation was observed for the 3-phase-WW configuration, in which the absence

of water worsened the separation of olive oil from the pt (Fig. 13). Consequently, the oil

lost in the husk in the 3-phase-WW configuration was lower than the oil content in the husk

produced by the 3-phase-WA configuration (Fig. 14).

Table 3 shows water content in the husk (Hh) and water content in the waste water (Hw)/

water content in the pt (Hp) versus Mp data.

For these parameters, no significant changes in humidity were noted, except for the Hw value.

In addition, comparing Hh in both configurations (Fig. 15), no significant differences were

observed, whereas there were significant differences between Hw and Hp based on mass flow

rate (Fig. 16). This finding demonstrates that 1) the addition of process water to olive paste

does not create a significant increase in Hh between the two configurations and 2) the process
water added into the decanter is necessary to improve the sedimentation of the light solids

from wastewater. Using the 3-phase-WA configuration, the addition of process water allows

for better sedimentation of light solids from the wastewater to the husk, whereas the 3-phase-

WW configurations lack of added process water does not allow for efficient sedimentation of

light solids, which instead flow into the pt.

Finally, it is noted that switching between the two configurations occurred in both directions

in less than one minute which was accomplished by adjusting the radius of the

wastewater/pt weir using a manual tool and opening the process water valve. Additionally,

switching between the two configurations occurred without stopping the machine.

3.3. Composition of the input and output of the decanter

Table 6 show the results of the solid/liquid composition of the input/output of the decanter

calculated from 100 kg of olives, for both configurations.

From the 3-phase-WA configuration to the 3-phase-WW configuration, the percentage of soft

solids shifted from 12.8% in the wastewater to 18.2% in the pt and increased the percentage

of the rigid solids in the husks (from 28.9% to 33.6%). In addition, no differences in the soft

solid contents in the oil between the two configurations were observed, and no rigid solids

were found in the pt. These results indicate that (1) the absence of added water does not

compromise the clarity of the oil obtained from the solids and (2) the absence of added water

and the optimization of the process parameters of the decanter do not compromise the rigid

solid sedimentation that flows in the husks that do not leave a residue in the pt.

Finally, it is important to emphasize that between the 3-phase-WA configuration and the 3-

phase-WW configuration, average water savings amounting to 14% of the weight of the

olives was attained.


4. Conclusions

Due to the strict laws for the disposal of mills wastewater, to the regularly changing rules,

increased disposal costs, and environmental problems combined with the development of new

technologies for biogas production from wastes, millers require decanters having high

flexibility configuration (with/without process water added) and ability to modulate the

output wastes composition without stopping the decanter.

Satisfying this last point would allow millers to modulate wastewater composition based on

its intended use without stopping the decanter. Currently, the typical decanter used for virgin

olive oil extraction does not achieve this point.

In this study the decanter Eureka developed show excellent quantitative performance and can

run in two configurations (i.e., with and without added water) with high extraction efficiency,

producing solids sufficiently dry and olive oil clarified by light solids. Additionally, it is

possible to switch from one configuration to the other without stopping the machine. For

example, in the 3-phase-WA configuration, the decanters discharge wastewater may

currently be disposed of into soil according to local regulations with consideration of disposal

costs. Instead, in the 3-phase-WW configuration without added water, the decanters

discharge pt could be used for the production of biogas or other uses, which could be

valuable. This configuration would also avoid the cost of disposing of wastewater.

Mathematical models developed for the prediction of functional performance as a function of

the olive paste mass flow show excellent generalization capabilities and would be useful to

determine the efficient use of a decanter. This permits to rapidly adjust the decanters settings

in order to obtain the exact desired performances.

This type of decanter is an important mechanical innovation that could convert wastewater

with no current economic value to a by-product with economic value when coupled with
improving technologies for the use of pt for the production of biogas or for human or

animal feeding. This new process could avoid or significantly reduce the influence of

wastewater management on the production costs of oil and on the environment.

Acknowledgements

The authors would like to acknowledge the support of BARRACANE S.r.l. and their

technical staff, who provided insight and expertise that significantly improved this study.

References

Alburquerque, J. A., Gonzlvez, J., Garca, D., & Cegarra, J. (2004). Agrochemical

characterisation of Alperujo a solid by-product of the two-phase centrifugation

method for olive oil extraction. Bioresource Technology, 91(2), 195-200.

Altieri, G. (2010). Comparative trials and an empirical model to assess throughput indices in

olive oil extraction by decanter centrifuge. Journal of Food Engineering, 97, 4656.

Altieri, G., Di Renzo, G.C., & Genovese, F. (2013). Horizontal centrifuge with screw

conveyor (decanter): Optimization of oil/water levels and differential speed during olive

oil extraction. Journal of Food Engineering, 119(3), 561-572.

Amirante, P., Clodoveo, M.L., Leone, A., Tamborrino, A., & Patel, V.B. (2010). Influence of

different centrifugal extraction systems on antioxidant content and stability of virgin

olive oil. In V.R. Preedy, & R.R. Watson (Eds.), Olives and olive oil in health and

disease prevention (pp. 8593). Academic Press, London.

Anlauf, H. (2007). Recent developments in centrifuge technology. Separation and

Purification Technology, 58(2), 242246.


Ayr, U., Tamborrino, A., Catalano, P., Bianchi, B., & Leone, A. (2015). 3D computational

fluid dynamics simulation and experimental validation for prediction of heat transfer in

a new malaxer machine. Journal of Food Engineering, 154, 30-38.

Baccioni, L., & Peri, C. (2014). Centrifugal separation. In C. Peri (Ed), The Extra-Virgin

Olive Oil Handbook (pp. 139 154). John Wiley & Sons, Ltd, UK.

Bell, G.R.A., Symons, D.D., & Pearse, J.R. (2014). Mathematical model for solids transport

power in a decanter centrifuge. Chemical Engineering Science, 107, 114-122.

Berk, Z. (2013). Centrifugation, In Z. Berk (ed.), Food Process Engineering and Technology

(pp. 241-257), Academic Press, San Diego.

Catalano, P., Pipitone, F., Calafatello, A., & Leone, A. (2003). Productive Efficiency of

Decanters with Short and Variable Dynamic Pressure Cones. Biosystems Engineering,

86(4), 459464.

Catania, P., Vallone, M., Pipitone, F, Inglese, P, Aiello, G. & La Scalia, G., (2013). An

oxygen monitoring and control system inside a malaxation machine to improve extra

virgin olive oil quality. Biosystems Engineering, 114, 1-8.

Caponio, F., Summo, C., Pardiso, V. M. & Pasqualone, A. (2014). Influence of decanter

working parameters on the extra virgin olive oil quality. European Journal of Lipid

Science and Technology, 116, 16261633.

Cherubini, C., Migliorini, M., Mugelli, M., Viti, P., Berti, A., Cini, E., & Zanoni, B. (2009).

Towards a technological ripening index for olive oil fruits. Journal Science of Food

Agriculture, 89, 671-682.

Corner-Walker, N., & Records, F.A. (2000). The dry solids decanter centrifuge: conveyor

torque and differential. Filtration & Separation. 37(8),1823.


Dermeche, S., Nadour, M., Larroche, C., Moulti-Mati, F., & Michaud, P. (2013). Olive mill

wastes: Biochemical characterizations and valorization strategies. Process

Biochemistry, 48(10), 1532-1552.

IOOC. (2001). Trade standard applying to olive oil and olive pomace oil. COI/T.15/NC no.

2/Rev. 10.

Leone, A., Tamborrino, A., Romaniello, R., Zagaria, R., & Sabella, E. (2014a). Specification

and implementation of a continuous microwave-assisted system for paste malaxation in

an olive oil extraction plant. Biosystems Engineering, 125, 24-35.

Leone, A., Tamborrino, A., Zagaria, R., Sabella, E., & Romaniello, R. (2014b). Plant

innovation in the olive oil extraction process: A comparison of efficiency and energy

consumption between microwave treatment and traditional malaxation of olive pastes.

Journal of Food Engineering, 146, 44-52.

Leone, A., Romaniello, R., Zagaria, R., Sabella, E., De Bellis, L., Tamborrino, A. 2015.

Machining effects of different mechanical crushers on pit particle size and oil drop

distribution in olive paste. European Journal of Lipid Science Technology.

DOI:10.1002/ejlt.201400485.

Letki, A.G. (2007). Centrifugation | Decanters. In I.D. Wilson (Ed.), Encyclopedia of

Separation Science, (pp. 1-10). Academic Press, Oxford. ISBN 9780122267703.

Letki, A.G., & Corner-Walker, N. (2003). Centrifugal separation. Kirk-Othmer Encyclopedia

of Chemical Technology. Wiley InterScience, New York.

Leung, W.W.F. (2007). Decanter Centrifuge. In W.W.F. Leung (ed.), Centrifugal Separations

in Biotechnology (pp. 95107). Academic Press, Oxford.

Madsen, B. (1989). Flow and sedimentation in decanter centrifuges. Institute of Chemical

Engineers Symposium, No. 113.


Pastore, G., DAloise, A., Lucchetti, S., Maldini, M., Moneta, E., Peparaio, M., Raffo, A., &

Sinesio, F. (2014). Effect of oxygen reduction during malaxation on the quality of extra

virgin olive oil (Cv. Carboncella) extracted through two-phase and three-phase

centrifugal decanters. LWT - Food Science and Technology, 59, 163-172.

Records, A., & Sutherland, K. (2001). Decanter Centrifuge Handbook (1st ed.). Elsevier

Science Ltd., Oxford, UK.

Reif, F., Stahl, W., & Langeloh, T., (1990). Optimising decanter centrifuges. Filtration &

Separation. 27(6), 408410.

Roig, A., Cayuela, M.L., & Sanchez-Monedero, M.A. (2006). An overview on olive mill

wastes and their valorization methods. Waste Management, 26, 960969.

Salvador, M.D., Aranda, F., Gmez-Alonso, S., & Fregapane G. (2003). Influence of

extraction system, production year and area on Cornicabra virgin olive oil: a study of

five crop seasons. Food Chemistry, 80, 359366.

Tamborrino, A., Leone, A., Romaniello, R., Catalano, P., & Bianchi, B. (2015). Comparative

trials to assess the operative performance of an innovative horizontal centrifuge

working in a continuous olive oil plant. Biosystems Engineering, 129, 160-168.

Tamborrino, A., Pati, S., Romaniello, R., Quinto, M., Zagaria, R., & Leone, A. (2014a).

Design and implementation of an automatically controlled malaxer pilot plant equipped

with an in-line oxygen injection system into the olive paste. Journal of Food

Engineering, 141, 1-12.

Tamborrino, A., Romaniello, R., Zagaria, R., & Leone, A. (2014b). Microwave-assisted

treatment for continuous olive paste conditioning: Impact on olive oil quality and yield.

Biosystems Engineering, 127, 92-102.


Fig. 1: Scheme of 3-phase processing decanter (a) and 2-phase processing decanter (b).

Fig. 2: Scheme of decanter model Eureka, Barracane s.r.l.: 3-phase-WA configuration (a) and

3-phase-WW or 2-phase configuration (b).

Fig. 3: Scheme of a decanter centrifuge model Eureka, Barracane s.r.l..

Fig. 4: Plant extraction efficiency versus olive paste flow mass rate, 3-phase-WA

configuration.

Fig. 5: Plant extraction efficiency versus olive paste flow mass rate, 3-phase-WW

configuration.

Fig. 6: Oil content in the husk versus olive paste flow mass rate, 3-phase-WA configuration.

Fig. 7: Oil content in the husk versus olive paste flow mass rate, 3-phase-WW configuration.

Fig. 8: Water content in the wastewater versus olive paste flow mass rate, 3-phase-WA

configuration.

Fig. 9: Water content in the pt versus olive paste flow mass rate, 3-phase-WW

configuration.

Fig. 10: Extraction efficiency observed versus predicted values, 3-phase-WA configuration.

Fig. 11: Extraction efficiency observed versus predicted values, 3-phase-WW configuration.

Fig. 12: Extraction efficiency versus olive paste flow mass rate based on post-hoc test;

Fig. 13: Oil content in the waste water/pt versus olive paste flow mass rate based on

statistical significance test.

Fig. 14: Oil content in the husk versus olive paste flow mass rate based on statistical

significance test.

Fig. 15: Water content in the husk versus olive paste flow mass rate based on statistical

significance test.

Fig. 16: Water content in the waste water/pt versus olive paste flow mass rate based on

statistical significance test.


Tab. 1: Experimental plan

Decanter n Olive paste flow Cavity pump Wastewater/pt


configuration (rpm) mass rate motor frequency discharge radius

(kg h-1) (Hz) (mm)

24.47 3480 21.40 288.5

24.49 4591 28.85 288.5

25.61 5640 36.30 288.5

3-phase-WA* 25.65 6868 43.35 296.1

26.64 8040 50.40 296.1

26.69 8728 55.20 297.9

27.28 9480 60.00 297.9

24.50 3480 21.40 288.5

24.51 4591 28.85 288.5

25.62 5640 36.30 288.5

3-phase-WW 25.66 6868 43.35 296.1

26.61 8040 50.40 296.1

26.68 8728 55.20 297.9

27.27 9480 60.00 297.9

*Process water added was equal to 14.4% w/w of olive oil.


Tab. 2: Dimensions and operating parameters of the decanter model Eureka, Barracane s.r.l.

Parameter Value
Bowl rotation speed ( ) 1900-2380 rpm
Central pinion shaft speed () 600-1350 rpm
Differential speed (n) 9-30 rpm
Length of the bowl cylindrical section (cyl) 1514.0 mm
Length of the bowl conical section (con) 1335.0 mm
Bowl radius (rb) 325.0 mm
Solis discharge radius (rs) 125.5 mm
Oil discharge radius (ro) variable
Wastewater/pt discharge radius (rwp) variable
Cone half-angle () 8.5
Tab. 3. Decanter performance data

Decanter Husk AV pat


Cavity oil mass
configuration Mass water mass mass mass mass 1
pump flow Efficiency Oh Oh Ow/2Op Hh 3
Hw/4Hp
flow rate flow rate flow flow flow
frequency rate
rate rate rate

(kg h-1) (Hz) (kg h-1) (kg h-1) (kg h-1) (kg h-1) (kg h-1) (%) (% w.m.) (% d.m.) (% d.m.) (%) (%)
1 3
3480 21.40 501 630 2089 1262 - 87.3 0.8 ab 4.0 0.1 a 9.2 0.3 a 3.1 0.7 e 56.4 1.4 a 90.2 1.8 a
1 3
4591 28.85 661 840 2573 1839 - 88.8 0.7 a 3.1 0.1 b 7.1 0.2 b 12.0 0.2 d 56.3 0.8 a 87.5 1.7 ab
1 3
5640 36.30 812 1050 2965 2437 - 88.9 0.5 a 2.6 0.1 c 6.0 0.1 c 14.1 0.4 c 56.5 1.2 a 85.1 2.2 ab
3-phase-WA
1 3
6868 43.35 989 1253 3245 3359 - 88.1 1.1 ab 2.4 0.1 d 5.2 0.1 d 16.2 0.4 b 53.4 1.6 a 84.6 1.5 b
1 3
8040 50.40 1158 1470 3682 4046 - 87.8 1.3 ab 2.6 0.1 c 5.5 0.1 cd 18.6 0.4 a 53.0 2.0 a 83.9 1.2 b
1 3
8728 55.20 1257 1546 3973 4466 - 85.6 1.1 bc 2.6 0.1 c 5.5 0.1 cd 19.3 0.6 a 52.8 1.2 a 83.4 2.5 b
1 3
9480 60.00 1365 1685 4409 4751 - 84.5 0.7 c 2.7 0.1 c 5.8 0.1 c 19.8 0.5 a 53.4 1.2 a 83.3 2.5 b
2 4
3480 21.40 - 670 1629 - 1181 91.4 1.1 ab 2.4 0.1 b 5.2 0.3 a 19.0 0.8 c 54.0 1.4 a 78.4 2.0 a
2 4
4591 28.85 - 890 2115 - 1586 92.5 1.0 a 2.1 0.1 c 4.5 0.1 b 18.6 0.8 c 53.7 2.6 a 77.9 0.9 a
2 4
5640 36.30 - 1085 2555 - 2000 92.2 1.4 a 1.8 0.1 d 3.9 0.1 c 19.8 0.9 c 53.4 1.9 a 77.3 2.9 a
2 4
3-phase-WW 6868 43.35 - 1290 3149 - 2429 90.1 1.3 ab 1.6 0.1 e 3.4 0.1 d 23.1 0.9 b 53.2 2.0 a 77.1 1.0 a
2 4
8040 50.40 - 1480 3796 - 2764 88.3 1.2 b 1.9 0.1 d 4.1 0.1 bc 23.8 0.9 b 53.5 0.8 a 76.9 2.0 a
2 4
8728 55.20 - 1550 4091 - 3087 84.9 0.9 c 2.4 0.1 b 5.1 0.2 a 26.2 1.0 a 52.9 1.7 a 76.7 2.5 a
2 4
9480 60.00 - 1640 4542 - 3298 82.3 0.9 c 2.7 0.1 a 5.8 0.2 a 27.0 0.9 a 53.2 2.1 a 76.3 2.7 a
43

Tab. 4 Goodness of fit for the mathematical models developed

Equation coefficients
Parameter Decanter
SSE R2 R2adj RMSE
(y) configuration y p1 (Mp)2 p2 Mp p3

p1 = -3.15E-09

E 3-phase-WA p2 = 3.59E-05 1,38E-03 7,73E-01 7,48E-01 8,75E-03

p3 = 7.87E-01

p1 = -4.82E-09

E 3-phase-WW p2 = 4.17E-05 2.02E-03 9.31E-01 9.23E-01 1.06E-02

p3 = 8.09E-01

p1 = 2.29E-09

Oh 3-phase-WA p2 = -3.49E-05 1.06E-04 9.71E-01 9.67E-01 2.42E-03

p3 = 1.84E-01

p1 = 2.18E-09

Oh 3-phase-WW p2 = -2.76E-05 1.22E-04 9.06E-01 8.96E-01 2.60E-03

p3 = 1.24E-01

p1 = -5.17E-09

Ow 3-phase-WA p2 = 9.18E-05 2.60E-03 9.58E-01 9.54E-01 1.20E-02

p3 = -2.14E-01

p1 = 1.58E-09

Op 3-phase-WW p2 = -5.64E-06 1.88E-03 9.16E-01 9.06E-01 1.02E-02

p3 = 1.85E-01
44

Tab. 5. Statistical parameters of the mathematical model validation

Decanter
Parameter MPE MBE EF RSME 2 cal
configuration

E 3-phase-WA -1.45E-01 1.26E-03 4.50E-03 8.95E-01 5.57E-04

E 3-phase-WW 1.25E-01 -1.23E-03 1.01E-02 9.19E-01 2.75E-03

Oh 3-phase-WA 6.36E-01 -4.06E-04 3.20E-03 9.15E-01 4.15E-03

Oh 3-phase-WW 1.21E+00 -6.99E-04 2.68E-03 8.75E-01 3.89E-03

Ow 3-phase-WA 4.76E-01 -1.36E-03 5.55E-03 9.85E-01 5.61E-03

Op 3-phase-WW 1.60E-01 1.60E-01 1.60E-01 1.60E-01 1.60E-01


45

Tab. 6. Mass balance of decanters input and output materials, calculated considering Mp of 5640 kg

h-1.

Input Output

Olive Process water Oil Husk WW

(kg) (kg) (kg) (kg) (%) (kg) (%)

Oil 20.50 -- 18.38 1.37 2.6 0.91 2.1

Vegetation water 52.30 14.40 0.15 29.70 56.5 36.78 85.1

Soft solids 12.00 -- 0.10 6.29 12.0 5.53 12.8


3-phase-WA*
Rigid solids 15.20 -- -- 15.20 28.9 -- --

Sum 100.00 14.40 18.63 52.56 100.0 43.22 100.0

Sum input/output 114.40 114.40

Oil 20.50 - 18.89 0.82 1.8 1.60 4.5

Vegetation water 52.30 - 0.25 24.19 53.4 27.41 77.3

Soft solids 12.00 - 0.10 5.09 11.2 6.45 18.2


3-phase-WW
Rigid solids 15.20 - - 15.20 33.6 - --

Sum 100.00 - 19.24 45.30 100.0 35.46

Sum input/output 100.00 100.00

* Mw=812 Kg h-1
46

Highlights

The decanters performances were mathematically modelled


The decanters performances were tested by adding/not adding process water
The developed mathematical models shows a good generalization capability
Good correlation between predicted and observed data were obtained
Extraction efficiency resulted high in both decanters configuration tested

Anda mungkin juga menyukai