You are on page 1of 8

JOURNAL OF APPLIED PHYSICS 117, 244304 (2015)

Design of nanoporous materials with optimal sorption capacity


Xuan Zhang,1 Koki Urita,2 Isamu Moriguchi,2 and Daniel M. Tartakovsky1,a)
1
Department of Mechanical and Aerospace Engineering, University of California, San Diego,
9500 Gilman Drive, La Jolla, California 92093, USA
2
Department of Applied Chemistry, Graduate School of Engineering, Nagasaki University,
Nagasaki 852-8521, Japan
(Received 11 May 2015; accepted 16 June 2015; published online 25 June 2015)
Modern technological advances have enabled one to manufacture nanoporous materials with a
prescribed pore structure. This raises a possibility of using controllable pore-scale parameters
(e.g., pore size and connectivity) to design materials with desired macroscopic properties
(e.g., diffusion coefficient and adsorption capacity). By relating these two scales, the homogeni-
zation theory (or other upscaling techniques) provides a means of guiding the experimental
design. To demonstrate this approach, we consider a class of nanoporous materials whose pore
space consists of nanotunnels interconnected by nanotube bridges. Such hierarchical nanoporous
carbons with mesopores and micropores have shown high specific electric double layer capaci-
tances and high rate capability in an organic electrolyte. We express the anisotropic diffusion
coefficient and adsorption coefficient of such materials in terms of the tunnels properties (pore
radius and inter-pore throat width) and their connectivity (spacing between the adjacent tunnels
and nanotube-bridge density). Our analysis is applicable for solutes that undergo a non-
equilibrium Langmuir adsorption reaction on the surfaces of fluid-filled pores, but other homoge-
neous and heterogeneous reactions can be handled in a similar fashion. The presented results can
be used to guide the design of nanoporous materials with optimal permeability and sorption
capacity. VC 2015 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4923057]

I. INTRODUCTION with a variety of upscaling techniques, including volume


averaging15 and homogenization via multiple-scale expan-
Small pore sizes and large surface porosity are some of
sions.1618 We use the latter approach to predict anisotropic
the unique properties of nanoporous materials, which can be
diffusion coefficients and effective adsorption rates for a
used in a wide range of applications involving ion exchange,
class of nanoporous materials,19 whose synthesis allows one
adsorption, sensing, and catalysis.13 Nanoporous materials
to control their pore structure. It is worthwhile emphasizing
play central role in biomedical diagnostics, such as combina-
that the use of various homogenization techniques to derive
torial biochemistry on-a-chip, chromatography, biosensors,
effective (continuum- or Darcy-scale) models of reactive
cell manipulation, and DNA analysis.46 Because of their
transport in porous media (primarily subsurface environ-
high electrical conductivity and wettability towards electro-
ments) is not new.2023 The novelty of our study comes from
lytes, nanoporous materials might significantly boost the per-
using such techniques to optimize macroscopic properties of
formance of energy storage devices.79
a material by using its microscopic properties as decision
Desirable properties of manufactured (nano)porous
variables.
materials can be enhanced by controlling their pore structure
The effective reaction rate in a given solid-solution sys-
and physicochemical properties of emerging pore net-
tem is an essential factor for sorption-system design. Our anal-
works.10 For instance, introduction of active sites onto pore
ysis is applicable for solutes that undergo a non-equilibrium
surfaces of mesoporous metallosilicate materials is a promis-
Langmuir-type adsorption reaction on the surfaces of fluid-
ing venue for designing novel adsorbents and catalysis.11
filled pores, but other homogeneous and heterogeneous reac-
Embedment of microporous tubes12 between layers of
tions can be handled in a similar fashion. The presented results
macro- and nanopores results in new materials with superior
can be used to guide the design of nanoporous materials with
mechanical and thermal properties.13 For example, hierarchi-
optimal permeability and sorption capacity.
cal nanoporous carbons with mesopores and micropores
In Sec. II, we formulate a non-equilibrium pore-scale
have shown high specific electric double layer capacitances
and high rate capability in an organic electrolyte. model, which is based on the first-order theory of sorption
Quantitative predictions of macroscopic properties of kinetics.2426 The corresponding continuum-scale model is
such new materials (e.g., their sorption capacity or thermal derived in Sec. III. This model relates the macroscopic prop-
conductivity) require the ability to relate these properties to erties of a nanoporous material (i.e., its diffusion coefficient
their pore-scale counterparts, such as a pore-size distribution and adsorption capacity) to its pore-scale counterparts (e.g.,
and pore-network connectivity.14 This can be accomplished the radii of micro- and macropores, and the size of nano-
bridges). In Sec. IV, we present simulations results and dis-
cuss their implications for material design. Our findings are
a)
Electronic mail: dmt@ucsd.edu summarized in Sec. V.

0021-8979/2015/117(24)/244304/8/$30.00 117, 244304-1 C 2015 AIP Publishing LLC


V
244304-2 Zhang et al. J. Appl. Phys. 117, 244304 (2015)

II. PROBLEM FORMULATION K^ c^


s^eq ; (4)
Consider a nanoporous material X ^ with a characteristic 1 K^ c^
^
length L. Let P denote the part of this material occupied by
where K^ (L3/mol) is the adsorption equilibrium constant.
nanopores whose characteristic length scale, e.g., a typical pore
The derivation of the reduced-complexity model is facili-
diameter, is l such that ! ! l=L " 1. The impermeable solid
^ occupies the rest of the nanoporous material, i.e., tated by rewriting this pore-scale boundary-value problem
skeleton S
^ ^ ^ The (multi-connected, smooth) boundary between (BVP) in the dimensionless form. Let us introduce dimen-
X P [ S.
^ and the solid skeleton S ^ is denoted by C.
^ sionless variables and parameters
the pore space P
^
The pore space P is occupied by a fluid, which contains a
x^ D^t c^ q^m
solute with concentration c^^x ; ^t [mol/L3], where x^ denotes a x ; t ; c ; qm ;
^ and ^t is time. The solute diffuses throughout the L L2 cin Lcin
point in P
^ and undergoes a heterogeneous reaction at the ^ L2 c
pore space P K c^in K; Da ; (5)
^ both taking place in isothermal condi- D
solid-fluid interface C,
tions. The former process is described by a diffusion equation where the Damkohler number Da represents the ratio of reac-
@^
c tion and diffusion time-scales. We show in Appendix A that,
^ 2 c^;
Dr ^;
x^ 2 P ^t > 0; (1)
@ ^t for nanoporous materials characterized by Da " 1, Eqs.
(2)(4) give rise to a nonlinear (but decoupled) boundary
where D [L2/T] is the diffusion coefficient. The pore spaces condition at the solid-fluid interface C
considered in our analysis consist of both interconnected
nanopores and nanotubes. The diffusive transport in the for- Kc
mer is due to molecular diffusion (with molecular diffusion &n ' rc Da qm ; x 2 C; t > 0: (6a)
1 Kc
coefficient Dm), while Knudsen diffusion (with diffusion
coefficient Dk) might take place in the latter. While the anal- The remainder of the pore-scale BVP takes the form of a dif-
ysis presented below is capable of handling spatially variable fusion equation
and anisotropic diffusion coefficients, we use the constant
diffusion coefficient D to simplify the presentation. @c
r2 c; x 2 P; t>0 (6b)
Mass conservation along the impermeable solid-fluid @t
interface C^ requires the normal component of the solute
mass flux &Dr^ ^ c to be balanced by the rate of change in the subject to the initial condition cx; 0 1 for x 2 P and the
adsorbed solute q^, i.e., corresponding boundary conditions on the bounding surface
of the nanoporous material X.
^ c @ q^ q^m @^
&Dn ' r^
s
; ^
x^ 2 C; ^t > 0; (2)
Solving this BVP in a large interconnected network of
@t ^ @ ^t nanopores, P, is neither feasible nor necessary. Instead,
one is often interested in the emerging (macroscopic) prop-
where q^ [mol/L2] is the adsorption amount per unit area of erties of the nanoporous material X as a whole. While its
the solid-fluid surface C; ^ q^m [mol/L2] is the maximal adsorp-
pore structure is described by diameters and a spatial
tion amount, s^^ ^ and
x ; ^t [-] is the fractional coverage of C, arrangement of individual nanopores and nanotubes
^
nx is the unit normal vector of C. Equation (1) is also sub- bridges, its macroscopic properties are characterized by
ject to an initial condition c^x; 0 cin , in which the initial bulk (average) quantities such as porosity /, effective dif-
concentration cin assumed to be spatially uniform. Finally, fusion coefficient Deff , and effective adsorption rate con-
appropriate boundary conditions are imposed on the bound- stant ceff . Connecting these two levels of description allows
ing surface of the nanoporous material X. ^
us to relate such macroscopic properties to the underlying
Previous efforts to derive effective (continuum-scale) microscopic (pore-scale) characteristics. This, in turn, ena-
descriptions of processes, which admit the pore-scale bles one to manufacture nanoporous materials with desired
representations (1) and (2), dealt with either equilibrium macroscopic attributes.
adsorption27,28 or general non-equilibrium adsorption By way of example, we consider nanoporous materi-
described by a coupled system of reaction-diffusion equa- als in which rows of interconnected uniform nanopores of
tions for c^ and s^.2023,29 While the former are not applicable radius R are bridged by regularly placed microporous
for many nanoporous phenomena, the latter are too com- tubes. The spacing between two neighboring pores in a
plex to be used in materials design. Instead, we derive a single row is expressed in terms of the angle h that we
reduced-complexity model based on the Lagergren refer to as an intersection factor. A number n of micropo-
(pseudo-first-order) rate equation3032 rous tubes (diameter d 0.7 nm) per unit cell is referred
s
d^ to as a microporous tube density. In the first class of
cs^eq & s^; (3) materials, the rows of interconnected nanopores are per-
d^t
fectly aligned (Fig. 1). In the second, the rows are shifted
where c (T&1) is the first-order adsorption rate constant and relative to each other by the distance R=2 (Fig. 6). These
s^eq is the adsorption coverage fraction at equilibrium. The two settings provide an adequate representation of the
latter is assumed to follow the Langmuir adsorption isotherm nanoporous materials.19
244304-3 Zhang et al. J. Appl. Phys. 117, 244304 (2015)


D " #
Deff I ry v dy: (10)
kUk PU

Here, I is the identity matrix, and the closure variable vy


is a U-periodic vector defined on P U . It satisfies

1
r2y v 0; y 2 P U ; hvyi ! vxdy 0 (11)
kUk P U

subject to the boundary condition along the fluid-solid seg-


ments CU of the boundary of P U
FIG. 1. A schematic representation of Material 1 (left) and the correspond-
ing unit cell U (right). The unit cell is a 2yc1 ) yc2 rectangle, with yc1 n ' ry v &n ' I; y 2 CU : (12)
R cos h and yc2 2R l.
Along the remaining (fluid) segments Cf of the boundary
III. MACROSCOPIC PROPERTIES OF NANOPOROUS of P U (see Fig. 1), the U-periodicity of vx is enforced.
MATERIALS Accounting for the normalization condition hvyi 0, in
the case of the rectangular unit cell U (Fig. 1), this yields
Macroscopic representations of a nanoporous material X
treat it as a continuum, without separating it into the pore " # " #
v1 &yc1 ; y2 v1 yc1 ; y2 0;
space P and the solid skeleton S. Macroscopic solute con-
@v1 @v " # (13a)
centration Cx; tthe microscopic concentration cy; t y1 ; 0 1 y1 ; yc2 0
averaged over a (representative elementary) volume V cen- @y2 @y2
tered at point x 2 Xis defined as and

1 " #
Cx; t ! cy; tdy; (7) v2 y1 ; 0 v2 y1 ; yc2 0;
kP V k P V x
@v2 " c # @v2 " c # (13b)
where kP V k is the total volume of pores contained in V. &y1 ; y2 y ; y2 0:
@y1 @y1 1
We show in Appendix B that, for nanoporous materials
composed of periodic arrangements of unit cells U, the mac- Here, yc1 R cos h and yc2 2R l. Figure 2 exhibits a nu-
roscopic concentration Cx; t satisfies a reaction-diffusion merical solution vy v1 ; v2 > of the BVP (11)(13)
equation defined on P U in Fig. 1. This solution and all the numerical
results reported below are obtained with COMSOL.
@C KC
/&1 r ' Deff rC & ceff qm ; (8)
@t 1 KC
where / ! kP V k=kVk kP U k=kUk is the porosity (with IV. SIMULATION RESULTS
P U denoting the pore space of the unit cell U); the effective
The upscaled relations derived above allow one to infer
reaction constant ceff is related to kCV k, the total surface
macroscopic properties of a nanomaterial, e.g., the (effec-
area of the pores contained in V (or to kCU k, the total surface
tive) anisotropic diffusion coefficient Deff and the effective
area of the pores contained in U) by
reaction constant ceff , from its microscopic characteristics.
kCV k kCU k Unless specified otherwise, the subsequent simulations deal
ceff ; (9) with the effects of these microscopic parameters on the mac-
kP V k kP U k
roscopic properties of Material 1 (Fig. 1). The external con-
and the effective diffusion tensor Deff defined as centration gradient is applied in the x1 direction.

FIG. 2. The horizontal v1 (left) and


vertical v2 (right) components of the
closure variable vy v1 ; v2 >
defined on the fluid-domain P U of the
unit cell U in Fig. 1. In this example,
the four vertical nanobridges are
spaced D 0:3 nm apart.
244304-4 Zhang et al. J. Appl. Phys. 117, 244304 (2015)

FIG. 3. Dependence of effective longi-


tudinal diffusion coefficient D11 (left)
and effective reaction constant ceff
(right) on pore radius R and intersec-
tion angle h.

A. Nanopore radius in isolated nanotunnels material, the microporous tube bridges significantly alter its
diffusive properties, giving rise to the non-zero transverse dif-
In the absence of nanobridges (n 0), diffusion through
fusion coefficient D22. They also nearly double the effective
the nanoporous Material 1 (Fig. 1) occurs only in the x1
reaction constant ceff .
direction along the disconnected nanotunnels. The latter con-
sist of stacked nanopores of uniform radius R, which varies
C. Nanobridge density
from R 20 nm to 60 nm. The effective (macroscopic) longi-
tudinal diffusion coefficient D11, normalized by molecular This numerical experiment deals with nanotunnels
diffusion coefficient D, and the effective reaction constant formed by pores of radius R 40 nm, which are intersected at
ceff are shown in Fig. 3; the transverse diffusion coefficient angle h 20+ ; the neighboring nanotunnels are connected by
is D22 0. The effective diffusion coefficient D11 increases microporous tubes of length l 10 nm. In the simulations
with pore radius R and intersection angle h. The reaction reported below, we set the distance between individual micro-
constant ceff decreases with both. porous tubes to either D 1:4 nm or D 0:3 nm. For the
given microporous tube length of l 10 , 11 nm and D
B. Intersection angle h 0:3 nm, the maximum possible number of microporous tubes
in the unit cell is nmax 10. If no limits are placed on the
Intersection angle h quantifies the distance between the
microporous tube length (i.e., if it is allowed to vary between
centers of any two adjacent spherical pores in a given nano-
l 10 , 50 nm), then placing microporous tubes D 1:4
tunnel (Fig. 1), such that h decreases as this distance increases.
nm apart would allow for the maximum number of micropo-
Figure 4 exhibits the impact of h on the effective properties of
rous tubes in the unit cell is nmax 34. The behavior of the
Material 1 (Fig. 1), in which nanotunnels composed of pores
effective transport properties of these nanoporous materials as
of radius R 40 nm are connected by microporous tubes of
a function of nanobridge density is shown in Fig. 5. While the
length l 10 nm. As h increases, the throats between the adja-
number of microporous tubes in a bridge, n, does not appreci-
cent pores become large, giving rise to a significant increase
ably affect the materials diffusive properties (D11 and D22), it
in the longitudinal component of the effective diffusion tensor
significantly influences its adsorbing capacity (ceff ).
(D11). The impact of h on both the transverse diffusion (D22)
and the effective reaction constant (ceff ) is less pronounced. A
D. Alternative material topology
slight (*3%) increase of the longitudinal diffusion coefficient
D11 caused by the addition of n 10 nanotubes is due to the To investigate the effects of spatial arrangement of
rise in the materials connectivity and porosity. Despite their nanotunnels and nanobridges comprising a nanoporous ma-
negligible contribution to the porosity of this nanoporous terial, we consider the second class of materials (Material 2

FIG. 4. Dependence of the longitudinal


(D11) and transverse (D22) components
of the effective diffusion tensor (left)
and effective reaction constant ceff
(right) on intersection angle h, for
materials with disconnected (n 0)
and interconnected (by n 10 micro-
porous tubes) nanotunnels. The latter
are composed of pores with uniform
radius R 40 nm.
244304-5 Zhang et al. J. Appl. Phys. 117, 244304 (2015)

FIG. 5. Dependence of the longitudinal


(D11) and transverse (D22) components
of the effective diffusion tensor (left)
and effective reaction constant ceff
(right) on the number of microporous
tubes (n) comprising the nanobridges
between the neighboring nanotunnels.
The latter are composed of pores with
uniform radius R 40 nm, which are
intersected at angle h 20+ .

significantly more sensitive to the number of microporous


tubes in the nanobridge; the transverse diffusion coefficients
(D22) of the two materials are practically the same. Although
not shown here, a local distribution of microporous tubes
within a unit cell also affects a materials diffusive proper-
ties. A symmetric arrangement of nanobridges, which con-
sists of an equal number of nanotubes arranged at acute and
obtuse angles (as in Fig. 6), yields the higher values of D11
and D22 than its asymmetric counterpart, in which only the
left or the right bridge is present, does.
FIG. 6. A schematic representation of Material 2 and the corresponding unit
cell U (right).
V. CONCLUSIONS
in Fig. 6). The neighboring nanotunnels in these materials Modern technological advances have enabled one to
are shifted by R=2, while in Material 1 (Fig. 1) they are manufacture nanoporous materials with a prescribed pore
aligned. structure. This raises a possibility of using controllable pore-
For a given structure of the nanotunnels (e.g., the pore scale parameters (e.g., pore size and connectivity) to design
radius R 40 nm and the intersection angle h 20+ ) and materials with desired macroscopic properties (e.g., diffusion
nanobridges (e.g., the microporous tube spacing D 0:3 nm coefficient and adsorption capacity). By relating these two
and the number of microporous tubes n), Eq. (9) suggests scales, the homogenization theory (or other upscaling techni-
that Materials 1 and 2 have the same adsorbing capacity (the ques) provides a means of guiding the experimental design.
effective reaction constant ceff ). The topologic differences To demonstrate this approach, we considered a class of nano-
between the two materials manifest themselves in the macro- porous materials whose pore space consists of nanotunnels
scopic diffusive properties (Fig. 7). Material 2 possesses the interconnected by carbon nanotubes (microporous tubes).
higher longitudinal diffusion coefficient (D11) that is We express the anisotropic diffusion coefficient and adsorp-
tion coefficient of such materials in terms of the tunnels
properties (pore radius and inter-pore throat width) and their
connectivity (spacing between the adjacent tunnels and
microporous tube-bridge density). Our analysis leads to the
following major conclusions.
(1) Upscaling (e.g., by means of the homogenization theory)
provides a general framework for optimal design of
nanoporous materials. In the case of adsorption-diffusion
systems, this approach provides a venue for manufactur-
ing nanoporous materials with desired anisotropic diffu-
sion coefficient Deff and reaction constant ceff , which can
be used, for example, in the design of filters.
(2) Intersection angle h quantifies the distance between the
centers of any two adjacent spherical pores in a given
nanotunnel. As h increases, the throats between the adja-
FIG. 7. Dependence of the longitudinal (D11) and transverse (D22) compo- cent pores become larger, giving rise to a significant
nents of the effective diffusion tensor on the number of microporous tubes
(n) for Materials 1 and 2. In both materials, pores of uniform radius increase in the longitudinal component of the effective
R 40 nm are intersected at angle h 20+ . diffusion tensor (D11). The impact of h on both the
244304-6 Zhang et al. J. Appl. Phys. 117, 244304 (2015)


transverse diffusion (D22) and the effective reaction con- 1 1
hAi ! Ady; hAiP ! Ady;
stant (ceff ) is less pronounced. kUk P U x kP U k P U x
(3) Despite their negligible contribution to the porosity of (B1)
this nanoporous material, the microporous tube bridges hAiC ! Ady;
CU x
significantly alter its diffusive properties, giving rise to
the non-zero transverse diffusion coefficient D22. They such that hAi /hAiP where / kP U k=kUk is the poros-
also nearly double the effective reaction constant ceff . ity. We use the multiple-scale expansion technique16,17 to
(4) While the number of microporous tubes in a bridge, n, derive effective continuum scale equations for average
does not appreciably affect the materials diffusive prop- concentration Cx; t ! hcx; ti. The method postulates
erties (D11 and D22), it significantly influences its adsorb- that concentration exhibits both large-scale (across the po-
ing capacity (ceff ). rous material, denoted by the coordinate x) and small-scale
(inside individual pores, denoted by the coordinate y) spa-
ACKNOWLEDGMENTS tial variability, such that y !&1 x with ! " 1; the corre-
sponding temporal scales are denoted, respectively, by t
This research was supported in part by Air Force Office
and sr Da t where Da is the Damkohler number.
of Scientific Research (DE-FG02-07ER25815) and National Let us introduce a membership (indicator) function
Science Foundation (EAR-1246315). Px Px=!,
$
APPENDIX A: DERIVATION OF THE BOUNDARY 1; x2P
CONDITION
Px (B2)
0; x 2 S:
Rewriting (2) and (3) in terms of the dimensionless
quantities (5) yields For the nanoporous materials under consideration, the func-
tion Px Px=! is periodic on the unit cell U. This
@s ds allows one to define the pore-scale diffusion equation (6b)
&n ' rc qm ; and Da seq & s: (A1) on the whole unit cell U (rather than on the multi-connected
@t dt
subdomain P U occupied the liquid)
In a typical application of nanoporous materials, e.g., meso-
porous catalyst membranes, the characteristic macroscopic @c
r ' Prc; y 2 U; t > 0: (B3)
length (membrane size) is33 L , 100 lm and the characteris- @t
tic pore scale is11 l , 20 & 100 nm. Hence, the length-scale
ratio is ! ! l=L , 10&4 . Moreover, the characteristic time Replacing the concentration cx; t with cx; y; t; sr and sub-
scale of adsorption processes in mesoporous membranes, stituting into Eq. (B3) yields
e.g., for methylene blue, is11,34 1=c , 40 & 80 min, while its h % &i
diffusion coefficient is35 D , 10&5 cm2 s & 1. Hence, the @c @c
Da rx ' P rx c !&1 ry c !&1 ry
Damk ohler number is Da , O10&3 , which allows its use @t @sr
h % &i
as a small perturbation parameter. ' P rx c !&1 ry c ; y 2 U: (B4a)
Consider an asymptotic expansion of sx; t in the
powers of Da; s s0 Da s1 ODa2 . Substituting this
The interfacial condition (6a) takes the form
expansion into the second equation in (A1) and collecting
the terms of equal powers of Da yields % & Kc
&n ' rx c !&1 ry c Da qm ; y 2 CU : (B4b)
ds0 ds1 1 Kc
0 and seq & s0 : (A2)
dt dt
In the multiple-scale expansion method, the concentra-
For the homogenous initial condition, sx; 0 0, this yields tion cx; y; t; sr is represented by an asymptotic series
s0 0 and ds1 =dt Kc=1 Kc. Hence, the first-order
X
1
approximation of the sorbed concentration, s * s0 Da s1 , cx; y; t; sr !m cm x; y; t; sr : (B5)
gives rise to m0

ds Kc Substituting Eq. (B5) into Eq. (B4), approximating the


* Da : (A3)
dt 1 Kc Langmuir adsorption isotherm with a series
Substituting this approximation into (A1) yields (6a). Kc 1
1& 1 & 1 & K c0 !c1
1 Kc 1 Kc
" #
APPENDIX B: HOMOGENIZATION OF TRANSPORT K 2 c20 2!c1 c0 . O!2 ;
EQUATIONS
Kc0 & K 2 c20 !Kc1 & 2K 2 c0 c1 O!2 ; (B6)
Three types of local average of a quantity Ax; t are
defined and recalling from Appendix A that Da , O!1 , we obtain
244304-7 Zhang et al. J. Appl. Phys. 117, 244304 (2015)

&!&2 ry ' Pry c0 &!&1 frx ' Pry c0 &qm Kc0 & K 2 c20 & n ' rx c1 ry c2 0; y 2 CU :
ry ' Pry c1 rx c0 .g (B14b)
$
0 @c0 ' " #(
! & rx ' P rx c0 ry c1 Integrating over U with respect to y, applying the interfacial
@t conditions (B14b), and accounting for the periodicity of
)
' " #( Py on the boundary of the unit cell U, we obtain
& ry ' P r x c 1 ry c 2 O!1 ; y2U (B7)
@c0 % " # &
& rx ' /&1 h I ry v irx c0
and @t " #
ceff qm Kc0 & K 2 c20 0; (B15)
&!&1 n ' ry c0 & !0 n ' rx c0 ry c1 .
&!1 qm Kc0 & K 2 c20 where ceff is defined in Eq. (9), / is the porosity and effec-
tive diffusion coefficient Deff =D hI ry vi is defined
n ' rx c1 ry c2 . O!2 ; y 2 CU : (B8) in (10).
Rewriting the linearized form of the Langmuir isotherm,
Collecting terms of the equal powers of ! yields BVPs for Kc0 & K 2 c20 , in its original form, Kc0 =1 Kc0 ; approximat-
cm x; y; t; sr (m 0; 1; ). ing the solute concentration c * c0 ; and defining its average
Leading-order term, O!&2 : Collecting the O!&2 over the cell as Cx; t ! hcx; ti leads to the homogenized
terms yields a BVP for the leading-order term in expansion continuum-scale diffusion-reaction equation for the average
(B5) concentration (8).

ry ' Pry c0 0; y 2 U; &n ' ry c0 0; y 2 CU : 1


J. S. Seo, D. Whang, H. Lee, S. Im Jun, J. Oh, Y. J. Jeon, and K. Kim, A
(B9) homochiral metalorganic porous material for enantioselective separation
and catalysis, Nature 404, 982986 (2000).
2
C.-M. Yang, Y.-J. Kim, M. Endo, H. Kanoh, M. Yudasaka, S. Iijima, and
This BVP has a trivial solution, which implies that c0 is inde- K. Kaneko, Nanowindow-regulated specific capacitance of supercapaci-
pendent of y. tor electrodes of single-wall carbon nanohorns, J. Am. Chem Soc. 129,
Term of order O!&1 : Since Da , O!1 , collecting 2021 (2007).
3
T. Ohkubo, J. Miyawaki, K. Kaneko, R. Ryoo, and N. A. Seaton,
the O!&1 terms yields a BVP for the first-order term in Adsorption properties of templated mesoporous carbon (CMK-1) for
expansion (B5) nitrogen and supercritical methane experiment and GCMC simulation,
J. Phys. Chem. B 106, 65236528 (2002).
ry ' Prx c0 ry c1 . 0; y2U (B10a) 4
T. A. Desai, D. J. Hansford, L. Kulinsky, A. H. Nashat, G. Rasi, J. Tu, Y.
Wang, M. Zhang, and M. Ferrari, Nanopore technology for biomedical
subject to the interfacial condition applications, Biomed. Microdevices 2, 1140 (1999).
5
S. P. Adiga, L. A. Curtiss, J. W. Elam, M. J. Pellin, C.-C. Shih, C.-M.
&n ' rx c0 ry c1 0; y 2 CU : (B10b) Shih, S.-J. Lin, Y.-Y. Su, S. D. Gittard, and J. Zhang, Nanoporous materi-
als for biomedical devices, JOM 60, 2632 (2008).
6
A. J. Rosenbloom, D. M. Sipe, Y. Shishkin, Y. Ke, R. P. Devaty, and
Equations (B10), which form a BVP for c1 x; y; t; sr , define W. J. Choyke, Nanoporous SiC: A candidate semi-permeable material
a local problem. It depends only on the geometry of the unit for biomedical applications, Biomed. Microdevices 6(4), 261267
cell. We represent its solution as16,17 7
(2004).
Basic research needs for electric energy storage, Technical Report
c1 x; y; t; sr vy ' rx c0 x; t; sr c"1 x; t; sr : (B11) (United State Department of Energy, 2007).
8
S. Kondrat, C. R. P#erez, V. Presser, Y. Gogotsi, and A. A.
Kornyshev, Effect of pore size and its dispersity on the energy
Substituting this into Eq. (B10a) yields an equation for the storage in nanoporous supercapacitors, Energy Environ. Sci. 5,
closure variable (a vector) vy 64746479 (2012).
9
M. P. Klein, B. W. Jacobs, M. D. Ong, S. J. Fares, D. B. Robinson, V.
ry ' I ry vrx c0 0; y 2 PU : (B12) Stavila, G. J. Wagner, and I. Arslan, Three-dimensional pore evolution of
nanoporous metal particles for energy storage, J. Am. Chem. Soc. 133,
This equation is subject to hvi 0 and the boundary 91449147 (2011).
10
M. E. Davis, Ordered porous materials for emerging applications,
condition
Nature 417, 813821 (2002).
11
I. Moriguchi, M. Honda, T. Ohkubo, Y. Mawatari, and Y. Teraoka,
n ' I ry vrx c0 0; y 2 CU ; (B13) Adsorption and photocatalytic decomposition of methylene blue on mes-
oporous metallosilicates, Catal. Today 90, 297303 (2004).
which is obtained by substituting Eq. (B11) into Eq. (B10b). 12
H. Yamada, H. Nakamura, F. Nakahara, I. Moriguchi, and T. Kudo,
Terms of order O!0 : Collecting the terms of order ! Electrochemical study of high electrochemical double layer capacitance
(B7) yields of ordered porous carbons with both meso/macropores and micropores,
J. Phys. Chem. C 111, 227233 (2007).
13
L. L. Zhang and X. S. Zhao, Carbon-based materials as supercapacitor
@c0 ' " #( electrodes, Chem. Soc. Rev. 38, 25202531 (2009).
& rx ' P rx c0 ry c1 14
@t G. Severino, G. Dagan, and C. J. van Duijn, A note on transport of a pulse
' " #( of nonlinearly reactive solute in a heterogeneous formation, J. Comput.
& ry ' P rx c1 ry c2 0; y2U (B14a)
Geosci. 4, 275286 (2000).
15
S. Whitaker, The Method of Volume Averaging (Kluwer Academic
subject to the interfacial condition Publishers, Netherlands, 1999).
244304-8 Zhang et al. J. Appl. Phys. 117, 244304 (2015)

16 26
U. Hornung, Homogenization and Porous Media (Springer, New York, Z. Reddad, C. Gerente, Y. Andres, and P. Le Cloirec, Adsorption of sev-
1997). eral metal ions onto a low-cost biosorbent: Kinetic and equilibrium stud-
17
I. Battiato and D. M. Tartakovsky, Applicability regimes for macroscopic ies, Environ. Sci. Technol. 36, 20672073 (2002).
27
models of reactive transport in porous media, J. Contam. Hydrol. S. Saraji, L. Goual, and M. Piri, Adsorption of asphaltenes in porous
120121, 1826 (2011). media under flow conditions, Energy Fuels 24, 60096017 (2010).
18 28
G. Allaire, A. Mikelic, and A. Piatnitski, Homogenization approach to C. A. Grattoni, R. A. Dawe, and M. S. Bidner, On the simultaneous deter-
the dispersion theory for reactive transport through porous media, SIAM mination of dispersion and nonlinear adsorption parameters from displace-
J. Math. Anal. 42, 125144 (2010). ment tests by using numerical models and optimisation techniques, Adv.
19
I. Moriguchi, Y. Shono, H. Yamada, and T. Kudo, Colloidal crystal- Water Resour. 16, 127135 (1993).
29
derived nanoporous electrode materials of cut SWNTs-assembly and J. Bear and A.-D. Cheng, Modeling Groundwater Flow and Contaminant
TiO2/SWNTs nanocomposite, J. Phys. Chem. B 112, 1456014565 Transport (Springer, 2010), Vol. 23.
30
(2008). W. Plazinski, W. Rudzinski, and A. Plazinska, Theoretical models of
20
C. G. Enfield, C. C. Harlin, and B. E. Bledsoe, Comparison of five kinetic sorption kinetics including a surface reaction mechanism: A review, Adv.
models for orthophosphate reactions in mineral soils, Soil Sci. Soc. Colloid Interface Sci. 152, 213 (2009).
31
Am. J. 40, 243249 (1976). Y.-S. Ho and G. McKay, Pseudo-second order model for sorption proc-
21
C. J. Van Duijn and P. Knabner, Travelling waves in the transport of re- esses, Process Biochem. 34, 451465 (1999).
32
active solutes through porous media: Adsorption and binary ion exchange- G. Limousin, J.-P. Gaudet, L. Charlet, S. Szenknect, V. Barthes, and M.
Part 1, Transp. Porous Media 8, 167194 (1992). Krimissa, Sorption isotherms: A review on physical bases, modeling and
22
W. J. Weber, P. M. McGinley, and L. E. Katz, Sorption phenomena in measurement, Appl. Geochem. 22, 249275 (2007).
33
subsurface systems: Concepts, models and effects on contaminant fate and M. J. Pellin, P. C. Stair, G. Xiong, J. W. Elam, J. Birrell, L. Curtiss, S. M.
transport, Water Res. 25, 499528 (1991). George, C. Y. Han, L. Iton, and H. Kung, Mesoporous catalytic mem-
23
G. Allaire, R. Brizzi, A. Mikelic, and A. Piatnitski, Two-scale expansion branes: Synthetic control of pore size and wall composition, Catal. Lett.
with drift approach to the Taylor dispersion for reactive transport through 102, 127130 (2005).
34
porous media, Chem. Engrg. Sci. 65, 22922300 (2010). D. E. Giammar and J. G. Hering, Time scales for sorption-desorption and
24
S. Azizian, Kinetic models of sorption: A theoretical analysis, J. Colloid surface precipitation of uranyl on goethite, Environ. Sci. Technol. 35,
Interface Sci. 276, 4752 (2004). 33323337 (2001).
25 35
C. Cheung, J. Porter, and G. McKay, Sorption kinetics for the removal of K. L. Kostka, M. D. Radcliffe, and E. von Meerwall, Diffusion coeffi-
copper and zinc from effluents using bone char, Sep. Purif. Technol. 19, cients of methylene blue and thioflavin T dyes in methanol solution,
5564 (2000). J. Phys. Chem. 96, 22892292 (1992).

You might also like