Anda di halaman 1dari 64

Accepted Manuscript

Title: A Reactive Distillation Process for Co-Hydrotreating of


Non-Edible Vegetable Oils and Petro-Diesel Blends to
Produce Green Diesel Fuel

Authors: Eduardo S. Perez-Cisneros, Mauricio Sales-Cruz,


Ricardo Lobo-Oehmichen, Tomás Viveros-Garcı́a

PII: S0098-1354(17)30018-2
DOI: http://dx.doi.org/doi:10.1016/j.compchemeng.2017.01.018
Reference: CACE 5674

To appear in: Computers and Chemical Engineering

Received date: 3-9-2016


Revised date: 11-1-2017
Accepted date: 12-1-2017

Please cite this article as: Perez-Cisneros, Eduardo S., Sales-Cruz, Mauricio.,
Lobo-Oehmichen, Ricardo., & Viveros-Garcı́a, Tomás., A Reactive Distillation
Process for Co-Hydrotreating of Non-Edible Vegetable Oils and Petro-Diesel
Blends to Produce Green Diesel Fuel.Computers and Chemical Engineering
http://dx.doi.org/10.1016/j.compchemeng.2017.01.018

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.
A Reactive Distillation Process for Co-Hydrotreating of Non-Edible
Vegetable Oils and Petro-Diesel Blends to Produce Green Diesel Fuel

Eduardo S. Perez-Cisnerosa*, Mauricio Sales-Cruzb, Ricardo Lobo-Oehmichena and Tomás


Viveros-Garcíaa

a
Universidad Autónoma Metropolitana - Iztapalapa. Departamento de Ingeniería de Procesos e Hidráulica. San Rafael
Atlixco No. 186, Colonia Vicentina. Delegación Iztapalapa. C.P. 09340, México D.F. México
b
Universidad Autónoma Metropolitana – Cuajimalpa. Departamento de Procesos y Tecnología. Avenida Vasco de
Quiroga No.4871, Colonia Santa Fe Cuajimalpa. Delegación Cuajimalpa de Morelos, C.P. 05300, México, D.F. México

* Corresponding Author: Eduardo S. Perez Cisneros

E-mail: espc@xanum.uam.mx
Highlights

 A co-hydrotreating RD process for production of green diesel is proposed

 Thermodynamic analysis of the reactive mixtures was performed by PC-SAFT EoS

 Effect of the triglycerides and FFA content on the RD performance is investigated

 Simulations reveal that bio-oils with high FFA content are suitable to be hydrotreated.
Abstract

A reactive distillation (RD) process for the hydrotreating (HDT) of vegetable oils and sulphured

petro-diesel to produce green diesel is developed. Process intensification (PI) of a reactive

separation unit that combines both, the hydrodeoxigenation (HDO) of triglycerides and free fatty

acids and the hydrodesulfurization (HDS) of petroleum-diesel reactions, is carried out. PI considers

the thermodynamic analysis of model mixtures of vegetable oils with hexadecane and hydrogen to

determine the appropriate operating conditions (temperature, pressure and composition blends) of

the RD process. Two different hydrotreating RD column configurations are proposed. The

simulation of the hydrotreating RD processes is performed with the Aspen Plus environment using

the PC-SAFT option to modelling the phase equilibria. Simulation results show that the

performance of the hydrotreating RD process is more energy efficient and higher yields are attained

when blends of vegetable oils with high free fatty acids content and petro-diesel are premixed and

further hydrotreated.

Keywords: Process intensification, reactive distillation, vegetable oils, green Diesel, hydrotreating
1. Introduction

After a century where petroleum has dominated as the preferred energy source and as raw material

for the production of organic chemicals, the continuous reduction of petroleum reservoirs, its

increasing extraction cost, the emergence of new economies where demand is increasing

exponentially and natural disasters in producing areas are forcing the petroleum-based fuels

consumer countries to redefine their energy strategies [1]. Such concerns have promoted much

research on reliable alternatives to conventional fuels. Biodiesel, made from vegetable oils or

animal fats, has a renewable nature and a lower contribution to global warming due to its almost

closed carbon cycle. This biofuel is conventionally produced through batch or continuous

transesterification of highly refined vegetable oils with methanol by using homogeneous or

heterogeneous catalysts [2]. An excellent review of the different integrated reactive separation

processes for producing biodiesel has been presented by Kiss and Sorin Bildea [3]. This biofuel has

many benefits, however, new biodiesel plants must be built requiring a large capital investment [4,

5]. The economics of biodiesel production depend on selling by-product glycerol, and increased

biodiesel production would cause the price for glycerol to decrease [6]. Also, despite fatty-acid

methyl esters (FAME) represent the first generation of biodiesel, unfortunately, the use of FAME

can cause problems in vehicles due to its high oxygen content, rendering it a less than ideal fuel for

conventional engines [7].

In this way, a renewable diesel or green diesel has emerged as a substitute for mineral diesel fuels.

Its production is not limited to seed oils, but it is flexible in that it can use a number of different

feedstocks. Renewable or green diesel is produced by hydrotreating (HDT) of vegetable oils,

whereby the oxygenates in bio-feedstocks are hydrogenated into hydrocarbons. HDT may take

place in existing diesel hydrotreaters reactors in a co-processing scheme, where the biofeed

component is deoxygenated and the fossil diesel component is simultaneously desulphurised, or it

can take place in a standalone unit processing 100% biofeed. In both cases, conversion occurs over
HDT catalysts and in conditions similar to those used in hydrotreaters. Thus, among the cheapest

technological options for green diesel production is to use the existing hydrodesulphurization

(HDS) processes comprised in the petroleum refineries [8, 9]. Petroleum refineries are already built

and, using existing infrastructure for biofuels production requires low capital cost investment. The

HDT of vegetable oils as well as suitable waste and residue fat fractions to produce green diesel is a

quite new but already commercial scale manufacturing process. Companies, including Neste Oil

[10–12], UOP/Eni [13–15], Syntroleum [16, 17], ConocoPhillips [18, 19], PETROBRAS [20, 21],

and Haldor Topsøe [22] produce biofuels through the processing of pure (or used) vegetable oils or

mixtures of vegetable oils and petro-diesel using existing HDT technology. For example, Haldor

Topsøe Hidroflex coprocessing technology has been successfully applied in the Preem Refinery in

Göteborg, Sweden, where 30% of raw tall oil is co-processed with a gasoil stream [22].

Despite the booming industrial HDT processes to produce green diesel fuels from vegetable oils,

there are technical challenges for the co-HDT of vegetable oils which could be overcome by

process intensification (PI): 1) Reactivity of the oxygen-containing species. Vegetable oils are much

more reactive than refractory sulphur compounds (i.e., dibenzothiophene). Thus, PI is required to

guarantee deep hydrodesulfurization of petro-diesel; 2) Hydrogen consumption. PI is needed to

reduce hydrogen consumption with maximum conversion of sulphured compounds and vegetable

oils; 3) Production of undesirable gases. PI must be considered to release the gases produced by the

HDO reactions in an optimal reaction-separation configuration; 4) Catalyst deactivation. High

temperatures may lead to accelerated catalyst deactivation, thus PI is required for the appropriate

catalyst selection and distribution along the hydrotreating unit under RD operating conditions; 5)

Green diesel properties. PI should be carried out to produce a green diesel with the diesel properties

established by the international standards.


Currently, the green diesel is produced based on the fixed packed bed co-HDT reactors installed in

the refineries modifying its original designs, depending on the raw material profile. However, most

of the technological developments from the industry to overcome the challenges for co-HDT of

vegetable oils remain undisclosed and the know-how to produce green diesel, this is, the operating

conditions and design details of each specific industrial reactor are not open to public.

Therefore, in this work, a new alternative technological approach for producing green diesel,

dealing with the challenges mentioned above and considering the diversity of vegetable oils-petro-

diesel blends, is proposed. The new alternative process is based on the use of a reactive distillation

(RD) process already present in most of the refinery systems. Actually, the use of a reactive

distillation (RD) process for HDS of napthas has been successfully applied by CDTECH Company

in Mexico [24] and this Company has also patented a reactive distillation process for deep

hydrodesulphurization (HDS) of diesel [25]. From the computational PI research, Perez-Cisneros

and co-workers, in a series of papers [26, 27, 28, 29] have performed the conceptual design of a

reactive distillation process for deep HDS of diesel and studied the multiplicity and operational

aspects of the RD process. From these works, it was concluded that deep HDS of diesel can be

achieved even at a reduced total pressure (~ 30 atm) using two fixed catalytic beds into the RDC

and an hydrogen excess feed of 3:1. On the other side, RD process has been thoroughly investigated

for biodesel production from different vegetable oils [30-32]. Kiss and Sorin Bildea [3]

demonstrated that, for biodiesel production, the preferred choice has been reactive distillation and it

has been successfully applied for both, esterification and trans-esterification reactions catalyzed by

heterogeneous (e.g. ion exchange resins, sodium ethoxide, tungstated or sulfated zirconia, niobium

oxide, mixed metal oxides) or homogeneous catalysts (e.g. NaOH, KOH). Recently, Glisic et al.

[33] performed a process and techno-economic analysis of green diesel production from waste

vegetable oil and they made the comparison with ester type biodiesel production. The authors

conclude that waste vegetable oils and/or lower quality triglyceride type oils could provide viable
renewable source for cost efficient diesel fuel substitute. The sensitivity analysis performed

indicates that a large capacity hydroprocessing unit located in a refinery, where most of the

supporting process infrastructure is already available, is the most cost effective way to process

waste vegetable oils and similar low cost bio-oils. However, it should be pointed out that the study

performed by Glisic et al. [33] considers triolein as the key component of the vegetable oil and the

hydrotreating process only accounts for the hydrodeoxigenation(HDO) of wasted vegetable oil

(triolein). Consequently, the hydrotreating process proposed does not deal with the simultaneous co-

hydrotreating (HDO-HDS) of vegetable oils and sulphured diesel.

Therefore, based on the RD column design developed by Perez-Cisneros and co-workers [29] for

deep HDS of diesel and the well established RD technology for biodiesel production [3], a new RD

technological alternative for producing green diesel is proposed. Thus, the objective of the present

work is to develop a conceptual design a reactive distillation process for the HDT of vegetable oils

and sulphured petro-diesel blends to obtain an improved green diesel fuel. In order to reach such

objective a systematic approach for the simulation, design and analysis of the process is proposed.

The systematic approach considers: 1) the thermodynamic analysis of the reactive system using

simplified model reacting and product mixtures, 2) analysis of the reaction rates and kinetic

expressions for the HDO-HDS reactive system and, 3) the simulation of the reactive-separation

process to determine the optimal configuration design and operating conditions for the

hydrodesulfurization (HDS) of sulphured diesel occurring simultaneously with the

hydrodeoxigenation (HDO) of vegetable oils.

2. The HDO-HDS Reactive System

Hydrotreating (HDT) of vegetable oils has previously been used to produce straight chain alkanes

ranging from n-C15–n-C18, from a fatty acid fraction of tall oil (produced during kraft refining), and

other vegetable oils [34-36]. The aim of the hydrodeoxigenation (HDO) process is to upgrade
vegetable oils by removing the oxygen content as water. The process includes treatment of

vegetable oils at high pressures and moderate temperatures over heterogeneous catalysts. The use of

vegetable oils, mainly non-edible vegetable oils, as feedstock for the HDO process is highly

favorable because their hydrocarbon content is in the same range as that of fossil fuels, such as

kerosene and diesel. The fatty acids in vegetable oils are composed of straight chains of carbon

atoms, and HDO produces n-alkanes as the primary products. Thus, triglycerides in the vegetable

oils are transformed into alkanes with water and propane as by-products. Carbon dioxide or carbon

monoxide is also obtained as a by-product when the reaction follows decarboxylation under HDO

conditions. The rate of the HDO reaction and its mechanism are influenced by the nature of the

feedstock [34]. Studies by Prasad and Bakshi [35] explained the mechanism of the HDO reactions

and the formation of by-products. The route of the reaction and products formed depend on the type

of catalyst used [36]. The mechanism involves simple HDO via an adsorbed enol intermediate, and

the product is a straight-chain hydrocarbon with water and propane as the by-products. The process

of deoxygenation of triglycerides in a simplified form is shown in Figure 1.

Figure 1. Possible deoxygenation pathways of triglycerides

Figure 2. Reaction pathways for dibenzothiophene hydrodesulfurization

On the other side, it is well known that there are two possible reaction pathways for sulfur removal

from the organo-sulphur compound dibenzothiophene (DBT) [28] (see Figure 2). The first pathway

is the sulphur atom direct extraction (hydrogenolysis) from the sulphured molecule. The second

pathway is the hydrogenation of one aromatic ring followed by the sulphur atom extraction. Van

Parijs and Froment [37] suggested a reaction network in which thiophene (Th) is

hydrodesulphurized to give hydrogen sulphide (H2S), 1-butene, cis and trans-2-butene, followed by

the secondary hydrogenation of the butenes into butane. For benzothiophene (BT), Van Parijs et al

[38] proposed a parallel pathway similar to Th, indicating also that such reactions occur at different
sites. In fact, to obtain an ULSD, most recalcitrant sulphur compounds must be eliminated from the

hydrocarbon feed, this is, DBT and 4,6-DMDBT. A detailed explanation of the reaction mechanism

for HDS of diesel is given by Withehurst et al. [39].

3. The Systematic Approach

1) Thermodynamic Analysis

Hydroconversion of vegetable oils involves highly asymmetric mixtures of hydrocarbons and

triglycerides which exhibit complex phase behavior. The mixture from the hydroconversion

reaction step contains unreacted triglycerides, long and short chain alkanes, water and gaseous

compounds such as hydrogen, propane, carbon monoxide and carbon dioxide. Obviously separation

steps must be carried out to separate long chain alkanes from triglycerides and gaseous components.

The process development needs a model to describe phase behavior of such a complex mixture and,

for this aim, a φ - φ approach based on equations of state seems the most suitable. Such an approach

requires preliminary study of binary or ternary sub-systems which characterize mixtures involved in

the hydroconversion process. In particular, in these mixtures it is possible to recognize binary sub-

systems which exhibit complex phase behaviors and give rise to thermodynamic modeling

difficulties. For example, highly asymmetric mixtures of hydrogen and long chain hydrocarbons

(C15-C18) and hydrogen-triglycerides. The first ones exhibit solvent-solute size asymmetry and

require models able to relate molecular features to macroscopic thermodynamic behavior [40]. On

the other hand, hydrogen or propane-triglyceride systems are characterized by a complex vapor-

liquid-liquid phase equilibrium behavior [41]. In this work, the implementation of the Statistical

Associating Fluid Theory (SAFT) was employed to describe such complex phase behavior.

a) Pure component properties of the reactive system

Due to the great variety of esters species and, in particular, triglycerides and fatty acids, group

contribution and corresponding state approaches are the usual methods for the estimation of pure
component properties. Sales-Cruz et al. [42] predicted critical properties of fatty acids, fatty acid

methyl esters, and triglycerides using group contribution approaches and applied them to density

and v s os t pr t ons r -Mikkelsen and Stenby [43] developed a group-contribution model

for the prediction of melting points and enthalpies of fusion for saturated triglycerides. Ceriani et al.

[44] presented a group-contribution method for the estimation of the vapor pressures and heats of

vaporization of fatty compounds. Díaz-Tovar et al. [45] estimated a variety of pure-component

properties for edible oils and biodiesel-related compounds, based on the approaches of Ceriani et al.

[44] and Marrero and Gani [46]. In the present work, the pure component properties were

determined using the Marrero and Gani group contribution method incorporated into the Integrated

Computer Aided System (ICAS) Platform (CAPEC, 2015). Table 1 lists the values of the pure

component properties and the PC-SAFT EoS parameters used for the phase equilibrium

calculations.

Table 1. Critical properties of the pure components and PC-SAFT equation of state
parameters. Properties determined using the Marrero and Gani [46] method included in
ICAS Platform (CAPEC, 2015)

b) Thermodynamic Modelling

The thermodynamic modelling of vegetable oils, sulphured diesel and hydrogen reacting mixtures

was performed by means of PC-SAFT equation of state. The PC-SAFT EoS model has showed its

capacity to depict chain molecule phase behaviour providing better predictions for all the system

investigated and it is recommended as reference EoS to calculate phase equilibria of asymmetrical

mixtures of vegetable oils and short-chain species [41]. SAFT is one of the most important

asso at on th or wh h lon s to th n ral p rtur at on at or It s as on th W rth m’s

[47] first order perturbation theory and it has been implemented into useful expressions in several

studies. SAFT EoS considers molecules as chains composed of freely joined spherical segments and

the basic idea underlying the model is the definition of an interaction potential to represent

attraction and repulsion forces between segments. The molecular model can be refined by assuming
the segments to have association sites and partial charges enabling the chains to associate or repulse

by hydrogen bonds and dipole interactions. Consequently, the residual Helmholtz free energy, is

made up of terms stemming from each of these contributions. The PC-SAFT EoS main innovation

relies on considering perturbation respect to spherical segments which are connected into rigid

chains (hard-chain term) rather than disconnected segments. Obviously, this means considering

interactions between chains instead of single segments and this would allow to capture the

behaviour of large chain molecules, such as hydrocarbons, polymers and vegetable oils, more

realistically. In this way, the residual Helmholtz free energy can be written as:

Ares Ahc Adisp Aassoc


= + +
NkT NkT NkT NkT (1)

In Equation (1) the hard-chain (hc) reference term accounts for hard-sphere (hs) and chain

formation contributions

Ahc =Ahs  xi mi + xi (1-mi )ln giihs (σii )


i i (2)

Where mi and σi correspond to the number of segments which constitute the chain of the i-th

component and to its diameter, respectively, while xi is the component mole fraction. The dispersion

term (disp) in Equation (1), according to Barker and Henderson [48], consists of first and second

order contributions

Adisp A1 A2 ε 
= + =-2πρI1 (η,m) xi x jmi m j  ij  σij3 -
NkT NkT NkT i j  kT 
-1 2
(3)
 Zhc   εij  3
-πρm 1+Zhc +ρ  I2 (η,m) xi x jmi m j   σij
 ρ  i j  kT 

where Zhc is the hard-chain contribution expressed in terms of compressibility factor, η is a packing

factor and I1 and I2 are integral terms. Expressions of these terms depend on the choice of the

interaction potential, which in the PC-SAFT EoS model is assumed to be a modified version of the

classical square-well. The interaction potential is then characterized by the three parameters, mi, σi
and εi, which is the well potential depth. For a detailed derivation of all the terms appearing in

Equations (1-3) the reader is referred to the paper of Gross and Sadowsky [49]. Extension to

provided in Eq (2) requires definition of mixing rules for a pair of unlike segments, which

according to Berthelot- Lorentz, results may be given as:

1
σij = (σi +σ j ), εij = εiε j (1-kij ) (4)
2

In this work, binary interaction parameters kij were not included.

c) Phase behaviour of the reacting model mixtures

In order to design a feasible reactive separation process to perform the HDO of triglycerides and

fatty acids and the HDS of a sulphured diesel, a thermodynamic analysis of the phase behavior of

the reacting mixtures is highly relevant. However, before to proceed with the thermodynamic

analysis, the predicted pure component properties and PC-SAFT EoS parameters estimated in this

work must be validated. Reliability of PC-SAFT EoS parameters was evaluated through comparison

of model results, relative to equilibrium results reported by Annesini et al. [41] where experimental

data available in the literature was used to obtain their regressed PC-SAFT EoS parameters.

Furthermore, PC-SAFT EoS parameters predicted using ICAS platform were also validated with the

LI-LII-V flash calculation example proposed by Annesini et al. [41]. Table 2 gives the results

reported and those calculated in the present work. It can be noted in Table 2 that the phase

distribution, this is, the flow rate values of the exit streams and their corresponding component mass

fraction are quite similar. Also, the binary phase diagrams for propane-C16 and propane-Triolein

mixtures were computed without significant differences.

Table 2. Validation of the pure component properties and PC-SAFT parameters with results

reported by Annesini et al. [41]. Mass fractions of the outlet streams for a three-phase flash

separator
Once the pure component properties and PC-SAFT EoS parameters were validated, four simplified

asymmetrical model mixtures were selected for representing the HDS-HDO reactive system. That

is, a triglyceride (triolein) representing a vegetable oil (i.e., virgin vegetable oil) and a fatty acid

(oleic acid) representing a used vegetable oil and n-hexadecane (n-C16) representing a petro-diesel

cut. Thus, the reacting vegetable oils-petrodiesel-hydrogen model mixtures are: a) triolein (TO) - n-

hexadecane - H2 and b) oleic acid (OA) – n-hexadecane – H2. Also, for the reaction products, we

define the mixtures: a) triolein (TO)- n-hexadecane - H2O and b) oleic acid (OA)- n-hexadecane -

H2O. It should be clear that the presence of water in the liquid phase together with hydrocarbons

and non converted vegetable oil, could lead to the existence of two different liquid phases.

Figure 3. Hydrogen Solubility at 30, 50, and 80 atm. a) triolein (TO) - n-hexadecane - H2 and

b) oleic acid (OA) – n-hexadecane – H2

Figure 3 shows the triangular liquid phase equilibrium compositions diagrams for the model

reacting mixtures at three different total pressures (30, 50 and 80 atm). It can be observed in Fig. 3a

(triolein as vegetable oil) that, as the pressure is increased the hydrogen solubility increases as well.

Also, it can be noted that the liquid composition of TO reaches a maximum value at the three

different pressures. In Figure 3b it can be observed that the hydrogen solubility is diminished when

OA is present. However, this mixture does not show the maximum values of TO composition at 30

atm of total pressure. This different behavior between TO and OA is very important in order to

design a reactive zone where the concentrations of triglycerides or fatty acids and hydrogen are

suitably high to have reasonable reaction yields. Figure 4 shows the liquid compositions of the

reacting model mixtures as a function of temperature at three different pressures. It can be seen in

Fig. 4a (triolein (TO) - n-hexadecane - H2) that the maximum triolein liquid composition appears at

30 atm and 800 K with a hydrogen liquid mole fraction of 0.18. Also, it should be observed that
above 1000 K only vapor phase exists. Figure 4b shows the liquid compositions of the reacting

mixture oleic acid (OA) – n-hexadecane – H2 as a function of temperature at the three different

pressures. It can be noted in Fig. 4b that at P=30 atm the liquid composition of OA increases

continuously while for 50 and 80 atm it decreases and above 780 K only vapor phase exists. The

above thermodynamic results lead to conclude that, in order to have a feasible reactive separation

process to carried out the HDO reactions, the optimal operating conditions are: i) the temperature

should not be greater than 760 K and 2) total pressure of P=30 atm is advised.

Figure 4. Equilibrium liquid compositions at 30, 50 and 80 atm as function of temperature. a)

triolein (TO) - n-hexadecane - H2 and b) oleic acid (OA) – n-hexadecane – H2

Figure 5 shows the ternary L-L-V equilibrium maps for the two product model mixtures at 30 and

50 atm, respectively. It can be noted in figure 5a that two liquid phases are present at 30 atm even

with small amounts of triolein. Certainly, an organic aqueous phase may be generated with the

unconverted triolein, water and n-hexadecane. Therefore, for any hydrotreating reactive separation

process, it should be mandatory the full releasing of water by vaporization to avoid the two liquid

phase phenomena affecting the solid catalyst activity. Table 3a shows the tie line compositions for

the L-L-V phase behaviour of the model mixture triolein (TO)- n-hexadecane - H2O. Also, it can be

observed in Figure 5a the existence of a binary azeotrope between triolein and water. The azeotrope

compositions and the temperature are listed in Table 3b. It should be pointed out that this azeotropic

condition would severely affect the performance of any hydrotreating reactor.

Figure 5. Ternary L-L-V Equilibrium compositions at 30 atm for the product mixtures: a)

triolein (TO)- n-hexadecane - H2O at 50 atm and b) oleic acid (OA)-n-hexadecane-H2O at 30

atm
Table 3. LI-LII-V equilibrium compositions for the product mixture triolein (TO)- n-

hexadecane - H2O at 30 atm. (a) Tie line compositions, (b) Azeotrope compositions and

temperature

Similarly, considering that oleic acid is not completely converted in the hydrotreating reactive

separation process, Figure 5b shows that two liquid phases are formed at 50 atm. Thus, in order to

have an efficient operation of the hydrotreating device, the full release of water together with the

complete conversion of the vegetable oil should be achieved in the hydrotreating process. Table 4a

gives the tie line compositions for the phase behaviour of the model mixture oleic acid (OA)- n-

hexadecane - H2O. Also, it can be noted in Figure 5b the existence of one binary azeotrope between

n-hexadecane and water and a ternary azeotrope of the model mixture. The azeotrope compositions

and the respective temperatures are given in Table 4b.

Table 4. LI-LII-V equilibrium compositions for the product mixture oleic acid (OA)- n-

hexadecane - H2O at 50 atm. (a) Tie line compositions, (b) Azeotrope compositions and

temperature

Finally, as a validation of the above simulation results, the existence of two liquid phases has been

reported by Toth et al [50]. The authors reported that, in some experiments, the residual fraction

contained the unconverted triglycerides, the formed or rather not converted di- and mono-

glycerides, hydrocarbons with higher carbon numbers, carboxylic acids and esters.

2) HDS-HDO catalyst, reaction rates and kinetics expressions

a) Hydrodesulfurization catalyst and kinetic expressions

The liquid-phase HDS reactions for DBT and 4,6-DMDBT have been thoroughly studied and

revisited by several authors [51-55]. In conventional HDS process several types of commercial

catalytic reactors (fixed-bed, moving-bed, LC-fining, etc.) are used and they operate under similar
principles. However, the fixed-bed reactor is preferred to process light feeds, while the moving-bed

reactor is selected for hydroprocessing of heavy feeds. The catalysts are chosen depending on the

feed properties and usually include supported molybdate and tungstate catalysts promoted by either

Ni or Co. γ-Alumina, zeolites, silica and silica aluminates are the typical supports. The reaction

network for HDS of DBT is shown in Figure 2. Experiments by Vanrysselberghe et al. [55] showed

that the most recalcitrant sulfur compound 4,6-DMDBT is mainly converted through the

hydrogenation reaction pathway, while DBT is mostly converted by the hydrogenolysis reaction

pathway. In general, when alkyl substituents are attached to the carbon atoms adjacent to the sulfur

atom, the rate for direct sulfur extraction is diminished whereas the sulfur removal rate via the

hydrogenation pathway is relatively unaffected. Tables 5 and 6 list the kinetic expressions and

kinetic parameters used for the thipophene (Th), benzothiophene (BT) and di-benzothiophene

(DBT) hidrodesulfurization reactions.

Table 5. Reaction rate expressions, reaction rate coefficients, and adsorption equilibrium

constants for the thiophene and benzothiophene HDS.

Table 6. Reaction rate expressions, reaction rate coefficients, and adsorption equilibrium

constants for the dibenzothiophene HDS.

b) Hydrodeoxygenation catalyst and kinetic expressions

The commonly used hydrodeoxygenation catalysts are supported noble and sulfided or reduced

metal catalysts. Sulfided CoMo/Al2O3 and NiMo/Al2O3 are widely used for hydrodeoxygenation,

and it is reported that sulfidation generates active sites on the catalyst [56-58]. However,

hydrodeoxygenation using sulfided catalysts is less favorable due to the formation of sulphur-

contaminated products and the reduction of the catalysts in the low-sulfur reactant. In addition,
catalytic activity depends on the type of feedstock used, reaction conditions and nature of the

catalysts. The acidic or basic sites on the catalyst support play the main role in activating the

reactants and determining product selectivity [59]. The product selectivity and extent of

hydrodeoxygenation also depend on the reaction temperature, pressure, H2 to oil volume ratio and

liquid hourly space velocity (LHSV). It is observed that the presence of water decreases the

stability, activity and selectivity of the catalyst, which, in turn, decreases the degree of

deoxygenation [60]. Sunflower oil has been used as an oil feedstock for hydrodeoxygenation over

various catalysts. Krár et al. [61] examined the catalytic hydrodeoxygenation of light gas oil

containing 10% sunflower oil over an alumina-supported transition metal catalyst. The products

formed were characterized by low sulfur and aromatic content. The paraffinic products showed

good cold flow properties with a high cetane number. They showed that sunflower oil could be

effectively hydrotreated over reduced NiMo/γAl2O3 catalysts to obtain high paraffinic products at a

high temperature range (360–380 °C) and pressure (6 MPa) and low space velocity of the oil

feedstock. The resulting products consisted of n-C15–n-C18, which showed that the deoxygenation

proceeded via both hydrodeoxygenation and decarboxylation, and also exhibited a high cetane

number and poor low-temperature properties. Also, Tóth et al [50] investigated the production of

green diesel with hydrocarbons in the diesel range by deoxygenation of a gas oil–sunflower oil

mixture. It was established that products with less sulphur and less aromatic content were formed

under the reaction conditions of 360–380 °C, 8 MPa, LHSV=0.75 h-1 and hydrogen/feedstock

volume ratio=600 N m3/m3 with varying vegetable oil content (0–15%). The process was carried

out over sulfided NiMo/Al2O3 catalyst at a temperature range of 300–380 °C and H2 pressure 6–8

MPa. When the amount of vegetable oil was >15%, HDS of gas oil decreased.

Despite the above experimental results explaining the reaction routes on a specific catalyst, sparse

information on the kinetics of HDO of vegetable oils is available in the open literature. Sheu et al.

[62] investigated the kinetics of HDO of pine bio-oil between 300–400 °C over Pt/Al2O3/SiO2, Co–
MoS2/Al2O3, and Ni–MoS2/Al2O3 catalysts in a packed bed reactor and their results were evaluated

on the basis of a kinetic expression as follows:

dwoxy
  kwoxy
m
Pn (5)
dZ

Here, woxy is the mass of oxygen in the product relative to the oxygen in the raw pyrolysis oil, Z is

the axial position in the reactor, k is the rate constant given by an Arrhenius expression, P is the

total pressure (mainly H2), m is the reaction order for the oxygen, and n is the reaction order for the

total pressure. Their experimental results are interesting, however, the rate term of Eq. (5) has a

non-fundamental form as the use of mass related concentrations and especially using the axial

position in the reactor as time dependency makes the term very specific for the system used. Thus,

correlating the results to other systems could be difficult. Furthermore, the assumption of a general

first order dependency for woxy is a very rough assumption when developing a kinetic model.

Recently, Zhang et al. [63] proposed the kinetics of HDO of waste cooking oil (WCO) with

unsupported CoMoS catalysts. The authors claim that decomposition of glycerides to fatty acids is a

very fast reaction and glycerol, a by-product of the decomposition, can be quickly converted to

propane and propene over the catalyst. Thus, decomposition of glycerides to fatty acids is assumed

to be an irreversible reaction. Fatty acids are then further transferred to hydrocarbons through

different routes since CO, CO2 and H2O are detected in gas and liquid products. Given sufficient

hydrogen gas, the reaction where fatty acids are hydrodeoxygenated to C18 is assumed to be pseudo-

first-order-kinetics. The HDC reaction is considered as an irreversible reaction.

Deactivation of catalyst is taken into account for both HDO and HDC reactions. A commonly used

second-order decay law is applied on both routes. Nevertheless, the kinetic expressions proposed by

Zhang et al. [62] for the simplified HDO and HDC reaction scheme is valid only for the specific

experimental conditions and the specific bio-oil feedstock. In general, it can be concluded that
kinetics of HDO is rather complex due to the nature of a real bio-oil feed. Thus, it seems that a

better understanding of the complex chemistry of the HDO reaction pathways is needed. Therefore,

in the present work, due to the lack of reliable HDO kinetic expressions it is assumed that the HDO

reaction achieve a target conversion of 99.99 % at the two reactive zones of the hydrotreating RD

column.

3) Simulation of the HDO-HDS Reactive Distillation Process

a) Simplified reactive system

The reactions considered for the HDS of the sulphured diesel can be written in a simplified form as:

Thiophene + 2 H2  Butadiene + H2S (7)

Benzothiophene + 3 H2  Ethylbenzene + H2S (8)

DBT + 2 H2  Biphenyl + H2S (9)

It should be noted that Eq. 9 only considers the hydrogenolysis reaction pathway. For the

hydrodeoxigenation of triglycerides (triolein in this case) and fatty acids (oleic acid in this work)

the simplified reactions are:

Triolein + 15 H2  3 C18H38 + C3H8 + 6 H2O (10)

Oleic Acid + 4 H2  C18H38 + 2 H2O (11)

b) Reference deep HDS reactive distillation processes for HDO and HDS

Viveros-García et al. [27] developed a conceptual design of a reactive distillation column (RDC) for

deep HDS of diesel through a thermodynamic analysis considering the following aspects: (i) the

volatility of the organo-sulfur compounds; (ii) the different reactivities of the organo-sulfur

compounds; and (iii) the computation of non-reactive and reactive residue curve maps for DBT

elimination. In this work the nonidealities of the vapor and liquid phases are calculated through the

PC-SAFT equation of state. The complete set of kinetic expressions and physical properties for

vapor liquid equilibrium calculations and the column configuration details used for the simulations
are given in Tables 1, 5, 6, 7 and 8. Thus, the reference RDC configuration used here consisted of

14 stages with two reactive zones (stages 5-7 and stages 9-11). A target conversion of 99% for the

DBT and complete elimination of Th and BT were assumed as part of the design specifications.

Also, the operating pressure of 30 atm and a H2 to hydrocarbon (HC) feed ratio of 3 were the

optimal values to reach deep HDS of petro-diesel in the RDC [27]. Table 7 gives the feed

compositions of the sulphured petro-diesel considering the hydrogenolysis reaction pathway [25].

Table 7. Design specifications of the hydrotreating reactive distillation column.

Table 8. Sulphured petro-diesel feed composition

4. Results and Discussion

Based on the reference RDC for deep HDS of diesel, two modified reactive separation processes to

perform the HDS and HDO reactions are proposed. Figure 6 shows a simple modification of the

basic HDS-RDC, where a mixer is added before feeding the RDC with the sulphured diesel and the

vegetable oil (TO or OA). Figure 7 shows a process that does not consider a mixer before feeding

the RDC, rather, the vegetable oil is separately fed below the catalytic bed of reactive zone II.

Figure 6. Hydrotreating RDC design diagram 1. Vegetable oil pre-mixed with sulphured
Petro-diesel fed at stage 9

Figure 7. Hydrotreating RDC design diagram 2. Triolein fed at stage 12 and sulphured Petro-
diesel fed at stage 9

In order to determine the best co-hydrotreating RDC configuration to overcome the technical

challenges for hydrotreating of vegetable oils described above, the effect of different design and

operating variables on the performance of the RDC were investigated and analyzed considering two

study cases: i) pure triolein representing the vegetable oil and, ii) pure oleic acid, a fatty acid,
meaning the bio-oil. The simulations of the co-hydrotreating RDC were carried out in Aspen-Plus

environment considering equilibrium stages. Thus, no heat and mass transfer phenomena and

mixing issues were taken into account.

4.1 Triolein study case

a) Effect of the triolein feed location

Figure 8 shows the reactive distillation column profiles considering that triolein is pre-mixed with

the sulphured diesel and fed at stage 9, assuming that a pre-mixer is added to the original HDS-RD

process proposed by Viveros et al. [27]. The co-hydrotreating process simulation begin with a small

amount of triolein (1 kmol/hr) in the feed mixture and it was continuously increased until 5.4

(kmol/hr). Figure 8a shows the variation of temperature along the reactive distillation column

considering the reference temperature profile without triolein in the feed. It can be seen that at stage

9, where the triolein-petro-diesel blend is fed, a temperature drop occurs and, oppositely, at stage 7

the temperature increases. This is, a hotspot at the end of the reactive zone I (stage 7) is generated

while a cold point is observed at the reactive zone II (stage 9). This behaviour of the RDC can be

explained as follows: as the temperature of the feed is lower (240 °C) than the temperature at stage

10 (~375 °C), part of the triolein fed is vaporized and converted at stage 7, generating the high

boiling point compound (n-C18) and promoting a temperature rise at stage 7. It should be clear that

n-C18 species is produced in the condensed liquid phase at stage 7. This explanation can be

corroborated by seeing Figure 8b, where the n-C18 generation profile is shown. It can be noted in

Figure 8b that, as the amount of triolein in the feed is increased, the production of n-C18 sharply

increases at stage 7, as well as the temperature does at the same stage. Figure 8c shows the effect of

increasing the triolein feed flow on the conversion of DBT. It can be noted in Figure 8c that for a

triolein feed flow greater than 3 (kmol/hr) deep HDS of diesel is not achieved. This reduced

conversion is very important, since deep HDS of diesel is a mandatory issue for any hydrotreating
process. The effect of high vegetable oil concentrations on the conversions of DBT was

experimentally reported by Tóth et al. [55].

Figure 8. RDC profiles at P=30 atm and triolein-petro-diesel blend fed at stage 9. (a)

temperature profile; (b) HDO product n-C18, liquid composition profile; (c) HDS of DBT,

liquid composition profile

After the above unconvincing simulation results, the feed stage location of triolein was modified

considering that triolein and the HDO product (n-C18) are the highest boiling point species and that

the n-C18 produced should leave at the bottom exit stream. Therefore, the triolein feed stage was

moved to stage 12, keeping the petro-diesel feed at stage 9. It should be clear that the petro-diesel

feed location at the middle of the RDC is because at the reactive zone I, deep HDS of thiophene

(Th) and benzo-thiophene (BT), the lightest sulphur compounds is carried out and, at the reactive

zone II, deep HDS of DBT is attained [27]. Thus, in order to do not affect the deep HDS of diesel at

the reactive zones, triolein is fed close to the bottom of the RDC, this is, at stage 12.

Figure 9 shows the reactive distillation column profiles considering that triolein is fed at stage 12

and the sulphured diesel fed at stage 9. Figure 9a shows a typical RDC temperature profiles where

neither, hotspots and cold points are present, and the trend of the temperature profiles are very

similar to those for deep HDS of diesel without triolein. It should be noted in Figure 9a that triolein

feed flow was varied from 1 to 15 (kmol/hr) and the temperature decreases as the triolein feed is

augmented. Also, it can be observed that the temperature at the bottom of the RDC (~415 °C) is

lower than that with triolein fed at stage 9 (~460 °C). Figure 9b shows the produced n-C18 liquid

profile along the RDC. It can be noted that most of the n-C18 is produced at the reactive zone II and

it is accumulated at the bottom of the column. Also, the concentration of n-C18 is decreased to zero

at the reactive zone I and this allows the hydrogenation of the organic-sulphur compounds (Th, BT,
DBT) for deep HDS. Figure 9c shows that, even with a higher triolein feed flow (15 kmol/hr), deep

HDS of DBT is achieved. Obviously, a lager hydrogen make up is required. Results shown in

Figure 9 reveal that, by a simple relocation of the vegetable oil feed (stage 12), an efficient and

flexible hydrotreating RD process is intensified.

Figure 9. RDC profiles at P=30 atm and triolein-petro-diesel blend fed at stage 12. (a)

temperature profile; (b) HDO product n-C18,liquid composition profile; (c) HDS of DBT,

liquid composition profile

b) Hydrogen consumption

Another important challenge to overcome is the appropriate management of the hydrogen required

for the simultaneous HDS and HDO reactions occurring in the co-hydrotreating reaction-separation

process. Figure 10 shows the flow rate of hydrogen required as a function of the triolein feed flow

at stages 9 and 12, respectively. It can be observed in Figure 10a that for a feed flow of triolein

greater than 3 (kmol/hr) a sharp increase in the flow rate of hydrogen occurs, passing from 300 to

310 (kmol/hr). This change corresponds to the unanticipated temperature profile showed in Figure

8a. It must be observed that, from the controllability point of view, a control link between the

vegetable oil feed flow and the hydrogen supplied could be unstable and a fine tune control

structure should be implemented. On the other side, when triolein is fed at stage 12, a quasi-lineal

function for the hydrogen supplied is observed (see Figure 10b). It should pointed out that, despite

the hydrogen consumption is higher whit triolein fed at stage 12, this RDC configuration allows to

manage the hydrogen make-up in a simple manner because the linear relationship of the hydrogen

supplied and the triolein feed flow.

Figure 10. Hydrogen make up. (a) triolein fed at stage 9. (b) triolein fed at stage 12
c) Hydrocarbon distribution

The quality of the final product obtained as an improved ultra-low sulphur diesel is very important,

as it defines the operating conditions of a second isomerization step to produce the green diesel.

This is, for the different triolein feed stage and feed flows configurations proposed, the hydrocarbon

distribution along the RDC must guarantee an optimal green diesel quality, i.e., most of the heaviest

hydrocarbons products should leave at the bottom exit stream of the RDC.

Figure 11 shows three hydrocarbon liquid composition profiles for different triolein feed flows and

feed stage locations. Figure 11a shows the hydrocarbon liquid composition profile considering a

triolein feed flow of 5 (kmol/hr) at stage 9. It can be noted that the liquid compositions of the

lightest hydrocarbon compounds (n-C11 and n-C12) increase at stage 9 and decrease at stage 7. This

strange behavior is due to the cold point and hotspot at such stages, where the vaporization of

triolein promotes the n-C18 generation from the HDO reaction. At the cold point (stage 9) the heat

from the stage 10 (higher temperature, ~330 °C) is used to vaporize some triolein, increasing the

liquid composition of the lightest hydrocarbon compounds. Figure 11b shows the hydrocarbon

distribution for triolein fed at stage 12. It is observed that, for the same triolein feed flow (5

kmol/hr), the liquid composition of n-C18 at the bottom of the column is lower (xn-C18 = 0.22) than

for the case of triolein fed at stage 9 (xn-C18 = 0.34). This composition difference is understandable

since part of the unconverted sulphured compounds (DBT) are present at the bottom of the column.

However, Figure 11c shows that similar liquid composition of n-C18 (xn-C18 = 0.34) can be obtained

with a triolein feed flow of 10 (kmol/hr) at stage 12 without affecting the deep HDS of diesel. Thus,

for a better hydrocarbon quality at the bottom of the co-hydrotreating RDC, the configuration

considering triolein fed at stage 12 is recommended.

Figure 11. Hydrocarbon liquid composition profile. (a) Triolein fed (5 kmol/hr) at stage 9; (b)

triolein fed (5 kmol/hr) at stage 12; (c) triolein fed (10 kmol/hr) at stage 12
d) Liquid water production

A very important factor to determine the appropriate triolein feed location and the operating

conditions of the hydrotreating process is the liquid water distribution along the reactive separation

process. This is due to three main reasons: i) liquid water may affect severely the catalyst activity,

ii) the presence of CO (or CO2) generated can form carbonic acid if liquid water is present and

carbonic corrosion in the air cooler and cold separator is possible, iii) high liquid water

concentrations promote the existence of two liquid phases and azeotropes, which may lead to

operating difficulties of the HDT reactive separation process.

Figure 12 shows the liquid water mole fraction profiles for different triolein feed flows and

locations. Figure 12a shows that the highest liquid water concentrations are at the reactive zone I

and at the top of the RDC. It is interesting to observe in Figure 12a that most of the water is

produced at reactive zone I, where n-C18 is produced by the HDO reaction and certainly, where the

hotspot appears. However, in Figure 12b a flat liquid water distribution (almost constant mole

fraction) along the RDC is shown. In this case, water is mainly produced at the reactive zone II and

it is readily vaporized at such stages. It should be pointed out that the amount of water in the

reactive zones is less that 10 % of the total molar liquid amount, that is, the maximum liquid water

mole fraction with triolein feed flow of 15 (kmol/hr) is xH2O=0.062. Consequently, with such low

liquid water concentrations along the RDC catalyst deactivation and corrosion problems are

negligible.

Figure 12. Liquid water profile along the RDC. (a) triolein fed at stage 9; (b) triolein fed at

stage 12

d) Release of generated gases


In this work, the gases produced by HDS and HDO reactions are: propane, steam and H2S. In the

commercial processes these gases are removed by a gas cleaning step or by increasing the purge gas

rate in any HDT packed bed reactor [56]. In the case of the hydrotreating RDC, the gases produced

are mostly driven in the vapour phase at each equilibrium stage and released from the partial

condenser at the top of the column. Tables 9 and 10 give the simulation results of the input and exit

streams indicating the molar flows, temperature and mass fractions of each stream. It can be noted

in the fourth column of Table 9 (5.4 kmol/hr of triolein fed at stage 9), that stream 4 (see Figure 6)

contains the most volatile compounds, that is, most of the propane, steam and H2S produced are

released at the exit stream 4. Also, it must be observed that stream 4 is further separated by a second

flash step to obtain a rich n-C11 and n-C12 hydrocarbon liquid phase. Table 10 gives similar results

for the light gases at stream 4 for 15 (kmol/hr) of triolein fed at stage 12. However, it should be

noted that the amount of n-C18 is drastically reduced (mass fractions: 4.32E-06, 8.81E-05) in stream

4 and 5 compared to that with triolein fed at stage 9 (mass fractions: 0.006, 0.105314). Thus, the

RDC design configuration considering triolein fed at stage 12 guarantee that the light gases

produced are released at the top of the column and that the enhanced diesel is completely

accumulated at the bottom of the column (stream 6).

Table 9. Stream simulation results for the Hydrotreating RDC with Triolein Fed (5.4

Kmol/hr) at stage 9 (see Figure 6).

Table 10. Stream simulation results for the Hydrotreating RDC with Triolein Fed (15

Kmol/hr) at stage 12 (see Figure 7).

4.2 Oleic acid study case

Simulation of the two RDC configurations considering only oleic acid as the renewable bio oil fed

to the RDC were performed. Experimental data of HDT of this feedstock type have been reported

by Marker at al. [64], where a brown grease is used as the vegetable oil raw material to be
hydrotreated and the fatty acid content was: 25% stearic acid, 45% oleic acid, 25 linoleic acid and

5% linolenic acid. In the present work, only oleic acid is considered to represent the vegetable oil. A

sulphured diesel and oleic acid mixture is fed at stage 9 of the RDC (see Figure 6). Simulation

results analysis for oleic acid is similar as for triolein study case.

Figure 13 shows the temperature, n-C18 and DBT liquid composition profiles along the RDC

considering that oleic acid is pre-mixed with the sulphured diesel and fed at stage 9. The HDT

process begins with a small amount of oleic acid (1 kmol/hr) in the feed mixture and it was

continuously increased (0.5 Kmol/hr) to 75 (kmol/hr). Figure 13a shows the variation of

temperature along the reactive distillation column. It can be noted that, contrary to the triolein

example, there are not hotspot or cold points along the RDC. Also, it should be observed in Figure

13a that, by increasing the oleic acid content in the mixture fed, a decrease in temperature at the

reactive zone I is observed. This result is not surprising, since the oleic acid boiling temperature is

greater than the boiling temperature of most of the hydrocarbons present in the sulphured diesel (see

Table 1). Figure 13b shows the n-C18 liquid composition profile along the RDC. It can be seen that,

most of n-C18 is produced in the reactive zone II and accumulated at the bottom of the column.

Figure 13c shows the effect of the oleic acid content in the feed mixture on the deep HDS of diesel.

It can be noted in Figure 13c that, even for a high oleic acid content (75 kmol/hr) deep HDS of

diesel is achieved. The above results lead to conclude that the performance of the RDC is enhanced

when a vegetable oil with high fatty acid content (oleic acid in this case) is used as the vegetable oil

raw material. It should be pointed out that this kind of vegetable oils with (high fatty acid content)

are the cheapest raw material in the bio-oil market [33].

Figure 13. RDC profiles at P=30 atm and oleic acid-petro-diesel blend fed at stage 9. (a)

temperature profile; (b) HDO product n-C18, liquid composition profile; (c) HDS of DBT,

liquid composition profile


Figure 14 shows the flow rate of hydrogen required as a function of the feed flow of oleic acid at

stage 9. It can be observed in Figure 14 that, for an oleic acid feed flow greater than 20 (kmol/hr),

the temperature of the condenser is reduced to 215 °C, and for a feed flow of 75 kmol/hr, a sharp

increase in the flow rate of hydrogen occurs (300 to 310 kmol/hr) and the temperature of the

condenser must be decreased to 185 °C. It should pointed out that, if the temperature at the

condenser is not diminished, convergence of the RDC simulation is not reached. In this case, from

the controllability point of view, a control loop between the vegetable oil feed (oleic acid) and the

hydrogen supplied should be linked to the temperature of the condenser.

Figure 14. Hydrogen make up. Oleic acid feed at stage 9. At OA feed flow of 20 and 75

(kmol/hr) a reduction of the condenser temperature is required

Figure 15 shows the hydrocarbon liquid composition distribution and the liquid water profile along

the RDC. It can be noted at Figure 15a that the hydrocarbon distribution is similar to that obtained

when triolein is independently fed at stage 12 (see Figure 11c). Thus, a good quality hydrocarbon

mixture is attained at the bottom of the RDC. Figure 15b shows the liquid water mole fraction

profiles for different oleic acid feed flows. Figure 15b shows that water is produced in reactive zone

II (reactive stage 9-11) and the liquid water mole fraction increases continuosly from stage 9 to the

top of the column as the amount of oleic acid in the feed mixture augmented. Thus, to avoid catalyst

deactivation and corrosion problems, the oleic acid feed flow should not be greater than 50 kmol/hr.

Figure 15. RDC profiles for oleic acid-petrodiesel blend fed at stage 9 at 30 atm. (a)

Hydrocarbon liquid composition profile with oleic acid feed flow = 50 kmol/hr. (b) Water

liquid composition profile


Finally, from the results given in Table 11 shows that for 75 (kmol/hr) of oleic acid pre-mixed and

fed at stage 9, most of the light gases are released at the exit stream 4. It should be noted from the

data given in Table 11 that the amount of n-C18 at the top of the column is irrelevant (mass

fractions: 1.59E-06, 4.01E-05) at the exit streams 4 and 5, respectively. Therefore, the RDC

configuration considering oleic acid pre-mixed and fed at stage 9 render a higher production of

henhanced ultra low sulphur diesel at the bottom of the column (exit stream 6).

Table 11. Stream simulation results for the Hydrotreating RDC with Oleic Acid Feed (75

Kmol/hr) at stage 9 (See Figure 6).

Finally, a comparison of the energy consumption of the HDO of vegetable oils process proposed by

Glisic et al. [33] with the results obtained in the present work is performed. It should mentioned,

however, that the hydrotreated reactor proposed by Glisic et al. [33] does not consider the deep

HDS of diesel. The simulation results [33] show that for a product capacity of 100,000 tonnes/year

the energy consumption is 22720 MW/year rendering an energy consumption of 0.74 KW/Kg of

product, while for the co-HDT RD process presented in this work the energy required is 0.0165

KW/Kg of product for a capacity of 164,588 Kg/year. It should pointed out that more energy is

required in the co-HDT RD process due to the sulphured petro-diesel fed to the RD column.

5. Conclusions

A new reactive distillation process for hydrotreating of vegetable oils and petro-diesel blends

has been developed. By performing process intensification of the new HDT-RD process, the

technical challenges found in the conventional packed bed hydrotreating reactors could be

overcome. Based on the simulation results of the two hydrotreating RDC configurations it is

proposed that the following conditions should be fulfilled in order to obtain an optimal green

diesel production:
1) To manage large amounts of vegetable oils containing high concentration of triglycerides,

the optimal vegetable oil feed stage location at the RDC should be below the reactive zone

II (stage 12) and, for the case when the vegetable oil contents are mostly fatty acids (waste

vegetable oils), the vegetable oil should be premixed with the petro-diesel and fed at stage 9.

2) Catalyst deactivation and corrosion problems due to the liquid water production must be

avoided, thus, water should be vaporized and together adverse gases produced should be

vaporized and released at the top of the RDC.

3) Phase equilibrium calculations reveal that an operating pressure of P=30 atm leads to safe

operation of the reaction and separation in the L-V region at each stage of the RD column.

This low pressure also favors the low hydrogen consumption and low energy consumption

to perform the separation. At higher total pressures (80-100 atm) the existence of LI-LII-V

phase behavior is predicted and operating and controllability of the RDC may be extremely

difficult. Therefore, high pressure operations should be avoided.

4) The hydrocarbon distribution along the RDC shows that is feasible to obtain an enhanced

ultra low sulfur diesel or green diesel at the bottom of the column. However, the following

operating conditions must be satisfied: P= 30 atm, petro-diesel fed at stage 9, vegetable oil

(high triglycerides content) separately fed at stage 12 and/or vegetable oil (high fatty acids

content) pre-mixed with petro-diesel and fed at stage 9.

In spite of the above observations on the HDT-RD process, it should be pointed out that the well

established knowledge of the reaction rates and kinetic expressions for deep HDS of diesel is

necessary to ensure that the correct interaction between the fluid phases and the catalyst surface

is considered in process modeling and associated simulations. However, in the case of the HDO

reaction network, a deeper understanding of the reaction pathways is needed and more

experimental data should lead to more reliable reaction rate and kinetic expressions.
Aknowledgements

The authors aknowledge the financial support provided by CONACyT to the Project 256787.

6. References

[1] J.H. Van Gerpen. Biological and Agricultural Engineering. University of Idaho, Moscow, ID,

USA (Website: www.deq.state.mt.us.—Biodiesel Economics. Montana Economics 2007)

[2] T. M. Yunus khan. A.E. Atabani, Irfan Anjum Badruddin, Ahmad Badarudin, M.S. Khayoon, S.

Triwahyono, Recent scenario and technologies to utilize non-edible oils for biodiesel

production, Renewable and Sustainable Energy Reviews, 2014, 37, 840-851.

[3] A. A. Kiss and C. Sorin Bildea, A review of biodiesel production by integrated reactive

separation Technologies, J Chem Technol Biotechnol 2012; 87: 861–879

[4] G.W. Huber, S. Iborra, A. Corma, Chem. Rev. 106 (2006) 4044.

[5] G. Knothe, J. Krahl, J.V. Gerpen, The Biodiesel Handbook, AOCS Press, Champaign, IL, 2005.

[6] E. Lotero, Y. Liu, D.E. Lopez, K. Suwannakarn, D.A. Bruce, J.G. Goodwin, Ind. Eng. Chem.

Res. 44 (2005) 5353.

[7] M. McCoy, Chem. Eng. News 83 (2005) 19.

[8] G W Hu r, P O’Connor, A Corma, Appl Catal A G n 329 (2007) 120

[9] P Pr l, D Ku ˇka, L C ˇ ap k, Bastl, P R sˇa´n k, Appl Catal A G n 397

[10] J. Jakkula, P. Aalto, V. Niemi, U. Kiiski, J. Nikkonen, S. Mikkonen, O. Piirainen WO

2004/022674 A1.

[11] G. Ondrey, Chemical Engineering 112 (2005) 14.

[12] www.nesteoil.com/default.asp?path=1,41,535,547,557,12361.

[13] F. Baldiraghi, M. Di Stanislao, G. Faraci, C. Perego, T. Marker, C. Gosling, P. Kokayeff, T.

Kalnes, R. Marinangeli, Ecofining: new process for green diesel production from vegetable

oil, G. Centi, F. Trifiro` , S. Perathoner, F. Cavani (Eds.), Sustainable Industrial Process,

Wiley–VCH, Weinheim, 2009, pp. 427–438.


[14] T. Kalnes, T. Marker, D.R. Shonnard, Int. J. Chem. React. Eng. 5 (2007) Article A48.

[15] www.uop.com/renewables/10010.html.

[16] http://www.b2i.us/profiles/investor/fullpage.asp?f=1&BzID=2029&to=cp&Nav=0&LangID=1&s=0&ID=11923.

[17] R. Abhari, L. Tomlinson, P. Havlik, N. Jannasch US Patent 7,968,757 B2.

[18] www.conocophillips.com/EN/tech/initiatives/Pages/index.aspx.

[19] D.R. Ghonasgi, E.L. Sughrue II, J. Yao, X. Xu US Patent 7,955,401 B2.

[20] http://www2.petrobras.com.br/tecnologia/ing/hbio.aspN.

[21] J.R. Gomes, J.L. Zotin, M.E. Pacheco, R.M.C.F. Silva US Patent 2010/0270207 A1.

[22] R. Egeberg, K. Knudsen, S. Nystro¨m, E. Lind, K. Efraimsson, PTQ Q3 (2011) 59.

[23] http://www.topsoe.com/sites/default/files/novel_hydrotreating_technology_for_production_of_green_diesel.ashx_.pdf

[24] http://www.abb.com/cawp/seitp202/8caa932d262fb41cc12572110055f30c.aspx

[25] http://www.google.com.mx/patents/US6416659

[26] Eduardo S. Perez-Cisneros, Salvador A. Granados-Aguila, Pedro Huitzil- Melendez and Tomas

Viveros-Garcia. European Symposium on Computer Aided Process Engineering – 12 J.

Grievink and J. van Schijndel (Editors) pp. 301-306 (2002).

[27] Viveros-García, T., Ochoa-Tapia, J. A., Lobo-Oehmichen, R., de los Reyes-Heredia, J. A., &

Pérez-Cisneros, E. S. (2005). Chemical Engineering Journal, 106, 119

[28] Granados-Aguilar, A. S., Viveros-García, T., & Pérez-Cisneros, E. S. (2008). Chemical

Engineering Journal, 143, 210.

[29] Cárdenas-Guerra J.C., Teresa López-Arenas, Ricardo Lobo-Oehmichen, Eduardo S. Pérez-

Cisneros. (2012). Computers and Chemical Engineering, 34, 196–209

[30] Pryor RW, Hanna MA, Schinstock JL, Bashford LL. Soybean oil fuel in a small diesel engine.

Transactions of the ASAE 1982;26:333–8.

[31] Moser BR., Camelina (Camelina sativa L.) oil as a biofuels feedstock: golden opportunity or

false hope. Lipid Technology 2010; 22: 270–273.


[32] Sharma YC, Singh B. A hybrid feedstock for a very efficient preparation of biodiesel. Fuel

Processing Technology 2010;91:1267–73.

[33] Sandra B. Glisic, Jelena M. Pajnik, Aleksandar M. Orlovic, Process and techno-economic

analysis of green diesel production from waste vegetable oil and the comparison with ester

type biodiesel production, Applied Energy 170 (2016) 176–185

[34] Ku ˇka D, B j lova´ M, Vlk J Conv rs on of v ta l o ls nto h ro ar ons over

CoMo/MCM-41 catalysts. Topics in Catalysis 2010;53:168–78.

[35] Prasad YS, Bakhshi NN. Effect of pretreatment of HZSM-5 catalyst on its performance in

canola oil upgrading. Applied Catalysis A: General 1985;18: 71–85.

[36] Palanisamy S, Gevert BS. Thermal treatment of rapeseed oil. Bioenergy technology. World

renewable energy congress 2011-Sweden.

[37] Van Parijs, I. A., & Froment, G. F. (1986). Kinetics of hydrodesulfurization on a CoMo/_-

Al2O3 catalyst. 1. Kinetics of the hydrogenolysis of thiophene. Industrial and Engineering

Chemistry Product Research and Development, 25, 431.

[38] Van Parijs, I. A., Hosten, L. H.,&Froment, G. F. (1986). Kinetics of hydrodesulfurization on a

CoMo/_-Al2O3 catalyst. 2. Kinetics of the hydrogenolysis of benzothiophene. Industrial and

Engineering Chemistry Product Research and Development, 25, 437.

[39] D.D. Whitehurst, T. Isoda, I. Mochida, Present state of the art and future challenges in the

hydrodesulfurization of polyaromatic sulfur compounds, Adv. Catal. 42 (1998) 345–471.

[40] P.D. Ting, Joyce P.C., Prasana K.J., Chapman W.G., Thies M.C., Phase equilibrium modeling

of mixtures of long-chain and short-chain alkanes using Peng-Robinson and SAFT , Fluid

Phase Equilibria, 2003, 206, 267-286

[41] Maria C. Annesini, F. Gironi and W. Guerani, Phase equilibria of highly asymmetric mixtures

involved in biodiesel production, Chemical Engineering Transactions, 2014, 38, 67-72

[42] M. Sales-Cruz, Aca-Aca G., S n h -Da a O , L p -Arenas T. Predicting critical properties,

density and viscosity of fatty acids, triacylglycerols and methyl esters by group contribution
methods. In 20th Europ an S mpos um on Comput r A Pro ss En n r n −

ESCAPE20; Pierucci, S., G. Buzzi Ferraris, Eds.; Elsevier B.V.: Amsterdam, 2010; Poster

101.

[43] C K r -Mikkelsen, Stenby, E. H. Predicting the melting points and the enthalpies of

fusion of saturated triglycerides by a group contribution method. Fluid Phase Equilib. 1999,

162, 7−17

[44] R. Ceriani, M r ll s, A J A Pr t n vapor−l qu qu l r a of fatty systems. Fluid Phase

Equilib., 2003, 215, 227−236

[45] C.-A. Díaz-Tovar, Gani, R.; Sarup, B. Lipid technology: Property prediction and process

design/analysis in the edible oil and biodiesel industries. Fluid Phase Equilib., 2011, 302,

284−293

[46] J. Marrero, Gani, R. Group-contribution based estimation of pure component properties. Fluid

Phas Equ l 2001, 183−184, 183−208

[47] M.S. Wertheim, Fluids with Highly Directional Attractive Forces. II. Thermodynamic

Perturbation Theory and Integral Equations, Journal of Statistical Physics, 1984,35, 35-47.

[48] Barker JA, Henderson D. Perturbation theory and equation of state for fluids: a successful

theory of liquids. J Chem Phys 1967;47:2856–61.

[49] J. Gross and G. Sadowsky, Perturbed-Chain SAFT: An Equation of State Based on a

Perturbation Theory for Chain Molecules, Industrial and Engineering Chemistry Research,

2001, 40, 1244-1260

[50] Tóth C, Baladincz P, Kova´ cs S, Hancsk J. Producing clean diesel fuel by cohydrogenation of

vegetable oil with gas oil. Clean Technologies and Environmental Policy 2011;13:581–5.

[51] D.H. Broderick, B.C. Gates, Hydrogenolysis and hydrogenation of dibenzothiophene catalyzed

by sulfided CoO-MoO3/_-Al2O3: the reaction kinetics, AIChE J. 27 (4) (1981) 663–673.


[52] Froment, G. F., Depauw, G. A., & Vanrysselberghe, V. (1994). Kinetic modeling and reactor

simulation in hydrodesulfurization of oil fractions. Industrial and Engineering Chemistry

Research, 33, 2975.

[53] M. Houalla, N.K. Nag, A.V. Sapre, D.H. Broderick, B.C. Gates, Hydrodesulfurization of

dibenzothiophene catalyzed by sulfided CoO-MoO3/_-Al2O3: the reaction network, AIChE J.

24 (6) (1978) 1015–1021.

[54] Vanrysselberghe, V., & Froment, G. F. (1996). Hydrodesulfurization of dibenzothiophene on a

CoMo/Al2O3 catalyst: Reaction network and kinetics. Industrial and Engineering Chemistry

Research, 35, 3311.

[55] V. Vanrysselberghe, R.L. Gall, G.F. Froment, Hydrodesulfurization of 4-

methyldibenzothiophene and 4,6-dimethyldibenzothiophene on a CoMo/_- A12O3 catalyst:

reaction network and kinetics, Ind. Eng. Chem. Res. 37 (1998) 1235–1242.

[56] Senol OI, Viljava TR, Krause AOI. Hydrodeoxygenation of methyl esters on sulfided NiMo/g-

Al2O3 and CoMo/gAl2O3 palm diesel: green and renewable fuel from palm oil. Palm oil

developments catalysts. Catalysis Today 2005;100:331–5.

[57] Topsøe H, Clausen BS, Massoth FE. Hydrotreating catalysts: science and technology, 22.

Germany: Springer; 1996 141–144.

[58] Ferrari M, Bosmans S, Maggi R, Delmon B, Grange P. CoMo/carbon hydrodeoxygenation

catalysts: influence of the hydrogen sulfide partial pressure and of the sulfidation temperature.

Catalysis Today 2001;65: 257–264.

[59] Zuo H, Liu Q, Wang T, Ma L, Zhang Q, Zhang Q. Hydrodeoxygenation of methyl palmitate

over supported Ni catalysts for diesel-like fuel production. Energy Fuels 2012;26:3747–55.

[60] Furimsky E, Massoth FE. Deactivation of hydroprocessing catalysts. Catalysis Today

1999;52:381–495.
[61] Krár M, Kova´ cs S, Boda L, Leveles L, Thernesz A, Wa´hlne I, et al. Fuel purpose

hydrotreating of vegetable oil on NiMo/gAl2O3 catalysts. Hungarian Journal of Industrial

Chemistry 2009;37:107–11.

[62] Y.E. Sheu, R.G. Anthony, E.J. Soltes, Fuel Process. Technol. 19 (1988) 31–50.

[63] Haiping Zhang, Hongfei Lin, Weizhi Wang, Ying Zheng, Peijun Hu, Hydroprocessing of waste

cooking oil over a dispersed nano catalyst: Kinetics study and temperature effect, Applied

Catalysis B: Environmental 150– 151 (2014) 238– 248

[64] http://www.osti.gov/scitech/servlets/purl/861458
Table 1. Critical properties of the pure components and PC-SAFT equation of state parameters.
Properties determined using the Marrero and Gani (2001) method included in ICAS Platform
(CAPEC, 2015)
Critical Properties PC-SAFT Parameters
Species MW Tb Tc Pc w m  /k
(gr/mol) (K) (K) (Bar) (A) (K)

Petro-Diesel
Thiophene 84.14 357.3 579.4 56.843 0.1969 2.4013 3.5401 299.01
Benzothiophene 134.2 493.1 754.0 41.340 0.2954 2.9313 3.8360 363.28
Di-Benzothiophene 184.26 604.6 897.0 38.503 0.3982 3.7015 3.9596 386.77
H2 2.02 20.5 33.3 13.071 -0.2159 1.0 2.986 19.2775
Butadiene 54.09 268.8 425.0 43.164 0.1950 2.2979 3.5395 239.01
Ethylbenzene 106.17 409.4 617.2 36.072 0.3034 3.0799 3.7974 287.35
BiPhenyl 154.21 528.2 773.0 33.741 0.4082 3.8877 3.8151 327.42
H2S 34.08 212.9 373.5 89.571 0.0941 1.6941 3.0214 226.79
n-C11 156.31 469.1 639.0 19.454 0.5303 4.9082 3.8893 248.82
n-C12 170.34 489.5 658.0 18.137 0.5763 5.3060 3.8959 249.21
n-C13 184.37 508.6 675.0 16.719 0.6173 5.6877 3.9143 249.78
n-C14 198.39 526.7 693.0 15.604 0.6430 5.9002 3.9396 254.21
n-C16 226.45 560.0 723.0 13.983 0.7174 6.6485 3.9552 254.7
n-C18 254.49 589.3 747.7 12.92 0.7946 7.3271 3.9668 256.2

Vegetable Oil and Biodiesel


Triolein 885.43 878.74 985.99 2.68 2.549 23.4907 3.925683 264.6772
Methyl Oleate 296.49 618.93 768.69 11.81 0.967 8.5753 3.850432 256.3849
Oleic Acid 282.47 647.04 810.37 13.40 1.150 10.5829 3.522527 244.0751
Table 2. Validation of the pure component properties and PC-SAFT parameters with results
reported by Annesini et al. (2014). Mass fractions of the outlet streams of a three-phase flash
separator.
Streams T (K) Flow ratio Triglycerides Fatty acids Alkanes Water Propane Gaseous
(kg/hr) (C15-C18) (H2,CO2,CO)
Inlet 1 0.255 0.084 0.538 0.047 0.032 0.044

Vapor
Phase 313.2 0.044 - - - 0.007 0.267 0.726
(V) This 0.0526 6.322e-26 8.887e-09 2.44e-05 0.00501 0.2630 0.732
work
353.2 0.054 - - - 0.035 0.313 0.625
This 0.06082 5.105e-22 5.142e-07 3.98e-04 0.0294 0.3126 0.658
work
383.2 0.066 - - 0.002 0.122 0.312 0.564
This 0.0681 1.444e-19 5.610e-06 0.0020 0.0826 0.319 0.5964
work

Oil-rich
liquid 313.2 0.910 0.281 0.093 0.591 0.001 0.021 0.013
Phase This 0.902 0.2828 0.0932 0.5967 0.0011 0.02016 0.006
(LII) work
353.2 0.905 0.282 0.094 0.594 0.005 0.016 0.009
This 0.897 0.284 0.0937 0.599 0.00325 0.0145 0.0055
work
383.2 0.907 0.282 0.093 0.592 0.014 0.012 0.007
This 0.896 0.2845 0.0937 0.6001 0.0064 0.0115 0.0038
work

Water-
rich 313.2 0.046 - - - 0.99901 2e-04 7.9e-04
liquid This 0.0457 3.78e-20 1.24e-20 7.97e-20 0.99993 6.98e-07 7e-05
Phase work
(LI) 353.2 0.041 - - - 0.99894 3e-04 7.6e-04
This 0.0423 3.82e-20 1.26e-20 8.06e-20 0.99987 5.52e-06 1.3e-04
work
383.2 0.027 - - - 0.99888 4e-04 7.2e-04
This 0.0356 3.73e-20 1.23e-20 7.86e-20 0.99979 1.818e- 2.1e-04
work 05
Table 3. LI-LII-V equilibrium compositions for the product mixture triolein (TO)- n-hexadecane -
H2O at 30 atm. (a) Tie line compositions, (b) Azeotrope compositions and temperature
(a) Tie line equilibrium compositions for Triolein-n-C16-H2O product mixture at P= 30 atm

Tie Line xITRIOLEIN xIH20 xIn-C16 xIITRIOLEIN xIIH2O xIIn-C16 yTRIOLEIN yH2O y n-C16
1 0 0.9999 8.09E-05 0.0000 0.6652 0.3348 0 0.9923 0.0077
2 2.27E-11 0.9999 6.90E-05 0.0391 0.7126 0.2483 9.40E-12 0.9935 0.0065
3 6.38E-11 0.9999 5.16E-05 0.0781 0.7564 0.1654 2.58E-11 0.9952 0.0048
4 1.32E-10 1.0000 2.85E-05 0.1177 0.7992 0.0831 5.20E-11 0.9973 0.0027
5 2.37E-10 1.0000 0 0.1581 0.8419 0.0000 9.01E-11 1.0000 0.0000

(b) Azeotrope compositions and temperature P=30 atm


No. xIITRIOLEIN xIIH20 xIIn-C16 Temperatura K
1 0.64 0.36 0 1174.041
Table 4. LI-LII-V equilibrium compositions for the product mixture oleic acid (OA)- n-hexadecane -
H2O at 50 atm. (a) Tie line compositions, (b) Azeotrope compositions and temperature
(a) Tie line equilibrium compositions for Oleic Acid-n-C16-H2O product mixture at P= 50 atm
Tie Line xIOleic Acid xIH20 xIn-C16 xIIOleic Acid xIIH2O xIIn-C16 YOleic Acid yH2O y n-C16
1 0.0000 0.9997 0.0003 0.0000 0.7703 0.2297 0.0000 0.9902 0.0098
2 0.0006 0.9991 0.0003 0.0213 0.7843 0.1944 0.0000 0.9911 0.0088
3 0.0016 0.9982 0.0003 0.0467 0.8054 0.1479 0.0001 0.9924 0.0075
4 0.0033 0.9964 0.0003 0.0734 0.8411 0.0855 0.0002 0.9942 0.0056
5 0.0705 0.9202 0.0093 0.0087 0.9911 0.0002 0.0003 0.9978 0.0019

(b) Azeotrope compositions and temperature at P=50 atm


No xIIOleic Acid xIIH2O xIIn-C16 Temperature K
1 0.0000 0.9902 0.0098 535.9240
2 0.6726 0.3272 0.0001 911.4946
Table 5. Reaction rate expressions, reaction rate coefficients, and adsorption equilibrium constants for the thiophene and benzothiophene HDS.

Component Thiophene (Van Parijs and Froment, 1986) Benzothiophene (Van Parijs et al., 1986)

kTh, pTh pH2 kBT , KBT , KH2 , pBT pH2


rTh,  3
rBT ,  3
   pH2S 
 
Rate expression pH2S 1/2
1  KH2 , p1/2
H2  KTh, pTh  K H 2 S ,  1  KH2 , pH2  KBT , pBT  KH2S , 
 pH2   pH2 

Reaction rate coefficients

 125.1014x106 J / kmol   73.6877 x106 J / kmol 


kTh,  52200exp    kmol / kgcat s kBT ,  0.94exp    kmol / kgcat s
 RT   RT 
Adsorption equilibrium constants

Kinetic parameters
 44.7687 x106 J / kmol 
KTh,  5.5268x109 exp   m / kmol  KBT ,  19.047 x105 m3 / kmol 
3

 RT 
KH2 ,  3.5332x106 m3 / kmol 
KH2 ,  5.2899x10 6
m3 / kmol 

KH2S ,  20.824x104 m3 / kmol 


KH2S ,  9.0x104 m3 / kmol 

Where T is in (K) and R is in (J/kmol K)


Table 6. Reaction rate expressions, reaction rate coefficients, and adsorption equilibrium constants for the dibenzothiophene HDS.

Reaction pathway Hydrogenolysis Hidrogenolysis and hydrogenation

Component Dibenzothiophene (Froment et al., 1994) Dibenzothiophene (Vanrysselberghe et al., 1998)

k1KDBT , KH2 , CDBT CH2


rDBT ,  kDBT , KH , KDBT ,CDBT CH2
 
3
rDBT , 
Rate expression CH S
1 K 
3
1  KDBT , CDBT  KH2 , CH2  KH2S , 2  CDBT  KH ,CH2  KBPH ,CBPH  KH2S ,CH2S  K4,6DMDBT ,C4,6DMDBT
DBT ,

 CH2 

Reaction rate coefficients

 158x106 J / kmol   122.77 x106 J / kmol 


k1  7.84x108 exp    kmol / kgcat s kDBT ,  6.7871x106 exp    kmol / kgcat s a

 RT   RT 
Adsorption equilibrium constants

 113.232x106 J / kmol 
 3.3631x10 exp   m / kmol 
11 3 a
KH ,
Kinetic
 RT 
parameters KDBT ,  5.31 m3 / kmol 
KDBT ,  7.5687x101 m3 / kmol  a

KH2 ,  4.02 m3 / kmol 


 48.214x106 J / kmol 
KBPH ,  3.8498x104 exp   m / kmol 
3 a

KH2S ,  1.72  RT 

 105.67 x106 J / kmol 


 1.4712x10 exp   m / kmol 
8 3 a
KH2S ,
 RT 

Where Ci is the concentration of species i, T is in (K) and R is in (J/kmol K). a Data taken from (Vanrysselberghe and Froment, 1996)
Table 7. Design specifications of the hydrotreating reactive distillation column.

Specification Value Specification Value

Number of total stages 14 H2/HC feed ratio 3

Stages of reactive zone I 5-7 Distillate flowrate (kmol/hr) 340

Stages of reactive zone II 10-12 Reflux ratio 0.5

HC feed stage 9 Column pressure (atm) 30

H2 feed stage 12 Partial condenser temperature (°C) 225

HC feed flowrate (kmol/hr) 100 Holdup per stage (kg of catalyst) 1000

43
Table 8. Sulphured diesel feed composition

Reaction pathway Hydrogenolysis

Feed composition
Component
(Mole fraction)

Th 0.0087

BT 0.0087

DBT 0.1000

n-C11 0.4966

n-C12 0.3166

n-C13 0.0089

n-C14 0.0015

n-C16 0.0589

Total sulfur content (wt %) 2.25

44
Table 9. Stream simulation results for the Hydrotreating RDC with Triolein Feed (5.4
Kmol/hr) at stage 9 (see Figure 6).
Triolein Sulphured Hydrogen
Stream: Feed diesel Feed Feed <4> <5> <6>
Mole Flow (kmol/sec)
H2 0 0 8.667E-02 5.613E-02 7.562E-04 6.104E-04
H2S 0 0 0 2.956E-03 2.271E-04 3.150E-05
Thiophene (Th) 0 2.417E-04 0 0 0 6.330E-22
Benzothiophene (BT) 0 2.417E-04 0 4.920E-09 1.900E-08 1.030E-07
Di-benzothiophene (DBT) 0 2.778E-03 0 1.990E-11 5.180E-10 4.580E-05
Butadiene 0 0 0 2.081E-04 3.360E-05 3.640E-08
Ethylbenzene 0 0 0 1.098E-04 1.278E-04 3.870E-06
Biphenil 0 0 0 1.445E-04 1.187E-03 1.400E-03
n-C16 0 1.636E-03 0 1.150E-05 2.865E-04 1.338E-03
n-C14 0 4.170E-05 0 1.680E-06 1.970E-05 2.030E-05
n-C13 0 2.472E-04 0 1.960E-05 1.550E-04 7.280E-05
n-C12 0 8.794E-03 0 1.188E-03 6.220E-03 1.386E-03
Water 0 0 0 7.660E-03 1.340E-03 1.290E-09
n-C11 0 1.380E-02 0 2.799E-03 9.972E-03 1.027E-03
n-C18 0 0 0 2.620E-05 1.370E-03 3.110E-03
Propane 0 0 0 1.360E-03 1.433E-04 5.420E-11
Triolein 1.500E-03 0 0 0 0 1.350E-26

Mass Fraction
H2 0 0 1 1.019E-01 4.618E-04 7.092E-04
H2S 0 0 0 9.077E-02 2.340E-03 6.200E-04
Thiophene (Th) 0 4.379E-03 0 0 0 3.070E-20
Benzothiophene (BT) 0 6.984E-03 0 5.940E-07 7.720E-07 7.960E-06
Di-benzothiophene (DBT) 0 1.102E-01 0 3.300E-09 2.890E-08 4.870E-03
Butadiene 0 0 0 1.014E-02 5.500E-04 1.140E-06
Ethylbenzene 0 0 0 1.050E-02 4.110E-03 2.366E-04
Biphenyl 0 0 0 2.008E-02 5.550E-02 1.240E-01
n-C16 0 7.978E-02 0 2.340E-03 1.970E-02 1.750E-01
n-C14 0 1.780E-03 0 3.010E-04 1.180E-03 2.320E-03
n-C13 0 9.815E-03 0 3.260E-03 8.650E-03 7.730E-03
n-C12 0 3.226E-01 0 1.820E-01 3.209E-01 1.360E-01
Water 0 0 0 1.243E-01 7.313E-03 1.340E-08
n-C11 0 4.644E-01 0 3.941E-01 4.721E-01 9.252E-02
n-C18 0 0 0 6.007E-03 1.053E-01 4.558E-01
Propane 0 0 0 5.390E-02 1.914E-03 1.380E-09
Triolein 1 0 0 0 0 6.890E-24

Total Flow (kmol/sec) 1.500E-03 2.778E-02 8.667E-02 7.261E-02 2.183E-02 9.043E-03


Total Flow (m3/sec) 1.926E-02 7.742E-03 1.275E-01 9.919E-02 5.772E-03 6.966E-03
Temperature (°C) 2.399E+02 2.399E+02 2.599E+02 225 225 4.566E+02
Pressure (atm) 3.000E+01 3.000E+01 3.000E+01 3.000E+01 3.000E+01 3.000E+01
Vapor Fraction 0 0 1 1 0 0
Enthalpy (J/kmol) 3.695E+08 -1.950E+08 6.891E+06 -3.343E+07 -2.230E+08 -7.130E+07
Entropy (J/kmol-K) 8.854E+05 -9.110E+05 -1.135E+04 -6.608E+04 -8.823E+05 -8.238E+05
Density (kg/m3) 6.900E+01 6.000E+02 1.370E+00 1.120E+01 5.720E+02 2.490E+02

45
Table 10. Stream simulation results for the Hydrotreating RDC with Triolein Feed (15

Kmol/hr) at stage 12 (see Figure 7).

Triolein Sulphured Hydrogen


Stream Feed Diesel Feed Feed <4> <5> <6>
Mole Flow (kmol/sec)
H2 0 0 1.222E-01 5.086E-02 2.330E-04 1.873E-03
H2S 0 0 0 3.163E-03 7.780E-05 1.900E-05
Thiophene (Th) 0 2.417E-04 0 6.270E-35 8.700E-35 2.330E-22
Benzothiophene (BT) 0 2.417E-04 0 1.210E-07 1.300E-07 9.300E-08
Di-benzothiophene (DBT) 0 2.778E-03 0 7.840E-11 5.290E-10 5.890E-07
Butadiene 0 0 0 2.294E-04 1.150E-05 7.670E-07
Ethylbenzene 0 0 0 1.692E-04 5.690E-05 1.520E-05
Biphenyl 0 0 0 6.290E-05 1.390E-04 2.575E-03
n-C16 0 1.636E-03 0 3.330E-07 2.070E-06 1.634E-03
n-C14 0 4.440E-05 0 3.100E-07 9.370E-07 4.320E-05
n-C13 0 2.472E-04 0 9.920E-06 2.060E-05 2.167E-04
n-C12 0 8.794E-03 0 1.333E-03 1.870E-03 5.592E-03
Water 0 0.000E+00 0 2.188E-02 1.197E-03 1.920E-03
n-C11 0 1.379E-02 0 4.683E-03 4.552E-03 4.560E-03
n-C18 0 0 0 3.000E-08 3.790E-07 1.250E-02
Propane 0 0 0 3.767E-03 1.255E-04 2.737E-04
Triolein 4.167E-03 0 0 8.610E-25 1.300E-28 1.370E-07

Mass Fraction
H2 0 0 1 5.786E-02 4.297E-04 6.606E-04
H2S 0 0 0 6.085E-02 2.426E-03 1.131E-04
Thiophene (Th) 0 4.379E-03 0 2.980E-33 6.700E-34 3.430E-21
Benzothiophene (BT) 0 6.984E-03 0 9.150E-06 1.590E-05 2.180E-06
Di-benzothiophene (DBT) 0 1.102E-01 0 8.150E-09 8.920E-08 1.900E-05
Butadiene 0 0 0 7.003E-03 5.691E-04 7.260E-06
Ethylbenzene 0 0 0 1.014E-02 5.525E-03 2.833E-04
Biphenil 0 0 0 5.471E-03 1.961E-02 6.949E-02
n-C16 0 7.978E-02 0 4.260E-05 4.297E-04 6.473E-02
n-C14 0 1.899E-03 0 3.470E-05 1.701E-04 1.500E-03
n-C13 0 9.815E-03 0 1.032E-03 3.474E-03 6.991E-03
n-C12 0 3.226E-01 0 1.281E-01 2.914E-01 1.667E-01
Water 0 0 0 2.225E-01 1.974E-02 6.051E-03
n-C11 0 4.643E-01 0 4.132E-01 6.510E-01 1.247E-01
n-C18 0 0 0 4.320E-06 8.810E-05 5.566E-01
Propane 0 0 0 9.376E-02 5.063E-03 2.112E-03
Triolein 1 0 0 4.300E-22 1.050E-25 2.120E-05

Total Flow (kmol/sec) 4.167E-03 2.778E-02 1.222E-01 8.616E-02 8.286E-03 3.122E-02


Total Flow (m3/sec) 1.534E-02 7.742E-03 1.799E-01 1.151E-01 1.949E-03 1.590E-02
Temperature (°C) 4.000E+02 2.399E+02 2.599E+02 2.250E+02 2.250E+02 4.146E+02
Pressure (atm) 3.000E+01 3.000E+01 3.000E+01 3.000E+01 3.000E+01 3.000E+01
Vapor Fraction 0 0 1 1 0 0
Enthalpy (J/kmol) -1.870E+09 -1.950E+08 6.891E+06 -7.491E+07 -2.290E+08 -1.385E+08
Entropy (J/kmol-K) -6.401E+07 -9.110E+05 -1.135E+04 -8.852E+04 -7.874E+05 -8.685E+05
Density (kg/m3) 2.405E+02 5.998E+02 1.370E+00 1.540E+01 5.607E+02 3.594E+02

46
Table 11. Stream simulation results for the Hydrotreating RDC with Oleic Acid Feed (75

Kmol/hr) at stage 9 (see Figure 6).

Mixed Feed Sulphured Hydrogen


Stream Stream Diesel Feed Feed <4> <5> <6>
Mole Flow (kmol/sec)
H2 0 0 1.333E-01 4.568E-02 2.870E-05 3.083E-03
H2S 0 0 0 3.238E-03 1.170E-05 1.030E-05
Thiophene (Th) 2.417E-04 2.417E-04 0 1.770E-33 3.860E-35 0
Benzothiophene (BT) 2.417E-04 2.417E-04 0 1.410E-07 2.550E-08 5.990E-07
Di-benzothiophene (DBT) 2.778E-03 2.778E-03 0 8.290E-12 1.020E-11 4.260E-07
Butadiene 0 0 0 2.392E-04 1.800E-06 6.730E-07
Ethylbenzene 0 0 0 2.059E-04 1.110E-05 2.390E-05
Biphenyl 0 0 0 1.700E-05 6.490E-06 2.754E-03
n-C16 1.636E-03 1.636E-03 0 5.100E-08 5.370E-08 1.636E-03
n-C14 4.170E-05 4.170E-05 0 6.630E-08 3.340E-08 4.160E-05
n-C13 2.472E-04 2.472E-04 0 3.200E-06 1.100E-06 2.429E-04
n-C12 8.794E-03 8.794E-03 0 7.437E-04 1.710E-04 7.880E-03
Water 0 0 0 3.836E-02 3.243E-04 2.076E-04
n-C11 1.380E-02 1.380E-02 0 4.663E-03 7.381E-04 8.397E-03
n-C18 0 0 0 1.110E-08 2.420E-08 1.944E-02
Oleic Acid 1.944E-02 0 0 8.810E-21 9.040E-20 3.170E-12

Mass Fraction
H2 0 0 1 5.153E-02 3.770E-04 7.342E-04
H2S 0 0 0 6.175E-02 2.603E-03 4.150E-05
Thiophene (Th) 2.006E-03 4.379E-03 0 8.340E-32 2.120E-32 0
Benzothiophene (BT) 3.200E-03 6.984E-03 0 1.060E-05 2.230E-05 9.490E-06
Di-benzothiophene (DBT) 5.050E-02 1.102E-01 0 8.550E-10 1.220E-08 9.280E-06
Butadiene 0 0 0 7.240E-03 6.351E-04 4.300E-06
Ethylbenzene 0 0 0 1.223E-02 7.703E-03 2.994E-04
Biphenyl 0 0 0 1.465E-03 6.526E-03 5.017E-02
n-C16 3.655E-02 7.978E-02 0 6.460E-06 7.930E-05 4.377E-02
n-C14 8.155E-04 1.780E-03 0 7.360E-06 4.320E-05 9.743E-04
n-C13 4.497E-03 9.815E-03 0 3.297E-04 1.322E-03 5.291E-03
n-C12 1.478E-01 3.226E-01 0 7.089E-02 1.900E-01 1.586E-01
Water 0 0 0 3.867E-01 3.810E-02 4.418E-04
n-C11 0.21277184 0.46443447 0 4.078E-01 7.525E-01 1.551E-01
n-C18 0 0 0 1.590E-06 4.010E-05 5.846E-01
Oleic Acid 0.54186897 0 0 1.390E-18 1.660E-16 1.060E-10

Total Flow (kmol/sec) 4.722E-02 2.778E-02 1.333E-01 9.315E-02 1.294E-03 4.372E-02


Total Flow (m3/sec) 1.532E-02 7.742E-03 1.962E-01 1.198E-01 2.669E-04 2.358E-02
Temperature (°C) 2.399E+02 2.399E+02 2.599E+02 2.150E+02 2.150E+02 414,59
Pressure (atm) 3.000E+01 3.000E+01 3.000E+01 3.000E+01 3.000E+01 3.000E+01
Vapor Fraction 0 0 1 1 0 0
Enthalpy (J/kmol) -3.943E+08 -1.950E+08 6.891E+06 -1.070E+08 -2.397E+08 -1.437E+08
Entropy (J/kmol-K) -1.136E+06 -9.110E+05 -1.135E+04 -7.581E+04 -7.173E+05 -9.395E+05
Density (kg/m3) 6.618E+02 5.998E+02 1.370E+00 1.491E+01 5.743E+02 3.590E+02

47
Figure Captions

Figure 1. Possible deoxygenation pathways of triglycerides


Figure 2. Reaction pathways for dibenzothiophene hydrodesulfurization
Figure 3. Hydrogen Solubility at 30, 50, and 80 atm. a) triolein (TO) - n-
hexadecane - H2 and b) oleic acid (OA) – n-hexadecane – H2
Figure 4. Equilibrium liquid compositions at 30, 50 and 80 atm as function of
temperature. a) triolein (TO) - n-hexadecane - H2 and b) oleic acid (OA)
– n-hexadecane – H2
Figure 5. Ternary L-L-V Equilibrium compositions at 30 atm for the product
mixtures: a) triolein (TO)- n-hexadecane - H2O at 50 atm and b) oleic
acid (OA)-n-hexadecane-H2O at 30 atm
Figure 6. Hydrotreating RDC design diagram 1. Vegetable oil pre-mixed with
sulphured Petro-diesel fed at stage 9
Figure 7. Hydrotreating RDC design diagram 2. Triolein fed at stage 12 and
sulphured Petro-diesel fed at stage 9
Figure 8. RDC profiles at P=30 atm and triolein-petro-diesel blend fed at stage 9.
(a) temperature profile; (b) HDO product n-C18, liquid composition
profile; (c) HDS of DBT, liquid composition profile
Figure 9. RDC profiles at P=30 atm and triolein-petro-diesel blend fed at stage 12.
(a) temperature profile; (b) HDO product n-C18,liquid composition
profile; (c) HDS of DBT, liquid composition profile
Figure 10. Hydrogen make up. (a) triolein fed at stage 9. (b) triolein fed at stage 12
Figure 11. Hydrocarbon liquid composition profile. (a) Triolein fed (5 kmol/hr) at
stage 9; (b) triolein fed (5 kmol/hr) at stage 12; (c) triolein fed (10
kmol/hr) at stage 12
Figure 12. Liquid water profile along the RDC. (a) triolein feed at stage 9; (b)
triolein feed at stage 12
Figure 13. RDC profiles at P=30 atm and oleic acid-petro-diesel blend fed at stage
9. (a) temperature profile; (b) HDO product n-C18, liquid composition
profile; (c) HDS of DBT, liquid composition profile
Figure 14. Hydrogen make up. Oleic acid fed at stage 9. With OA feed flows of 20
and 75 (kmol/hr) a reduction of the condenser temperature is required
Figure 15. RDC profiles for oleic acid-petrodiesel blend fed at stage 9 at 30 atm. (a)
hydrocarbon liquid composition profile, (b) liquid water composition
profile
Figure 1.
Figure 2.
Xn-Hexadecane
0,0
(a) 1,0
0,1
0,9
0,2
0,8
0,3
0,7
0,4
0,6
0,5
0,5
0,6
0,4
0,7
0,3
0,8
P=50 atm 0,2
0,9 P=80 atm
P=30 atm 0,1
1,0
0,0
XH
2 0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1,0 X Triolein

Xn-Hexadecane
0,0
(b) 1,0
0,1
0,9
0,2
0,8
0,3
0,7
0,4
0,6
0,5
0,5
0,6
0,4
0,7
0,3
0,8
0,2
0,9 P=80 atm P=50 atm P=30 atm
0,1
1,0
0,0
XH2 XOleic Acid
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1,0

Figure 3.
0,8
(a)
P=30 atm
Triolein

X (Liquid mole fraction) 0,6 50 atm

80 atm

0,4
n-Hexadecane
P=30 atm
80 atm

0,2
50 atm
80 atm 50 atm

P=30 atm
0,0 H2

200 400 600 800 1000 1200

Temperature (K)

0,8
(b)
Oleic Acid P=30 atm
P=50 atm
0,6
P=80 atm
X (Liquid mole fraction)

0,4
n-Hexadecane P=30 atm
P=50 atm
P=80 atm

0,2
P=50 atm
P=80 atm
H2
P=30 atm
0,0

200 300 400 500 600 700 800

Temperature (K)

Figure 4.
(a)

(b)

Figure 5.
<5> H2, H2O,
H2S

T=225 oC

Oleic Acid
<2>
T=240 oC
Triolein R=0.5 <6> Light HC
<2> (350 Kmol/hr)
T=240 oC Column sizing details (Triolein Feed Flow 5.4 Kmol/h)
Parameter size
<1> Number of theoretical stages
HC Feed stage
14
9
Reactive H2 Feed stage 12
Petro-Diesel
Zone I Triolein Feed stage 9
(100 Kmol/hr)
(stages 5-7) Pressure drop per stage 0 atm
mole fraction
Total Pressure of the column 30 atm
Th= 0.087
Reflux ratio 0.5
BT= 0.087
Number of reactive zones 2
DBT= 0.1
n-C11= 0.4966
Number reactive stages per zone 3
Number of nonreactive stages 8
n-C12= 0.3166
n-C13= 0.0089
<3> Holdup per reactive stage (kg of catalyst) 1000 kg
n-C14= 0.0016 Condenser duty -1.287E+10 J/hr
n-C16= 0.0589 Reboiler duty 1.54E+10 J/hr
Reactive
Stream Enthalpy and temperatures:
Zone II
(stages 9-11) stream Temperarure (°C) Enthalpy (J/kmol)
<1> 240 -1.950E+08
<4> <2> 240 -3.943E+08
H2 Feed <4> 260 6.891E+06
(300 Kmol/hr) <5> 225 -7.49E+07
T=260 oC <6> 225 -2.29E+08
<7> 414.6 -1.39E+08

<7> Enhanced
Ultra-Low
Sulfur Diesel

Figure 6.
<4> H2, H2O,
H2S

T=225 oC

R=0.5 <5> Light HC


(350 Kmol/hr)
Column sizing details (Triolein Feed Flow 15 Kmol/h)
Parameter size
Number of theoretical stages 14
HC Feeding stage 9
mole fraction Reactive H2 Feeding stage 12
Th= 0.087 Zone I Triolein Feeding stage 12
BT= 0.087 (stages 5-7) Pressure drop per stage 0 atm
DBT= 0.1 Total Pressure of the column 30 atm
n-C11= 0.4966
Reflux ratio 0.5
n-C12= 0.3166
n-C13= 0.0089
Number of reactive zones 2
n-C14= 0.0016 Number reactive stages per zone 3
Petro-Diesel n-C16= 0.0589 Number of nonreactive stages 8
(100 Kmol/hr) Holdup per reactive stage (kg of catalyst) 1000 kg
<1> Condenser duty -2.247E+10 J/hr
Reboiler duty 9.242E+09 J/hr
Reactive
Zone II Stream Enthalpy and temperatures:
H2 Feed (stages 9-11)
stream Temperarure (°C) Enthalpy (J/kmol)
(300 Kmol/hr) <2> <1> 240 -1.950E+08
<2> 260 6.891E+06
<3> 400 -1.870E+09
Triolein
or
<3> <4> 215 -1.07E+08
<5> 215 -2.40E+08
Wasted- <6> 414.6 -1.44E+08
Vegetal
Oil
<6> Enhanced
Ultra-Low
Sulfur Diesel

Figure 7.
475
(a) Reactive Reactive
450 Zone I Zone II

425

400
Temperature ( C)

375
o

350

325
Triolein Feed
300 Flow
5.4 Km ol/hr
275 5 k m ol/hr
4 Km ol/hr
250 3 Km ol/hr
2 Km ol/hr
225 1 Km ol/hr
No Triolein
200
0.55 (b) 5.4 Km ol/hr
5.0 Km ol/hr
0.50 4.0 Km ol/hr
X n-C18 ( (Liquid mole fraction)

3.0 Km ol/hr
0.45 2.0 Km ol/hr
1.0 Km ol/hr
0.40

0.35

0.30

0.25

0.20

0.15

0.10

0.05

0.00

(c)

0.06
XDBT (Liquid mole fraction)

1 Km ol/hr
2 Km ol/hr
3 Km ol/hr
0.04 4 Km ol/hr
5 Km ol/hr
5.4 Km ol/hr

0.02

0.00

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Stages

Figure 8.
440
(a) Reactive Reactive
420 Zone I Zone II
400

380

360
Temperature ( C)
o

340

320

300

280 Triolein feed Flow


at stage 12
260 1 Kmol/hr
240 5 Kmol/hr
10 Kmol/hr
220 15 Kmol/hr
200
(b)
0,4
Xn-C18 (Liquid mole fraction)

0,3
Triolein feed Flow
at stage 12
0,2 1 Kmol/hr
5 Kmol/hr
10 Kmol/hr
15 Kmol/hr
0,1

0,0

0,018
(c) Triolein feed Flow
0,016 at stage 12
1 Kmol/hr
0,014
5 Kmol/hr
XDBT (Liquid mole fraction)

0,012 10 Kmol/hr
15 Kmol/hr
0,010

0,008

0,006

0,004

0,002

0,000

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Stages

Figure 9.
314
(a)
312
H2 Feed Flow on stage 12 (Kmol/hr)

310

308

306

304

302

300

298
1 2 3 3.5 4 4.5 5 5.4
Triolein Feed Flow at stage 9 (Kmol/hr)
450
440 (b)
430
Flow rate of H2 on stage 12 (Kmol/hr)

420
410
400
390
380
370
360
350
340
330
320
H2 Feed on stage 12 (T=533 K)
310
Triolein Feed at stage 12 (T=673.2 K)
300
290
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Triolein Feed Flow at stage 12 (Kmol/hr)

Figure 10.
0,5
Triolein Feed Flow at stage 9
(a) (5 Kmol/hr)
n-C11
0,4
n-C18
X (Liquid mole fraction)

0,3
n-C12

0,2

0,1
n-C16
n-C13
0,0
n-C14

0,6 (b)
Triolein Feed Flow at stage 12
n-C11
(5 Kmol/hr)
0,5
X (Liquid mole fraction)

0,4

n-C12
0,3
n-C18

0,2

0,1 n-C16

n-C13
0,0
n-C14

(c) n-C11 Triolein Feed Flow at stage 12


0,5 (10 Kmol/hr)

0,4
X (Liquid mole fraction)

n-C18

0,3

n-C12
0,2

BPH
0,1
n-C16
n-C13
0,0
n-C14
Reactive Reactive
Zone I Zone II

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Stages

Figure 11.
0,07
(a) Triolein Feed Flow
0,06 at stage 9
1 Kmol/hr
2 Kmol/hr
XH2O (Liquid mole fraction)

0,05
3 Kmol/hr
4 Kmol/hr
0,04 5 Kmol/hr
5.4 Kmol/hr
0,03

0,02

0,01

0,00

(b)
0,14 Triolein feed Flow
at stage 12
0,12
1 Kmol/hr
XH2O (Liquid mole fraction)

5 Kmol/hr
0,10 10 Kmol/hr
15 Kmol/hr
0,08

0,06

0,04

0,02

0,00

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Stages

Figure 12.
450
(a) Reactive Reactive
425
Zone I Zone II
400

375

350
Temperature ( C)
o

325

300
Oleic Acid feed flow
275 at stage 9
250 75 Kmol/hr
70 kmol/hr
225 50 Kmol/hr
20 Kmol/hr
200
10 Kmol/hr
175 1 Kmol/hr

150
(b)
0.4 Oleic Acid Feed Flow
at stage 9
Xn-C18 (Liquid mole fraction)

1 Kmol/hr
0.3 10 Kmol/hr
20 Kmol/hr
50 Kmol/hr
70 Kmol/hr
0.2

0.1

0.0

0.014
(c) Oleic Acid Feed Flow
at stage 9
0.012
1 Kmol/hr
10 Kmol/hr
XDBT (Liquid mole fraction)

0.010
20 Kmol/hr
50 Kmol/hr
0.008 70 Kmol/hr

0.006

0.004

0.002

0.000

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Stages

Figure 13.
550

o
TCond = 185 C
H2 Feed Flow on stage 12 (Kmol/hr)

500

450

400 TCond = 215 oC

350

300

TCond = 225 oC
250
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80

Oleic Acid Feed Flow at stage 9 (Kmol/hr)

Figure 14.
0,7
(a)
Olec Acid Feed Flow at stage 9
0,6 n-C11
(50 Kmol/hr)

0,5
X (Liquid Mole Fraction)

n-C18

0,4

0,3
n-C12
0,2

0,1 BPH n-C16

n-C13
0,0
n-C14

0,35
(b)
Oleic Acid feed Flow at stage 9
0,30
1 Kmol/hr
10 Kmol/hr
XH O (Liquid mole fraction)

0,25 20Kmol/hr
50 Kmol/hr
70 Kmol/hr
0,20
75 Kmol/hr

0,15

0,10
2

0,05

0,00

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Stages

Figure 15.

Anda mungkin juga menyukai