Adsorption (2021) 27:129–145
https://doi.org/10.1007/s10450-020-00281-w
Accurate blank corrections for zero length column experiments
Maarten Verbraeken1 · Alessio Centineo1 · Luigi Canobbio1 · Stefano Brandani1
Received: 14 October 2020 / Revised: 26 October 2020 / Accepted: 29 October 2020 / Published online: 18 November 2020
© The Author(s) 2020
Abstract
In this study we present a new methodology for correcting experimental Zero Length Column data, to account for contributions to the measured signal arising from extra-column volumes and the detector. The methodology considers the experimental setup as a series of mixing volumes with diffusive pockets whose contributions to the overall measured signal can
be accurately described by simple model functions. The composite effect of the individual contributions is subsequently
described through the method of convolution. It is shown that the model parameters are closely related to the physical characteristics of the setup components and as such they remain valid over a range of process conditions. The methodology is
firstly validated through fitting to experimental experiments without adsorbent present. The inverse procedure of deconvolution can in turn be applied to experimental data with adsorbent, to yield corrected data which can readily be modelled using
standard tools for equilibrium and kinetic analysis. A number of case studies is finally presented exemplifying the effect of
applying accurate blank corrections, demonstrating also the application to a nonlinear adsorption system.
Keywords Blank correction · Zero length column experiment · Deconvolution · Adsorption equilibrium · Adsorption
kinetics
Abbreviations
Apipe
Cross sectional area of diffusive pipe (m2)
Dimensionless concentration
C
Cin
Dimensionless inlet concentration
c
Concentration at time t (mol m–3)
c0
Concentration at time zero (mol m–3)
Concentration at infinite time (mol m–3)
c∞
Total gas concentration (mol m–3)
cT
D
Diffusivity (m2s–1)
F
Flowrate (m3 s–1)
Fin
Flowrate entering ZLC (m3 s–1)
Function defined by Eq. (15)
F(t)
Laplace transfer function
G(t)
LMS
Dimensionless parameter described by Eq. (21)
Dimensionless parameter described by Eq. (12)
LT1
Ldiff
Diffusion length in slab side pocket (m)
P
Pressure (Pa)
Electronic supplementary material The online version of this
article (https://doi.org/10.1007/s10450-020-00281-w) contains
supplementary material, which is available to authorized users.
* Stefano Brandani
[email protected]
1
School of Engineering, University of Edinburgh, Robert
Stevenson Road, The King’s Buildings, Edinburgh EH9 3FB,
UK
Q
Q
q∗
R
T
t
u
VCSTR
Vdiff
Vf
Vmix
Vs
x
yin
yout
Dimensionless concentration in diffusive side
pocket
Average dimensionless concentration in diffusive
side pocket
Equilibrium amount adsorbed (mol m–3)
Ideal gas constant (J mol–1 K–1)
Temperature (K)
Time (s)
Convolution integrand (s)
Volume of well mixed cells (m3)
Volume of diffusive side pocket (m3)
Fluid volume ZLC (m3)
Volume in mixing cell connected to diffusive
side pocket (m3)
Volume of adsorbent (m3)
Spatial coordinate (m)
Mole fraction of adsorbate entering ZLC
Mole fraction of adsorbate leaving ZLC
Greek symbols
𝛼
Correction factor for diffusive length (m)
𝛽slab,n
Roots of Eq. (15)
𝛽sphere,n Roots of Eq. (24)
𝛾MS
Dimensionless parameter described by Eq. (22)
Dimensionless parameter described by Eq. (13)
𝛾T1
𝜆CSTR
Inverse time constant of mixing volume (s–1)
13
Vol.:(0123456789)
130
𝜆MS
𝜆T1
𝜏MS
Adsorption (2021) 27:129–145
Inverse time constant for the mass spectrometer/
detector (s–1)
Inverse time constant of diffusive side pocket
(s–1)
Time constant for the mass spectrometer/detector
(s–1)
1 Introduction
Adsorption is a widely used technology for separation,
purification and drying, due to its simplicity, reliability and
scalability. In industrial settings, the technology is run in
a cyclical fashion, e.g. pressure swing adsorption (PSA),
vacuum swing adsorption (VSA), temperature swing adsorption (TSA), etc. Key to running these processes efficiently is
having a minute understanding of the process’ equilibrium
and kinetic properties. Whereas equilibrium data can be
obtained on either bulk or small sample quantities, kinetic
data is often obtained on lab scale sample quantities. One
method for obtaining such information is the Zero Length
Column (ZLC) technique, which uses very small amounts
of sample (typically less than 15 mg)(Brandani and Mangano 2020). Breakthrough experiments, which can be run
on a range of sample sizes, are also routinely used for this
purpose and kinetic properties may be extracted by fitting
experimental data to simulations, where the kinetic properties are adjustable modelling parameters (Ruthven 1984;
Wilkins et al. 2020).
Separation has become ever more important in recent
years, in particular due to an effort to curtail global warming and limit emissions of greenhouse gases. For this reason, many groups are now pursuing the development of next
generation adsorbents, such as MOFs, flexible zeolites, hierarchical silicas, etc. Naturally, these materials are initially
synthesised on a small scale in the lab and so establishing
their kinetic properties requires column experiments on a
similarly small scale. Such small column experiments are
liable to exhibiting non-negligible extra-column effects,
which can lead to erroneous interpretation of data and thus
inaccurate kinetic information. This effect from the extracolumn-volume (ECV) has long been recognised in the field
of liquid chromatography, in which non-square pulse inputs
and detector effects need to be accounted for in the final
analysis (Delley 1986; Kaltenbrunner et al. 1997; Gritti et al.
2006). A small number of studies also address this problem
in step-response techniques (Wilkins et al. 2020). In the case
of a linear response of the ECV, a point-by-point correction
may be applied (Guntuka et al. 2008), whereby the gas phase
concentration of a blank run at time t is subtracted from the
13
uncorrected adsorbent run, performed under identical conditions, at the same time t. In order to avoid having to run
blank experiments under identical conditions for each experiment with adsorbent, Rajendran et al. use a tanks-in-series
model to account for ECV and subtract its contribution from
the sample runs to leave a corrected signal (Rajendran et al.
2008). Joss and Mazzotti use a similar approach based on
dispersed plug-flow and by adding a stagnant volume to their
ECV model manage to account for mass transfer and heat
effects to accurately describe extra-column contributions for
a range of pressures and flow rates (Joss and Mazzotti 2012).
As their model is partly empirical and does not attempt
to capture the physics of their experimental setup, model
parameters require adjusting for different process conditions.
In this latter study, the ECV contributions are added to the
simulated signal of the column to match experimental data,
an approach also used by Friedrich et al., who add a detector model to their simulations. The detector model is based
on Langmuir adsorption with LDF mass transfer limitation
and is fit to experimental blank runs (Friedrich et al. 2015).
In this study, we take inspiration from several previous
studies, but with the aim to describe the experimental setup
based on its physical components. This has the advantage of
requiring a minimum of modelling parameters and making
the model valid over a large range of process conditions,
i.e. without the need to modify parameters. The contributions of several components with different time constants
to the experimental signal will yield a composite effect,
which can mathematically be derived through the method
of convolution, an approach also used in refs. (Delley 1986;
Kaltenbrunner et al. 1997). The basis of this approach is
the fact that each component will have a transfer function
that converts the inlet concentration to an outlet response.
Assuming linearity allows the solution of each model in the
Laplace domain and the composite transfer function can then
be obtained from the product of the individual terms, as
all components are in series. This is a standard result from
process dynamics and control (Stephanopoulos 1983). The
model for the extra-column effects is parametrised by fitting experimental blank runs, i.e. column experiments with
no adsorbent present. By applying the inverse procedure
of deconvolution on signals from experimental runs with
adsorbent, a corrected signal is obtained which corresponds
to the dynamic response that the ZLC would have if a perfect step input could be achieved and the detector had an
ideal response. The key assumption of the deconvolution
procedure is linearity and this is typically the case for the
dynamic response of all the components before the ZLC
and the detector.
Adsorption (2021) 27:129–145
131
2 Experimental
2.1 Materials
Gases used in this study are helium (BOC, CP grade,
99.999% purity), carbon dioxide (BOC, 99.8% purity) and
nitrogen (BOC, 99.998% purity). The adsorbents used in the
case studies will be described in the corresponding section.
2.2 Equipment
1/16″. The column and Tee-junction 2 have an outer diameter of 1/8″ and are thus connected to the piping through
reducing unions. Although the experimental setup in ref.
(Centineo and Brandani 2020) is modified for measuring
vapour adsorption, the part which is relevant for our modelling purposes, i.e. the section situated between the solenoid valves and the detector, is essentially identical to that
described in Fig. 1.
3 Theory
The experimental setups under consideration have been
described in detail in refs. (Hu et al. 2015) and (Centineo
and Brandani 2020). A simplified diagram of the experimental setup in ref.(Hu et al. 2015) is depicted in Fig. 1. The
dosing mixture is prepared by mixing appropriate quantities
of carrier gas (i.e. He or N2) and the dosing component in
four 1 L stainless steel canisters and leaving the mixture to
equilibrate for at least 3 h. The flow rates of both the inert
carrier and the dosing gas are controlled by mass flow controllers (MFC, Brooks Instrument). Thereafter the gases pass
through solenoid valves (SV1 to SV4) towards the column
and outlet with detector (MS; Dycor mass spectrometer,
Ametek Process Instruments). A back pressure regulator
(BPR, Brooks Instrument) at the outlet allows for operation at elevated pressures. The column is housed within a
Carbolite oven, with a Eurotherm temperature controller.
The solenoid valves are operated in an alternating fashion,
such that either of the two gas streams are fed to the column
and by activation allow a step change in the composition
of the gas phase. A differential pressure gauge between the
dosing and carrier lines allows for balancing the gas pressures, so that no pressure spikes occur upon switching the
solenoid valves. All the components after the solenoid valves
are connected by stainless steel pipes with outer diameter of
Fig. 1 Schematic diagram of the
experimental setup
3.1 Simulation of blank experiments
through the method of convolution
For our modelling purposes, we firstly consider the experimental setup as described in Fig. 1. Since the pressures
between the dosing and carrier gas lines are balanced we
assume that the concentrations at each solenoid valve are
constant. We can therefore limit our section of interest to
what is happening after solenoid valves SV1 and SV2. This
section contains a number of mixing volumes and diffusive pockets, as described in detail in Fig. 2, where we use
more accurate graphical representations of the actual setup
components (Swagelok 2020). A particular area of interest is highlighted by the dashed box and concerns the Teejunction between the solenoid valves. When the solenoid
valves change position, a certain volume ceases to be in flow
and thus becomes a diffusive pocket, shown in Fig. 3. More
precisely, Tee-junction 1 is a mixing volume, whereas the
combined volume of the solenoid valve and pipe connecting to this tee junction is a diffusive volume. The remaining
reducing unions, ZLC and Tee-junction 2 are all considered
as simple mixing volumes in flow. The following assumptions are additionally made:
BPR
MFC
Outlet
MFC
MS
Diff P gauge
Dosing mixture – HeCO2/N2
Carrier - He
Column
Tee 2
SV3
SV4
SV1
SV2
Tee 1
Vent
13
132
Adsorption (2021) 27:129–145
VCSTR
(
)
dC
= F Cin − C
dt
(1)
where C is expressed in terms of a deviation function, i.e.
C=
c − c0
c∞ − c0
(2)
Converting to the Laplace domain yields:
sC̃ − C(0) =
)
F (̃
Cin − C̃
(3)
VCSTR
And after re-arranging and considering that C(0) = 0, this
yields:
𝜆CSTR
C̃
̃ CSTR (s) =
=G
̃
s + 𝜆CSTR
Cin
with 𝜆CSTR =
F
.
VCSTR
(4)
Equation 4 is the so-called transfer func-
tion for a CSTR. The equivalent function in the time domain
is:
G(t) = 𝜆CSTR e−𝜆CSTR t
Fig. 2 The experimental section of interest for our modelling purpose. The dashed box is shown in more detail in Fig. 3
The mass balance for Tee-junction 1 is assumed to be
that of a CSTR, with volume Vmix , connected to a diffusive
pocket with slab geometry (i.e. one dimensional diffusion
along length Ldiff ) and volume, Vdiff :
Vdiff
Diffusive volume
(
)
dQ
dC
+ Vmix
= F Cin − C
dt
dt
(6)
where C is as previously defined and Q is the average dimensionless concentration in the slab, i.e.
Direcon of flow
=
(5)
∙
Tee 1
Q=
q − q0
q∞ − q0
(7)
The diffusion equation for a slab is
Fig. 3 Diagram of diffusive volume (dashed box), composed of volume in the solenoid valve and the 1/16″ pipe. The total diffusive
length is expected to be larger than the length of 1/16″ pipe, due to
the ink-bottle configuration
•
•
•
•
All gases are ideal
The pressure drop in the system is negligible
The system behaves isothermally
Axial dispersion is negligible in pipes with convective
flow
• Radial dispersion is negligible
For modelling purposes, the mixing volumes in flow in
the experimental setup are treated as well-mixed cells, or
CSTRs, for which the following mass balance applies:
13
dQ
𝜕2Q
=D 2
dt
𝜕x
(8)
and the average concentration is defined as
( )
dQ
D 𝜕Q
=
dt
Ldiff 𝜕x x=Ldiff
(9)
Solving these equations in the Laplace domains yields:
̃
̃ T1 (s) = C(s) =
G
C̃ in (s) √
where
LT1 𝜆T1
�� �
s
s𝜆T1 tanh
+ s𝛾T1 + LT1 𝜆T1
𝜆
T1
(10)
Adsorption (2021) 27:129–145
𝜆T1 =
133
D
2
Ldiff
(11)
Cin (t) and G(t), acting in tandem, the resulting function,F(t),
is defined as:
t
LT1 =
𝛾T1 =
F 1
Vdiff 𝜆T1
0
Vmix
Vdiff
∞
∑
n=1
(13)
2
−2LT1 𝛽slab,n
𝜆T1
−𝛽 2 𝜆T1 t
(
)2 e slab,n
2
2
2
+ 𝛽slab,n
+ 𝛾T1 𝛽slab,n
LT1 + 𝛾T1 𝛽slab,n
− LT1
(14)
where 𝛽slab,n are the roots of the following equation:
(
)
𝛽slab,n tan 𝛽slab,n + 𝛾T1 𝛽slab,n − LT1 = 0
(15)
Transfer functions can be combined to obtain concentrations leaving various parts. In the Laplace domain they may
simply be multiplied; for instance for Tee-junction 2:
̃ T1 (s)G
̃ RedUnion (s)G
̃ ZLC (s)G
̃ T2 (s)
C̃ T2 (s) = C̃ in (s)G
(16)
By switching between a pure carrier and dosing gas, it
can be assumed that the input concentration for Tee-junction
1 is a step function, i.e.
Cin (t) =
0, t < 0
1, t ≥ 0
(17)
which in the Laplace domain becomes C̃ in (s) = 1s .
Alternatively, the effects on the concentration profile of
the various parts in the time domain can be calculated using
the convolution theorem. This states that for two functions,
(18)
As all the transfer functions in this study are exponential
functions, it is straightforward to find analytical expressions
to Eq. (18). Using the step function in Eq. (17) as the initial
value for Cin and using the resultant function F(t) as the
inlet function for the subsequent convolution integral, the
cumulative effect of the various setup parts can be calculated
as shown in Fig. 4. Here, an arbitrary sequence has been
used, i.e. low volume tee, 1/16″–1/8″ reducing union, 1/8″
straight union, 1/8″ tee. It is important to note however that
the convolution process can be performed in any order, as
the procedure is symmetrical. This symmetry stems from the
fact that the convolution integral is in essence a linear operator and is therefore, strictly speaking, only valid for linear
systems. When viewed in a semi-log plot, the diffusive tail
from Tee-junction 1 becomes more evident. The concentration profile in all these components is affected by flowrate;
its effect can be seen in Fig. 5.
The modelling parameters can mostly be obtained by
inspection of the dimensions of the various setup components, such as diameter and length of pipes and columns (as
listed in Table 1) and through literature values of diffusivities for various species in the carrier gas, typically helium
and nitrogen (Arora and Dunlop 1979; Ellis and Holsen
1969; Massman 1998; Paganelli and Kurata 1977; Schwertz
and Brow 1951; Walker et al. 1960). The inner volume of the
solenoid valves can be estimated by analysis of experiments
using an empty column, a so-called blank run.
A number of blank runs for a typical mixture of 1% CO2
in helium are shown in Fig. 6. It can be seen that the curves
are essentially composed of two exponential decays: the
1.0
1
Convolution Low V Tee
Convolution 1/16" - 1/8" Red. Union
Convolution 1/8" Union
Convolution 1/8" Tee
0.8
Convolution Low V Tee
Convolution 1/16" - 1/8" Red. Union
Convolution 1/8" Union
Convolution 1/8" Tee
0.6
c/c0
c/c0
0
(12)
Inverting Eq. (10) to the time domain, finally gives:
G(t) =
t
F(t) = Cin (t) × G(t) = ∫ G(t − u)Cin (u)du = ∫ G(u)Cin (t − u)du
0.1
0.4
F = 5 cc/min
F = 5 cc/min
0.2
0.01
0.0
0
10
20
0
10
t (s)
20
30
40
t (s)
Fig. 4 Cumulative effect of the various setup components on the gas phase concentration profile
13
134
Adsorption (2021) 27:129–145
1
1
Low V tee
1/8" union
1/8" tee
Flowrate 5 cc/min
Flowrate 10 cc/min
Flowrate 20 cc/min
0.1
0.1
(b)
c/c0
c/c0
(a)
0.01
0.01
Convolution Tee 2
0.001
0.001
0
10
20
30
40
50
0
20
40
t (s)
60
80
t (s)
Fig. 5 Effect of flow rate on concentration profile at 1/8″ Tee-junction 2 (a) and at various components (dashed line at 2 cm3/min and solid line
at 5 cm3/min) (b)
Table 1 Key dimensions for
setup components
Component
Inner diameter
(mm)
Low volume tee
Straight union (1/8″)
Regular tee (1/8″)
Reducing union (1/16″ to 1/8″)
Pipe (1/16″)
1 cc/min
2 cc/min
5 cc/min
10 cc/min
c/c0
0.1
0.01
0.001
1E-4
0
100
200
300
400
500
600
700
800
t(s)
Fig. 6 Blank runs in a mixture of 1% CO2 in helium, clearly showing two exponential decays. The initial decay corresponds to mixing
volumes in flow, whereas the long-time tail corresponds to diffusive
pockets within the setup
initial decay corresponds to the part of the setup that comprises mixing volumes in flow, whereas the long-time tail
is due to a diffusive process. When the curves are plotted
versus Flowrate × time, the area underneath the curves correspond to volume of the setup. The assumption of ideal
13
Source
0.280
41.7
119
12.2
-
(Swagelok 2020)
(Swagelok 2020)
(Swagelok 2020)
(Swagelok 2020)
(Swagelok 2020)
Table 2 Volumes as calculated from integration of blank runs at different flow rates
1
T = 35°C
1% CO2
2.29
2.29
1.27
0.876
Volume (mm3)
Flow rate
Volume in flow (mm3)
Diffusive
volume
(mm3)
1
2
5
10
478
503
485
585
178
197
285
288
plug flow through the 1/16″ pipes simply causes a delay in
the detector response, which can easily be removed from the
analysis. Consequently, the volume in flow and that of the
diffusive pocket can be estimated by integrating the relevant
parts of the curves. The results are shown in Table 2. It is
evident that the total mixing volume is larger than the sum of
the components in Table 1, but critically there is very little
variation, as is also clear from Fig. 7, in which the initial part
of the curves at different flow rates all overlap.
The experimental blank runs can now be simulated
using the expressions in Eqs. (5, 14 and 18) using appropriate parameters (which can be found in the supplementary material, Table S1), as shown in Fig. 8. Due to the
fact that the total diffusive volume at the tee-junction (i.e.
Adsorption (2021) 27:129–145
135
1
c/c0
1 cc/min
2 cc/min
5 cc/min
10 cc/min
T = 35°C
1% CO2 in He
0.0
0.2
0.4
0.6
0.8
1.0
Ft(cc)
Fig. 7 Blank runs in a mixture of 1% CO2 in helium; c/c0 versus
eluted volume (Flowrate × time). It is evident that the initial part of
the curves overlap, as expected, since the mixing volumes of the
setup do not change with changing the flowrate
Vdiff = Vvalve + Vpipe) is large relative to Vpipe, in effect creating an ‘ink bottle’ type diffusive configuration (see Fig. 3),
it was found that Ldiff in Eq. (11) is larger than the actual
length of the connecting pipe. Due to some mixing in the
valve itself, we are using the following approximation for
Ldiff , which works well for modelling purposes:
Ldiff = 𝛼
Vdiff
Apipe
whereas the simulated blanks in Fig. 8 show a decent fit at
low flowrates, it is evident that significant deviation occurs
at higher flowrates. The simulations do not capture the shape
of the blanks at longer times and underestimate the total
volume in flow. Therefore, a final optimisation of the simulated curves is obtained by including a response from the
mass spectrometer, which is likely to cause some accumulation and have a kinetic limitation. This effect only becomes
apparent at higher flow rates (F > 10 cm3/min). The mass
spectrometer effect is similar to that of the low volume tee
(Tee-junction 1), i.e. it can be modelled as a combination
of a mixing volume with a diffusive side pocket. The inlet
to the ionising volume of the mass spectrometer is through
a fused silica capillary of approximately 1 m length, which
continuously samples gas from the outlet of the ZLC at a
flowrate of about 0.02 ± 0.01 cm3/min. The diffusive pocket
is also more likely to be multidimensional; we are assuming
a spherical geometry. The transfer function for this part is:
̃
LMS 𝜆MS
C(s)
̃
=
G(s)
=
�� �
C̃ in (s) √
s
𝜆MS s coth
− 𝜆MS + s𝛾MS + LMS 𝜆MS
𝜆
MS
(19)
where
𝜆MS =
1
𝜏MS
(20)
LMS =
1 F 1
3 Vdiff 𝜆MS
(21)
𝛾MS =
1 Vmix,MS
3 Vdiff ,MS
(22)
with 𝛼 chosen such that:
Vdiff
Apipe
≥ Ldiff ≥ Lpipe
1
1
1 cc/min
2 cc/min
5 cc/min
/
/
T = 35°C
1% CO2
simulated
0.1
c/c0
c/c0
0.1
10 cc/min
20 cc/min
40 cc/min
/
/
Simulated
T = 35°C
1% CO2
0.01
0.001
0.01
0.001
1E-4
1E-4
0
200
400
600
800
1000
t(s)
0
100
200
300
400
500
t(s)
Fig. 8 Experimental blank runs and simulated fits, using convolution theorem
13
136
Adsorption (2021) 27:129–145
Here it is recognised that 𝜏MS corresponds to the time constant of the mass spectrometer, which may not necessarily
involve a diffusion process, but could instead be due to a
capacitive effect of the Faraday cup. Inverting Eq. (19) to
the time domain, finally gives:
G(t) =
∞
∑
n=1
2
𝜆MS
−2LMS 𝛽sphere,n
(
)2
2
2
2
LMS + 𝛾MS 𝛽sphere,n
+ 𝛽sphere,n
+ 𝛾MS 𝛽sphere,n
− LMS + 1 − 1
2
e− 𝛽sphere,n 𝜆MS t
(23)
where 𝛽sphere,n are the roots of the following equation:
(
)
2
𝛽sphere,n cot 𝛽sphere,n − 𝛾MS 𝛽sphere,n
− 1 + LMS = 0
(24)
Assuming that the flowrate into the mass spectrometer
is constant, a single set of parameters, i.e. LMS , 𝜆MS and 𝛾MS
will be used for simulating this effect on the shape of the
blank runs. The resulting fits when the detector is included
in the convolution model are shown in Fig. 9. The modelling
parameters are listed in Table 3.
scale by the square root of the molecular masses, as expected
from a process described by molecular diffusion, i.e.:
�
1 �
�
√
DCO2 −He 2
MwCO2
28.01
L
=√
= ��
= 0.80
DN2 −He� 2
1
44.01
L
MwN2
Blanks were also run at elevated pressures and as Fig. 12
shows, the model can easily capture this effect by adjusting
the diffusivity of CO2 in helium. These experiments were
performed prior to a major system upgrade and therefore the
parameters describing the low volume Tee-junction 1 have
different values. The model parameters at different pressures
can be found in the supplementary material, Table S2.
Table 3 Key parameters in the convolution model
Model parameter
1% CO2 in He 10% CO2 in He 1% N2 in He
Volumes in flow
Straight union (cm3)
3.2 Effect of changing process conditions
7.09 × 10−2
7.09 × 10−2
7.09 × 10−2
Reducing union (cm ) 2.50 × 10
Tee-junction 2 (cm3) 2.02 × 10−1
−2
2.50 × 10−2
−1
2.02 × 10
2.02 × 10−1
2.80 × 10−4
2.80 × 10−4
2.80 × 10−4
−1
−1
2.46 × 10−1
−3
3
To validate the current model, it was applied on blank runs
under a range of different process conditions. When using
different concentrations of CO2 or changing the dosing gas
to N2 in He, the only parameters expected to change are
the value for the diffusivity in Tee-junction 1 (in the case
of using N2) and the mass spectrometer parameters. The
various diffusive volumes and those in flow should remain
unchanged. As can be seen from Figs. 10 and 11 and the
modelling parameters in Table 3 this is indeed the case.
Moreover, the diffusivities in tee-junction 1 for the models
describing blanks in 1% N2/He and 1% CO2/He atmospheres
Slab diffusive pocket
Vmix(cm3)
3
Vdiff (cm )
2.46 × 10
−3
Mass spectrometer
λMS (s–1)
γMS
LMS
1.90 × 10
1.90 × 10
2.39 × 10−3
9.0 × 10−5
0.3
2800
1.5 × 10−4
1.0
6000
9.0 × 10−5
1.5
1500
1
1 cc/min
2 cc/min
5 cc/min
/
/
T = 35°C
1% CO2
10 cc/min
20 cc/min
40 cc/min
Simulated blanks
T = 35°C
1% CO2
simulated
0.1
c/c0
0.1
c/c0
2.50 × 10
2.46 × 10
2
D∕Ldiff
(s–1)
1
−2
0.01
0.001
0.01
0.001
1E-4
1E-4
0
200
400
600
t(s)
800
1000
1200
0
100
200
300
400
500
600
700
800
900
1000
t(s)
Fig. 9 Experimental blank runs at 1% CO2 in He and simulated fits using convolution theorem, now including kinetic limitations from the mass
spectrometer
13
Adsorption (2021) 27:129–145
137
1
1
1 cc/min
2 cc/min
5 cc/min
/
/
T = 35°C
1% N2
Simulated
0.1
c/c0
c/c0
0.1
10 cc/min
20 cc/min
40 cc/min
/
/
Simulated
T = 35°C
1% N2
0.01
0.001
0.01
0.001
1E-4
1E-4
0
200
400
600
0
100
200
t(s)
300
400
500
t(s)
Fig. 10 Experimental blank runs at 1% N2 in He and simulated fits using the convolution theorem
1
1
1 cc/min
2 cc/min
5 cc/min
/
/
T = 35°C
10% CO2
0.1
simulated
10 cc/min
20 cc/min
40 cc/min
/
/
Simulated blanks
0.1
c/c0
c/c0
T = 35°C
10% CO2
0.01
0.001
0.01
0.001
1E-4
1E-4
0
300
600
900
1200
0
100
200
300
400
t(s)
500
600
700
800
900
1000
t(s)
Fig. 11 Experimental blank runs at 10% CO2 in He and simulated fits using the convolution theorem
1
1
5 Ncm3/min
10 Ncm3/min
20 Ncm3/min
/
/
Simulated blanks
5 Ncm3/min
10 Ncm3/min
20 Ncm3/min
/
/
Simulated blanks
0.1
0.1
T = 35°C
33% CO2
p = 3.0 bar
c/c0
c/c0
T = 35°C
50% CO2
p = 2.1 bar
0.01
0.01
0.001
0.001
0
100
200
300
400
500
600
700
800
t(s)
0
100
200
300
400
500
600
700
800
t(s)
Fig. 12 Blanks at p = 2.1 and p = 3.0 bar and model fits. The only model parameter which requires adjusting is the diffusivity of CO2/He in teejunction 1
3.3 Model transferability
The advantage of using the method of convolution is that
it is based on the physical properties of the experimental
setup, i.e. dimensions of parts, molecular diffusivities etc.,
and should therefore be transferable to describe results
obtained on other setups with a similar configuration. Figure 13a shows a blank response in 1% CO2/He on a different
13
138
Adsorption (2021) 27:129–145
1
1
5 cc/min
10 cc/min
20 cc/min
/
/
Simulated
Alternative setup
T = 35°C
1% CO2 in He
0.1
(b)
c/c0
c/c0
0.1
(a)
10 cc/min
20 cc/min
40 cc/min
/
/
Simulation
Alternative setup
T = 30°C
pH2O = 2.1 kPa
0.01
0.01
0.001
1E-4
0.001
0
100
200
300
400
0
300
600
t(s)
900
1200
1500
t(s)
Fig. 13 Blank runs on a different setup with a similar configuration, but different components in 1% CO2/He (a) and 2% H2O/He atmospheres
(b)
Table 4 Model parameters for a different experimental setup, comprising narrow bore 1/16″ pipes
Model parameter
1% CO2 in He
2% H2O in He
Total volume in flow (cm3)
2.50 × 10−1
2.50 × 10−1
Slab diffusive pocket
Mixing volume (cm3)
2.80 × 10−4
Diffusive volume (cm3)
4.26 × 10−1
2
–1
D/L (s )
Mass spectrometer
λMS (s–1)
γMS
LMS
3.4 Deconvolution of the experimental signal
−4
1.03 × 100
3.49 × 10
2.50 × 10−4
9.0 × 10−5
1.0
7.0 × 10−5
0.24
1.5 × 104
9.0 × 102
experimental setup in our lab, which comprises narrow bore
1/16″ pipes and a different mass spectrometer, for obvious
reasons (although same make and model as previously). Due
to the narrow bore pipes, the diffusive flux into Tee-junction
1 is reduced, as evidenced by a shallower long time slope of
the diffusive tail. The model fits the data well, after adjustment of the relevant parameters, see Table 4. This particular
setup is also routinely used for water vapour studies. Figure 13b shows the blank runs for a 2% H2O/He mixture. It
is evident that these blanks look significantly different from
those run with typical gas mixtures. They exhibit less flow
rate dependence and more accumulation (the overall blank
volume of H2O is roughly an order of magnitude larger than
that of CO2). The model cannot capture the behaviour of this
blank in its entirety, but does fit the data accurately at short
and long times, the latter being of particular importance for
kinetic analysis. The discrepancy at intermediate times is
probably due to water accumulating on the surface of pipes,
which would introduce some axial dispersion. Such an effect
could be introduced to the model if necessary. In order to fit
the data, the model parameters reflect extra accumulation
13
in the solenoid valves as well as in the mass spectrometer,
as seen in Table 4. The fits resulting from not including this
additional accumulation can be found in the supplementary
material.
Evidently, the presence of diffusive pockets within the setup
can lead to erroneous interpretation when measuring kinetics or equilibrium behaviour of adsorbents. However, now
that the system and the effects of its constituent parts on the
concentration profile have been accurately modelled, these
effects can be removed from any experimental signal by the
process of deconvolution.
As an example, we are describing the process of deconvolution of the mass spectrometer effect (MS). For convenience(we
) will call the concentration at the MS at time ti , Fi =
CMS ti , whereas the concentration leaving Tee Junction
2, CT2 , is in this case the quantity of interest. And recognising that the∞transfer function is simply a summation of inte∑
an e−bn t , we have for integral n at time ti:
grals, i.e.
n=1
ti
−bn ti
Fi = an e
∫
ebn u CT2 (u)du
(25)
0
It is similarly easy to derive that at time ti+1 = ti + Δt:
ti+1
−bn ti+1
Fi+1 = an e
∫
ebn u CT2 (u)du
(26)
0
ti+1
⎡
−bn Δt ⎢ Fi
Fi+1 = an e
⎢ an
⎣
+
⎤
ebn (u−ti ) CT2 (u)du⎥
∫
⎥
⎦
ti
(27)
Adsorption (2021) 27:129–145
139
We can now approximate the concentration leaving Tee
Junction 2 by a linear function, which should be accurate for
small time steps, Δt , i.e.:
( )
( )
( ) CT2 ti+1 − CT2 ti (
)
(28)
CT2 (u) = CT2 ti +
u − ti
Δt
Then through integration by parts:
Fi+1
clearly show the effect of the two diffusive elements. The
deconvolution process can be carried out consecutively, to
remove both contributions. Figure 14b shows how after consecutively removing both diffusive contributions, the only
remaining part are now the empty (no adsorbent) mixing
volumes in flow, as evidenced by essentially a single exponential decay (i.e. a straight line in the semi-log plot). This
ti+1
⎧
⎫
�
�ti+1
F
⎪
1
1 bn (u−ti ) dCT2 ⎪
i
−bΔt
bn (u−ti )
= an e
e
−
du⎬
CT2 (u)
⎨a + b e
∫ bn
du
n
ti
⎪ n
⎪
ti
⎩
⎭
{
−bn Δt
Fi+1 = an e
}
( )
( )
)
( )
( )] CT2 ti+1 − CT2 ti ( b Δt
Fi
1 [ bn Δt
e CT2 ti+1 − CT2 ti −
+
en − 1
an b n
b2n Δt
Extending this to N exponentials finally yields:
Fi+1 =
(29)
N
∑
{
( ) −b Δt an [ ( )
( )
]
Fi n e n +
CT2 ti+1 − CT2 ti e−bn Δt
bn
n=1
}
( ( )
( ))
an CT2 ti+1 − CT2 ti (
)
−bn Δt
1−e
−
b2n Δt
(31)
Fi+1 and Fi are known from the experimental signal, whereas
CT2 (t = 0) = 1. CT2 (t = Δt) can thus be calculated initially,
and consecutive concentrations at Tee Junction 2 can be
solved iteratively. This method of deconvolution can also
be found in ref.(Brandani 1998).
Figure 14a shows how both the mass spectrometer effect
and the effect of the low volume tee can be removed from
a model blank curve using this methodology. Here, a different set of modelling parameters has been used, to more
(30)
contribution from elements in flow may also be removed
as desired in a similar fashion, but this is less critical for
kinetic analysis. When ZLC or breakthrough experiments
are performed to rapidly assess an adsorbent’s capacity for
adsorption, this gas volume in flow can easily be subtracted
from the overall eluted volume. For experiments at low flow
rates however, which can be used to measure equilibrium
data (Brandani et al. 2003), it may be desirable to remove it,
as the initial curvature of the desorption curve will lead to
erroneous calculation of isotherm points (see supplementary
material).
As mentioned previously, the process of convolution and
thus deconvolution is a symmetrical one, which means that
the order in which various contributions to the overall signal
are removed is inconsequential as long as each component
is a linear system. When dealing with actual adsorbents
1
1
Original signal
Signal after deconvolution of Low V Tee
Signal after deconvolution of MS
Original signal
Signal after deconvolution of LVT/MS
After 2 consecutive deconvolutions
/
/
0.1
0.1
F = 10 cc/min
c/c0
c/c0
F = 10 cc/min
0.01
0.001
0.01
0.001
(a)
(b)
1E-4
1E-4
0
20
40
60
80
100
t (s)
Fig. 14 Separate deconvolution of mass spectrometer and low volume
tee contributions from model blank curve (a). Consecutive deconvolution of both mass spectrometer and low volume tee contributions
0
20
40
60
80
100
t (s)
from model blank curve in any order yields a simple exponential
decay, corresponding to mixing volumes in flow (b)
13
140
Adsorption (2021) 27:129–145
however, some non-linearity may arise from the isotherms
under consideration, which means that the ZLC itself is not
a linear system. All the flow components remain linear, and
signal linearity is an essential requirement for any detector.
The effect of non-linearity will be discussed in a following
section.
4 Deconvolution of actual experimental
data—case studies
In order to apply the deconvolution procedure to actual
experimental data without creating numerical issues, the
raw data were first smoothened using polynomial fits. This
can be done with predefined functions in most numerical
software tools. For example in Mathcad the built in ‘regress’
function can be used to this end on a number of data segments, with a minimal polynomial order of 5. The final longtime approach to zero is always an exponential function,
which can be identified easily. Naturally, this initial data
treatment procedure results in curve smoothing, eliminating
most of the noise, which is typical for un-corrected data.
4.1 Case study 1—CO2 adsorption on zeolite 13X,
an adsorbent with large uptake
The first case study is of experimental adsorption data
on zeolite 13X, measured in 10% CO2/He atmosphere at
35 °C, at ambient pressure. The sample is a single bead with
a diameter of 2.0 mm from UOP, Honeywell. Under the
measurement conditions, this material has appreciable CO2
uptake, i.e. in excess of 2 mol kg–1 (Park et al. 2016). As
1
Original signal (norm.)
Corrected signal (norm.)
Blank
Zeolite 13X
T = 35°C
10% CO2
c/c0
0.1
F = 10 cc/min
shown in Fig. 15 the difference between the response of such
a sample with a typical mass of ~ 6 mg and that of the blank
run is substantial, and as such the response is not affected to
a significant extent by extra-column effects. Consequently,
carrying out the deconvolution procedure only leads to small
changes in the desorption curve for the sample. Importantly
too, the long-time asymptote remains unchanged. This part
is often critical in determining diffusional time constants,
especially in non-linear systems, and thus using the original normalised data would not lead to an erroneous interpretation (Hu et al. 2014). An additional system with large
uptake, i.e. silica gel in water vapour, can be found in the
supplementary material.
4.2 Case study 2—CO2 adsorption on HISIV 3000,
small uptake, fast kinetics
In the second case study we consider a sample with small
uptake of adsorbate under the measurement conditions.
HISIV 3000, a commercial silicalite from UOP, Honeywell
(UOP 2020), is a high silicon molecular sieve material and
due to its low aluminium content, interaction with polar
molecules, such as CO2 is low. The normalised desorption
curves at different flow rates, as seen in Fig. 16, all show a
long-time tail, which at first sight could be interpreted as
being related to kinetics of CO2 transport through the sample. Also, when plotted versus Flowrate × time, a so called
Ft plot, Fig. 16, it is clear that the curves are not overlapping,
even at low flow rates, which would be indicative of flow
rate dependency and thus kinetic limitations. However, when
the curves are plotted together with the blank runs under the
same conditions, it becomes evident that this long-time tail
is caused by the experimental setup itself, Fig. 17. Performing the deconvolution procedure removes this contribution
and leaves the actual gas phase concentration leaving the
ZLC column, allowing for unambiguous determination of
the sample’s actual kinetics. Now, when plotting the deconvoluted signals versus Flowrate × time , it is clear that the
desorption process is in fact independent of flow rate and
thus controlled by equilibrium conditions, and therefore too
fast to be determined under the measurement conditions.
0.01
4.3 Case study 3—water adsorption on SBA‑15;
effect of isotherm shape and different sample
masses
0.001
0
200
400
600
800
t(s)
Fig. 15 ZLC desorption for zeolite 13X in 10% CO2 in He) at 35 °C.
Due to significant adsorption capacity under these conditions, the
measured signal is far removed from the blank run without adsorbent.
The deconvolution procedure therefore only has a small effect on the
normalised signal
13
Equilibrium and kinetic water adsorption measurements
can be time consuming, especially when done gravimetrically or volumetrically. The ZLC method was shown to be
a good alternative technique, which can be up to an order
of magnitude faster, especially when small sample masses
are used (Centineo and Brandani 2020). Figure 18a shows
the desorption runs at a flowrate of 0.5 cm3 min–1 for two
Adsorption (2021) 27:129–145
141
1
1
2 cc/min
5 cc/min
10 cc/min
T = 20°C
10% CO2
1 cc/min
2 cc/min
5 cc/min
T = 20°C
10% CO2
c/c0
c/c0
0.1
0.1
0.01
0.001
0.01
0
100
200
300
0
2
4
t (s)
6
8
10
Ft (cc)
Fig. 16 Uncorrected ZLC desorption data for HISIV 3000 in 10% CO2 at 20 °C showing long-time tail (left). Uncorrected curves in the Ft plot
are not overlapping, suggesting measurements are performed under kinetic control (right)
1
1
Original signal (norm.)
Corrected signal (norm.)
Blank
1 cc/min
2 cc/min
5 cc/min
T = 20°C
10% CO2
T = 20°C
10% CO2
0.1
c/c0
c/c0
F = 5 cc/min
0.1
0.01
0.001
0.01
0
50
100
150
0
1
2
t (s)
3
4
5
Ft (cc)
Fig. 17 Original and corrected signal with blank run at 5 cc/min,
showing diffusive long-time tail is predominantly caused by extra column effects and is removed after deconvolution procedure (left). The
Ft plot of deconvoluted curves are overlapping, indicating measurements are under equilibrium control (right)
1
1
0.5 mg - no deconvolution
1.5 mg - no deconvolution
Blank run
0.1
SBA-15
T = 35°C
(a)
(b)
SBA-15
T = 35°C
c/c0
c/c0
0.1
0.5 mg - no deconvolution
0.5 mg - after deconvolution
1.5 mg - no deconvolution
1.5 mg - after deconvolution
0.01
0.01
0.001
0.001
0
1
2
3
t (h)
4
5
0
1
2
3
4
5
t (h)
Fig. 18 Desorption for SBA-15 in 5% H2O / He atmosphere at 35 °C for two sample masses. The desorption data for 0.5 mg sample is close to
the blank run at long times and so needs to be corrected
13
142
Adsorption (2021) 27:129–145
1
0.6
0.5 mg - No deconvolution
0.5 mg - After deconvolution
1.5 mg - No deconvolution
Aquadyne
(Na,TEA)-ZSM-25
T = 35°C
20% CO2
F = 1 cc/min
0.1
0.4
SBA-15
T = 35°C
0.3
c/c0
n (kg kg —1)
0.5
0.01
0.2
0.1
0.001
0.0
0.0
0.2
0.4
0.6
0.8
p/p0
Fig. 19 Calculated isotherms from ZLC data. Using uncorrected data
for the 0.5 mg sample leads to erroneous behaviour in the Henry’s
Law region. Applying deconvolution yields the correct behaviour as
found for by gravimetric means (Aquadyne) or a larger sample mass
different amounts of mesoporous SBA-15 at 35 °C, 0.5 mg
and 1.5 mg, respectively (details for this adsorbent can be
found in ref. (Centineo and Brandani 2020)). Under these
conditions, water desorption is an equilibrium process and
the data can thus be used to calculate the relevant isotherm.
Upon inspection however, the run for 0.5 mg shows significant overlap with the blank run at low concentrations. When
performing deconvolution on the original signals (Fig. 18b),
it is clear that the long-time slope is affected, whereas the
initial part of the desorption remains largely unchanged.
Deconvolution does not affect the experimental signal for
the larger sample of 1.5 mg, which is as expected since it
is always far removed from the blank run. Figure 19 shows
the calculated isotherms from both un-corrected and corrected desorption signals, as well as an isotherm measured
by gravimetric means. Whereas the isotherm from the uncorrected 1.5 mg sample corresponds well with the gravimetric data, the Henry’s Law region of the un-corrected
0.5 mg sample shows some deviation. After correction,
all isotherms match well. In this particular case, due to the
adsorbent’s unusual isotherm, caused by hydrophobicity at
low concentrations with subsequent capillary condensation,
uptake alone is not sufficient in judging whether the correction procedure is necessary and visual comparison with the
blank run is a better indicator.
4.4 Case study 4—effect of non‑linearity
on deconvolution procedure
As mentioned earlier, the convolution integral is a linear
operator and as long as the various contributions to the overall system are linear in nature, the convolution integral yields
13
0
500
1000
1500
2000
t(s)
Fig. 20 Uncorrected desorption data for (Na,TEA)-ZSM-25 at 35 °C
in 20% CO2 / He. The low flow rate of 1 cc/min was used, at which
the desorption is assumed to occur under equilibrium conditions
the correct response, as measured at the outlet of the system, i.e. at the detector. This is evidently true for the blank
runs, as these were shown to consist of linear processes such
as CSTRs, and diffusing elements with linear isotherms.
When considering actual measurements including a column
packed with adsorbent, non-linearity may now arise from the
adsorption system being studied. This will have the effect
that the order in which the various contributions take place
will actually matter. This is not seen in the previous example, as the deconvolution affects only the initial part of the
isotherm which is linear. To illustrate the effect of nonlinearity, we consider another non-linear adsorption system, i.e.
CO2 on templated (Na,TEA)-ZSM-25. This system exhibits
a stepped isotherm and constitutes a significant deviation
from linearity at 20% CO2 and 35 °C (as described in detail
in Min et al. 2018; Verbraeken et al. 2020)). The ZLC desorption data for a measurement at low flow rate is shown
in Fig. 20. The system’s response post-column is linear, so
we may calculate the outlet concentration of the ZLC by
removing the contributions of the detector and Tee-junction
2 alone through the deconvolution procedure. Similarly, the
column’s inlet concentration can be calculated by applying
the convolution integral to all the contributions occurring
prior to the column, i.e. the step change in concertation and
Tee-junction 1. Due to the low flow rate used during desorption, the gas phase concentration remains in equilibrium
with the adsorbed phase concentration and in turn, each
point on the ZLC curve would correspond to a point on the
isotherm. The inlet and outlet concentration may thus be
used to calculate the isotherm for this adsorption system
through the mass balance, in accordance with:(Brandani
et al. 2003)
Adsorption (2021) 27:129–145
143
3.0
1.0
Deconvolution of outlet signal only
Convolution of inlet concentration
Complete deconvolution
0.8
2.5
2.0
n (mol kg—1)
(a)
0.6
(Na,TEA)-ZSM-25
T = 35°C
c/c0
(Na,TEA)-ZSM-25
T = 35°C
0.4
0.2
1.5
(b)
1.0
0.5
Convolution of inlet / deconvolution of outlet
Complete deconvolution
0.0
0.00
0.0
0
2
4
6
8
0.05
0.10
0.15
0.20
pCO2 (bar)
Ft (cc)
Fig. 21 Two different ways of calculating isotherms from equilibrium desorption data (a). A ‘straight’ deconvolution of all extra column components and subsequent integration of the area is indicated
by the green curve and area. The red curve indicates the signal after
deconvolution of the components post-column, whereas convolution
of the components pre-column yield the inlet concentration shown by
the blue line; subsequent integration is indicated by the red patterned
area. The resulting isotherms are shown in b
∞
⎧ ∞
⎫
�
Fin cT ⎪ yout − yin
yout − yin ⎪ Vf �
q =
dt
−
dt
− cT yout − yin
⎨
⎬
∫ 1 − yout ⎪ Vs
Vs ⎪∫ 1 − yout
t
⎩0
⎭
0.4
∗
0.3
Full deconvolution
Convolution inlet /
deconvolution outlet
Golden/Sircar
Here q∗ is the equilibrium amount adsorbed in mol m–3,
P
Fin the inlet flowrate, cT = RT
, Vs volume of the solid, Vf the
fluid volume, and yout and yin the mole fraction leaving the
column, respectively. The resulting isotherm can be found
in Fig. 21. Also shown in this figure is the isotherm calculated by simply removing all extra-column contributions
through the deconvolution procedure, treating the system
as a linear one. It can be seen that the difference between
the two methods is in fact small even for a system which
is highly non-linear, indicating that the full deconvolution
procedure should be suitable for most adsorption systems.
It is finally important to note that the long-time behaviour
is always in the linear regime, due to the low concentrations
involved. The effect of non-linearity and thus order of carrying out the deconvolution will therefore have an insignificant
impact on this part of the experimental data. The long-time
analysis on data which has received a full deconvolution
procedure should thus lead to the correct kinetics/Henry’s
law constants.
In contrast, when these two methods for isotherm calculation are carried out for similar desorption data for HISIV
3000 in 10% CO2 / He at 20 °C, it can be seen that the resulting two isotherms are practically identical, Fig. 22. This is
to be expected, as this adsorption system is essentially linear for the concentration range studied. Isotherm data on
silicalite from literature is also shown in Fig. 22, showing
that the corrected data yields the correct isotherm (a binder
content of 23 wt.% has been used (Brandani et al. 2016) to
n (mol/kg)
(32)
0.2
HISIV 3000
T = 20°C
0.1
0.0
0.00
0.02
0.04
0.06
0.08
p (bar)
Fig. 22 Isotherms calculated by two methods for HISIV 3000 in
CO2 at 20 °C. Both isotherms are identical as this adsorption system
is linear under the measurement conditions. Included are calculated
isotherm data by Golden and Sircar (Golden and Sircar 1994), multiplied by 0.77 to account for 23 wt.% binder content
compare the isotherm measured on the pellets with the absolute isotherm in silicalite crystals (Golden and Sircar 1994)).
5 Conclusions
We have developed a correction procedure for adsorption
column measurements, which accurately takes into account
the extra-column effects due to the experimental setup configuration. The model is based on the physical components
and layout of the setup and it is shown that their composite
13
144
behaviour can be described by the method of convolution.
The physical meaningfulness of the model is exemplified by
the fact that the model is valid for a range of conditions (e.g.
flow rates, pressures, gas compositions) and is transferable
to other experimental setups. By carrying out the convolution procedure in reverse on actual experimental data with
adsorbent present, the corrected signal is obtained which
can be modelled using standard tools for analysing kinetics
and equilibrium.
Carrying out this correction procedure is of particular
importance for experiments involving samples with small
uptake. This includes samples which inherently have small
uptake of the dosing component, but is also applicable when
experiments are run with small quantities of adsorbent,
which may be required for kinetic studies. Since the correction procedure with the algorithm of Brandani (1998)
is computationally very light, it would be advisable to perform it for any measurement for which the ratio of eluted
volumes with and without adsorbent is smaller than 4 – 5.
Under these conditions, the extra-column effect will start to
affect the shape of the desorption curves, which may lead to
erroneous interpretations. It has also been shown how the
procedure can be applied to systems that exhibit non-linear
isotherms.
In this study, we have exclusively shown how the procedure can be applied to ZLC experiments, but it should
be noted that it is equally applicable to columns of larger
dimensions. Indeed, the convolution model successfully fits
blank experiments using our extended columns (described in
refs. Georgieva et al. 2019; Lozinska et al. 2012)), without
the need for additional terms.
As a final consideration, the deconvolution procedure can
be applied to the experimental data to provide a corrected
set of inlet and outlet concentrations for a ZLC experiment.
These can then be used to extract physical parameters. For
nonlinear systems, where this has to be done from a numerical solution of the model equations, it is also possible to
add the parametrised models of the individual components,
including the detector, to the numerical solution and regress
the raw data directly, as similarly done in ref.(Friedrich et al.
2015). While the two approaches should be equivalent, the
fact that the deconvolution process removes the noise from
the signal makes it preferable.
Acknowledgements The authors would like to acknowledge funding
from the Engineering and Physical Sciences Research Council, Grants
EP/N033329/1 (Cation controlled gating for Selective Gas Adsorption over Adaptable Zeolites) and EP/N024613/1 (Versatile Adsorption
Processes for the Capture of Carbon Dioxide from Industrial Sources
– FlexICCS).
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long
as you give appropriate credit to the original author(s) and the source,
13
Adsorption (2021) 27:129–145
provide a link to the Creative Commons licence, and indicate if changes
were made. The images or other third party material in this article are
included in the article’s Creative Commons licence, unless indicated
otherwise in a credit line to the material. If material is not included in
the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will
need to obtain permission directly from the copyright holder. To view a
copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
References
Arora, P.S., Dunlop, P.J.: The pressure dependence of the binary diffusion coefficients of the systems He–Ar, He–N2, He–O2, and He–
CO2 at 300 and 323 K: tests of Thorne’s equation. J. Chem. Phys.
71, 2430–2432 (1979). https://doi.org/10.1063/1.438648
Brandani, F., Ruthven, D., Coe, C.G.: Measurement of adsorption
equilibrium by the zero length column (ZLC) technique part 1:
single-component systems. Ind. Eng. Chem. Res. 42, 1451–1461
(2003). https://doi.org/10.1021/ie020572n
Brandani, S.: Analysis of the piezometric method for the study of diffusion in microporous solids: isothermal case. Adsorption. 4, 17–24
(1998). https://doi.org/10.1023/A:1008831202564
Brandani, S., Mangano, E.: The zero length column technique to measure adsorption equilibrium and kinetics: lessons learnt from 30
years of experience. Adsorption. (2020). https://doi.org/10.1007/
s10450-020-00273-w
Brandani, S., Mangano, E., Sarkisov, L.: Net, excess and absolute
adsorption and adsorption of helium. Adsorption. 22, 261–276
(2016). https://doi.org/10.1007/s10450-016-9766-0
Centineo, A., Brandani, S.: Measurement of water vapor adsorption
isotherms in mesoporous materials using the zero length column technique. Chem. Eng. Sci. 214, 115417 (2020). https://doi.
org/10.1016/j.ces.2019.115417
Delley, R.: Modifying the Gaussian peak shape with more than one
time constant. Anal. Chem. 58, 2344–2346 (1986)
Ellis, C.S., Holsen, J.N.: Diffusion coefficients for He-N2 and N2-CO2
at elevated temperatures. Ind. Eng. Chem. Fundam. 8, 787–791
(1969). https://doi.org/10.1021/i160032a030
Friedrich, D., Mangano, E., Brandani, S.: Automatic estimation
of kinetic and isotherm parameters from ZLC experiments.
Chem. Eng. Sci. 126, 616–624 (2015). https://doi.org/10.1016/j.
ces.2014.12.062
Georgieva, V.M., Bruce, E.L., Verbraeken, M.C., Scott, A.R., Casteel, W.J., Brandani, S., Wright, P.A.: Triggered gate opening and
breathing effects during selective CO2 adsorption by merlinoite
zeolite. J. Am. Chem. Soc. 141, 12744–12759 (2019). https://doi.
org/10.1021/jacs.9b05539
Golden, T.C., Sircar, S.: Gas adsorption on silicalite. J. Colloid
Interface Sci. 162, 182–188 (1994). https ://doi.org/10.1006/
jcis.1994.1023
Gritti, F., Felinger, A., Guiochon, G.: Influence of the errors made in
the measurement of the extra-column volume on the accuracies of
estimates of the column efficiency and the mass transfer kinetics
parameters. J. Chromatogr. A 1136, 57–72 (2006). https://doi.
org/10.1016/j.chroma.2006.09.074
Guntuka, S., Farooq, S., Rajendran, A.: A- and B-site substituted lanthanum cobaltite perovskite as high temperature oxygen sorbent 2
column dynamics study. Ind. Eng. Chem. Res. 47, 163–170 (2008)
Hu, X., Brandani, S., Benin, A.I., Willis, R.R.: Development of a
semiautomated zero length column technique for carbon capture
applications: rapid capacity ranking of novel adsorbents. Ind. Eng.
Chem. Res. 54, 6772–6780 (2015). https://doi.org/10.1021/acs.
iecr.5b00513
Adsorption (2021) 27:129–145
Hu, X., Mangano, E., Friedrich, D., Ahn, H., Brandani, S.: Diffusion
mechanism of CO2 in 13X zeolite beads. Adsorption. 20, 121–135
(2014). https://doi.org/10.1007/s10450-013-9554-z
Joss, L., Mazzotti, M.: Modeling the extra-column volume in a small
column setup for bulk gas adsorption. Adsorption. 18, 381–393
(2012). https://doi.org/10.1007/s10450-012-9417-z
Kaltenbrunner, O., Jungbauer, A., Yamamoto, S.: Prediction of the preparative chromatography performance with a very small column.
1996 Int. Symp. Prep. Chromatogr. 760, 41–53 (1997)
Lozinska, M.M., Mangano, E., Mowat, J.P.S., Shepherd, A.M., Howe,
R.F., Thompson, S.P., Parker, J.E., Brandani, S., Wright, P.A.:
Understanding carbon dioxide adsorption on univalent cation
forms of the flexible zeolite rho at conditions relevant to carbon
capture from flue gases. J. Am. Chem. Soc. 134, 17628–17642
(2012). https://doi.org/10.1021/ja3070864
Massman, W.J.: A review of the molecular diffusivities of H2O, CO2,
CH4, CO, O3, SO2, NH3, N2O, NO, and NO2 in air, O2 and N2
near STP. Atmos. Environ. 32, 1111–1127 (1998). https://doi.
org/10.1016/S1352-2310(97)00391-9
Min, J.G., Kemp, K.C., Lee, H., Hong, S.B.: CO2 adsorption in the
RHO family of embedded isoreticular zeolites. J. Phys. Chem. C
122, 28815–28824 (2018). https://doi.org/10.1021/acs.jpcc.8b099
96
Paganelli, C.V., Kurata, F.K.: Diffusion of water vapor in binary
and ternary gas mixtures at increased pressures. Spec. Issue
HonorProfr. Hermann Rahn. 30, 15–26 (1977). https ://doi.
org/10.1016/0034-5687(77)90018-4
Park, Y., Ju, Y., Park, D., Lee, C.-H.: Adsorption equilibria and kinetics
of six pure gases on pelletized zeolite 13X up to 1.0MPa: CO2,
CO, N2, CH4, Ar and H2. Chem. Eng. J. 292, 348–365 (2016).
145
Rajendran, A., Kariwala, V., Farooq, S.: Correction procedures for
extra-column effects in dynamic column breakthrough experiments. Chem. Eng. Sci. 63, 2696–2706 (2008). https ://doi.
org/10.1016/j.ces.2008.02.023
Ruthven, D.M.: Principles of Adsorption and Adsorption Processes.
Wiley, New York (1984)
Schwertz, F.A., Brow, J.E.: Diffusivity of water vapor in some common gases. J. Chem. Phys. 19, 640–646 (1951). https ://doi.
org/10.1063/1.1748306
Stephanopoulos, G.: Chemical Process Control: An Introduction to
Theory and Practice. PTR Prentice Hall, New Jersey (1983)
Swagelok: Swagelok (R), www.swagelok.com
UOP, H.: Honeywell, UOP, uop.honeywell.com
Verbraeken, M.C., Mennitto, R., Georgieva, V.M., Bruce, E.L., Greenaway, A.G., Cox, P.A., Gi Min, J., Bong Hong, S., Wright, P.A.,
Brandani, S.: Understanding CO2 adsorption in a flexible zeolite through a combination of structural, kinetic and modelling
techniques. Sep. Purif. Technol. (2020). https://doi.org/10.1016/j.
seppur.2020.117846
Walker, R.E., deHaas, N., Westenberg, A.A.: Measurements of multicomponent diffusion coefficients for the CO2–He–N2 system using
the point source technique. J. Chem. Phys. 32, 1314–1316 (1960).
https://doi.org/10.1063/1.1730915
Wilkins, N.S., Rajendran, A., Farooq, S.: Dynamic column breakthrough experiments for measurement of adsorption equilibrium
and kinetics. Adsorption. (2020). https://doi.org/10.1007/s1045
0-020-00269-6
Publisher’s Note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.
13