Report TKHT
Report TKHT
Member List
Full name Student ID
Nguyễn Ái Minh Châu 2252089
Huỳnh Ngọc Hân 2252198
Đinh Nguyễn Phương Nghi 2252515
Nguyễn Ngọc Như Quỳnh 2252704
Trương Hữu Tài 2252729
TABLE OF CONTENTS
1. INTRODUCTION.....................................................................................................................2
1.1. Type of Reaction: Heterogeneous – Fluid-Fluid Reaction.....................................2
1.2. Reaction Equipment................................................................................................4
1.3. Reactions: Mechanism, Reaction Types, and Kinetics...........................................4
2. KINETICS.................................................................................................................................6
2.1. Study of chemical kinetics of urea synthesis from CO2 absorption in aqueous NH3
solutions 6
2.2. Development of the rate expression model for urea production from extent of reaction. .10
2.3. Determination of rate constants and equilibrium constants...............................................14
3. REACTORS............................................................................................................................18
3.1. Descriptions of reactors.....................................................................................................18
3.2. Determination of the optimal kinetic parameters from the application of optimization
model 18
3.3. Development of the transient mole fraction and temperature models of the urea reactors19
3.4. Input parameters for simulation.........................................................................................22
3.5. Results and discussion.......................................................................................................23
1
1. INTRODUCTION
There are several commercial processes for the production of urea, generally
classified into once-through, partial recycle, and total recycle processes. Among them, the
total recycle process is the most widely adopted in industry, where all unconverted ammonia
is recycled back into the reactor, achieving an overall ammonia conversion efficiency of
approximately 99%.
2
distinguishes them fundamentally from homogeneous reactions, where all reactants are in a
single phase—either all gases, all liquids, or all solids dissolved uniformly in a solution—
and thus react throughout the entire volume without phase boundaries.
Apart from urea synthesis, heterogeneous fluid-fluid reactions are extensively used in
various industrial fields. In petroleum refining, processes such as hydrocracking involve
gas-liquid reactions where hydrogen gas interacts with liquid hydrocarbons in the presence
of a solid catalyst. Catalytic converters in automobile exhaust systems are another example,
where gas-phase pollutants are converted into less harmful substances through reactions at
solid surfaces, although this includes both gas-solid and gas-liquid interactions. In
environmental engineering, gas-liquid scrubbers remove acidic gases such as SO₂ or CO₂
by absorbing them into alkaline liquid solutions, another typical example of heterogeneous
fluid-fluid chemistry.
3
equipment such as packed towers, tray columns, or stirred-tank reactors with enhanced
mixing to ensure sufficient phase contact. In contrast, homogeneous reactions generally
proceed faster due to uniform mixing and are easier to model kinetically, but they often
suffer from more complex downstream separation and recycling processes.
In chemical engineering, the Continuously Stirred Tank Reactor (CSTR) and the
Plug Flow Reactor (PFR) are two essential reactor models. A CSTR maintains perfect
mixing, ensuring uniform composition and temperature throughout the reactor. Reactants
are continuously fed in and products removed at the same rate. While easy to operate and
control, a single CSTR often requires a large volume to achieve high conversion, especially
for first-order reactions.
In contrast, CSTRs in parallel divide the flow among multiple identical units
operating simultaneously. While this setup offers operational flexibility and redundancy, it
does not improve conversion efficiency like a series configuration does. Compared to both,
PFRs still offer the best conversion for many reactions, though they are less flexible and
more complex to control.
The overall urea production process consists of three principal chemical reactions:
4
(1) Formation of Ammonium Carbamate (Heterogeneous Reaction):
Mechanism: Fast, exothermic reaction between gaseous ammonia and carbon dioxide
forming a liquid phase product.
Equilibrium: Assumed to reach equilibrium at reactor outlet conditions due to high
residence times.
Kinetics: Reaction rate is fast enough to be treated as an equilibrium process in
modeling.
5
Mechanism: Slow and endothermic, occurring under conditions of high temperature,
high urea concentration, and low ammonia concentration.
Reaction Type: Homogeneous.
Kinetics: The formation of biuret is an undesired side reaction, and its kinetics are
crucial for process optimization to maintain biuret concentration at acceptable levels
(typically less than 1%).
2. KINETICS
The approach used to replicate the experimental outcomes is based on the classical
heterogeneous model, which assumes isothermal and isobaric conditions in line with the
experimental setup. Driven by the concentration gradient, carbon dioxide diffuses from the
gas bubbles into the liquid phase. In order to model this phenomenon, the two films theory
is applied. Rigorously, the reactive absorption process should incorporate both the film and
bulk regions in each phase, along with the corresponding mass balance equations. Thus, the
dynamic variation of the species concentrations results in a system of partial differential
equations that takes into account the interface flow rate and the reactions taking place in the
liquid phase. This model must be complemented with the relevant boundary and initial
conditions, related to bulk concentrations, the interface equilibrium and the molar transfer
rate taking place at the gas/ liquid interface.
6
Fig 1: Representation of the mass transfer process
where (CCO2) is the carbon dioxide concentration into the liquid bulk and (C CO2* )
represents its equilibrium concentration while the surface where mass transfer occurs is
indicated by S.
7
To accurately describe the capture process, ammonia losses should also be
considered and modelled via eq. (2):
where P is the total pressure of the capture system, y i is the molar fraction of the ith
species within the gas phase and Hij is the Henry constant for the i th species in the jth
solvent.
The Henry constant of a specific gas is strictly dependent on solvent properties and
system temperature. Eqs. (4) and (5) show the correlations adopted to evaluate the Henry
constant of CO2 (HCO2H2O) and NH3 (HNH3H2O) in water, respectively as a function of
temperature:
8
approach proposed in the literature is taken into account where the Hatta number, Ha (eq.
(9)), and the enhancement factor infinite, E inf(eq. (10)), are obtained to firstly determine the
reactive regime occurring:
where kapp is the apparent kinetic constant of the reactive phenomena that determine
the CO2 consumption into the liquid phase , Di is the diffusion coefficient of the ith species
and NH3 is the ammonia stoichiometric coefficient in the reaction involving CO 2
consumption. Under the experimental conditions considered in this study, two different
regimes are possible, i.e. the pseudo first order one that occurs when ammonia is in excess
compared to CO2 and the fast intermediate regime that takes place when CO 2 accumulates in
the liquid phase. When the pseudo-first-order condition is met (3 < Ha≪Einf), the
enhancement factor can be computed through eq. (11):
As the process evolves, the absorption changes into the fast intermediate regime,
during which the enhancement factor can be approximate by the eq. (12):
9
2.2. Development of the rate expression model for urea production from extent of
reaction
The synthesis of urea involves an equilibrium reaction where NH 3 and CO2 combine
to form an intermediate compound known as ammonium carbamate. This compound
gradually dissociates, leading to the formation of urea and water in the solution. The rate
model is described as a function of the reaction extent (reaction coordinates), based on the
mass action law and thermodynamic models, which incorporate the van't Hoff equation and
fugacity functions to express the individual component rates for urea synthesis. The
synthesis reaction proceeds according to an equilibrium process
where, k1 is the rate constant for the forward reaction, k 2 is the rate constant for the
backward reaction, and k3 is the rate of reaction for the dissociation of ammonium
carbamate to yield urea and water.
Reaction (1) represents a consecutive reaction that is split into exothermic and
endothermic parts, which are detailed in reactions (2a) and (2b), respectively
It has been found that the rate of the formation of ammonium carbamate (reaction 2a)
is much faster than the rate of its dehydration (reaction 2b). In the first stage we have a rapid
synthesis process of NH4COONH2, the concentration of which in a very short time reaches
its maximum value, whereas the concentration of (NH4)2CO and H2O grow rather slowly. In
the second stage, the synthesis of (NH4)2CO takes place, which leads to a decrease of the
concentration of NH4COONH2 to an equilibrium value, whereas the molar concentration of
10
NH3 and CO2 do not undergo any essential changes. The first stage is the shorter the
quicker the reaction (2a) in relation to reaction (2b).
As the result, the overall reaction rate was investigated through stage II, where the
total change in heat of reaction equalled - 33 kcal/mol, derived from the Individual heat of
reactions of Hr = -38.5kcal/mol and Hr = + 5kcal/mol in reactions (2a) and (2b),
respectively. The algorithm the development of the rate expression are as follows:
(1) The rates of carbamate and urea formation (reactions 2a and 2b) are expressed in
terms of relative reaction rates, following the principle of mass action, described in
equations (13a) and (13b), respectively
where, (-r1) is the rate of depletion of ammonia, (-r 2) is the rate of depletion of CO2
and (-r3) is the rate of formation of ammonium carbonate.
where (-r3) is the rate of dissociation of ammonium carbamate, (-r4) is the rate of
formation of urea, and (-r5) is the rate of formation of water. Equations (13a) and (13b) are
combined to obtain the overall rate of urea formation as follows:
where (-r3)t is the overall rate of ammonium carbamate formation, (-r3)1 the rate of
formation of ammonium carbamate from the exothermic equilibrium reaction (2a)), and (-
r3)2 is the rate of dissociation of ammonium carbamate from the reaction (2b).
11
(ii) CO2 is the limiting reactant as the reaction stops immediately its amount is
finished within the reactor and expresses the rate of formation of ammonium carbamate and
urea in terms of concentrations of the reactants gives equations (14a) to (14d) as
Applying equation (14c) to equation (13c) gives the rate of urea for-mation in the
reactor as follows:
Expressing equation (15) in terms of Dalton's law of partial pressure and the mole
fraction gives
12
reaction process, and PT is the total pressure of the system [atm]. The partial pressure of the
gas is related to the mole fraction and the total pressure of the system as follows:
where dni,, is the change in the amount of i-component [mol], vi,j the stoichiometric
coefficient of the i-components in the y-coordinate and j is the extent of the reaction in the j-
coordinate [mol].
In terms of the reaction coordinates 1 and 2 the components are assigned digits as,
1(NH3), 2(CO2), 3 (carbamate), 4 (urea), and 5 (water), and the rate stoichiometrically in
terms of the extent of reaction or reaction coordinates is:
Assuming the reaction begins with n1,0 moles of NH3 and n2,0 moles of CO2, the
mathematical expressions for vi's and yi's as functions of 1 and 2 provide the mole
fractions of the various species in the equilibrium reaction process, as summarized in Table
1. Additionally, the rate of urea formation, in terms of the extent of the reaction, is given by
equation (20), respectively.
13
(20)
The procedure outlined below was employed to examine the rate constants and
equilibrium constants involved in the synthesis process.
(i) The equilibrium constant KC is mathematically expressed as the ratio of the forward
reaction rate constant k1 to the reverse reaction rate constant k2 for the equilibrium reaction
described in equation (2b).
(ii) The thermodynamic equilibrium constant expressed through the change in standard
Gibbs free energy G0:
Equation (23) gives an expression for the equilibrium constant after manipulation (22) as,
14
where G0 is the change in Gibbs free energy [kJ/mol] and, R is the gas constant for an
ideal gas [J/mol. K] and T is the absolute temperature of the mixture [K].
(iii) The equilibrium constant can be expressed in terms of fugacity, fugacity coefficient,
partial pressure, and the extent of reaction of the involved species.
From a thermodynamic perspective, the equilibrium constant is derived using the fugacity f
and the extent of reaction . Given that the process occurs in the gas phase under elevated
pressures and temperatures, fugacity is applied as a parameter representing high-pressure
conditions in the reacting system. Accordingly, the equilibrium constant is expressed in
terms of fugacity, with units of [atm], as follows:
where f-vi,j is an operator based on the quotient of fugacity product raised to the respective
coefficients v of the j-component to the fugacity of the reactants to the exponent of
coefficient v of the i-component defined clearly as:
where fvj represents the fugacity of the products raised to the power of their
respective stoichiometric coefficients, while fvi denotes the fugacity of the reactants, also
raised to their corresponding stoichiometric values. By applying equations (24) and (25), we
obtain,
where f1 is the fugacity of ammonia, f2 is the fugacity of CO2 and f3 is the fugacity of
ammonium carbamate. Fugacity is measured in units of pressure, typically atmospheres.
The fugacity coefficient is defined as the ratio of the fugacity to the pressure of component
k, and is given by the following expression::
15
Here, represents the fugacity coefficient, f k is the fugacity of component k, P k is the
partial pressure of component k, and k refers to components 1, 2, 3, ..., up to n, as defined
within the reaction system.
(31)
Recognizing that in a high-pressure system the pressure drops due to the formation of
an aqueous product, the fugacity coefficient approaches unity, allowing equation (31) to be
simplified as follows:
16
(32)
(iv) The rate constants, equilibrium constant, and extent of reaction are interconnected to
derive an expression for the reaction coordinate, also known as the extent of reaction.
The extent of reaction is determined by utilizing the relationship between the equilibrium
constant—expressed in terms of Gibbs free energy—and the procedure outlined in
algorithm "ii", through the combination of equations (22) and (32), as follows:
(33)
where 2 =21 = , P=PT is the pressure of the system [atm].
Equation (33) is solved either through the quadratic formula or by using the method of
completing the square to determine the extent of reaction. Once the extent of reaction and
the equilibrium constant are known, the rate constants can be calculated using equations
(21), (22), and (33), following the relationships provided in equations (34) and (35).
(34), (35)
17
3. REACTORS
18
the urea synthesis process, the optimization model is used to the transient reactor models
created in this work to determine the optimal kinetic parameters.
Ei(j+1) = Eij + ⍺∆
Ai(j+1) = Aij + ⍺∆
where :
Ei(j+1), Ai(j+1): the optimal activation energies and Arrhenius constants of i-components in the
equilibrium reaction process based on the convergence criterion set
Eij, Aij: are the current or input values of the activation energies and Arrhenius constants
⍺: a scalar constant ranged from 0.35 ≤ α ≤ 0.95
∆: the incremental change defined mathematically as, ∆ = -inv(JJT).J, in which inv(JJT) is
the inverse of the product of matrix J and its transpose and J is an n x p matrix.
3.3. Development of the transient mole fraction and temperature models of the urea
reactors
Predicting and describing the production process within reactors under different
beginning and boundary circumstances using mathematical expressions derived from
material and energy balance principles applied to the reactors is the fundamental function of
modeling. These models' performance and error analysis determine their acceptance and
dependability.
19
conservation to these urea reactors, models for their mole and temperature profiles were
developed using unsteady-state material and energy balance equations.
This equation is applied to CSTR in sequence for a single mole of reactants entering
the reactor to create urea. Using the rate expression model and expressing the urea process
in terms of mole fraction in an unstable condition:
where yi,j denotes the mole fraction of the i-component in the j-reactor (i.e., i = NH 3, CO2,
Carbamate, urea or water and j = 1, 2 or 3).
Similarly, for one mole of reactants to the reactor, the material balance equation (35)
is applied to PFR in terms of mole fraction and extent of reaction as
20
where yi denotes the mole fraction of the i-component and v is the superficial velocity [m/s].
Similarly to the material balance model, the general energy balance equation used to
develop the temperature unsteady state models for the CSTR in series and PFR, is stated as
The temperature models below for the CSTR in series and PFR respectively, are
obtained by applying the energy balance equation above
21
where
Ti,j: the output temperature of the i-component in the j-reactor [K]
J = (-Hr)Cp [m3K]
T0: the input temperature [K]
1: the space velocity or reciprocal of space time, 1=UACpVR=v0VR
TC: the coolant temperature [K]
where
Ti,j: the output temperature [K]
J = (-Hr)Cp [m3K]
1: the space velocity or reciprocal of space time, 1=UACpVR=v0VR
TC: the coolant temperature [K]
The input conditions used for the simulation and validation of the models were
obtained from Indorama and Notore plants.
22
3.5. Results and discussion
As shown in Fig. 3, these dynamics match with the lowest rate expression result for
urea synthesis, maximizing yields inside the reactors. These results are consistent with
values documented in the literature, suggesting that the rate expression model that was
constructed may be dependable for the process of producing urea from ammonia and CO2.
23
The urea yield in the Indorama reactor, which employs a Continuous Stirred Tank
Reactor (CSTR) configuration, has been evaluated and compared with both plant data and
values reported in the literature. This yield is determined based on a mole fraction model
developed for the reactor, which incorporates a kinetic rate expression. The urea yield is
influenced by the rate of urea formation, reaching a maximum value of 0.726 at a residence
time of 2.5 minutes, corresponding to the point of lowest urea decomposition. Beyond this
point, the yield declines due to the effects of consecutive reactions. When benchmarked
against previously reported values, the modeled yield of 0.726 exceeds the literature-
reported yield of 0.60 and the initially reported plant value of 0.34, and aligns closely with
the more recent plant-reported yield of 0.64.
24
0.53. This elevated yield suggests that the developed kinetic model provides a reliable and
accurate prediction of the urea synthesis equilibrium. The modeling approach employed
offers a more precise representation of the process compared to existing methodologies
reported in plant operations and the literature, making it a valuable tool for both industrial
application and academic research. Although the Indorama reactor demonstrates a slightly
higher urea yield than the Notore reactor, the results are closely aligned, with a small
absolute deviation of 0.032 between the two.
25
The presents refined values of activation energies, Arrhenius constants, and rate
constants. Validation of these kinetic parameters with plant data resulted in error values
ranging from 0.0004 to 0.0491 for both RMSE and deviations, indicating high accuracy,
reliability, and acceptability for control, modeling, and optimization of the reactors to
enhance the description of the urea process and performance of the controller.
V= = 10.66
V= = 7.20
26