A New Phase-Equilibrium Model For Simulating Industrial Nylon-6 Production Trains
A New Phase-Equilibrium Model For Simulating Industrial Nylon-6 Production Trains
Thomas N. Williams
Honeywell International, Inc., 15801 Woods Edge Road, Colonial Heights, Virginia 23834
Chau-Chyun Chen
Aspen Technology, Inc., Ten Canal Park, Cambridge, Massachusetts 02141
This paper presents a thermodynamically consistent model for the phase equilibrium of water/
caprolactam/nylon-6 mixtures, based on the POLYNRTL (polymer nonrandom two-liquid)
activity-coefficient model. Our model predicts phase equilibrium for any binary or ternary mixture
containing water, -caprolactam, and nylon-6 at industrially relevant temperatures and pressures
with an average error of 1%. This paper also demonstrates its application to simulate a melt
train and a bubble-gas kettle train for industrial production of nylon-6. Prior literature makes
simplifying assumptions about liquid-phase (molten polymer) activities of water and -capro-
lactam; these assumptions are shown to be unrealistic.
k2
polycondensation Pm + Pn y\ z Pm+n + W
k ′ 2
k3
polyaddition of caprolactam CL + Pn y\ z Pn+1
k ′ 3
k4
ring opening of cyclic dimer W + CD y\ z P2
k ′ 4
Figure 2. Nylon-6 molecules of degree of polymerization n
k5 existing as two types: unterminated (above) and terminated by
polyaddition of cyclic dimer CD + Pn y\ z Pn+2
k ′ AA (below). B-ACA represents nylon-6 repeat segments, T-NH2
5
represents amine end groups, T-COOH represents carboxylic acid
a P is aminocaproic acid. P is a nylon-6 molecule with a degree
end groups, and T-AA represents AA end groups.
1 n
of polymerization n.
(T-AA). Table 4 presents the reactions in Table 2
expressed using segment notation.
2.1. Nylon-6 Polymerization Kinetics. Arai et al.18 Table 4 also gives the associated reaction rates for
have presented the accepted standard regarding the each equilibrium reaction. Table 5 gives the species
chemistry and kinetics of the hydrolytic polymerization conservation equations corresponding to the reaction
of -caprolactam. Their reaction mechanism includes the rates of Table 4.
ring opening of caprolactam (CL), polycondensation, POLYMERS PLUS estimates the concentrations of
polyaddition of CL, ring opening of a cyclic dimer (CD), oligomers of degree of polymerization 2 and 3 below:
and polyaddition of a CD. Table 2 lists these five
equilibrium reactions.
Table 3 gives the accompanying rate constants ki,
again from Arai et al.18
[P2] ) [T-COOH] ( [T-NH2]
[B-ACA] + [T-NH2] ) (2)
( )
We add one more equilibrium reaction to this kinetic
scheme: the termination reaction with a monofunctional [B-ACA]
[P3] ) [T-COOH]
acid, such as acetic acid (AA). [B-ACA] + [T-NH2]
AA + Pn y\
k
k2′
z Pnx + W
2
(1) ( [T-NH2]
[B-ACA] + [T-NH2] ) (3)
We assume that this reaction follows the same kinetics We compute the number-average degree of polymeri-
as the polycondensation reaction, as in ref 5. zation, DPn, by considering the distribution of polymer
We ignore the analysis of cyclic oligomers higher than chain lengths. This distribution contains m different
dimers for the sake of simplicity. chain lengths, with each chain length i characterized
We use Aspen Technology’s (Cambridge, MA) com- by a population of Ni chains with degree of polymeri-
mercial simulation package POLYMERS PLUS to simu- zation DPi:
late nylon-6 polymerizations. This package implements
m
∑
the above reaction-kinetics model using the segment-
based methodology detailed in the patent by Barrera NiDPi
i)1
et al.19 This methodology tracks the polymer concentra- DPn ) (4)
tion and number-average degree of polymerization by m
tracking the concentrations of their constitutive seg-
ments. Figure 2 shows the segmental breakdown of
∑
i)1
DPi
nylon-6 molecules that can be either unterminated or
terminated by AA. In terms of segments, the numerator of eq 4 represents
Nylon-6 segments include the nylon-6 repeat segment the total concentration of segments that count as repeat
(B-ACA) and the end groups terminal amine (T-NH2), units, or ([B-ACA] + [T-COOH] + [T-NH2] + [P1]). The
terminal carboxylic acid (T-COOH), and terminal AA denominator represents the concentration of polymer
rate constant
expression ki ) A0i exp -( ) E0i
RT ( )
+ Aci exp -
Eci
RT
[T-COOH]
equilibrium constant
expression Ki )
ki
ki′
) exp (
∆Si - ∆Hi/T
R )
i A0i (kg/mol‚s) E0i (J/mol) Aci (kg2/mol2‚s) Eci (J/mol) ∆Hi (J/mol) ∆Si (J/mol‚K)
1 1.66 × 102 8.32 × 104 1.20 × 104 7.87 × 104 8.03 × 103 -33.01
2 5.26 × 106 9.74 × 104 3.37 × 106 8.65 × 104 -2.49 × 104 3.951
3 7.93 × 105 9.56 × 104 4.55 × 106 8.42 × 104 -1.69 × 104 -29.08
4 2.38 × 108 1.76 × 105 6.47 × 108 1.57 × 105 -4.02 × 104 -60.79
5 7.14 × 104 8.92 × 104 8.36 × 105 8.54 × 104 -1.33 × 104 2.439
Ind. Eng. Chem. Res., Vol. 42, No. 17, 2003 3903
( )
k2 [B-ACA]
P1 + T-COOH y\
k ′)k /K
z T-COOH:B-ACA + W R3 ) k2[P1][T-COOH] - k2′[W][T-COOH]
2 2 2 [B-ACA] + [T-NH2]
T-NH2 + P1 y\
k ′)k /K 2
k2
z T-NH2:B-ACA + W
2 2
R4 ) k2[T-NH2][P1] - k2′[W][T-NH2] ( [B-ACA]
[B-ACA] + [T-COOH] )
( )
k2 [B-ACA]
T-NH2 + T-COOH y\
k ′)k /K
z B-ACA:B-ACA + W R5 ) k2[T-NH2][T-COOH] - k2′[W][B-ACA]
2 2 2 [B-ACA] + [T-NH2]
k3
Polyaddition of Caprolactam (CL + Pn y\ z Pn+1)
k ′ 3
k3
P1 + CL y\
k ′)k /K
z T-NH2:T-COOH R6 ) k3[P1][CL] - k3′[P2]
3 3 3
T-NH2 + CL y\
k ′)k /K
z T-NH2:B-ACA
3
k3
3 3
R7 ) k3[T-NH2][CL] - k3′[T-NH2] ( [B-ACA]
[B-ACA] + [T-COOH] )
k4
Ring-Opening of Cyclic Dimer (W + CD y\ z P2 )
k ′
k4 4
CD + W \
yk ′)k /K z T-COOH:T-NH2 R8 ) k4[CD][W] - k4′[P2]
4 4 4
k5
Polyaddition of Cyclic Dimer (CD + Pn y\ z Pn+2)
k ′ 5
k5
P1 + CD y\
k ′)k /K
z T-NH2:B-ACA:T-COOH R9 ) k5[P1][CD] - k5′[P3]
5 5 5
T-NH2 + CD y\
k ′)k /K
z B-ACA:B-ACA:T-NH2
5
k5
5 5
R10 ) k5[T-NH2][CD] - k5′[T-NH2] ( [B-ACA]
[B-ACA] + [T-COOH] )
k3
Polycondensation of Acetic Acid (Pn + AA y\ z Pn,T-AA)
k ′ 3
P1 + AA y\
k ′)k /K
2
k2
z T-AA:T-COOH + W
2 2
R11 ) k2[AA][P1] - k2′[W][T-AA] ( [T-COOH]
[T-COOH] + [B-ACA] )
T-NH2 + AA y\
k ′)k /K
z B-ACA:T-AA + W
2
k2
2 2
R12 ) k2[AA][T-NH2] - k2′[W][T-AA] ( [B-ACA]
[T-COOH] + [B-ACA] )
a A colon represents a covalent bond between segments.
Table 5. Species Conservation Equations for the lymerization, nor do we compute higher moments of the
Reaction Rates in Table 4 distribution of the DP.
functional 2.2. Polymer Equilibrium. 2.2.1. Previous At-
group time rate of change tempts at Developing a Phase-Equilibrium Model
W d[W]/dt ) R2 + R3 + R4 + R5 + R11 + R12 - (R1 + R8) for Water/Caprolactam/Nylon-6. Researchers typi-
CL d[CL]/dt ) -(R1 + R6 + R7) cally use one of three phase-equilibrium models in
CD d[CD]/dt ) -(R8 + R9 + R10) simulating nylon-6 polymerizations: the Jacobs-Sch-
AA d[AA]/dt ) -(R11 + R12) weigman model,20 the Fukumoto model,21 or the Tai et
P1 d[P1]/dt ) R1 - (2R2 + R3 + R4 + R6 + R9 + R11) al. model.12 These models only predict water concentra-
B-ACA d[B-ACA]/dt ) R3 + R4 + 2R5 + R7 + R9 + 2R10 + R12
T-NH2 d[T-NH2]/dt ) R2 + R6 + R8 + R9 - (R5 + R12) tions in reacting mixtures of water, caprolactam, and
T-COOH d[T-COOH]/dt ) R2 + R6 + R8 + R9 + R11 - (R5) nylon-6.
T-AA d[T-AA]/dt ) R11 + R12 The Jacobs-Schweigman model20 is the most sim-
plistic of the three models. It consists of an empiricism
chains, or ([T-COOH] + [P1]). Therefore, we rewrite the based on experimental equilibrium data for a VK
number-average degree of polymerization below: (Vereinfacht Kontinuierliches) tube reactor. The model
predicts the concentration of water [W] as a function of
[B-ACA] + [T-NH2] + [T-COOH] + [P1] temperature T:
DPn ) (5)
[T-COOH] + [P1]
1.76 - 0.006T (°C)
[W] (mol/kg) ) (6)
We do not compute the weight-average degree of po- 1.8
3904 Ind. Eng. Chem. Res., Vol. 42, No. 17, 2003
Table 6. Liquid-Phase, Water Mole-Fraction Predictions 2. POLYNRTL takes advantage of the already exist-
at 1 atm for the Jacobs-Schweigman,20 Fukumoto,21 and ing database of NRTL binary interaction parameters.
Tai et al.12 Phase-Equilibrium Models
On the basis of these suggestions, we choose to model
temp (K) Jacobs-Schweigman Fukumoto Tai et al. nylon-6 phase equilibrium using the POLYNRTL prop-
473 0.006 0.017 0.103 erty method. This method uses (1) the polymer NRTL
493 0.004 0.008 0.050 activity-coefficient model for the liquid phases, (2) the
513 0.003 0.004 0.025 Redlich-Kwong26 equation of state for the vapor phase,
533 0.002 0.002 0.013 (3) the van Krevelen27 model for the liquid properties
(enthalpy, entropy, Gibbs free energy, heat capacity, and
Because this expression does not contain pressure as a molar volume), and (4) Henry’s law for any supercritical
free variable and was developed using VK tube equi- components.
librium data, it should not be used at pressures deviat- The POLYNRTL activity-coefficient model combines
ing from atmospheric pressure. the traditional NRTL model with the FH description for
Adding the system pressure P to the Jacobs-Sch- configurational entropy. It essentially calculates the
weigman model results in the Fukumoto model21 based Gibbs free energy of mixing in a polymer solution as
on vapor-liquid equilibrium (VLE) data: the sum of two contributions: (1) the entropy of mixing
( )
8220 from the FH activity-coefficient model; (2) the enthalpy
[W] (mol/kg) ) P (mmHg) exp - 24.0734 (7) of mixing from the NRTL activity-coefficient model.28
T (K) These activity coefficients include binary parameters to
The third model is the Tai et al. model,12 which model the interactions between two components. POLYN-
explicitly considers the partial pressures of volatile RTL represents the temperature dependence of the
components. It consists of a system of three equations binary interaction parameters (τij) by eq 11. We set the
to solve for the mole fraction of water, xW, given the
bij
temperature T and total pressure P: τij ) aij + + cij ln T (11)
T
log
[ xW
PW (mmHg) ] )
3570
T (K)
- 11.41 (8) nonrandomness factor Rij to 0.3, as suggested by Praus-
nitz et al.29
Therefore, we model binary interactions in equilibri-
4100
log[PCL (mmHg)] ) - + 9.6 (9) um mixtures by specifying the coefficients aij through
T (K) cij. We typically regress equilibrium data or use a
predictive model, such as UNIFAC, to obtain the values
P ) PW + PCL (10) of these parameters.
2.3. Equilibrium Data for Water/Caprolactam/
PW is the partial pressure of water, and PCL is the Nylon-6. There are two sources of phase-equilibrium
partial pressure of caprolactam. data that we use for regression. The first source
All three models are empirical. Therefore, we cannot characterizes the binary interactions between water and
be confident in their phase-equilibrium predictions at caprolactam. The second characterizes the interactions
conditions that deviate from those in which they are in the reactive, ternary system nylon-6/caprolactam/
based. water. All of the data sets appear in the Supporting
Significantly, these three models give conflicting Information.
predictions for identical process conditions. Consider, Maczinger and Tettamanti30 give four sets of low-
for example, the water liquid-mole-fraction predictions pressure, isobaric phase-equilibrium data for the binary
at atmospheric pressure for a mixture of water, capro- water/caprolactam. Tables 15-18 in the Supporting
lactam, and nylon-6 at a range of relevant processing Information show their data.
temperatures of 473-523 K (Table 6). Giori and Hayes31 present one set of isothermal
The Fukumoto model21 predicts mole fractions that phase-equilibrium data for the ternary system water/
are about twice those of the Jacobs-Schweigman20 caprolactam/nylon-6. They carry out nylon-6 polymer-
predictions. Furthermore, the Tai et al. model12 predicts izations at 543 K in a laboratory reactor and charac-
mole fractions that are an order of magnitude higher terize the resulting phase equilibrium. Table 19 in the
than both models. This difference casts considerable Supporting Information shows their experimental data.
doubt on the validity of at least two of these models.
2.2.2. Polymer Nonrandom Two-Liquid (POLYN- 3. Methodology
RTL) Model. Bokis et al.16 give an excellent description
of how to choose an appropriate thermodynamic equi- We first characterize the binary interactions between
librium model when simulating polymer processes. They water and caprolactam using the first four data sets30
suggest using activity-coefficient models, instead of (Tables 15-18 in the Supporting Information).
equations of state, for processes that involve low-to- We then simulate the Giori and Hayes experiments31
moderate pressures (pressure < 1 × 106 Pa) and/or to quantify the binary interactions between water/
nonideal components (e.g., polar compounds such as nylon-6 and caprolactam/nylon-6.
alcohols, water, and ketones). 3.1. Characterizing the Phase Equilibrium of
Furthermore, they state that the POLYNRTL activ- Water/Caprolactam. Here, we explain how to obtain
ity-coefficient model22 has the following advantages over the binary parameters for water/caprolactam. Our
the Flory-Huggins (FH)23,24 and universal quasi-chemi- handling of the data consists of testing the data for
cal functional group (UNIFAC)25 activity-coefficient thermodynamic consistency. We use two consistency
models: tests for our isobaric data sets: a Van Ness et al. test32
1. POLYNRTL covers large ranges of temperatures and a Wisniak test.33 The Wisniak test can be used in
and compositions accurately. a point or area mode; however, we use only the point
Ind. Eng. Chem. Res., Vol. 42, No. 17, 2003 3905
test because of the difficulty in accurately of computing mentally based and preferable to a combined Redlich-
integrals of “functions” of our experimental data. Kister34/Herington method35 for testing the consistency
3.1.1. Consistency Test One: Van Ness et al.32 In of isobaric phase-equilibrium data.
this test, we first obtain a redundant data set comprised For each data point i at system temperature Ti, we
of temperature, pressure, liquid-mole-fraction, and compute two functions Li and Wi:
vapor-mole-fraction measurements (T-P-x-y). We then
regress only the T-P-x data. After regressing these
data, we are able to predict the original pressure data Li )
∑T0kxk∆s0k - T (17)
i
∆s
(using T-x data) and vapor-mole-fraction data (using
(∑ )
T-P-x data). Consistency test one comes by examining
( xk ln γk) - w
the extent of corroboration between the measured and Wi ) RTi (18)
predicted vapor mole fractions. This test is an indirect ∆s
application of the Gibbs-Duhem equation because the
POLYNRTL model obeys the Gibbs-Duhem equation. For every species k, at the system pressure, there is a
The fundamental phase-equilibrium relation for this pure-component boiling point T0k, a liquid mole fraction
analysis is xk, and entropy of vaporization ∆s0k. In eq 18, ∆s is the
entropy of mixing of the mixture.
yiφiP ) xiγi(xi,T) Psat
i (T) exp [
vi
RT
(P - Psat
i ) ] (12)
We compute the boiling temperature T0k for each
component k by setting the pressure to the system
pressure in eq 13 and then backing out the temperature.
We compute the activity coefficient γi using the POLYN- We calculate the entropy of vaporization from the
RTL model; it is a function of the liquid composition xi enthalpy of vaporization ∆h0k:
and temperature T. Psat
( )
i is the vapor pressure, which
B+C(T0k/Tc,k)+D(T0k/Tc,k)2
we compute using an Antoine form: T0k
Ak 1 -
Psat
i ( Bi
) exp Ai + + Ci ln T + DiTEi
T ) (13) ∆s0k )
∆h0k
T0k
)
Tc,k
T0k
(19)
We have neglected vapor-phase nonidealities (fugacity The critical temperature Tc,k as well as the constants
coefficient φi equals 1) because most of the data are at Ak through Dk are tabulated in standard reference
vacuum conditions; however, one can easily estimate the sources (such as Daubert and Danner36) and differ for
degree of vapor-phase nonideality occurring at high each chemical species.
pressures using the Redlich-Kwong equation of state. We find the activity coefficients from the experimental
Furthermore, the Poynting pressure correction is neg- data using eq 12:
ligible (the exponential term is 1).
We eliminate the need for vapor-phase composition γi ) yiP/xiPsat
i (T) (20)
data by summing eq 12 over all species; for a binary
system containing species i and j, we have We compute the mixture entropy of vaporization for the
mixture ∆s using the following mixing rule:
P ) xiγi(xi,T) Psat
i (T) + xjγj(xj,T) Psat
j (T) (14)
( )
We perform the point test by computing the ratio of
Pobsd - Pcalcd 2
∑k
k k Li and Wi and plotting this as a function of liquid mole
SSE ) (15) fraction. Inconsistent data contain ratio values that are
Pobsd
k much different than 1 and do not scatter randomly
about 1.
Once we have the binary interaction parameters, we 3.1.3. Determining Binary Interaction Param-
predict the vapor composition using the original T-P-x eters from Phase-Equilibrium Data. The Van Ness
data and eq 12. Last, we compute the percent difference et al. test32 effectively regresses binary interaction
in vapor-mole-fraction data vs prediction for each liquid- parameters using T-P-x data only. If the data are
composition data point: judged inconsistent, Prausnitz et al.37 suggest that we
( )
assume that the vapor-phase composition data are
ycalcd - yobsd incorrect for two reasons:
percent deviation ) 100 (16)
yobsd (i) Accurately measuring y is typically more difficult
than measuring x and pressure.
We plot the percent deviation vs the liquid mole fraction (ii) Composition data at the ends of the composition
to characterize the consistency of the observed phase- scale (y ∼ 0 or 1) are likely to be the least accurate.
equilibrium data. If the data are inconsistent and we throw out the
3.1.2. Consistency Test Two: Wisniak.33 The vapor-mole-fraction measurements, then we have al-
second consistency test that we use, due to Wisniak, is ready obtained binary interaction parameters using the
based on the bubble-point equation for mixtures and data that we are most confident in (T-P-x data).
utilizes the Clausius-Clapeyron equation. It is funda- However, if the data are consistent, then we can use
3906 Ind. Eng. Chem. Res., Vol. 42, No. 17, 2003
Table 8. Regression Results for POLYNRTL Binary Table 9. Comparison of Model Predictions with Giori
Interaction Parameters for Water/Caprolactam (Based and Hayes Polymerization Data31
on T-P-x Data Only)
caprolactam liquid water liquid water vapor total pressure
POLYNRTL binary value (temperature mole fraction mole fraction mole fraction (Pa)
interaction parameter units are K)
Data
awater/CL -0.313 0.096 0.012 0.74 571 000
aCL/water 0.628 0.097 0.0249 0.85 974 000
bwater/CL -15.4 0.099 0.0387 0.886 1 400 000
bCL/water -13.7 0.101 0.0515 0.89 1 830 000
cwater/CL 0.0495 0.106 0.065 0.91 2 160 000
cCL/water -0.0898 Model Prediction
0.087 0.012 0.69 537 000
0.086 0.025 0.84 1 010 000
0.086 0.039 0.90 1 440 000
0.085 0.052 0.93 1 830 000
0.085 0.066 0.95 2 180 000
% Error
-9.38 0.00 -6.76 -6.02
-11.34 0.40 -1.18 4.02
-13.13 0.78 1.58 3.25
-15.84 0.97 4.49 0.28
-19.81 1.54 4.40 0.99
Table 10. Predictions for Liquid-Phase Water Mole Fractions of Previous Literature Models and the Ideal Model and
Comparison with Experimental Data of Giori and Hayes31
ideal (fugacity and
liquid-phase water activity coefficients
mole-fraction data Jacobs-Schweigman20 Fukumoto21 Tai et al.12 set to 1)
0.0119 0.0014 0.0101 0.0607 0.0129
0.0249 0.0014 0.0181 0.1100 0.0269
0.0387 0.0014 0.0248 0.1509 0.0411
0.0515 0.0014 0.0323 0.1974 0.0554
0.0650 0.0014 0.0382 0.2338 0.0698
× 10-1 parts semiterminated nylon-6. The molecular 6.1. Melt Train. We simulate a typical commercial
weight of the incoming polymer is 9.8 kg/mol. This melt train involving four reactors.10 The first two
reactor is operating at ca. 523 K and 4.67 kPa with a reactors carry out monomer conversion; the third and
residence time of 1740 s.10 fourth reactors carry out devolatilization and molecular
Our model predicts a caprolactam mass fraction in weight buildup. Figure 8 shows the process-flow dia-
the exit stream of 1.5%, while plant data show 2.7% for gram for this train.
similar operating conditions. This is an underprediction The makeup stream contains 5.35 × 10-5 kg/s of
by 44%. This again shows that our phase-equilibrium water, 0.0120 kg/s of caprolactam, and 2.01 × 10-5 kg/s
calculations are in the right direction considering mass- of AA terminator. We have assumed a production rate
transfer limitations. of about 0.0112 kg/s of polymer or 89 pounds per hour
Regarding the wiped-wall evaporator studies, one may (pph) and have calculated approximate flow rates for
ask the following valid question: may we approximate water and terminator based on refs 10 and 41-44.
the error of the phase-equilibrium model from these The first vessel is a continuous stirred-tank reactor
results, as in the condenser study? Unfortunately, this (CSTR), STAGE-1 in Figure 8. The residence time is
is not possible because we do not know to what extent 8100 s, the temperature is 488 K, and the pressure is
mass-transfer limitations are affecting the reactor 545 kPa.
performance. In other words, if there is no mass-transfer The second vessel is a plug-flow reactor (PFR),
limitation, the phase-equilibrium model is in error by STAGE-2 in Figure 8. The feed to this vessel enters a
-62 to -44%. If there is a strong mass-transfer limita- vapor headspace, which we simulate using a flash unit
tion, then the phase-equilibrium model may have little at reactor conditions (HD-SPACE in Figure 8). Vapor
or no error. from this flash enters the reflux condenser (RFLX-CND
in Figure 8). Most of the caprolactam is returned back
6. Simulating Integrated Industrial Nyon-6 to the reactor headspace, while a mixture of caprolactam
Production Trains and water goes to the train condenser (CONDENSE in
Figure 8). The reactor has a residence time of 28 800 s,
Previous attempts at nylon-6 integrated process mod- a temperature of 498 K, and a pressure of 66.7 kPa.
eling were hindered by the lack of a fundamental
The third reactor is a vapor-liquid plug-flow reactor,
thermodynamic phase-equilibrium model. For example,
EVAPORAT in Figure 8. The vapor phase is removed
Nagasubramanian and Reimschuessel39 simulated the
via a vacuum system (EVAP-FL in Figure 8) and enters
molecular weight buildup in the finishing stage by first
the train condenser CONDENSE. This reactor has a
“instantaneously” removing all water from the polym-
residence time of 1200 s, a temperature of 523 K, and a
erization mass while ignoring caprolactam devolatiliza-
pressure of 2.33 kPa.
tion. There is no corresponding unit operation that can
perform this separation in an actual plant. Further- The fourth reactor is a vapor-liquid plug-flow reactor,
more, this assumption has led Tirrell et al.40 to an FINISHER in Figure 8. The vapor phase is removed via
unusual conclusion: that it is preferable to perform a vacuum system (FIN-FL in Figure 8) and also enters
molecular weight buildup and monomer conversion in the train condenser CONDENSE. We used a sweep-
a single reactor. To have adequate molecular weight steam mass flow rate of ca. one-quarter of that of the
buildup, we need to enforce low water concentrations. polymer flow rate entering the reactor. This reactor has
However, to do this, we typically use severe operating a residence time of 7200 s, a temperature of 536 K, and
conditions, such as high vacuum in melt processes,10 to a pressure of 4 kPa.
remove most of the water. In doing so, we cannot avoid The train condenser CONDENSE recovers the unre-
losing a significant amount of the caprolactam to the acted monomer and sends it back to the inlet of the
vapor phase. Therefore, equilibrium thermodynamics train. Uncondensed species, mostly water, leave in a
clearly suggests that it is impractical to try to obtain waste stream. The condenser operates at an assumed
both molecular weight buildup and monomer conversion pressure of 101.325 kPa and an assumed temperature
in the same reactor. of 403 K.
Assuming an instantaneous and complete water Table 11 summarizes the vessel operating conditions.
removal was necessary in the past because researchers We model mass-transfer limitations in the evaporator
did not have access to a fundamental thermodynamic and finisher reactors using two-film diffusion theory.45
model. However, with our new phase-equilibrium model, The diffusion coefficient is relatively high for water in
we are able to simulate any unit operation that involves nylon-6 melts,38 and we assume that the evaporator and
caprolactam, water, and nylon-6 at multiphase condi- finisher reactors generate a large interfacial surface
tions. These include flash units and condensers. There- area. Therefore, we set the water mass-transfer coef-
fore, we simulate multiphase reactors with condensers ficient to 1 × 10-2 m/s and set the mass-transfer
that return unreacted monomer back to the inlet of the coefficient for caprolactam to 5 × 10-4 m/s.
train. In particular, we simulate two commercial tech- Table 12 shows the predicted mass flow rates for
nologies: a melt train and a bubble-gas kettle train. All water, caprolactam, cyclic dimer, and nylon-6 exiting
details such as the unit operation configuration, feed each reactor. Furthermore, the table reports number-
conditions, and operating conditions are available in the average molecular weight predictions (Mn).
patent literature. We see that all of the monomer conversion (about 81%
This section is not concerned with developing a in this case) takes place in the first two reactors.
validated simulation model for these processes; this However, because of elevated moisture levels, the mo-
endeavor is well beyond the scope of this paper. There- lecular weight only grows to about 10.7 kg/mol. In the
fore, we do not make detailed comparisons between third and fourth reactors, we remove nearly all of the
model predictions and plant data. We wish only to water and double the polymer molecular weight. Our
illustrate the utility of our phase-equilibrium model in final molten polymer product has a flow rate of 1.16 ×
simulating entire manufacturing trains. 10-2 kg/s, the polymer has a number-average molecular
3910 Ind. Eng. Chem. Res., Vol. 42, No. 17, 2003
Figure 8. Four-reactor nylon-6 manufacturing process. Most of the monomer conversion takes place in the first two reactors, while the
molecular weight buildup occurs in the third and fourth reactors. The second vessel has a reflux condenser, while the entire train has a
condenser to recycle the unreacted monomer.
Table 11. Operating Conditions for Each Reactor in the of monomer is devolatilized in the third and fourth
Four-Reactor Train Depicted in Figure 8 reactors (36% of incoming caprolactam). Previous analy-
residence ses, such as those by Nagasubramanian and Reims-
unit operation temp (K) pressure (kPa) time (s) chuessel39 and by Tirrell et al.,40 would assume that no
first reactor 488 545 8100 water is lost in the first two vessels (approximated
second reactor 498 66.7 28800 roughly by one vessel in their studies) and no caprolac-
reflux condenser 403 66.7 tam is lost in the third and fourth vessels (approximated
third reactor 523 2.33 1200 roughly by one vessel in their studies). Our simulation,
fourth reactor 536 4.00 7200
train condenser 403 101.325
which contains phase-equilibrium and mass-transfer
models, shows that these are poor assumptions for a
Table 12. Model Predictions for Liquid-Phase Mass Flow typical commercial melt train.
Rates Exiting Each Reactor (Process Flow Diagram in 6.2. Bubble-Gas Kettle Train. We now simulate a
Figure 8)
typical commercial bubble-gas kettle train.13 This train
cyclic has three agitated kettles as in Figure 9.
caprolactam water dimer nylon-6 Mn
reactor (kg/s) (kg/s) (kg/s) (kg/s) (kg/mol) A total of 2.08 × 10-3 kg/s of caprolactam enters the
1 2.85 × 10-4 9.36 × 10-3 7.39 × 10-6 5.34 × 10-3 2.9 first vessel, along with 7.73 × 10-4 kg/s of steam.
2 5.71 × 10-5 2.81 × 10-3 2.73 × 10-5 1.19 × 10-2 10.7 We model the first kettle, KETTLE1 in Figure 9, as
3 3.10 × 10-7 1.69 × 10-3 2.87 × 10-5 1.20 × 10-2 11.7 a vapor-liquid CSTR with exit liquid and vapor streams.
4 1.25 × 10-6 1.92 × 10-4 4.21 × 10-5 1.16 × 10-2 18.3
Steam and lactam are fed to this reactor. This kettle
weight of 18.3 kg/mol, and the caprolactam content is operates at 527 K and 579 kPa. This kettle has a
about 1.6%. residence time of 9300 s.
The molecular weight prediction is close to plant We assume that the second kettle, KETTLE2 in
experience of melt trains, about 18 kg/mol. The capro- Figure 9, is mass-transfer-limited. We again model it
lactam content is also close: plant experience ranges as a multiphase CSTR utilizing two-film theory to
from 0.7 to 1.5%. compute the evaporation rate of volatiles. A total of 4.91
Additional predictions now follow: 2.44 × 10-4 kg/s, × 10-5 kg/s of inert gas (assumed to be nitrogen) is
or 85%, of the incoming water is lost to the vapor phase bubbled through the reaction mixture to remove vola-
in the second reactor, while virtually all of the water is tiles. We assume that the mass-transfer coefficients are
vaporized in the third reactor. Almost no caprolactam 1 and 1 × 10-4 m/s for water and caprolactam, respec-
is lost in the second reactor because of the presence of tively. This kettle has a residence time of 13 400 s, a
the reflux condenser; however, about 2.83 × 10-3 kg/s temperature of 528 K, and a pressure of 101.325 kPa.
Ind. Eng. Chem. Res., Vol. 42, No. 17, 2003 3911
Figure 9. Three-reactor nylon-6 manufacturing process. Lactam and water are fed to the first multiphase kettle, while inert gas flow
devolatilizes the second and third vessels.
Table 13. Operating Conditions for Each Kettle Depicted that exists for every realistic manufacturing process.
in Figure 9 This high level of modeling detail is possible using our
temp pressure residence steam/nitrogen fundamental phase-equilibrium model for water/capro-
kettle (K) (kPa) time (s) flow rate (kg/s) lactam/nylon-6.
1 527 579. 9300 7.73 × 10-4 (steam)
2 528 101.325 13400 4.91 × 10-5 (nitrogen) 7. Conclusions
3 528 101.325 13400 4.91 × 10-5 (nitrogen) We have advanced nylon-6 process-simulation tech-
Table 14. Predicted Liquid-Stream Compositions
nology by developing and documenting a fundamental
Exiting Each Kettle (Process Flow Diagram Depicted in phase-equilibrium model. It represents a step forward
Figure 9) toward the generation of a single, consistent model that
cyclic
describes the phase behavior of all commercially sig-
caprolactam water dimer nylon-6 Mn nificant nylon polymerization-depolymerization tech-
reactor (kg/s) (kg/s) (kg/s) (kg/s) (kg/mol) nologies.
1 4.83 × 10-4 4.77 × 10-5 6.45 × 10-6 1.51 × 10-3 5.2 In creating this model, we use phase-equilibrium data
2 2.37 × 10-4 3.99 × 10-6 8.71 × 10-6 1.67 × 10-3 12.4 to characterize the ternary system, water/caprolactam/
3 1.89 × 10-4 1.34 × 10-6 9.56 × 10-6 1.69 × 10-3 20.6 nylon-6. We illustrate adequate treatment for inconsis-
tent thermodynamic data, including the generation of
The third kettle, KETTLE3 in Figure 9, is identical a coherent set of POLYNRTL binary interaction pa-
to the second kettle regarding operating conditions. rameters. We then show how to simulate the reactive
Table 13 summarizes the vessel operating conditions. system, water/caprolactam/nylon-6, to extract binary
Table 14 shows the liquid-phase flow rates, composi- interaction parameters for water/nylon-6 and caprolac-
tions, and polymer molecular weights coming out of each tam/nylon-6 segments. Our model predicts the liquid
kettle. mole fraction of water with an average error of 1%;
With this train, we produce 1.69 × 10-3 kg/s of nylon previous literature models sometimes generate predic-
(81% conversion), with 10% caprolactam coming out of tions that are more than an order of magnitude in error.
the third kettle. The molecular weight of the polymer After generating a complete model for the phase
is 20.6 kg/mol. Both the caprolactam content and equilibrium for water/caprolactam/nylon-6, we validate
molecular weight predictions are close to plant experi- the interaction parameters by performing exploratory
ence of bubble-gas kettle trains: about 9% and 21 kg/ simulations of commercial manufacturing processes. We
mol for caprolactam and molecular weight, respectively. simulate a condenser and match the split fraction for
In the first kettle, 75% of the incoming water exits in water (99%). We predict a caprolactam split fraction of
the vapor stream, along with 4% of the incoming 53%, which approximates the real split fraction of 49%.
caprolactam. In the second kettle, the corresponding We also simulate two commercial wiped-film evapora-
percentages are 98% and 18%, respectively. Lastly, the tors and show that the models underpredict the mass
third kettle devolatilizes 80% of the incoming water and fraction of caprolactam in the exit polymer stream. This
34% of the incoming caprolactam. is in accordance with expectations: we did not simulate
This example again shows that an incomplete devola- mass-transfer limitations and, therefore, the amount of
tilization based on a more detailed analysis of the phase caprolactam predicted to be in the liquid phase should
equilibrium is more realistic. Using the previous as- be low when compared with plant data.
sumptions regarding nylon phase behavior would have Last, we demonstrate two applications of our new
resulted in severe prediction inaccuracies. phase-equilibrium model that simulate two integrated
6.3. Example Summary. These two applications nylon-6 production processes: a melt train and a bubble-
illustrate that we are no longer bound by unrealistic gas kettle train. We simulate wide ranges of tempera-
assumptions, such as no or instantaneous water re- ture and pressure conditions, with temperature ranging
moval, while ignoring caprolactam devolatilization. In from 403 to 536 K and pressure ranging from 2.33 to
fact, we obtain a high level of detail in our simulations, 545 kPa. We perform fundamental kinetic, thermody-
including an analysis of the monomer-recovery portion namic, and mass-transfer calculations to make detailed
3912 Ind. Eng. Chem. Res., Vol. 42, No. 17, 2003
(29) Prausnitz, J. M.; Lichtenthaler, R. N.; de Azevedo, E. G. (38) Nagasubramanian, K.; Reimschuessel, H. K. Diffusion of
Molecular Thermodynamics of Fluid-Phase Equilibrium, 3rd ed.; Water and Caprolactam in Nylon-6 Melts. J. Appl. Polym. Sci.
Prentice Hall: Englewood Cliffs, NJ, 1999; p 261. 1973, 17, 1663-1677.
(30) Maczinger, J.; Tettamanti, K. Phase Equilibrium of the (39) Nagasubramanian, K.; Reimschuessel, H. K. Caprolactam
System Caprolactam/Water. Period. Polytech. 1966, 10, 183. Polymerization. Polymerization in Backmix Flow Systems. J. Appl.
(31) Giori, C.; Hayes, B. T. Hydrolytic Polymerization of Ca- Polym. Sci. 1972, 16, 929-934.
prolactam. II. Vapor-Liquid Equilibrium. J. Polym. Sci., Polym. (40) Tirrell, M. V.; Pearson, G. H.; Weiss, R. A.; Laurence, R.
Chem. Ed. 1970, 8, 351-358. L. An Analysis of Caprolactam Polymerization. Polym. Eng. Sci.
(32) Van Ness, H. C.; Byer, S. M.; Gibbs, R. E. Vapor-Liquid 1975, 15, 386-393.
Equilibrium: Part 1. An Appraisal of Data Reduction Methods. (41) Wagner, J. W.; Haylock, J. C. Control of Viscosity and
AIChE J. 1973, 19, 238-244. Polycaproamide Degradation during Vacuum Polymerization. U.S.
(33) Wisniak, J. A New Test for the Thermodynamic Consis- Patent RE28,937, 1976.
tency of Vapor-Liquid Equilibrium. Ind. Eng. Chem. Res. 1993, (42) Wright, W. H.; Bingham, A. J.; Fox, W. A. Method to
32, 1531-1533. Prepare Nylon-6 Prepolymer Providing a Final Shaped Article of
(34) Prausnitz, J. M.; Lichtenthaler, R. N.; de Azevedo, E. G. Low Oligomer Content. U.S. Patent 3,813,366, 1974.
Molecular Thermodynamics of Fluid-Phase Equilibrium, 3rd ed.; (43) Boggs, B. A.; Balint, L. J.; Ager, P. W.; Buyalos, E. J.
Prentice Hall: Englewood Cliffs, NJ, 1999; p 247. Wiped-Wall Reactor. U.S. Patent 3,976,431, 1976.
(35) Herington, E. F. G. Tests for the Consistency of Experi- (44) Saunders, L. V. J.; Rochell, D. Polymer Finisher. U.S.
mental Isobaric Vapour-Liquid Equilibrium Data. J. Pet. Inst. Patent 3,686,826, 1972.
1951, 37, 457-470. (45) Whitman, W. G. The Two-Film Theory of Gas Absorption.
(36) Daubert, T. E.; Danner, R. P. Physical and Thermodynamic Chem. Metall. Eng. 1923, 29, 146-148.
Properties of Pure Chemicals. Data Compilation; Hemisphere Received for review February 6, 2003
Publishing Corp.: New York, 1989; Vol. 1. Revised manuscript received June 9, 2003
(37) Prausnitz, J. M.; Lichtenthaler, R. N.; de Azevedo, E. G. Accepted June 10, 2003
Molecular Thermodynamics of Fluid-Phase Equilibrium, 3rd ed.;
Prentice Hall: Englewood Cliffs, NJ, 1999; p 249. IE030112+