Ding 2018

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Adsorption

https://doi.org/10.1007/s10450-018-9956-z

Optimization and analysis of the VPSA process for industrial-scale


oxygen production
Zhaoyang Ding1 · Zhiyang Han1 · Qiang Fu1 · Yuanhui Shen1 · Caixia Tian1 · Donghui Zhang1

Received: 29 September 2017 / Revised: 4 June 2018 / Accepted: 11 June 2018


© Springer Science+Business Media, LLC, part of Springer Nature 2018

Abstract
A two-bed, six-step vacuum pressure swing adsorption (VPSA) process using LiLSX zeolite for industrial-scale 80% oxygen
production with rapid cycling was investigated by numerical simulation and optimization. keeping the structural parameters
of the VPSA devices unchanged and ensuring the purity of oxygen never be lower than 80%, the operating variables were
optimized to achieve the target of minimum power consumption and maximum yield, which can maximize the devices’ poten-
tial performance for various production requirements. In the case of minimum power consumption, after optimization, the
minimum power consumption was reduced from 0.313 to 0.289 kWh Nm−3 ­O2, meanwhile, the productivity fell from 106.42
to 93.75 Nm3 h−1 ton−1. In the maximum productivity case, a high productivity of O­ 2 was achieved (117.87 Nm3 h−1 ton−1)
−3
but the power consumption rose to 0.324 kWh Nm O2. A subsequent detailed analysis was carried out to provide insights
into these observations and compare the impact of the decision variables on the process performance.

Keywords Vacuum pressure swing adsorption · Industrial scale air separation · Minimum power consumption ·
Optimization · Maximum yield

List of symbols dP Particle pore diameter, m


Dv Molecular diffusion volume, ­m2 s−1
Latin letters
F Molar flow rates, mol ­s−1
c Total gas phase concentration, mol m−3
h Heat transfer coefficient between gas and column
Cpg Constant pressure specific heat of the gag mixture,
wall, W m−2 K−1
kJ kg−1 K−1
Hb Bed height, m
Cps Specific heat of the adsorbent, kJ kg−1 K−1
ΔHi Isosteric heat of adsorption of component i,
Cv Value constant, kmol bar−1 s−1
kJ mol−1
ci Gas phase concentration of component i, mol m−3
IP1 Extended Langmuir model fitting parameters
Dax Axial dispersion coefficient, ­m2 s−1
IP2 Extended Langmuir model fitting parameters
Dc,i Effective diffusion coefficient of component i,
IP3 Extended Langmuir model fitting parameters
­m2 s−1
IP4 Extended Langmuir model fitting parameters
Dk,i Knudsen diffusion coefficient of component i,
KA Load factor
­m2 s−1
kg Effective axial thermal conductivity, W m−1 K−1
Dm Molecular diffusion coefficient, ­m2 s−1
Lbed Legend of adsorption bed, m
Lbuffer Legend of buffer tank, m
M Molecular weight of the gas, kg mol−1
Electronic supplementary material The online version of this
article (https​://doi.org/10.1007/s1045​0-018-9956-z) contains N Total number of component
supplementary material, which is available to authorized users. P Pressure, Pa
qi Adsorbed phase concentration of component i,
* Donghui Zhang mol kg−1
donghuizhang@tju.edu.cn
R Ideal gas constant, J mol−1 K−1
1
Collaborative Innovation Center of Chemical Science Rb Bed radius, m
and Engineering, The Research Center of Chemical Rbed Radius of adsorption bed, m
Engineering, School of Chemical Engineering Rbuffer Radius of buffer tank, m
and Technology, Tianjin University, Tianjin 300072, China

13
Vol.:(0123456789)
Adsorption

RP Particle radius, m and accuracy is indispensable in PSA simulations, which


t Time, s become possible because of improvements in computing
T Temperature, K technology concerning the numerical solutions of partial
Tfeed Temperature of feed gas, K differential equations. Effective optimization for the design
vg Superficial velocity, m s−1 and operation of PSA processes to improve performance is
wads Mass of adsorbent, kg becoming increasingly essential (Ruthven 2000). Significant
yi Molar fraction of component i contributions have been made toward the optimization of
z Axial direction the PSA process, as has been reported by many researchers.
Iv and Westerberg (1990) established a mixed integer
Greek letters
nonlinear programming model (MINLP) to determine the
𝜀b Bulk phase porosity
annualized-consumption-optimized design of a pressure
𝜀p Particle phase porosity
swing adsorption (PSA) for an air separation system with
𝜀m Volumetric efficiency
several decision variables (number of PSA beds, bed sched-
𝜌p Density of the adsorbent, kg m−3
uling, bed interconnection steps, and pressures of adsorp-
𝜌g Density of the gas phase, kg m−3
tion and desorption). The power consumption objection and
𝜂p Work efficiency of compressor and vacuum pump
decision variables were connected by a series of equations
𝜇 Gas viscosity, Pa s
with an index of artificial formulation. Results shown that
Subscripts the optimal value of the objective function (75.5 J mol−1)
in Gas enters the system from the outside was about 25% smaller than the base case (100.8 J mol−1).
out Gas is emitted from the system to the outside Nilchan and Pantelides (1998) introduced an optimization
i Species framework by proposing the complete discretization (CD) of
the partial differential and algebraic equations (PDAEs) and
efficiently solving a nonlinear programming (NLP) prob-
1 Introduction lem for PSA air separation. Ding et al. (2002) developed an
advanced optimization algorithm for the direct determina-
Oxygen-enriched technology (Hong et al. 2009) has been tion of the cyclic steady states (CSSs) of PSA and tempera-
widely used in industry for energy efficient and reducing pol- ture swing adsorption (TSA). Jiang et al. (2003) used a New-
lution, which would result in great economic and environ- tonian-based method with accurate sensitivities to achieve
mental benefits. This technique has become popular because fast and robust convergence of the CSS, and compared the
of the widespread application of oxygen in fish farming, results given by finite difference derivatives (FD) and accu-
environmental remediation, glass-blowing, and sewage rate sensitivities (AC) generated by DASPK. Meanwhile,
treatment (Sangjin Lee et al. 2007). Currently, various tech- there is no time and space dependency in CSS, which means
nologies, such as low-temperature distillation (Kansha et al. that the uncertain error caused by space coupled error and
2011), membrane separation (Burdyny and Struchtrup 2010) time discretization is nonexistent. Thus, this solution pro-
and pressure swing adsorption, are the main methods for vides a kind of optimization algorithm more than a solution
air separation as well as oxygen enrichment. Vacuum pres- to industrial problem. Cruz et al. (2005) used an adaptive
sure swing adsorption (VPSA) has increased significantly in multi-solution approach to implement the simulation of the
the oxygen enriched (Castle 2002) because VPSA is a kind PSA and (VSA) processes and to optimize the equalization
of gas separation technology with low power consumption, stage configuration to obtain a better process by comparing
high degree of automation, and large elasticity excellency. two types of adsorbents. The numerical error in space dis-
Furthermore, VPSA is convenient in construction with less cretization was reduced by using high-resolution schemes
investment. with a very fine mesh and a grid adaptation strategy, which
The commercial acceptance of this technology has been was based on the solution of hyperbolic equations. Agarwal
remarkable, and it is nowadays very competitive, which has et al. (2009) adapted a reduced order model making the opti-
brought many improvements over the original process. It has mization high-efficiency based on proper orthogonal decom-
reached a consensus that an effective numercial simulation position approximating a dynamic PDE-based model to a
for commercial design and optimization of cyclic adsorption DAE system of significantly lower order. The same authors
processes (CAPs) is instructive required. (Agarwal et al. 2010) proposed a systematic optimization-
The effectiveness of simulation has been proved by com- based formulation to synthesize PSA cycles for the post-
paring with successful industrial units, and experiments and combustion ­CO2 capture, which is a superstructure able to
the theory of pressure swing adsorption has been perfected predict a number of different PSA operating steps. Nikolić
by the contributions of many scientists (Ruthven 1984; Dong et al. (2009) introduced an optimization framework to opti-
et al. 1999; Sircar 1985). Meanwhile, the fast algorithms mize the number of beds, cycles configurations, and various

13
Adsorption

operating and design parameters. Khajuria and Pistikopoulos 2 Mathematical model and process
(2013) presented a detailed optimization-based approach for for the VPSA system
simultaneously incorporating PSA design, operational, and
control aspects under the effect of time variant and invariant 2.1 Mathematical model
disturbances. Daeho Ko applied of an optimization method
to scale-up the design to large commercial size PSA plants 2.1.1 Model of the adsorption bed
for carbon dioxide from coalbed methane without experi-
mental efforts. Yang et al. (2014) investigated a dynamic Notably, on the one hand, in the view of modeling adsorp-
optimization and sensitivity analysis for recovering C ­ H4 tion process precisely, the developed mathematical model
from mixture gas (70% ­N2/30% ­CH4), the decision variables should be able to capture the dynamic and transient behavior
were only the flow rates and Cv values instead of step time, in adsorption bed. On the other hand, in order to reduce the
the optimization objective is a function of power consump- computation cost, some assumptions have been made typically
tion, recovery rate and natural gas price. for the model of the adsorption bed:
The objective function of above optimization work is a
single variable, or a single relational, multi-objective opti- (1) The gas phase behaves as an ideal gas.
mization studies of adsorption processes have been per- (2) There are no radial variations in the gas concentration,
formed. Sankararao and Gupta (2007) employed a newly temperature, or pressure.
developed multi-objective optimization technique, modified (3) The pressure drop along the bed is calculated by the
MOSA-aJG, to solve and obtain the optimal values for the Ergun equation.
adsorption pressure, desorption pressure, adsorption, purge (4) Extended Langmuir isotherm models are used to
times, and input flow rates during the adsorption and purge describe the adsorption behaviors.
stages. Haghpanah et al. (2013a, b) performed the weighted (5) There is thermal equilibrium between the gas and solid
essentially nonoscillatory (WENO) finite volume scheme phases.
to achieve full multi-objective purity-recovery and power- (6) The LDF approximation model with lumped mass
productivity pareto fronts of the analyzed vacuum swing transfer coefficient is used.
adsorption (VSA) cycles. (7) Velocity change along the length of the column was
Existing works related to the simulation and optimization neglected in heat balance equation.
of gas separation processes have provided various and valu-
able ideas. Table 1 gives a brief overview of the optimization Based on the above assumptions, a series of equations for
of PSA (VPSA) for air separation. Most of these studies the adsorption bed were illustrated as follows. The calcula-
mainly focus on the VPSA unit with a pilot scale, which was tion of component and total mass balance with an estimated
much difference from industrial cases to a certain degree. In axial dispersion coefficient, are given by Eqs. (1a, 1b), while
some cases, the use of a simplified or reduced-order model empirical equations of axial dispersion coefficient are given
is inappropriate because of the resultant poor accuracy (e.g., by Eqs. (1c, 1d).
temperature). ( )
In this study, we aim at finding the optimal operation con- 𝜕 2 ci 𝜕 vg ci
−𝜀b Dax 2 +
(
+ 𝜀b + (1 − 𝜀b )𝜀p
) 𝜕ci (
+ ρp 1 − 𝜀b
) 𝜕qi
=0
ditions to achieve the goal of minimum energy consumption 𝜕z 𝜕z 𝜕t 𝜕t
and maximum productivity of an 80% VPSA oxygen process (1a)
used for nonferrous metallurgy and glass fiber industry (Pio- ( )
𝜕2c 𝜕 vg c ( ) 𝜕c 𝜕q
neer 2017; Praxair 2017). The 80% oxygen is enough for this −𝜀b Dax
𝜕z2
+
𝜕z
+ 𝜀b + (1 − 𝜀b )𝜀p
𝜕t
+ ρp (1 − 𝜀b )
𝜕t
=0
process while the energy consumption is much lower than (1b)
the oxygen process with 90% O ­ 2. The mathematical modules
of adsorption bed and accessory equipment were established 𝜈g Rp
Dax = 0.73Dm + ( )
based on the elementary theory of momentum transmission, 𝜀 (1c)
𝜀b 1 + 9.49 2𝜈bDRm
thermal transmission and the mass transfer during adsorp- g p

tion process. The VPSA process was constructed by combin-


ing the modules with certain scheme in software gPROMS, √( )
1
in which the accuracy of mathematic model was proved (Sun 0.01013T MA
+ M1
1.75

B
et al. 2015). The time of controllable steps, the flow factor Dm = ( ) (1d)
1∕ 1∕
(Cv) of key valve, the flow rates of the roots blower and P Dv,A3 + Dv,B3
vacuum pump were defined as decision variables. Addition-
ally, the purity of oxygen was set to be not lower than 80%,
which make the result consistent with practice.

13
13
Table 1  A brief overview of PSA optimization for air separation
Authors PSA model Decision variable Bounded variable Objective function Optimization technique

Nilchan and Pantelides RPSA with two-bed, six- Base Opt Lower Upper
(1998) step with zeolite 5A PF (bar) 2.12 1.99 O2 purity 89% 89% Average requirement NLP
dp (um) 302.5 347.8 Pro ­(m3 h−1) 3.60E−02 (kWh Nm−3 O2) 2.783
2.097
Lbed (m) 1 1.002
Tcycle (s) 3 4.7
Jiang and Biegler (2003) Single-bed 3-step non- Qfed ­(m3 s−1) 4.63 3.93 O2 purity 95% 95% Average requirement NLP, CSS condition (FD)
isothermal VSA and Qevacuation ­(m3 s−1) 6.78 6.9 (kWh Nm−3 O2) and accurate sensitivities
6-step industrial ­O2 VSA 0.1930.178 by DASPK (AC)
cv of purge 0.736 0.775
with one adsorbent and
two adsorbents t ads (s) 25 25.6
tpurge (s) 10 4.97
Cruz et al. (2005) Non-isothermal model θpre θpre 0.1 10 O2 recovery NLP with dimensionless
OXYSIV 5 UOP™ and θpro θpro 0.1 10 variables to reduce the
OXYSIV 7 UOP™ number of variables
θequ θequ 0.1 10 Average requirement
x x 0 1 (kWh Nm−3 ­O2)
φ φ 0.1 100
pTH/PTL pTH/PTL 1 10
Sankararao and Gupta A single fixed-bed Qfeed ­(m3 s−1) Qfeed ­(m3 s−1) 1.92 × 10−05 O2 purity Modified MOSA-aJG
(2007) adsorption column, Three tads (s) tads (s) 20 90 O2 recovery
multi-objective (two two-
Qpurge ­(m3 s−1) Qpurge ­(m3 s−1) 3.85 × 10−05 N2 purity
objective and one four
objectives) optimization tpurge (s) tpurge (s) 20 60 N2 recovery
zeolite 5A PH (atm) PH (atm) 2 4
PL (atm) PL (atm) 1 2
Liu and Sun (2013) Basic Skarstrom PSA cycle Qfeed ­(m3 s−1) – 0.0362 Qfeed ­(m3 s−1) 1.92E−05 O2 purity 3.28% and O
­ 2 Non-dominated sort-
corresponding flow rates t (s) – 28.46 20 90 recovery 1.59% ing genetic algorithm II
ads tads (s)
and time intervals for the (NSGA-II)
Qpurge ­(m3 s−1) – 0.0242 Qpurge ­(m3 s−1) 3.85E−05
four stages zeolite 5A
tpurge (s) – 31.68 tpurge (s) 20 60
PH (atm) – 3.8729 PH (atm) 2
PL (atm) – 1.0647 PL (atm) 1 2
Adsorption
Adsorption

In these equations, ci and qi are the gas and solid phase IP2i
IP1i × e T × Pi
concentrations of component i, respectively. Dax is the coef- q∗i = (5)
∑ IP4i
ficient of axial dispersion, and Dm is the mean molecular 1 + i IP3i × e T × Pi
diffusion coefficient. Rp is the particle radius, εb is the bed
void, and ρp is the particle density. Dv,A(B) is the molecular Here, Dc,i represents effective diffusion coefficient of
diffusion volume of gas A or B which diffuses each other, component i, and Dk,i is Knudsen diffusion coefficient of
while ­MA(B) is the molecular weight of gas A or B. vg is the component i. q∗i is the saturation adsorption capacity, IP1i ,
superficial velocity of the gas phase in the axial direction. IP2i , IP3i and IP4i are the fitting parameters of isotherm
The energy balance equation of the gas and solid phases model.
is shown in Eq. (2).
( N N
) N
( )∑ ( ) ∑ ( ) 𝜕T ∑ 𝜕T
𝜀b + (1 − 𝜀b )𝜀p ci Cpg,i − R + (1 − 𝜀b )𝜌p Cps + (1 − 𝜀b )𝜌p qi Cpg,i − R + vg 𝜌g Cpg,i
i=1 i=1
𝜕t i=1
𝜕z
N
∑ 𝜕qi T − Tamb ( ) 𝜕P 𝜕2T
+ (1 − 𝜀b )𝜌p ΔHi + 2h − 𝜀b + (1 − 𝜀b )𝜀p − kg 2 = 0 (2)
i=1
𝜕t Rb 𝜕t 𝜕z

Here, the heat transfer resistance between gas and solid 2.1.2 Model of accessory equipment
phases is neglected and the thin-wall model is to describe
the heat transfer to environment. Additionally, the tempera- The mathematical model of accessory equipment is as
ture of wall is equal to ambient and the wall heat balance is important as as the mathematical model of adsorption bed
neglected. In above equations, Cpg,i is the heat capacity of for the numerical simulation of adsorption process. How-
component i, Cps is the heat capacity of the adsorbent and ever, the rigorous models of accessory equipment are usually
ΔHi is the heat of adsorption component i. h is the heat- neglected in many simulation of PSA process, which would
transfer coefficient between the gas and environment, while be subject to inaccuracy of simulation results.
kg is effective axial thermal conductivities of the gas and In many papers, a multistage compression model has
solid. Tamb and T are the ambient temperature and tempera- been adopted to count power consumption of compressor
ture of the gas and solid, respectively. ρg is the gas density. and vacuum pump but it does not work well for roots blower
The momentum balance for the pressure drop along the with micro positive pressure and roots vacuum pump used
bed with a superficial gas velocity is calculated by the Ergun in VPSA process for air separation. When the performance
equation Eq. (3). curve and mechanical characteristics (O’Brien et al. 2005)
( )2 ( ) are adapted to calculate the power, higher accuracy can be
𝜕P 150𝜇 1 − 𝜀b 1 − 𝜀b 𝜌g | | achieved. Therefore, A model with the performance curve
− = ( )2 vg + 1.75 2R 𝜀 3 ||vg ||vg (3) and mechanical characteristic used to count the energy con-
𝜕z 𝜀 3 2R p b
b p
sumptions are expressed by Eqs. (7a–c).
Here, P is gas pressure, Rp is the particle radius, and µ is If Pout > Pin dW = KA Vin (−Pin )∕𝜂p 𝜀m (7a)
the gas mixture viscosity.
The description of mass transfer behavior is embodied If Pout < Pin dw = 0 (7b)
by the Linear Driving Force model, shown in Eqs. (4a–c), tcycle ∑
and the extended Langmuir model is used to describe the Wcycle = ∫ dW (7c)
adsorption ­isotherm19, as shown in Eq. (5). 0

Here, KA is the load factor (1.2–1.5), 𝜀m is the volumetric


𝜕qi 15DC,i ( ∗ )
= q i − q i (4a) efficiency, Vin is feed flowrate and 𝜂p is the effective factor.
𝜕t RP 2 As we known, the gas distributor is necessary to install
the top and bottom end of the bed for improving gas distri-
𝜀p Dk,i Dm bution in adsorption bed. Therefore, the both ends of bed
Dc,i = (4b) are usually not filled with adsorbent and the void space in
𝜏 Dk,i + Dm
adsorption bed cannot be neglected. In here, a buffer model
is used to represent this void space. In the sub-model of
√ buffer tank, compression effects are ignored. The energy
T
Dk,i = 48.5Dp (4c)
Mi

13
Adsorption

balance and mass balance of buffer model are described by Table 2  Physical parameter for Parameter Value
the following Eqs. (8a, b). system
Lbed ∕m 2
�∑ �
𝜕T ni Cpi � N
� �N
� Rbed/m 1.75
− Fi Ti yk Cpg,i + Fo To yk Cpg,i = 0 Tfeed/m 298.15
𝜕t i=1 o=1
(8a) Lbuffer ∕m 4
Rbuffer/m 2
𝜕ni ∑ ∑ εp 0.53
− Fi yi + FO y0 = 0 (8b)
𝜕t εb 0.36
Cpg/kJ kg−1 K−1 1.03
Here, Fi and FO represent the mole flow rates at the inlets
Cps/kJ kg−1 K−1 1.21
and outlets of the buffer model.
Rp/m 8.5 × 10−4
Valves are the most important units, which are typically
ρp/kg m−3 640.0
used to change the direction of gas flow to realize the differ-
DvN2/m2 s−1 1.30 × 10−4
ent step of VPSA cycle sequence. Sometimes, it is also used
DvO2/m2 s−1 1.29 × 10−4
to regulate the flow rates by adjusting valve constants. The
DvAr/m2 s−1 1.24 × 10−4
models used to describe the behavior of valve are presented
h/W m−1 K−1 0.3
below:
kg/W m−1 K−1 0.02452
Unidirectional valve∶ if Pin > Pout F = Cv (Pin − Pout )∕106 ks/W m−1 K−1 0.48
(9a)

else F = 0 (9b)
radius and packed with LiLSX zeolite. The detailed physi-
Bidirectional valve∶F = Cv (Pin − Pout )∕10 (9c) 6 cal parameters of adsorption bed and adsorbent are listed
Four performance indicators, named O­ 2 purity and recov- in Table 2, while the cycle sequence of VPSA process is
ery, unit productivity and energy consumption, were used to presented in Table 3. Moreover, Fig. 1 depicts the schematic
evaluate the process performances (Yin et al. 2015). diagram of VPSA process for air separation and Fig. 2 illus-
trates schematic diagram of two-bed six-step cycle sequence
t
∫0cycle Fout yout dt used in the simulation of VPSA process. Detailed descrip-
purityO2 = t (10a) tion of each step is described below:
∫0cycle Fout dt As shown in Fig. 2, in step 1, bed 1 is pre-pressured
with feed mixture supplied from the roots blower and the
t air buffer tank, while bed 2 is regenerated by the vacuum
∫0cycle Fout yout dt
recoveryO2 = (10b) pump and the vacuum buffer tank. In the front part of step
t
∫0cycle Fin yin dt 2, the feed mixture is pass through bed 1, the more strongly
adsorbate ­(N2) is retained in the bed and the enriched ­O2
t
steam leaves the top of bed as product. At the same time,
∫0cycle Fout yout dt the bed 2 is stilled regenerated by vacuum pump but the
productivityO2 = (10c)
tcycle wads vacuum buffer tank is unavailable because of the high pres-
sure achieved after step 1. In the last part of step 2, most of
the enriched O­ 2 steam is continuously withdrawn as product
t
∫0cycle dW and the remain is adopted as purge gas that was used to
WO 2 = t (10d) realize a complete regeneration of adsorbent in bed 2 under
∫0cycle Fout yout dt
evacuated pressure. For the sake of improvement of recov-
ery of light products and reduction of power consumption,
2.2 VPSA process pressure equalization is carried out between bed 1 and bed
2 with a top–top way during step 3. At the same time, the
A two-bed and six-step VPSA process for producing 80% feed mixture discharged from the roots blower was stored in
oxygen from the air was adopted in this study, and a rig- the air buffer tank, and the vacuum buffer tank is evacuated
orous model described at last section was established in by the vacuum pump to achieve a lower pressure. Notably,
framework of gPROMS software. The feed temperature the steps 4–6 was the same as the step 1–3 but with the
was set to be 298.15 K and the feed flow rate was design as beds interchanged. Remarkably, the buffer tanks mentioned
330 Nm3 min−1. Each bed was 2 m in length, and 1.75 m in above, are set to be 4 m in length and 2 m in radius.

13
Adsorption

Table 3  Schedule of the VPSA Step 1 2 3 4 5 6


process
Time (s) 5 6 5 3 5 6 5 3
Bed 1 FR↑ AD1↑※ AD2↑※ ED↑ VU↓※ PUR↓※ ER↓
Bed 2 VU↓※ PUR↓※ ER↓ FR↑ AD1↑※ AD2↑※ ED↑

FR feed repressurization step, AD1 and AD2 adsorption step, ED equalization-depressurization step, VU
vacuum step, PUR purge at vacuum pressure step, ER equalization–repressurization step, ※ for the con-
venience of description in the following analysis part, the cross overlapping steps are divided, the AD step
is partitioned into AD1 and AD2, the AD2 offering the source of purge gas is part of the adsorption step,
PUR is purge at low pressure step which flushed by the gas from AD2 step

Fig. 1  Flow diagram of the


process

The ­O2 buffer tank is designed to supply oxygen product pressures up to 2.5 bar. Additionally, the Langmuir model
continuous, while the feed buffer and vacuum buffer are used was used for correcting the experimental data and the cor-
to store the work of the blower and vacuum pump when they responding fitting parameters of model are listed in Table 4.
cut off from adsorption bed. Furthermore, the feed buffer The characteristic parameters of the power equipment were
tank can offer a large amount of feed mixture during the re- obtained from the manufacturer (AR series of Roots blower
pressurization step, which is benefit to shorten the step time 2017; TAR-295W two-stage wet Roots vacuum pump 2017).
and increase the unit productivity. Meanwhile, the use of the In this work, we proposed the following dynamic bound-
vacuum buffer tank is more helpful to realize a complete and ary conditions.
fast regeneration of the adsorbent in adsorption bed. At last, FR step (Z = 0, inlet; Z = L, closed)
in order to evaluate the impact of the added buffers on process
performances, comparisons of performance results are listed 𝜀b Dax
𝜕Ci || | | |
= − vg | (ci,in − ci ); 𝜋R2bed P| vg | = Fin,FR RTin ;
in Table S2 for both case with and without the added buffers. 𝜕z ||z=0 |z=0 |z=0 |z=0
In this study, a more real feed mixture considering three 𝜕T || | ∑ N
|
kg = − vg | 𝜌g Cpg,i | (Tin − T)
components (78% N ­ 2, 21% O
­ 2, 1% Ar) was selected as feed |
𝜕z |z=0 | z=0
i=1
|z=0
gas instead of two components (78% N ­ 2, 22% O
­ 2), even the (6a)
amount of computation increases when more components
are considered. The experimental isotherm data of N ­ 2, ­O2 𝜕ci || 𝜕T || |
and Ar depicted in Fig. 3 were measured by using the static | = 0; = 0; vg | = 0; (6b)
𝜕z |z=L 𝜕z ||z=L |z=L
volumetric method, in which the isotherms were determined
at temperatures of 293.15, 298.15, 303.15, and 308.15 K with AD step (Z = 0, inlet; Z = L, outlet)

13
Adsorption

Fig. 2  Schematic diagram of the cycle sequence used in the VPSA process simulations. The blue and gray regions denote the oxygen and nitro-
gen in the bed. (Color figure online)

𝜕Ci || | | |
1.4 N2-293.15K 𝜀b Dax = − vg | (ci,in − ci ); 𝜋R2bed P| vg | = Fin,AD RTin
N2-298.15K
𝜕z ||z=0 |z=0 |z=0 |z=0
N2-303.15K N
1.2 𝜕T || | ∑ |
Adsorption capacity (mol/kg)

N2-308.15K kg = − vg | 𝜌g Cpg,i | (Tin − T)


𝜕z ||z=0 |z=0 |z=0
1.0 i=1
(6c)
0.8
O2-293.15K Ar-293.15K 𝜕ci || 𝜕T ||
0.6 | = 0; = 0; P|z=L = Pout,AD ; (6d)
O2-298.15K Ar-298.15K
𝜕z |z=L 𝜕z ||z=L
O2-303.15K Ar-303.15K
0.4 O2-308.15K Ar-308.15K
ED step (Z = 0, closed; Z = L, outlet)
0.2 𝜕ci || 𝜕T || |
| = 0; = 0; vg | = 0; (6e)
0.0 𝜕z |z=0 𝜕z ||z=0 |z=0
0 50 100 150 200 250
Pressure (bar)
𝜕ci || 𝜕T ||
| = 0; = 0; P|z=L = Pout,ED (6f)
𝜕z |z=L 𝜕z ||z=L
Fig. 3  N2, ­O2 and Ar adsorption isotherms on LiLSX zeolite
VU step (Z = 0, inlet; Z = L, closed)
𝜕ci || 𝜕T ||
| = 0; = 0; P|z=0 = Pout,VU ; (6g)
Table 4  Extended Langmuir model fitting parameters 𝜕z |z=0 𝜕z ||z=0
Parameters N2 O2 Ar
𝜕ci || 𝜕T || |
−1 −1 −7 −6
6.354 × 10−6 | = 0; = 0; vg | = 0; (6h)
IP1 (mol kg kPa ) 7.107 × 10 6.861 × 10 𝜕z |z=L 𝜕z ||z=L |z=L
IP2 (K) 2910 1567 1634
IP3 ­(kPa−1) 2.563 × 10−5 4.625 × 10−5 4.325 × 10−5 PUR step (Z = 0, outlet; Z = L, inlet)
IP4 (K) 1612 441.3 450.6 𝜕ci || 𝜕T ||
ΔH (kJ mol−1) − 23.43 − 13.22 − 12.65 | = 0; P|z=0 = Pout,PUR ; =0 (6i)
𝜕z |z=0 𝜕z ||z=0

13
Adsorption

𝜕Ci || |
𝜀b Dax = − vg | (ci,in − ci );
𝜕z ||z=L |z=L
N
𝜕T || | ∑ | ( )
kg | = − vg | 𝜌g Cpg,i | T −T ;
𝜕z |z=L |z=L |z=L in
i=1
2 | | (6j)
𝜋Rbed P| vg | = Fin,PUR RTin
|z=L |z=L
ER step (Z = 0, closed; Z = L, inlet)
𝜕ci || | 𝜕T ||
| = 0; vg | = 0; =0 (6k)
𝜕z |z=0 | z=0 𝜕z ||z=0

𝜕Ci || |
𝜀b Dax = − vg | (ci,in − ci );
𝜕z ||z=L |z=L
N
𝜕T || | ∑ |
kg | = − vg | 𝜌g Cpg,i | (Tin − T)
𝜕z |z=L |z=L |z=L
i=1
| | (6l)
𝜋R2bed P| vg | = Fin,ER RTin
|z=L |z=L

inlet: steam inlet into the bed; closed: steam inlet or outlet
from the bed is closed; outlet: steam outlet from the bed. Fig. 4  The framework of process optimization
Once the process is operational, there are two extremes of
operation:
3 Optimization strategies
1. Minimize power consumption: This means operating the
device in the most consumption-effective manner, which 3.1 Optimization framework
may mean operating below maximum throughput and
focusing on minimizing power consumption, for exam- The goal programming (GP) method was chosen to establish
ple, by tuning recycle streams, etc. the framework for our optimization because it is useful when
2. Maximize productivity: This means maximizing produc- only one goal is selected as the objective function. Figure 4
tion, possibly at the consumption of higher operating shows the framework of process optimization. It mainly con-
consumptions. sists of three parts: the objective function, the control vari-
ables, and the constraints. The mathematical structure can be
The above two extremes of operation will be found by opti- expressed as:
mization and are denoted case 1 and case 2. To reiterate, case 1
minimizes power consumption, and case 2 maximizes produc- ∑ ( )
min or max f y, y′ , q
tivity (“original” represents the initial conditions).
The resulting mathematical model of VPSA process con- ( )
sists of a set of PDAEs, which are solved using the gPROMS s.t. F y, y� , q = 0
modeling tool. The boundary conditions of each step are
j(t, q) = 0
also described above. The centered finite difference method
(CFDM) was utilized to discretize the spatial domain with g(t, q) ≤ 0
n discretization points, and the method of lines (MOL) was
employed to solve the resulting system of differential–alge- qLB ≤ q ≤ qUB
braic equations (DAE). DASOLV, the DAE solver and direct Here, f represents the objective functions, and F repre-
sensitivity code in gPROMS were applied to simulate the sents the VPSA model established in the last chapter. j and g
cycle. The NLPSQP solver employs a reduced sequential represent the equality constraints and inequality constraints
quadratic programming (r-SQP) method for the solution of that the VPSA system must meet, while q represents the
an NLP problem for optimization, which will be the subject decision variables. The objective functions are the minimum
of Sect. 3. power consumption per N ­ m3 ­O2 and the maximum produc-
tivity and for optimal cases 1 and 2, respectively.

13
Adsorption

Table 5  The boundaries of the decision variables and performance by the four controllable steps, and the relationships are
constraints shown in Table S3. To obtain optimized and realistic results,
Decision variables Initial value LB UB the changes in the decision variables, which are independent
variables in the mathematical modeling corresponding to the
Process design
variables manipulated under real conditions, are specified
Cv1 (kmol bar−1 s−1) 0.015 0 50
within certain ranges, as listed in Table 5. There are lower
Cv2 (kmol bar−1 s−1) 0.40 0 100
and upper boundaries for these performance constraints, so
Cv3 (kmol bar−1 s−1) 0.40 0 100
the process meets the requirements to find the optimal solu-
Operational variables
tions that are listed in Table 5.
Vblower ­(m3 h−1) 330 315 360
To be more realistic, the minimum interval of step time
Vvacuum ­(m3 h−1) 560 530 580
was set to be 0.1 s−1, the minimum interval of the volume
tFR (s) 5 4 7
flow rate was set to be 0.01 m3 h−1, and the minimum inter-
tAD1 (s) 6 5 10
val of Cv value was set to be 0.001 kmol bar−1 s−1.
tPUR (s) 5 3 10
tED (s) 3 1 7
3.2 Cyclic steady state definition
Performance constraint
PH (kPa) 120 150
The CSS is critical for the optimization of the VPSA pro-
P (kPa)
L 35 80
cess, resulting in the spatial profiles at the beginning of the
Purity (yO2)% 80 100
cycle being almost equal to those at the beginning of next
LB lower boundaries, UB upper boundaries cycle. Successive substitution and direct determination are
two commonly used methods to define the CSS. As a tradi-
tional dynamic optimization approach, successive substitu-
The duration of each controllable step, the Cv value, as tion shows excellent stability and convergence, although it
well as flow rates of root blower and roots vacuum pump are may take more cycles and time to reach the CSS than the
selected as decision variables. direct determination approach. In addition, successive sub-
It is worthy of note that the flow rate of roots blower is stitution can provide a global observation for the change
slightly changed when the pressure is regulated within the of all variables. In this work, successive substitution was
allowable range. However, it is difficult to maintain a con- adopted to define the CSS and is expressed as:
stant flow rate of roots vacuum pump during the vacuum
| |
step, so the average value of Vvacuum ­(m3 min−1) is employed ecss = |ycycle=N − ycycle=N+1 | ≤ 𝜀css
| |
as a constant decision. Moreover, each step can be expressed

Table 6  The results of the Decision variables Results


optimization of cases 1 (min.
power) and case 2 (max. Initial value Min. power Max. productivity
productivity)
Process design
Cv1 (kmol bar−1 s−1) 0.015 0.023 0.011
Cv2 (kmol bar−1 s−1) 0.400 0.615 0.363
Cv3 (kmol bar−1 s−1) 0.400 0.602 0.231
Operational variables
Vblower ­(m3 min−1) 330.0 319.32 335.36
Vvacuum ­(m3 min−1) 560.0 531.57 563.83
tFR (s) 4.5 4.5 4.7
tAD1 (s) 6.0 6.5 6.8
tPUR (s) 5.5 5.9 5.3
tED (s) 2.5 1.7 3.0
Cycle time (s) 37 37.2 39.6
Performance variables
Purity 80.11% 80.08% 80.13%
Recovery 65.61% 57.55% 68.80%
Power (kWh Nm−3 O2) 0.313 0.289 0.324
Productivity ­(Nm3 h−1 ton−1) 106.42 93.75 117.87

13
Adsorption

Table 7  Solver and convergence statistics 160


AD Origin
Parameter Value Case1
140 Case2
Number of axial nodes 30
Discretization scheme Second-order 120

Preseeure (kPa)
central
finite differ- 100
ence
DAE solver absolute tolerance 10−5 80
DAE solver relative convergence tolerance 10−5
NLP solver convergence tolerance 10−5 60
VU1 PUR
Machine details Intel Core
i7-6700K 40
FR AD1 AD2 ED ER
VU
@3.40 GHz
0.0 0.2 0.4 0.6 0.8 1.0
Dimensionless cycle time

Here, y represents the significant variables used to moni- Fig. 6  Pressure history curves for one cycle of the three cases with
tor the CSS, such as the oxygen purity and recovery, and dimensionless time units
ecss is a small positive value that plays a critical role in the
evaluation of the number of cycles to reach the CSS. In this
work, the value of 𝜀css was ­10−5. are kept in mind for the analysis of the optimization results
as shown in Fig. 5. That is, if the reduction of power con-
sumption is wanted, the time of AD1, PUR should be pro-
4 Results and discussion longed, and the valve opening of PUR, AD and VP need
to be turned up, the flow rate of blower and vacuum pump
4.1 Optimization results should be minimized, the time of FR, ED is supposed to
shorten. When the improvement of productivity is preferred,
Table 6 shows the results of two optimization solu- the change trend of decision variables is the opposite except
tions under the solver and convergence statistics listed the time of AD1 step.
in Table 7. The CPU time required for optimization was
more than 7 days, longer than other optimizations. The 4.2 Process analysis
reason for the lengthy optimization is that the tolerance
was set to 1­ 0−5 instead of 1­ 0−3, the number of bed nodes In this paper, although the pressure for each step is not a
was 30, and successive substitution was used. Obviously, direct decision variable, it is the most intuitive parameter
the optimization results are significant. For minimum under the influence of the decision variables in this pro-
consumption in case 1, the goal was reduced to 0.289 cess. To better reflect the optimization results, the pressure
from 0.313 kWh Nm−3 ­O2, although the productivity also history curves of the original process, case 1, and case 2
dropped from 106.42 to 93.75 Nm3 h−1 ton−1. For maxi- are compared, and the time of each step is converted into a
mum productivity in case 2, a high productivity of ­O2 can dimensionless variable by division by the total cycle time.
be achieved (117.87 Nm3 h−1 ton−1) accompanied by lower Figure 6 shows the pressure history curves for one cycle
bound of purity restriction (80%), which is the same as the at the CSS. There is a significant difference in the typical
result of prior optimization, and the consumption rises to purpose. Obviously, higher pressure is better for increasing
0.324 kWh Nm−3 O2. the ­O2 productivity, while a lower pressure reduces power
It would be good that the effects of the variables on the consumption, which is the apparent verdict.
power consumption and productivity of oxygen production

Fig. 5  Effects of operation variables on the power consumption and productivity of oxygen production

13
Adsorption

50 a 312
Pressure drop AD Origin
Case1
40 310 Case2

308

Temperature (K)
Pressure drop (kPa)

30

306

20
304
VU1 PUR

10 302

FR AD1 AD2 ED VU ER
0 300
0 5 10 15 20 25 30 35 0.0 0.2 0.4 0.6 0.8 1.0
Dimensionless cycle time
Cycle time (s)

Fig. 7  Pressure drop for the original process over one cycle b

The pressure drops inside the bed, which will cause the
momentum loss of the gas, cannot be ignored. The pres-
sure drops in Fig. 7 for original process over one cycle.
As can be seen, the two optimization results shown in
Fig. S1 suffered almost the same pressure drop without
an effective reduction after optimization. The adsorption
step and the vacuum step remain at a 10 kPa pressure drop,
which reached 40 kPa at the beginning of the equalization-
repressurization (ER) and equalization-depressurization
(ED) steps, then falling to zero. As illustrated in Fig. 3, the
temperature has a great influence on the adsorption capac-
ity (And and Yang 1997), especially for large-scale VPSA
processes. Low temperatures are conducive to the adsorp-
tion step, while high temperatures benefit the desorption
steps. Figure 8 shows the temperature at the middle of the
bed in one cycle for the three cases (a) and Temperature
distribution for the original process over one cycle (b).
The temperature change is about 10 K in the same posi- Fig. 8  Temperature history curves for one cycle of the three cases
tion during one cycle, but the maximum temperature differ- with time dimensionless (a), temperature distribution for the original
ence is achieved to 22 K between the inlet at vacuum step process over one cycle (b)
and the adsorption front at adsorption step. The change
in temperature is reasonable compared with other simula-
tion. The optimization has not improved the phenomenon 4.2.1 The feed pressurization step (FR)
of temperature fluctuation.
Figure 9 shows the shaft power of a single bed (Bed1) In this step, there are two suppliers for increasing the
over one cycle. In the first half of the cycle, the roots blower pressure of the adsorption bed simultaneously: the roots
offers the gas for adsorption, and in the second half of the blower and the air buffer tank, which stores the feed gas
cycle, the roots vacuum works for regeneration. Overall, the from the blower at the equalization step. The residence
blower power consumption is half that of the vacuum pump. times for the original, case 1, and case 2 are 4.5, 4.5, and
The details will be discussed in the corresponding step when 4.7 s and the flow rates of the roots blower are 330.0,
the changes in the decision variables for the process optimi- 319.32, and 335.36 m3 min−1, respectively. The flow rate
zation are analyzed. optimization results are clearer than those of the duration,
as shown in Table 6. Figure 10a–c presents the changes in
the pressure of the three locations (the buffer tank, inlet,

13
Adsorption

900 and middle of the adsorption bed) and the flow rate from
Case1
800 Origin the blower and feed buffer tank. The trends in these three
Case2 graphs are similar, and a large amount of gas in the feed
700
Roots blower buffer flows into the bed, co-boosted by the gas from the
600 roots blower that supplies the constant flow. The flow
Power (KW)

500 rate from the buffer tank decreases as the pressure drop
400 FR between the buffer and bed decreases. The pressure at the
inlet of the bed soars sharply at the beginning of the pro-
300
cess because of the resistance of the molecular sieves, with
200 Roots vacuum the ­N 2 was adsorbed and ­O 2 pass through, the pressure
100 AD ED VU ER
reduces slightly by 10 kPa, maintaining for a moment, and
then continuing to rise. Consequently, the pressure in the
0
0.0 0.2 0.4 0.6 0.8 1.0 middle of the bed increases more smoothly. Because of the
Dimensionless cycle time
check valve at the outlet of feed buffer tank, the gas will
not return to the feed buffer tank when the flow from feed
Fig. 9  Shaft power for a single bed over one CSS cycle with dimen-
buffer reduces to zero if the pressure in the buffer is lower
sionless time units
than that in the bed as shown in Fig. 10a–c. Subsequently,
the air is supplied separately by the blower. The adsorbents
at the bottom of the bed tend to become saturated, the

a 18 b 14
Origin Inlet of bed pressure Case1 Inlet of bed pressure
16 Air buffer pressure 130 Air buffer pressure 110
12
Mid of bed pressure Mid of bed pressure
14
120 10 100
12
Flow rate (m /s)
Flow rate (m /s)

Pressure (kPa)
8
Pressure (kPa)
3

10 110
3

90
8 6
100
6 80
4
4 90
Flow rate from blower 2 Flow rate from blower
2 Flow rate from buffer Flow rate from buffer 70

0 80 0
60
0 1 2 3 4 0 1 2 3 4
Cycle time (s) Cycle time (s)
d 1.0
c before the step of Origin
22 Case2 Inlet of bed pressure 0.9
150 before the step of Case1
20 Air buffer pressure before the step of Case1
Mid of bed pressure 0.8
18 140
2
Concentration of gas N

16 0.7
130
14 0.6
Pressure (kPa)
Flow rate (m /s)

120
3

12
0.5
10
110
0.4
8
6 100 0.3
4 Flow rate from blower After the step of Origin
90 0.2
Flow rate from buffer After the step of Case1
2
0.1 After the step of Case2
0 80

0 1 2 3 4 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Cycle time (s) Axial distance (m)

Fig. 10  Pressure of the buffer tank and adsorption bed changes with flow rate from the blower and buffer for original (a), case 1 (b), and case 2
(c), and gas-phase ­N2 concentration profile before and after the feed pressurization step (d)

13
Adsorption

a 1.0 b 1.0
After the step of Origin
After the step of Origin
0.9 After the step of Case1 0.9 After the step of Case1
After the step of Case2
0.8 After the step of Case2
0.8
Concentration of gas N2

0.7 0.7

qN2 (mol/kg)
0.6 0.6
0.5
0.5
0.4
0.4
0.3
0.3 before the step of Origin
before the step of Origin
0.2 before the step of Case1 before the step of Case1
0.2
before the step of Case1 before the step of Case1
0.1
0.1
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Axial distance (m) Axial distance (m)

Fig. 11  Axial distribution of ­N2 on gas (a) and solid (b) phases at the end of the AD step

pressure continues to rise despite the absence of gas from blower, though case 1 remains at zero (shown in Fig. 9).
the buffer tank. This is consistent with the observation The pressure of case 1 is under the atmospheric pressure, so
from the gas-phase nitrogen concentration profile before there is no need to work on the gas. In some cases, to save
and after this step presented in Fig. 10d. energy, the blower can be waived when the pressure in the
At a higher equalizing pressure level, larger flow rate, bed is lower than the atmospheric pressure, the air can enter
and longer time, the pressure of bed in case 2 can reach the adsorption tower actively.
120 kPa, as shown in Fig. 10a–c, which agree with the
results in Table 6 that show the higher pressure for adsorp- 4.2.2 The adsorption step (AD1 and AD2)
tion contributing to the maximum productivity. With the
opposite optimizing condition, case 1 rises to atmospheric Air feeds into the adsorption bed at the AD step, and the
pressure after this step. Based on the above condition, the ­N2 concentration wave moves forward with time increasing
reason why the shaft power of case 2 is twice as original in proved in Fig. 11a. This step consists of two small steps,
the FR step is that the enhanced pressure in case 2 is as the AD1 and AD2 step: the entire product goes into the oxy-
original, which is proportional to the shaft power from the gen buffer tank as AD1 step, and, then the product purges
the other bed, as AD2 step. The adsorption times of the
original, case 1, and case 2 processes are 11.5, 12.4, and
12.1 s, respectively, which are the summation of tAD1 and
Origin
2.0
Case 1 tAD2 . Although the Cv2 value (0.615 kmol bar−1 s−1) of case
Case 2 1 is larger than that of the original (0.400 kmol bar−1 s−1)
and case 2 (0.363 kmol bar−1 s−1), the flow rates in Fig. 12
1.5
show the opposite trend which is same as the trend of pres-
Flow rate (m3/s)

purge step
sure drops. As the flow rate expressed by Eq. (9c) is con-
1.0 Origin
trolled by the Cv2 value and the pressure drop jointly and
Case 1
Case 2 the latter plays a leading role. The Cv value of outlet of the
product buffer tank (Cv3) is compatible with Cv2 , which
Pressure drop

0.5 controls the gas flow rates at the inlet and outlet of the buffer
tank together. The result is to maintain a certain flow rate
6 8 10 12
Cycle time (s)
14 16
production from buffer tank. The flow rates of the original
0.0 and case 2 processes start to increase at the beginning of
6 8 10 12 14 16
Cycle time (s) this step; However, that of case 1 remains zero until the
pressure in the adsorption bed exceeds that in the oxygen
Fig. 12  Flow rate towards to the oxygen buffer tank in the adsorption buffer tank via check valve AD1(2). When the purge step
step occurs, part of product gas flushes through another bed and

13
Adsorption

Origin 4.2.3 The purge step (PUR)


0.34
Case 1
Case 2 In the purge step, part of the product gas is used to purge the
0.32
adsorbents under the vacuum pressure in the final part of the
0.30
vacuum step of another bed. On the one hand, the product
Flow rate (m3/s)

0.28 flashes into the desorption bed, and the ­N2 partial pressure
in the gas phase decreases. In addition, the adsorbed amount
0.26 also decreases and the N ­ 2 will be taken off by the flow in
0.24
the meantime, so the desorption effect will be improved.
Moreover, the product gas should have gone to the oxygen
0.22 tank instead of flashing another bed, which would cause a
reduction in the recovery of product. Importantly, there is an
0.20
10 11 12 13 14 15 16 17 appropriate flow rate for each objective. The purge Cv values
Cycle time (s) (Cv1) of the original process, case 1, and case 2 are 0.015,
0.023, and 0.011 kmol bar−1 s−1, respectively, one order of
Fig. 13  The flow rate of three cases at the purge step magnitude less than Cv2 . Figure 13 shows the flow rate of
three case at the purge step. The purge flow rate of case 1
surpasses those of the original and case 2 making up for the
Table 8  The purge ratio of three cases short slab of the lower flow rate of the Roots vacuum pump
Item Origin Case1 Case2
(531.57 m3 min−1). The purge ratio is the performance of
the purge time and Cv. In practice, the implementation of
Purge ratio 0.051 0.062 0.033 purge ratio is also realized through adjustment of purge time
and Cv. Table 8 also shows that the purge ratio of case1 is
larger than others, however all of them is small. It can be
the growth trend of flow rate slows down shown in Fig. 12. inferred that the bed layer adsorption of nitrogen desorp-
In other words, the smaller Cv2 and slightly larger feed flow tion is mainly through vacuum to achieve, so the required
rate in case 2 result in a higher pressure and the adsorption products for flushing gas volume than the PSA process is
increases correspondingly, as shown in Fig. 11b. As a result, small and purge step is an auxiliary useful means of desorp-
the higher-pressure results in a higher yield. Conversely, tion, therefore, case 1 has a little better desorption effect,
case 1, which has a larger value of Cv and slightly lower as shown in Fig. 14, and the roots vacuum pump consumes
feed rate, has a lower adsorption pressure, thus consuming less power, although at the expense of sacrificing some of
less power. the recovery of oxygen production.

a 1.0 b 0.5
Origin
0.9 Origin
Case 1
0.4 Case 1
0.8 Case 2
Case 2
Concentration of gas N2

0.7
qN2 (mol/kg)

0.3
0.6

0.5
0.2
0.4

0.3
0.1
0.2

0.1 0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Axial distance(m) Axial distance (m)

Fig. 14  Axial distributions of ­N2 on gas (a) and solid (b) phases at the end of vacuum purge step

13
Adsorption

a 1.1 b 1.1
After the step of Origin
1.0 After the step of Case1 After the step of Origin
1.0
After the step of Case2 After the step of Case1
0.9 After the step of Case2
0.9
Concentration of gas N2

0.8
0.8

qN2 (mol/kg)
0.7
0.7
0.6
0.6
0.5

0.4 0.5

0.3 0.4 before the step of Origin


before the step of Origin
before the step of Case1 before the step of Case1
0.2 0.3 before the step of Case1
before the step of Case1
0.1 0.2
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Axial distance (m) Axial distance(m)

Fig. 15  Axial distribution of ­N2 in the gas (a) and solid (b) phases at the start and end of the ED step

a 1.0 b 0.6
before the step of Origin
0.9 After the step of Origin
before the step of Case1
After the step of Case1
before the step of Case1 0.5
0.8 After the step of Case2
Concentration of gas N2

0.7
0.4
qN2 (mol/kg)

0.6

0.5 0.3
0.4

0.3 0.2
before the step of Origin
After the step of Origin before the step of Case1
0.2
After the step of Case1 before the step of Case1
0.1
0.1 After the step of Case2

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Axial distance(m) Axial distance (m)

Fig. 16  Axial distribution of ­N2 in the gas (a) and solid (b) phases at the start and end of the ER step

4.2.4 The equalization‑repressurization axial distribution of ­N2 on gas (a) and solid (b) phases at
and equalization‑depressurization steps (ER and ED) the start and end of the ER and ED steps. In the ED step,
With the pressure of the adsorption bed decreasing, part of
The equalization step is the operation that the two beds the nitrogen leaves the adsorbents, so that the concentration
interact with each other to save power and retrieving some in the gas phase increases, as shown in Fig. 15a, while the
of the product gas. In this step, to improve the equalization adsorbed oxygen at the top of the bed desorbs and enters the
efficiency, the valve opening must be sufficiently large, so other bed. At the end of this stage, the concentration of N ­ 2
the Cv value is set as a constant. There is an optimal dura- in three cases is about 24% at outlet of the bed, so it will not
tion time for each case, so this is the optimized variable. pollute the other bed. Figure 15b shows the reduction in the
If the time is short, the pressure equalization is not suf- adsorption capacity at the bottom of the adsorption bed. As
ficient, reducing the product concentration and recovery. shown, the adsorption front moves forward. At the ER step,
Otherwise, if the time is long, the other tower would cause the ­O2 enriched gas from the ED bed pours in, resulting in
pollution because of the excessive amount of nitrogen. The a pressure rise and the ­N2 adsorbed by the adsorbents. As
appropriate time for the equalization of case 1 and case 2 shown in Fig. 16, the ­N2 concentration in the gas falls, and
are 1.7 and 3.0 s respectively. Figures 15 and 16 show the the main component of the gas is oxygen, which contributes

13
Adsorption

0.5 feed buffer tank. Figure 17 shows the axial net adsorption
Origin capacity of ­N2, which is the difference value of adsorption
Net adsorption capacity of N2(mol/kg)

0.4
Case 1 capacity of N­ 2 at the start and end of VU step. The reason
Case 2 for the curves rising near the outlet of the bed is that at the
equalizing step, certain nitrogen in the pressure drop of the
0.3
tower has been parsed into the rising tower, which preserves
the mechanical power of pressure and enhances recovery of
0.2 oxygen adequately.

0.1
4.3 The optimization of 93% ­O2 process

A process for 93%O2 product has been optimized, the con-


0.0 sumption can be reduced to 0.312 kWh Nm−3 ­O2 from
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Axial distance(m)
0.344 kWh Nm−3 O2 with the recovery of 52.32%, the
maximum productivity can reach to 98.65 m3 h−1 ton−1
productivity. The comparison of process performance with
Fig. 17  Net adsorption capacity of ­N2 on the solid phase at the start
and end of VU step commercial processes and data in patent is listed as flowing
Table 9. The optimal minimum power consumption is close
to Praxair (Wang 2003), but the productivity is 50% higher
to the adsorption front moving backwards ready for next than Praxair. The performance of Pioneer (2017) is near to
step. This performance indicates that the equalization time the origin process. Watson et al. (1995) offers the CaX-VSA
is optimal. process, the productivity is disadvantage.

4.2.5 The vacuum step (VU1 and PUR)


5 Conclusions
Vacuum desorption is a highly efficient way for regeneration
and its effect is dependent on the characteristic adsorption The systemic simulation and optimization of the VPSA pro-
curve of components. The residence time (the summation of cess for air separation gas by a six-step, two-bed operation
tFR , tAD1, and tAD2 ) and the flow rate are the decision varia- has been demonstrated and discussed. Models of the bed and
bles. After optimization, the appropriate time increases from sum-models that contain valves and compressors were mod-
16 to 16.9 s for case 1 and 16.8 s for case 2. The flow rate eled in gPROMS. Two modes of production were analyzed
of case 2 is 563.83 m3 min−1, which is similar to that of the as optimization goals (minimum consumption and maxi-
original scheme (560.0 m3 min−1), and the flow rate of case mum productivity) using the NLSQP method. The operating
1 dropped to 531.57 m3 min−1. Moreover, the shaft power, as variables were optimized while maintaining the structural
shown in Fig. 9, has the same trend as the flow rates. Actu- parameters of the adsorption bed constant. The minimum
ally, the effective desorption time should also include the consumption can be reduced to 0.289 kWh Nm−3 O2 from
time of equalization when the pump works on the vacuum 0.313 kWh Nm−3 O2 with the productivity dropping from
buffer tank, removing the gas from the desorption bed at the 106.42 to 93.75 m3 h−1 ton−1. For maximum productivity, a
beginning of the vacuum step. Based on the above analysis, high ­O2 productivity of 117.87 m3 h−1 ton−1 can be achieved
the true times of this step for the original scheme, case 1, with a lower bound of purity restriction (80%), as for the
and case 2 are 18.5, 18.6, and 19.8 s, respectively, and there previous optimization, although the consumption rises to
are also vacuum buffer tanks with the same function as the 0.324 kWh Nm−3 O2. Analysis and comparison were then

Table 9  The comparison of process performance with commercial processes and data in patents
Performance Origin Min power Max productivity Pioneer (2017) Watson et al. Praxair
consumption (1995) (Wang
2003)

Purity 93.0% 93.0% 93.0% 93.0% 93.0% 93.0%


Recovery 58.35 52.32% 65.73% 55.6% 52.0% 56.15%
Power consumption (kWh Nm−3 O2) 0.344 0.312 0.356 0.34 0.33 0.31
Productivity ­(Nm3 h−1 ton−1) 87.66 78.62 98.65 85.6 31.50 56.15

13
Adsorption

carried out to provide insights into the effects influencing the Kansha, Y., Kishimoto, A., Nakagawa, T., Tsutsumi, A.: A novel cryo-
decision variables and the process performance. Future work genic air separation process based on self-heat recuperation. Sep.
Purif. Technol. 77(3), 389–396 (2011)
will concentrate on combination of process with equipment Khajuria, H., Pistikopoulos, E.N.: Optimization and control of pressure
and optimization of structure parameters of the adsorption swing adsorption processes under uncertainty. AIChE J. 59(1),
tower. 120–131 (2013)
Liu, Y., Sun, F.: Parameter estimation of a pressure swing adsorp-
Acknowledgements The authors gratefully acknowledge the financial tion model for air separation using multi-objective optimisation
support from the Key Laboratory of Ethnic Affairs Commission and and support vector regression model. Expert. Syst. Appl. 40(11),
Ministry of Education of China (KF2015003). 4496–4502 (2013)
Nikolić, D., Kikkinides, E.S., Georgiadis, M.C.: Optimization of
multibed pressure swing adsorption processes. Ind. Eng. Chem.
Res. 48(11), 5388–5398 (2009)
Nilchan, S., Pantelides, C.C.: On the optimisation of periodic adsorp-
References tion processes. Adsorption 4(2), 113–147 (1998)
O’Brien, T.F., Bommaraju, T.V., Hine, F.: Chemical engineering prin-
ciples. Springer, Boston (2005)
Agarwal, A., Biegler, L.T., Zitney, S.E.: Simulation and optimization of Pioneer: VPSA-O2 Plant. http://www.vpsat​ech.com/produ​ct/vpsa-
pressure swing adsorption systems using reduced-order modeling. oxyge​n/ (2017). Accessed 10 Dec 2017
Ind. Eng. Chem. Res. 48(5), 2327–2343 (2009) Praxair: Praxair vacuum pressure swing adsorption VM80 oxygen
Agarwal, A., Biegler, L.T., Zitney, S.E.: A superstructure-based opti- system. http://www.praxa​ir.com/-/media​/docum​ents/speci​ficat​
mal synthesis of PSA cycles for post-combustion ­CO2 capture. ion-sheet​s-and-broch​ures/servi​ces/vpsa--vm80-xl.pdf (2017).
AIChE J. 56(7), 1813–1828 (2010) Accessed 10 Dec 2017
And, S.U.R., Yang, R.T.: Limits for air separation by adsorption with Ruthven, D.M.: Principles of adsorption and adsorption processes.
LiX zeolite. Ind. Eng. Chem. Res. 36(12), 5358–5365 (1997) AIChE J. 56(4), 168–179 (1984)
AR series of Roots blower. http://www.china​blowe​rs.com/EN/produ​ Ruthven, D.M.: Past progress and future challenges in adsorption
ct/496.html (2017). Accessed 21 Mar 2017 research. Ind. Eng. Chem. Res. 39(7), 2127–2131 (2000)
Burdyny, T., Struchtrup, H.: Hybrid membrane/cryogenic separation Sangjin Lee, J., Jung, J., Moon, J., Jee, A., Lee, C.: Parametric study
of oxygen from air for use in the oxy-fuel process. Energy 35(5), of the three-bed pressure—vacuum swing adsorption process for
1884–1897 (2010) high purity O
­ 2 generation from ambient air. Ind. Eng. Chem. Res.
Castle, W.F.: Air separation and liquefaction: recent developments and 46(11), 3720–3728 (2007)
prospects for the beginning of the new millennium. Int. J. Refrig. Sankararao, B., Gupta, S.K.: Multi-objective optimization of pressure
25(1), 158–172 (2002) swing adsorbers for air separation. Ind. Eng. Chem. Res. 46(11),
Cruz, P., Magalhães, F.D., Mendes, A.: On the optimization of cyclic 3751–3765 (2007)
adsorption separation processes. AIChE J. 51(51), 1377–1395 Sircar, S.: Book review. Principles of adsorption and adsorption pro-
(2005) cesses. Langmuir 1(4), 529–529 (1985)
Ding, Y., Croft, D.T., Levan, M.D.: Periodic states of adsorption cycles Sun, W., Shen, Y., Zhang, D., Yang, H., Ma, H.: A Systematic simula-
IV. Direct optimization. Chem. Eng. Sci. 57(21), 4521–4531 tion and proposed optimization of the pressure swing adsorption
(2002) process for ­N2/CH4 separation under external disturbances. Ind.
Dong, F., Lou, H., Kodama, A., Goto, M., Hirose, T.: The Petlyuk PSA Eng. Chem. Res. 54(30), 7489–7501 (2015)
process for the separation of ternary gas mixtures: exemplification TAR-295W two-stage wet Roots vacuum pump. http://www.csroo​
by separating a mixture of C ­ O2–CH4–N2. Sep. Purif. Technol. ts.com/html/AR-Wshis​hiluo​cizhe​nkong​beng/69.html (2017).
16(2), 159–166 (1999) Accessed 21 Mar 2017
Haghpanah, R., Majumder, A., Nilam, R., Rajendran, A., Farooq, S., Wang, X.: Non-isothermal simulation and optimization of pressure
Karimi, I.A., Amanullah, M.: Multiobjective optimization of a swing adsorption for separating oxygen from air. Doctoral Dis-
four-step adsorption process for postcombustion ­CO2 capture via sertation, Nanjing Tech University (2003)
finite volume simulation. Ind. Eng. Chem. Res. 52(11), 4249– Watson, C.F., Whitley, R.D., Agrawal, R., Kumar, R.: Air, oxygen. US
4265 (2013a) Patent US5411578, 02 May 1995
Haghpanah, R., Nilam, R., Rajendran, A., Farooq, S., Karimi, I.A.: Yang, H., Yin, C., Jiang, B., Zhang, D.: Optimization and analysis of a
Cycle synthesis and optimization of a VSA process for postcom- VPSA process for ­N2/CH4 separation. Sep. Purif. Technol. 134(1),
bustion ­CO2 capture. AIChE J. 59(12), 4735–4748 (2013b) 232–240 (2014)
Hong, J., Chaudhry, G., Brisson, J.G., Field, R., Gazzino, M., Gho- Yin, C., Sun, W., Yang, H., Zhang, D.: Optimization of three-bed
niem, A.F.: Analysis of oxy-fuel combustion power cycle utilizing VPSA system for biogas upgrading. Chem. Eng. Sci. 135, 100–
a pressurized coal combustor. Energy 34(9), 1332–1340 (2009) 108 (2015)
Iv, O.J.S., Westerberg, A.W.: The optimal design of pressure swing
adsorption systems. Chem. Eng. Sci. 46(12), 2967–2976 (1990)
Jiang, L., Biegler, L.T., Fox, V.G.: Simulation and optimization of
pressure-swing adsorption systems for air separation. AIChE J.
49(5), 1140–1157 (2003)

13

You might also like