We develop a semi-analytical model for transport in structured catalytic microreactors, where both reactant and product are compressible fluids. Using lubrication and Fick–Jacobs approximations, we reduce the three-dimensional governing equations to an effective one-dimensional set of equations. Our model captures the effect of compressibility, corrugations in the shape of the reactor, and an inhomogeneous catalytic coating of the reactor walls. We show that in the weakly compressible limit (e.g., liquid-phase reactors), the distribution of catalyst does not influence the reactor yield, which we verify experimentally. Beyond this limit, we show that introducing inhomogeneities in the catalytic coating and corrugations to the reactor walls can improve the yield.

Catalysts and catalytic reactors are ubiquitous in the modern-day world. They are used in the production of organic and inorganic chemicals,1 in crude oil refining,2 in energy conversion,3–5 and for environmental protection/green chemistry.6 The latter has become increasingly more relevant as the world deals with the consequences of climate change.7,8 Research into novel catalysts is ever ongoing, with great attention being given to recent developments such as supported ionic liquid phase (SILP) catalysis,9,10 supported catalytically active liquid-metal solutions (SCALMS),11,12 and solid catalysts with ionic liquid layer (SCILL).13 Catalysts such as these are deposited within a reactor through which fluid containing reactant is flowed. These reactors may be slabs through which channels run through,14 or compressed solid grains/particles.15 

A well-known use of catalysts is in the automotive industry. The functioning of motor vehicles leads to the production of environmentally dangerous chemicals, which are eliminated by catalysts rather than emitted into the atmosphere.22 A common reactor design is that of the so-called monolithic reactor, composed of a slab of material through which channels run through.14 The walls of these channels are coated with catalytic material. Catalytic porous media such as packed-bed reactors are also an industry-standard.15 Common models for the transport through such media rely on coarse-graining the porous structure. These models do not consider the pores explicitly, but instead treat the flow through the porous media as flow in free space with effective transport parameters.16,17 Such models often follow from averaging the flow quantities over a volume much larger than the typical pore size,18,19 and as such, fail when examining flow at the pore scale.20,21

Understanding the transport of reacting gases at the microscale is also vital for the development and study of miniaturized reactors, which have received great attention in the past decades.23–25 Research and development of new chemical processes benefit from the decreased amount of sample required in microfluidic devices, as well as from the heightened control over the chemical conditions inside the microreactor coming from higher surface-to-volume ratios.23–25 In addition to an increase in efficiency and a decrease in cost,25 conducting research in miniaturized reactors leads to a decrease in the production of environmentally unsafe chemicals.26 Indeed, an effort to make the chemical industry greener has taken root in the past decades, and it is expected to play a vital role in the mitigation of climate change.26 Large-scale production of chemicals also benefits from employing multiple microreactors in parallel rather than a single, large batch reactor. The increased control over the chemical properties inside the reactor may facilitate operation at very efficient conditions that are not easily enforceable in batch reactors.23–25 Furthermore, the increased control leads to the entire process becoming safer, as the chance of a malfunction decreases.25 

Here, we present a theory for the flow of a compressible fluid in a quasi-2D reactor/pore whose walls are coated with a catalyst. To obtain analytical results, we make use of the Fick–Jacobs approximation, which is valid for thin reactors. As the reactant flows through the reactor, it turns into the product in the presence of the catalyst. The theory enables the study of reactors with corrugated walls and a spatially varying reaction rate. As a result, we can optimize the reactor parameters such as the shape of the corrugation and the distribution of catalytic material. Achieving a homogeneous catalytic coating often requires involved experimental techniques.27 We show that these efforts may be unnecessary, as the distribution of catalytic material has no effect on the flow rate of the product in the incompressible limit (such as in liquid-phase reactors), and verify this result experimentally.

Consider a reactor defined by the space between two solid slabs. These slabs are symmetric with respect to the axis z = 0 and are homogeneous in the y direction. Thus, the reactor is quasi-3D. The slabs are corrugated in the x direction, with the separation between them defined as 2h(x), as shown in Fig. 1. It will be useful to define the average channel height,
(1)
The length of the reactor is L such that to the left (x < 0) and the right (x > L) of the slabs are two reservoirs of gas with pressure P0 and PL, respectively. The gas fills the reactor as well, with a spatially varying pressure P(x, z, t), such that
(2)
and
(3)
Due to the translational symmetry in the y direction, the pressure depends only on the x and z directions. The pressure is related to the number density of the gas via the virial expansion
(4)
where kB is the Boltzmann constant, T is the absolute temperature, ρ(x, z, t) is the number density of the gas (with dimensions meter−3), and Bi (i ≥ 1) are the virial coefficients. Many gasses (e.g., methane, water vapor, oxygen, and nitrogen) exhibit values of |B1| below 102cm3/mol for T ≈ 400–1000 K.28 Typically, gasses (including the four aforementioned examples) in that temperature range and for pressures around 1 bar exhibit densities 105mol/cm3.29 These densities are well below B11102mol/cm3, and thus, the ideal gas equation of state is a good approximation,
(5)
The pressure difference between the ends of the reactor leads to a non-zero flow velocity v(x, z, t) inside the reactor as the gas moves from one reservoir to the other. The flow velocity is governed by the (compressible) Navier–Stokes equation,
(6)
where ρf(x, z, t) is the mass density of the fluid, and η and ζ are the shear and bulk viscosity, respectively. No bulk force is applied to the fluid. The slabs, the reactor, and the reservoirs are kept at a constant temperature, and thus, the viscosities η and ζ are homogeneous in space and constant in time as well.30,31 Equation (6) is subjected to no-slip boundary conditions on the reactor walls,
(7)
Furthermore, the continuity equation applies,
(8)
The walls of the reactor are coated with a catalytic material that decomposes a reactant gas coming from the inlet (A particles) and converts it into a product gas (B particles). For the remainder of this work, A and B particles are physically indistinguishable (same mass, same size, etc.). Thus, the chemical reaction is merely a change of label. Since A and B particles have the same mass m, we write
(9)
and
(10)
where ρA(x, z, t) and ρB(x, z, t) are the number densities of species A and B, respectively. These number densities are governed by the advection-diffusion equations,
(11)
and
(12)
where jA(x, z, t) and jB(x, z, t) are the fluxes of species A and B respectively, and ξ(x, z, t) is the rate of reaction. For simplicity, we assume a first-order reaction rate,32,
(13)
where δ(z) is the Dirac delta function, which localizes the chemical reaction at the reactor walls. The function α(x) is a spatially varying reaction rate with unit meter/sec, as can be seen from the dimensional analysis of Eqs. (12) and (13).33 The fact that α(x) is spatially varying accounts for a possibly inhomogeneous concentration of catalytic material coating the reactor wall. It will be useful to define the spatially-averaged reaction rate α0 such that
(14)
and
(15)
where δα(x) is the deviation from the average reaction rate. Equations (11) and (12) are solved with the added constraint that gas cannot escape through the reactor walls,
(16)
(17)
and as such
(18)
(19)
The fluxes jA(x, z, t) and jB(x, z, t) can be decomposed into two components,
(20)
and
(21)
with the first term on the right-hand side of Eqs. (20) and (21) accounting for the motion of the fluid mixture as a whole, and the second term in the right-hand-side accounting for diffusion. The velocity v(x, z, t) is the center-of-mass/barycentric velocity of the mixture.34,35 In addition to this motion, A and B particles may move relative to another one, which is captured by the diffusive fluxes jD(A)(x,z,t) and jD(B)(x,z,t).34,36 Because the gas is well-approximated as ideal, the diffusive fluxes may be written as Fick’s law,35,
(22)
(23)
where D is the diffusion coefficient, which does not vary in space or time.35 Note that the sum of the diffusive fluxes is zero,
(24)
as is required from the fact that the diffusive fluxes are defined as the motion of A and B particles with respect to the center of mass of the binary mixture.34–36 We assume the absence of product in the inlet,
(25)
which together with the inlet pressure P0 and outlet pressure PL form the set of boundary conditions needed to solve Eqs. (11) and (12),
(26)
and
(27)
FIG. 1.

Longitudinal section of the quasi-2D catalytic reactor with length L and variable height h(x), the average of which is h0. The irreversible decomposition of a chemical species A to produce a chemical species B occurs solely at the inner reactor wall with rate αρA. The blue color indicates the value of the total number density and, therefore, of the pressure. The darker the blue, the higher the pressure. The pressure at inlet P0 is different from the pressure at outlet P0, leading to a flow in the x direction vx.

FIG. 1.

Longitudinal section of the quasi-2D catalytic reactor with length L and variable height h(x), the average of which is h0. The irreversible decomposition of a chemical species A to produce a chemical species B occurs solely at the inner reactor wall with rate αρA. The blue color indicates the value of the total number density and, therefore, of the pressure. The darker the blue, the higher the pressure. The pressure at inlet P0 is different from the pressure at outlet P0, leading to a flow in the x direction vx.

Close modal

Solving Eqs. (6), (11), and (12) in the general case presents significant difficulties. Instead, we specialize Eqs. (6), (11), and (12) to the case of a thin reactor (h0/L ≪ 1). Such a condition is often true for monolithic reactors.14,37,38 The separation of length scales leads to relaxation in the transverse direction, which is much faster than relaxation in the longitudinal direction. As a result, we can describe the dynamics in the longitudinal direction via an effective 1D equation. This is the so-called Fick–Jacobs approximation39–45 for transport in thin channels. Such a technique has been used to model the transport of colloids,46–50 polymers,51,52 electrolytes,53–55 and chemically active systems.56–59 This procedure has the added benefit of yielding simple formulas that provide analytical insight and are computationally cheap to evaluate.

As we are interested in the long-term functioning of the reactor, our starting point is to disregard transients in the dynamics and focus on the steady state. We will now solve for the steady-state flow of the mixture through the reactor. At steady state, the continuity equation [Eq. (8)] yields
(28)
where
(29)
is the total flux of the mixture. Integrating Eq. (28) in the transverse direction leads to
(30)
where we take the integration limits to be ± for ease of future calculations. We shall see that all results will reduce to integrals inside the reactor. Equation (30) reduces to
(31)
as there is no flux at infinity. Indeed, there is no flux outside of the reactor [Eqs. (18) and (19)], which leads to
(32)
We now define the total flow J(x) across the reactor as the integral of the flow density jx(x, z),
(33)
which inserted in Eq. (32) yields
(34)
as is required for a steady state. The total flux J is a flux of particles per unit length, as the integral in Eq. (30) is done over z only (due to the quasi-3D geometry of the reactor). As discussed earlier, we aim to reduce the quasi-3D problem to an effective one-dimensional problem by employing the Fick–Jacobs approximation. This approximation exploits the difference in timescales characterizing transverse and longitudinal transport.
As discussed, we assume a thin reactor
(35)
which allows for an approximated solution of Eq. (6) by employing the lubrication approximation.60 Such a procedure is done in detail in  Appendix A, and yields the flow velocity profile,
(36)
provided that the channel is thin enough so that
(37)
(38)
where we define the Reynolds number Re as
(39)
where ρ̄f and v̄x are the typical scales for mass density and longitudinal velocity, respectively. The condition in Eq. (37) is obtained in thin channels (h0L), provided that Re is itself not a large number. While the latter generally holds for flows in micrometric channels, Eq. (37) will be seen as imposing a maximum height for the reactor (given a certain reactor aspect ratio).
The condition in Eq. (38) requires the knowledge of the ratio of bulk viscosity to shear viscosity. For some gasses, the ratio ζ/η is smaller than 1 (e.g., carbon monoxide), or order one (e.g., methane),61 and Eq. (38) is fulfilled whenever Eq. (35) is fulfilled as well. However, it is also possible for the ratio ζ/η to take values as large as 4000, such as for carbon dioxide.61 The conditions in Eqs. (37) and (38) further imply that the variation of pressure along the transverse direction is negligible (more information in  Appendix A),
(40)
In addition to the advective contribution, the flow of product and reactant also contains a diffusive contribution as per Eqs. (20) and (21). Due to the length scale separation between reactor length and height, either diffusion or advection can dominate the transport in certain directions. The competition between diffusion and advection is quantified via two Péclet numbers (one for each direction). For the longitudinal direction, we have
(41)
and in the transversal direction,
(42)
where v̄z is the typical velocity in the z direction. As a result of Eq. (35), we have
(43)
which is rigorously derived in  Appendix A. To obtain an analytical result, we now specialize our theory to the case where advection is the dominant transport mechanism in the longitudinal direction, i.e.,
(44)
and so we approximate
(45)
(46)
Conversely, in the transversal direction, we assume diffusion to be the dominant transport mechanism, with
(47)
which by using Eq. (43) yields a maximum value for
(48)
In this regime, we obtain
(49)
(50)
It is worth noting that Eqs. (44) and (47) can be written as restrictions on h0 and L. By approximating v̄x via the plane Poiseuille law (averaged over the reactor height),62 one obtains
(51)
(52)
Using typical values of η ≃ 5 × 10−5 Pa s,63  D ≃ 10−5 m2/s,64 and ΔP ≃ 105 Pa, one finds (3P)1/2 ≈ 0.1 μm. As such, we expect our theory to be valid for reactors with micrometric width and relatively large aspect ratios L/h0.
Under the assumption of Eq. (47), the dependence of the concentration of reactant and product in the transversal direction is determined by diffusion and the reaction at the wall. Diffusive flows promote homogeneity in z, whereas the reaction promotes gradients in concentration. The competition of these two effects is quantified via the Damköhler number along the transversal direction,
(53)
In reactors that are small enough for Daz to be small, i.e.,
(54)
the concentration of both A and B particles may be approximated as independent of the transversal direction (more information in  Appendix A),
(55)
(56)
for |z| < h(x). The condition in Eq. (54) is fulfilled whenever the average reaction rate α0 is such that
(57)
which for h0 ≈ 1 μm yields the condition α0 ≪ 10 m/s. Note that this velocity quantifies the reaction rate and is not the velocity at which gas is flowing. Examples of environmentally and economically relevant reactions that fit this criterion are methane/air mixtures in contact with platinum (α0 < 10−1 m/s for temperature below 1100 K65), carbon monoxide/oxygen mixtures in contact with platinum (α0 < 1 m/s for temperature below 305 K66), and ammonia production in sustainable reactors (e.g., electrochemical cells, α0 < 10−5 m/s67,68).
Furthermore, note that decreasing the reaction rate of a catalyzed reaction can often be achieved by simply decreasing the amount of catalytic material coating the reactor wall. As such, we proceed with the condition in Eq. (54), from which Eqs. (55) and (56) result. Finally, the theory assumes a Knudsen number Kn small enough to justify using the no-slip boundary condition [Eq. (7)],
(58)
where λ is the mean free path of the gas particles and Kn* is a threshold value limiting our theory’s regime of validity. It will be seen that ultimately the velocity profile plays a role only via the volumetric flux per unit length,
(59)
Previous experimental work in rectangular channels with |ΔP|/P0 ≈ 0.46 shows that using the no-slip boundary conditions induces an error in the flow rate as little as 30% for Kn = 0.1.69 Thus, we suggest Kn* = 0.1. The mean free path may be computed using
(60)
where u is a numerical constant 0.5.70 For air at T ≈ 300 K and P = 1 atm, the above formula yields λ ≈ 10−2 μm, which plugged into Eq. (58) yields Kn ≈ 10−2 when assuming h0 = 1 μm. More generally, plugging Eq. (60) into Eq. (58) yields a condition on the minimum pressure for which the theory is still valid,
(61)
which for air around 300 K (η ≈ 10−5 Pa s,63  m ≈ 5 × 10−26 Kg63) flowing in a channel of h0 = 1 μm yields P ≳ 0.4 atm. As a summary, we now repeat all of the assumptions that define the regime of validity of our model,
We now calculate the flux of product at the reactor outlet.
We now have all the ingredients necessary to solve for the number densities and fluxes. Inserting Eq. (36) in Eqs. (45) and (46), we obtain the fluxes in the x direction, which can then be plugged into Eq. (34). This procedure yields
(62)
which is solved by
(63)
from which the pressure P(x) is obtained when plugged into the equation of state [Eq. (5)]. The dependence of the number density in the longitudinal direction is entirely handled by the first term inside brackets of Eq. (63).
We now compute the volumetric flux rate by combining Eq. (59) with Eq. (36), which is written as
(64)
yielding
(65)
It can be seen from Eqs. (33), (45), and (46) that
(66)
which together with Eq. (65) yields
(67)
which can be rewritten as
(68)
after defining the pressure drop
(69)
Henceforth, we will consider the pressure to be higher in the left (x < 0) reservoir than in the right (x > L) reservoir, and thus, ΔP < 0. This is done merely to facilitate the discussion by fixing the signs of the fluxes. Interchanging P0 with PLP → −ΔP and P0P0 + ΔP) merely results in a sign switch for J and JB(L). It is noteworthy that Eq. (68) includes also a quadratic contribution in ΔP.
As chemical species A and B exhibit the same physical properties, the total fluid motion is insensitive to the composition of the mixture, and hence, both the total flow J and total number density ρ(x) show no dependence on the reaction rate α(x). Nonetheless, we are interested in maximizing the flow of product, and as such, we require the solution of Eqs. (11) and (12). Employing the same technique as for the total number density, we integrate Eq. (11) along the z direction to obtain
(70)
which reduces to
(71)
as there is no flux of reactant at infinity, with ξ(x, z) being given by Eq. (13). Indeed, as there is no flux anywhere outside the reactor [Eq. (18)], we have
(72)
By using Eqs. (45) and (55), we write
(73)
and using the definition in Eq. (59), we write
(74)
which is solved by
(75)
and using Eq. (10), the density profile of products reads
(76)
Similar to Eq. (66), the flow of product per unit length is defined as
(77)
which, at the outlet, using Eq. (76), reads
(78)
We now identify a local timescale associated with transport across the reactor τT(x),
(79)
and a local timescale associated to the reaction τR(x),
(80)
which is the time associated with transport in the transverse direction with velocity α0. We identify the ratio of these timescales as the local Damköhler number in the longitudinal direction,
(81)
such that
(82)
Another variable of interest is the purity Π at the outlet, defined as the ratio of product density to total density,
(83)
By using Eqs. (66) and (77), we write
(84)
and using Eq. (78),
(85)
which depends only on the spatially averaged Damköhler number (in the longitudinal direction). In Sec. V, we analyze the equations we have just derived in the nearly incompressible limit.
In this section, we shed light on the role of compressibility on the output of a catalytic microreactor. Indeed, it is known that compressible fluids behave close to being incompressible when weakly driven, i.e., when the density varies only a little around an average value. We now show that our equations describe an incompressible fluid when this limit is considered. As such, we can describe both liquid-phase and gas-phase reactors and can pinpoint the role the fluid phase plays in the reactor output. As discussed, we study Eqs. (68) and (82) in the limit of weak density variations. Let the density ρ(x) be written as
(86)
where δρ(x) is the variation of density with respect to the inlet value. We now focus on the limit
(87)
and perform expansions of the fluxes in this small parameter, keeping only terms up to linear order. The total flow J in Eq. (68) yields
(88)
and by using the equation of state of Eq. (5), we rewrite the above equation as
(89)
We also obtain the volumetric flux Q to linear order from Eq. (65),
(90)
which does not vary in space. As such, Eqs. (68) and (82) to linear order in δρ(x)/ρ(0) describe the flow of an incompressible fluid, such as a liquid. Finally, we may obtain the flow of product at the outlet from Eq. (78) as
(91)
where only α0 enters the final result due to Q being homogeneous in space. As such, the output of a liquid-phase reactor is insensitive to spatial variations in the catalytic concentration. By comparing Eq. (91) and Eq. (82), we identify the average Damköhler number for liquids as
(92)
The purity [defined in Eq. (83)] may be obtained from Eqs. (84) and (91) as
(93)
We now relax the incompressibility and focus on the role of the shape of the reactor, as well as the distribution of catalytic material inside it. To do so, we apply our theory to the case of a sinusoidal reactor shape and catalytic coating density. Such functional forms are mathematically simple, but still capture the general effect of introducing corrugation to the shape and inhomogeneities to the catalytic coating.
In this section, we wish to investigate how the flow rate of product may be optimized by a suitable choice of reactor shape, or of catalytic coating. To this end, we now apply our theory to the concrete case of a reactor with a sinusoidal shape,
(94)
where h1 is a constant determining the height of the corrugation and
(95)
Note that the reactor shape is fore-aft symmetric. For values of h1 < 0, the reactor is narrow at the inlet/outlet and bulges in the middle. Conversely, for values h1 > 0, the channel is wide at the inlet/outlet and narrows in the middle. The value h1 = 0 corresponds to a flat reactor. See Fig. 2(a) for a sketch of the aforementioned reactor geometries. We further take a sinusoidal reaction rate,
(96)
with α0 being the average reaction rate and α1 determining the deviation from that average. Note that the reaction rate is, in general, not fore-aft symmetric, with the sign of α1 determining if more catalytic material is found near the inlet (α1 > 0) or near the outlet (α1 < 0). This choice of h(x) and α(x) is simple, yet captures the general effect of corrugation and inhomogeneities in the catalytic coating.
FIG. 2.

(a) Sketches of three different reactor shapes. Black regions indicate the reactor wall and blue regions indicate the reactor chamber filled with gas. (b) Black line: total flux J normalized by value for a flat (h1 = 0) reactor Jf (values reported on the left axis). Blue line: Damköhler number ⟨Dax normalized by value for a flat reactor ⟨Dax,f (values reported on the right axis). (c) Flow of product at reactor outlet JB(L) normalized by value for a flat reactor JB(L)f. (d) Purity at reactor outlet Π. For panels (c) and (d), ⟨Dax,f ∈ {10 (orange), 1 (green), 10−1 (violet)}.

FIG. 2.

(a) Sketches of three different reactor shapes. Black regions indicate the reactor wall and blue regions indicate the reactor chamber filled with gas. (b) Black line: total flux J normalized by value for a flat (h1 = 0) reactor Jf (values reported on the left axis). Blue line: Damköhler number ⟨Dax normalized by value for a flat reactor ⟨Dax,f (values reported on the right axis). (c) Flow of product at reactor outlet JB(L) normalized by value for a flat reactor JB(L)f. (d) Purity at reactor outlet Π. For panels (c) and (d), ⟨Dax,f ∈ {10 (orange), 1 (green), 10−1 (violet)}.

Close modal
The simplest case is that of incompressible flows (such as a liquid-phase reactor), as the fluid density is homogeneous. The effect of corrugations on the total flux J and in the Damköhler number ⟨Dax can be seen from Eqs. (68) and (92). Indeed, one may write
(97)
and
(98)
where Jf and ⟨Dax,f are the respective values of J and ⟨Dax for flat channels (h1 = 0). Note that Eq. (97) remains valid even for non-sinusoidal reactor shapes. Using Eq. (94), the integral on the right-hand-side of Eq. (97) can be computed numerically, leading to the results of Fig. 2(b). In this figure, we see that the net effect of the corrugation (h1 ≠ 0) is to decrease the total flux, with a corresponding increase in the Damköhler number. This decrease in J is due to the presence of a bottleneck that becomes narrower the larger the value of |h1|/h0. Indeed, for |h1|/h0 = 1, the bottleneck is shut and no gas can flow through (J = 0). Furthermore, J/Jf is insensitive to the sign of |h1|/h0. As a result, bulging reactors (h1 < 0) exhibit the same total flow as constricting reactors (h1 > 0). An analysis of Eq. (91) shows that the flow rate of product JB(L) normalized by its value for flat channels depends only on the numbers h1/h0 and ⟨Dax,f. We plot this data in Fig. 2(c), where it can clearly be seen that the shape that maximizes the flow rate of the product is the flat reactor for all values of ⟨Dax,f. The latter number controls only the width of the peak around h1 = 0. Conversely, Fig. 2(d) shows that, while flat channels exhibit the largest product fluxes, they also lead to the lowest purity. The large flow velocities that lead to high JB(L) also decrease the time that reactant particles stay inside the reactor and, as such, the time they have to react with the catalytic wall. We now relax the condition of incompressibility and study a general reactor with homogeneous catalytic coating.
In this section, we study compressible flow reactors. For clarity of discussion, we start with the relatively simple case of homogeneous catalytic coating. The first major difference between the incompressible and compressible regime is that Eq. (98) is not valid for the latter, as Q(x) [and ρ(x)] is now spatially varying. According to Eq. (81), we have
(99)
To analyze the behavior of ⟨Q−1(x)⟩, it is useful to first inspect the spatial variation of both the volumetric flow rate and the density. We may now combine Eqs. (63) and (94) to numerically obtain the density profile inside the reactor, which we plot in Fig. 3(a). As expected, the presence of corrugations (h1 ≠ 0) leads to density/pressure profiles that decrease more sharply near the bottleneck, which is located at z = L/2 for constricting channels (h1 > 0), and at z ∈ {0, L} for bulging channels (h1 < 0). This behavior is seen for both low relative pressure drops (|ΔP|/P0 = 0.1) and high relative pressure drops (|ΔP|/P0 = 0.9). By comparing these two plots for the case of flat channels (h1 = 0), one clearly sees the effect of compressibility. For low relative pressure drops, the pressure profile is nearly a straight line, as expected from incompressible flows in constant section channels. However, the pressure profile for constant section channels becomes curved at high relative pressure drops. The reason for this non-linearity is that the density profile controls not only the amount of gas inside the reactor but also the pressure profile and, thus, the flow velocity. Therefore, the condition for steady state (i.e., homogeneous total flux J) is non-linear in xρ(x) [see Eq. (62)]. In particular, to keep the total flow J constant in space, regions with smaller gas density require larger pressure gradients (in magnitude). One clearly sees that this non-linearity in xρ(x) leads to a spatially averaged density ⟨ρ(x)⟩ that changes with h1 at high relative pressure drops. In particular, negative values of h1 showcase higher values of ⟨ρ(x)⟩. Finally, note that all pressure profiles intersect at x/L = 0.5. This is a result of the fore-aft symmetry of the reactor shape h(x), as defined in Eq. (94).
FIG. 3.

(a) Difference between gas density ρ(x) and its value at outlet ρ(L) [normalized by the difference between inlet and outlet density, ρ(0) and ρ(L), respectively] as a function of position for sinusoidal reactors of different corrugation parameters h1/h0. (b) Volumetric flow rate Q(x) normalized by inlet value Q(0). For all panels, |ΔP|/P0 ∈ {0.1(), 0.9()}, and h1/h0 ∈ {−0.9, −0.2, 0, 0.2, 0.9}.

FIG. 3.

(a) Difference between gas density ρ(x) and its value at outlet ρ(L) [normalized by the difference between inlet and outlet density, ρ(0) and ρ(L), respectively] as a function of position for sinusoidal reactors of different corrugation parameters h1/h0. (b) Volumetric flow rate Q(x) normalized by inlet value Q(0). For all panels, |ΔP|/P0 ∈ {0.1(), 0.9()}, and h1/h0 ∈ {−0.9, −0.2, 0, 0.2, 0.9}.

Close modal

Having obtained the density profile, we may obtain the volumetric flow rate Q(x) from Eq. (65), which we plot in Fig. 3(b). Indeed, the latter and the former are inversely proportional to one another, per Eq. (66), and so all previous reasoning about the density profile applies to the volumetric flux. As with the gas density, the spatially averaged volumetric flow rate ⟨Q(x)⟩ is also dependent on h1 and is sensitive to its sign. Therefore, reactant particles spend more time inside bulging reactors (h1 < 0) than inside constricting reactors (h1 < 0). Such an effect implies a breaking of symmetry around h1 = 0 for JB(L) and Π, as shown in Fig. 4. While this effect is small (and for these parameters, only visible for |ΔP|/P0 = 0.9 and low purity reactors), it shows that the flow rate of corrugated reactors can outperform that of flat reactors. Next, we introduce an inhomogeneity in the catalytic coating in an attempt to increase this gain.

FIG. 4.

(a) Flow of product at reactor outlet JB(L) normalized by value for a flat reactor JB(L)f. (b) Purity at outlet Π. For both panels, |ΔP|/P0 ∈ {0.1(), 0.5(), 0.9()}, α1/α0 = 0, and 2α0L/Qf(0) ∈ {10 (orange), 1 (green), 10−1 (violet)}. The function Qf(x) is defined as Q(x) for h1 = 0.

FIG. 4.

(a) Flow of product at reactor outlet JB(L) normalized by value for a flat reactor JB(L)f. (b) Purity at outlet Π. For both panels, |ΔP|/P0 ∈ {0.1(), 0.5(), 0.9()}, α1/α0 = 0, and 2α0L/Qf(0) ∈ {10 (orange), 1 (green), 10−1 (violet)}. The function Qf(x) is defined as Q(x) for h1 = 0.

Close modal
In this section, we study compressible flow reactors with inhomogeneous catalytic coatings (α1 ≠ 0). According to Eq. (81), we have
(100)
Because Q(x) is a monotonically growing function [see Fig. 3(b)], the last term in Eq. (100) is largest when α1 = 1, keeping all other parameters unchanged. Thus, a reactor operating at a large relative pressure drop showcases the highest yields (both flow of product and purity) when the catalytic material is deposited closest to the inlet. Near the inlet, the flow velocities take their smallest values, so reactant particles spend the most time there. However, the case may arise where the catalyst distribution is more difficult to tune experimentally than the reactor corrugation. In that case, one observes that there is an optimum corrugation parameter as in the homogeneously coated case, but which now depends on the catalytic distribution [Figs. 5(a) and 5(b)]. Indeed, Fig. 5(b) shows that an inhomogeneous distribution of catalytic material may boost the gain one obtains by introducing corrugations. Note that this gain is the largest for the least optimal catalyst distributions (α1 < 0), for high relative pressure drops, and for low values of Π [Figs. 5(c) and 5(d)]. In Fig. 6(a), we sweep over all values of α1/α0 and show how the optimum corrugation parameter depends on the catalyst distribution. In general, bulging reactors (h1 > 0) are the best if the catalyst is most concentrated near the outlet (α1 < 0). Conversely, constricting reactors (h1 > 0) are the best if the catalyst is most concentrated near the inlet (α1 > 0). Finally, Fig. 6(b) shows that the largest gains (with respect to flat reactors) are obtained when the catalyst distribution is furthest away from being homogeneous (α1 = −1).
FIG. 5.

Flow of product at reactor outlet JB(L) normalized by value for a flat reactor JB(L)f and purity at outlet Π. (a) and (c) α1/α0 = 0.9. (b) and (d) α1/α0 = −0.9. For all panels, |ΔP|/P0 ∈ {0.1(), 0.5(), 0.9()}, and 2α0L/Qf(0) ∈ {10 (orange), 1 (green), 10−1 (violet)}. The function Qf(x) is defined as Q(x) for h1 = 0.

FIG. 5.

Flow of product at reactor outlet JB(L) normalized by value for a flat reactor JB(L)f and purity at outlet Π. (a) and (c) α1/α0 = 0.9. (b) and (d) α1/α0 = −0.9. For all panels, |ΔP|/P0 ∈ {0.1(), 0.5(), 0.9()}, and 2α0L/Qf(0) ∈ {10 (orange), 1 (green), 10−1 (violet)}. The function Qf(x) is defined as Q(x) for h1 = 0.

Close modal
FIG. 6.

(a) Optimum value of the corrugation parameter h1*/h0. (b) Optimum flow rate of product at outlet JB*(L) normalized by value JB(L)f for flat reactor (h1 = 0). Parameters are −ΔP/P0 ∈ {0.1(), 0.5(), 0.9()}, and 2α0L/Qf(0) ∈ {1 (green), 10−1 (violet), 10−2 (gray)}. The function Qf(x) is defined as Q(x) for h1 = 0.

FIG. 6.

(a) Optimum value of the corrugation parameter h1*/h0. (b) Optimum flow rate of product at outlet JB*(L) normalized by value JB(L)f for flat reactor (h1 = 0). Parameters are −ΔP/P0 ∈ {0.1(), 0.5(), 0.9()}, and 2α0L/Qf(0) ∈ {1 (green), 10−1 (violet), 10−2 (gray)}. The function Qf(x) is defined as Q(x) for h1 = 0.

Close modal

We performed experiments with catalytic membranes with heavily fore-aft asymmetric distribution of catalytic material (see  Appendix B for experimental details). These experiments aimed at verifying the prediction stated in Sec. V that incompressible flow reactors should exhibit the same flow rate of the product no matter how catalyst material is distributed on the reactor walls, provided that the total catalyst mass is preserved (same α0). To verify this prediction, we performed a catalytic study using anodic aluminum oxide membranes loaded with SCALMS nanoparticles, which act as the catalyst. As shown in Fig. 7, the loading is heavily fore-aft asymmetric. We flowed a liquid solution of methylene blue and ascorbic acid (the reactants) through the membrane, which led to the former reducing to leuco-methylene blue (the product). By measuring the light absorbance of the fluid mixture at 665 nm, we were able to monitor the density of the reactant in different positions of our experimental setup. In Fig. 8(a), we show that the membrane is, indeed, catalytically active as there is a decrease in absorbance when comparing samples upstream and downstream of the membrane. Furthermore, this activity is due to the catalyst and not to the support itself, as unloaded membranes do not affect the absorbance [violet lines in Fig. 8(a)]. When measuring a sample downstream of membranes loaded preferentially near the inlet or outlet, no difference is seen within the error margin [orange and green lines in Fig. 8(a)]. To strengthen this conclusion, we performed an additional experiment with the same catalytic membrane, which, after some time, was flipped around so that the inlet and outlet switch positions. Again, no significant difference is seen in the absorbances [Fig. 8(b)].

FIG. 7.

Characterization of a membrane coated with catalytic nanoparticles. (a) Scanning electron micrograph of the membrane in cross-section view. Cyan arrow describes the paths scanned for the data in panel (b). Catalytic nanoparticles visible as lighter spheres peppering the inside of the membrane. (b) Intensity I (in arbitrary units) given by cross-section energy-dispersive x-ray spectroscopy analysis measured along the pores [see cyan arrow in panel (a)]. Higher values of I indicate larger concentrations of a given element: aluminum (green), oxygen (orange), and gallium (violet). The latter is a marker for the nanoparticles and acts as the catalytic material.

FIG. 7.

Characterization of a membrane coated with catalytic nanoparticles. (a) Scanning electron micrograph of the membrane in cross-section view. Cyan arrow describes the paths scanned for the data in panel (b). Catalytic nanoparticles visible as lighter spheres peppering the inside of the membrane. (b) Intensity I (in arbitrary units) given by cross-section energy-dispersive x-ray spectroscopy analysis measured along the pores [see cyan arrow in panel (a)]. Higher values of I indicate larger concentrations of a given element: aluminum (green), oxygen (orange), and gallium (violet). The latter is a marker for the nanoparticles and acts as the catalytic material.

Close modal
FIG. 8.

Data for absorbance A as a function of wavelength λ. Higher values of A indicate larger concentrations of reactant. (a) Violet solid line (): measurement at a position upstream of the membrane. Violet dashed line (): measurement at a position downstream of the membrane (no catalyst). Orange line (): measurement at a position downstream of the membrane (catalyst loaded preferentially near inlet). Green line (): measurement at a position downstream of the membrane (catalyst loaded preferentially near the outlet). (b) Dashed lines (): measurement at a position upstream of the membrane. Solid lines (): measurement at space downstream of the membrane. Green lines: measurement at a time before flipping the membrane. Orange lines: measurement at a time after flipping the membrane.

FIG. 8.

Data for absorbance A as a function of wavelength λ. Higher values of A indicate larger concentrations of reactant. (a) Violet solid line (): measurement at a position upstream of the membrane. Violet dashed line (): measurement at a position downstream of the membrane (no catalyst). Orange line (): measurement at a position downstream of the membrane (catalyst loaded preferentially near inlet). Green line (): measurement at a position downstream of the membrane (catalyst loaded preferentially near the outlet). (b) Dashed lines (): measurement at a position upstream of the membrane. Solid lines (): measurement at space downstream of the membrane. Green lines: measurement at a time before flipping the membrane. Orange lines: measurement at a time after flipping the membrane.

Close modal

In this paper, we analyzed the performance of an isothermal slit-shaped gas-phase reactor whose walls are coated with a catalytic material and through which flow is driven via an externally imposed pressure drop. We derived a Fick–Jacobs theory, where the quasi-3D Navier–Stokes and advection-diffusion equations that govern the transport of the gaseous reactant and product were systematically reduced to an effective one-dimensional set of equations. These equations apply to corrugated reactors whose length far exceeds their average height and which are symmetric with respect to their centerline. Closed-form solutions for the pressure profile inside the reactor, the composition of the gas, and the flow of product and reactant are given for any corrugation that varies weakly. Furthermore, the density of the catalytic material is also taken as spatially varying and is the rate of the chemical reaction that turns the reactant into product.

To investigate the effect of corrugation and the inhomogeneous reaction rate, we studied a model reactor where both the reactor shape and reaction rate profile are sinusoidal. For small pressure drops relative to the inlet pressure, flat reactors exhibit the highest flow rate of the product when compared with corrugated reactors of the same volume. In this incompressible limit, the reactor yield is insensitive to the distribution of the catalytic material, which we verify experimentally. As such, the often laborious task of ensuring a homogeneous catalytic coating may be bypassed when dealing with incompressible flows, such as liquid-phase reactors. For compressible flows, we find an optimal non-flat reactor shape. This optimum shape is controlled by the inhomogeneity of the catalytic coating, as is the gain in product flow rate compared to a flat reactor. We find that reactors, where the catalytic material is mostly located near the reactor inlet, can be optimized by a shape that is narrowest in the center. Conversely, a reactor where catalytic material is mostly found near the outlet can be optimized by a shape that is widest at the center.

Thus, one may optimize the output of a reactor via its shape or via an appropriate distribution of catalytic material rather than by the addition of further costly catalysts. We listed three reactions for which the assumptions of our model hold: decomposition of methane and carbon monoxide and production of ammonia. All three are of environmental or economic relevance.67,71,72 Finally, our results for the highly compressible regime may also be relevant for the design of novel low-pressure gas chromatography systems. These devices frequently operate with near-vacuum conditions at the outlet,73,74 thus exhibiting values of relative pressure drop close to its maximum value.

We acknowledge the funding by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) (Project IDs 416229255-SFB 1411 and 431791331-SFB 1452). We thank the Helmholtz Association of German Research Centers (HGF) and the Federal Ministry of Education and Research (BMBF), Germany, for supporting the Innovation Pool project “Solar H2: Highly Pure and Compressed.”

The authors have no conflicts to disclose.

G. C. Antunes: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Writing – original draft (equal). M. Jiménez-Sánchez: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Writing – original draft (equal). P. Malgaretti: Conceptualization (equal); Formal analysis (equal); Supervision (equal); Writing – review & editing (equal). J. Bachmann: Conceptualization (equal); Funding acquisition (equal); Supervision (equal); Writing – review & editing (equal). J. Harting: Conceptualization (equal); Funding acquisition (equal); Supervision (equal); Writing – review & editing (equal).

The data that support the findings of this study is openly available in Zenodo at http://doi.org/10.5281/zenodo.10161303.

As stated in the main text, we now show how to solve Eq. (34) by a systematic reduction of the quasi-3D problem to an effective one-dimensional problem. We assume a long reactor such that the Péclet number Pex along the x direction is large,
(A1)
where v̄x is the typical velocity in the x direction. In such a reactor, advection is the dominant transport method in the longitudinal direction and
(A2)
(A3)
where jD(A)x(x,z) and jD(B)x(x,z) are the x components of the diffusive fluxes of A and B species, respectively. Thus, the fluxes in the x direction are obtained from Eqs. (20) and (21) as
(A4)
(A5)
where (jA)x(x,z) and (jB)x(x,z) are the x components of fluxes of A and B particles, respectively. Plugging Eqs. (A4) and (A5) in Eq. (34) yields
(A6)
We obtain vx(x, z) from solving the Navier–Stokes equation [Eq. (6)] for the steady state,
(A7)
the x and z components of which are
(A8)
(A9)
respectively. We now restrict ourselves to the case of a thin reactor,
(A10)
which greatly simplifies the solution of Eqs. (A8) and (A9). Let us normalize the x and z coordinates by the reactor length and average height, respectively. As such, we introduce the dimensionless variables x̂ and ẑ such that
(A11)
(A12)
Furthermore, we define the typical velocity scale v̄z in the z direction. As such, we introduce the dimensionless velocities
(A13)
(A14)
which are of order unity. Similarly, we introduce the typical pressure scale P̄ and mass density scale ρ̄f, along with the dimensionless pressure and mass density
(A15)
(A16)
which are numbers of order unity. Using these dimensionless variables, Eq. (A8) is written as
(A17)
and Eq. (A9) as
(A18)
We can relate the two velocity scales v̄x and v̄z using the continuity equation [Eq. (8)], which is rewritten as
(A19)
Equation (A19) then yields
(A20)
where the non-dimensional number K̂ is defined as
(A21)
Note that if ẑP̂(x̂,ẑ)v̂z(x̂,ẑ)=0, then according to Eq. (A19), so is x̂P̂(x̂,ẑ)v̂z(x̂,ẑ)=0, and Eq. (A20) still holds. Because the dimensionless velocities and positions are of order unity, so is K̂. Plugging Eq. (A20) in Eq. (A17) yields
(A22)
and in Eq. (A18) yields
(A23)
We now show that P̂ẑ(x̂,ẑ) is much smaller than P̂x̂(x̂,ẑ). To do so, we calculate
(A24)
which, when expanded in powers of the number h0/L (which is much smaller than 1), yields
(A25)
where we define the Reynolds number Re as
(A26)
which we take to be small such that
(A27)
Note that 2v̂x2ẑ(x̂,ẑ) is the largest contribution inside the curly brackets of Eq. (A24) only in regions of space away from the points where 2v̂x2ẑ(x̂,ẑ)=0. While such points cannot be ruled out a priori for generic flow fields (even assuming thin reactors), they are not expected to occur often. By combining Eqs. (A23) and (A25), we obtain
(A28)
Assuming that the reactor is thin enough so that
(A29)
one concludes from Eq. (A28) that
(A30)
Accordingly, we now perform the approximation of neglecting the gradient of pressure in the transverse direction,
(A31)
and according to Eq. (5)
(A32)
Having determined the pressure along the transverse direction, we will now turn to the number densities ρA(x, z, t) and ρB(x, z, t). As the sum of the two is proportional to the pressure, it is sufficient to determine one of them. Focusing on the number density of the reactant, we wish to quantify the variation of ρA(x, z, t) in the z direction and write Eq. (11) for the steady state (ρ̇A(x,z,t)=0),
(A33)
where the diffusive flux in the x direction has been neglected as per Eq. (A2). By applying the same reasoning as for the pressure, we write Eq. (A33) as
(A34)
where we have used Eq. (13). The Dirac deltas in Eq. (A34) are now non-dimensional, as the arguments ẑ+h(x̂)/h0 are also non-dimensional. Using Eq. (A20), we obtain
(A35)
which can be rewritten as
(A36)
Integrating along the z direction from − to ẑ yields
(A37)
To proceed, we assume the reactor to be so thin that the Péclet number Pez along the z direction is much smaller than 1,
(A38)
We also identify the Damköhler number in the transverse direction Daz. This number is the ratio of the timescale associated with diffusing along the cross-section h02/D and the timescale associated with the reaction h0/ᾱ,
(A39)
where ᾱ is the typical scale for the reaction rate. We also assume the channel to be thin enough so that Daz is small,
(A40)
and as a result,
(A41)
Thus, the number density of both reactant and product may be approximated as constant in the transverse direction,
(A42)
(A43)
With the approximation of Eq. (A31), we may rewrite Eq. (A22) as
(A44)
and as h0/L ≪ 1, we employ the lubrication approximation60 and keep only the terms on the right-hand-side that are of leading order in (h0/L)2 or Re(h0/L),
(A45)
where we have returned to dimensional variables for clarity of reading. Integrating both sides over the z direction twice yields
(A46)
where we have used
(A47)
coming from the no-slip boundary condition. The flow described in Eq. (A46) is Poiseuille-like with a local pressure gradient xP(x).

Reagents for the preparation of the samples were purchased from Sigma–Aldrich, Alfa Aesar, Strem, or VWR and used as received. A Millipore Direct-Q system was employed for the purification of water before use. Aluminum plates (99.99%) for the anodization process were purchased from Smart Membranes GmbH. The generation of the ordered system of parallel nanopores involves a two-step anodization procedure described in the literature.75,76 The specific anodization parameters used here include the applied voltage 195 V, the electrolyte solution 1 wt. % H3PO4, the anodization temperature 0 °C, and its duration 18 h. Thereafter, the remaining metallic aluminum substrate was removed by CuCl2 solution (0.7M in 10% HCl). Finally, the samples were treated with a 10% v/v H3PO4 solution for 60 min. at 45 °C to etch the Al2O3 barrier layer, open the lower pore extremities, and widen the pore diameter to ∼400 nm. These nanoporous membranes were finally laser-cut into individual nanoreactors with 4 mm diameter. Catalytically active SCALMS nanoparticles with a diameter of 30 ± 5 nm were synthesized by a hot injection method adapted from a published procedure.77 They were suspended in chloroform (1 mg/ml) and then infiltrated into the nanoporous reactors by placing 25 μl of suspension on a 4 mm reactor, then passing the liquid using vacuum filtration, and repeating the procedure four times in total. The sample was washed with acetone, isopropanol, and, finally, with distilled water, then left to dry under vacuum.

The catalytic model reaction adapts a procedure published by Rahim et al.,78 diluting the 100 mg/l methylene blue solution in 0.1M aqueous HCl further to 10 μM (or 3.6 mg/l, a dilution by a factor 28). 100 μl of a freshly prepared ascorbic acid solution (1 mg/ml in water) was added to a glass vial containing 30 ml of the methylene blue solution, then the mixture was flown at a rate of 100 μl/min through the nanoreactor, consisting of a SCALMS-loaded anodized alumina membrane of 4 mm diameter held in a Kapton ring and sandwiched between both parts of a 3D-printed PMMA holder. The holder consists of two parts in Swagelok VCR geometry that seal on the Kapton ring while supporting the porous nanoreactor on a backing grid. The reaction medium was collected in a vial in individual aliquots and analyzed by UV–Vis spectrometry.

The catalyst distribution and chemical composition along the nanopores were analyzed using a JSM-F100 field-emission scanning electron microscope from JEOL equipped with an EDX detector. These results were averaged over ten samples. The only nanoparticles that deposit close to the pore outlet are the ones that avoided deposition further upstream. As such, the closer to the outlet, the less likely it is for a nanoparticle to deposit, leading to a fore-aft asymmetric distribution of catalytic material inside the pore.

1.
J.
Hagen
,
Industrial Catalysis: A Practical Approach
(
Wiley VCH
,
2015
).
2.
A.
Tanimu
,
G.
Tanimu
,
H.
Alasiri
, and
A.
Aitani
, “
Catalytic cracking of crude oil: Mini review of catalyst formulations for enhanced selectivity to light olefins
,”
Energy Fuels
36
,
5152
5166
(
2022
).
3.
B.
Zhu
,
Z.
Liang
, and
R.
Zou
, “
Designing advanced catalysts for energy conversion based on urea oxidation reaction
,”
Small
16
,
1906133
(
2020
).
4.
Y.
Pei
,
Y.
Cheng
,
J.
Chen
,
W.
Smith
,
P.
Dong
,
P. M.
Ajayan
,
M.
Ye
, and
J.
Shen
, “
Recent developments of transition metal phosphides as catalysts in the energy conversion field
,”
J. Mater. Chem. A
6
,
23220
23243
(
2018
).
5.
D.
Shekhawat
,
J. J.
Spivery
, and
D.
Berry
,
Fuel Cells: Technologies for Fuel Processing
(
Elsevier Science
,
2011
).
6.
P. T.
Anastas
,
M. M.
Kirchhoff
, and
T. C.
Williamson
, “
Catalysis as a foundational pillar of green chemistry
,”
Appl. Catal., A
221
,
3
13
(
2001
).
7.
P. T.
Anastas
and
M. M.
Kirchhoff
, “
Origins, current status, and future challenges of green chemistry
,”
Acc. Chem. Res.
35
,
686
694
(
2002
).
8.
J. B.
Zimmerman
,
P. T.
Anastas
,
H. C.
Erythropel
, and
W.
Leitner
, “
Designing for a green chemistry future
,”
Science
367
,
397
400
(
2020
).
9.
A.
Riisager
,
R.
Fehrmann
,
M.
Haumann
, and
P.
Wasserscheid
, “
Supported ionic liquid phase (SILP) catalysis: An innovative concept for homogeneous catalysis in continuous fixed-bed reactors
,”
Eur. J. Inorg. Chem.
2006
,
695
706
.
10.
A.
Riisager
,
R.
Fehrmann
,
M.
Haumann
, and
P.
Wasserscheid
, “
Catalytic silp materials
,” in
Regulated Systems for Multiphase Catalysis
, edited by
W.
Leitner
and
M.
Hölscher
(
Springer
,
Berlin, Heidelberg
,
2008
), pp.
149
161
.
11.
N.
Taccardi
,
M.
Grabau
,
J.
Debuschewitz
,
M.
Distaso
,
M.
Brandl
,
R.
Hock
,
F.
Maier
,
C.
Papp
,
J.
Erhard
,
C.
Neiss
,
W.
Peukert
,
A.
Görling
,
H.-P.
Steinrück
, and
P.
Wasserscheid
, “
Gallium-rich Pd–Ga phases as supported liquid metal catalysts
,”
Nat. Chem.
9
,
862
867
(
2017
).
12.
G.
Rupprechter
, “
Popping up to the surface
,”
Nat. Chem.
9
,
833
834
(
2017
).
13.
U.
Kernchen
,
B.
Etzold
,
W.
Korth
, and
A.
Jess
, “
Solid catalyst with ionic liquid layer (SCILL)—A new concept to improve selectivity illustrated by hydrogenation of cyclooctadiene
,”
Chem. Eng. Technol.
30
,
985
994
(
2007
).
14.
T.
Boger
,
A. K.
Heibel
, and
C. M.
Sorensen
, “
Monolithic catalysts for the chemical industry
,”
Ind. Eng. Chem. Res.
43
(
16
),
4602
4611
(
2004
).
15.
Z. I.
Önsen
and
A. K.
Avci
,
Multiphase Catalytic Reactors: Theory, Design, Manufacturing, and Applications
(
John Wiley & Sons, Ltd.
,
2016
), pp.
171
212
.
16.
J. P.
du Plessis
and
S.
Woudberg
, “
Pore-scale derivation of the Ergun equation to enhance its adaptability and generalization
,”
Chem. Eng. Sci.
63
,
2576
2586
(
2008
).
17.
I. F.
Macdonald
,
M. S.
El-Sayed
,
K.
Mow
, and
F. A. L.
Dullien
, “
Flow through porous media-the Ergun equation revisited
,”
Ind. Eng. Chem. Fundam.
18
,
199
208
(
1979
).
18.
Y.
Bachmat
and
J.
Bear
, “
Macroscopic modelling of transport phenomena in porous media. 1: The continuum approach
,”
Transp. Porous Media
1
,
213
240
(
1986
).
19.
J.
Bear
and
Y.
Bachmat
, “
Macroscopic modelling of transport phenomena in porous media. 2: Applications to mass, momentum and energy transport
,”
Transp. Porous Media
1
,
241
269
(
1986
).
20.
A. L.
Negrini
,
A.
Fuelber
,
J. T.
Freire
, and
J. C.
Thoméo
, “
Fluid dynamics of air in a packed bed: Velocity profiles and the continuum model assumption
,”
Braz. J. Chem. Eng.
16
,
421
(
1999
).
21.
I.
Battiato
and
D. M.
Tartakovsky
, “
Applicability regimes for macroscopic models of reactive transport in porous media
,”
J. Contam. Hydrol.
120–121
,
18
26
(
2011
).
22.
R. M.
Heck
,
R. J.
Farrauto
, and
S. T.
Gulati
,
Catalytic Air Pollution Control: Commercial Technology
(
Wiley
,
2009
).
23.
J.
Ran
,
X.
Wang
,
Y.
Liu
,
S.
Yin
,
S.
Li
, and
L.
Zhang
, “
Microreactor-based micro/nanomaterials: Fabrication, advances, and outlook
,”
Mater. Horiz.
10
,
2343
2372
(
2023
).
24.
G.
Kolb
and
V.
Hessel
, “
Micro-structured reactors for gas phase reactions
,”
Chem. Eng. J.
98
,
1
38
(
2004
).
25.
S. H.
DeWitt
, “
Micro reactors for chemical synthesis
,”
Curr. Opin. Chem. Biol.
3
,
350
356
(
1999
).
26.
P.
Anastas
and
N.
Eghbali
, “
Green chemistry: Principles and practice
,”
Chem. Soc. Rev.
39
,
301
312
(
2010
).
27.
W.
Qing
,
F.
Liu
,
H.
Yao
,
S.
Sun
,
C.
Chen
, and
W.
Zhang
, “
Functional catalytic membrane development: A review of catalyst coating techniques
,”
Adv. Colloid Interface Sci.
282
,
102207
(
2020
).
28.
J.
Sengers
,
M.
Klein
,
J.
Gallagher
, and
National Bureau of Standards Washington D C Heat Div
,
Pressure-Volume-Temperature Relationships of Gases Virial Coefficients
(
Defense Technical Information Center
,
1971
).
29.
E. W.
Lemmon
,
I. H.
Bell
,
M. L.
Huber
, and
M. O.
McLinden
, “
Thermophysical properties of fluid systems
,” in
NIST Chemistry WebBook, NIST Standard Reference Database Number 69
, edited by
P. J.
Linstrom
and
W. G.
Mallard
(
National Institute of Standards and Technology
,
Gaithersburg, MD, 2089
) (retrieved 28 November 2023).
30.
M.
Le Bellac
,
F.
Mortessagne
, and
G. G.
Batrouni
,
Equilibrium and Non-Equilibrium Statistical Thermodynamics
(
Cambridge University Press
,
2004
).
31.
B.
Sharma
and
R.
Kumar
, “
Estimation of bulk viscosity of dilute gases using a nonequilibrium molecular dynamics approach
,”
Phys. Rev. E
100
,
013309
(
2019
).
32.
P.
Atkins
and
J.
de Paula
,
Atkins’ Physical Chemistry
(
Oxford University Press
,
2006
).
33.

It is worthwhile to note that δ(z) has unit meter−1, as can be seen from how the integral aaδ(z)dz=1 is dimensionless (for any value of a ≠ 0).79 

34.
S. H.
Lam
, “
Multicomponent diffusion revisited
,”
Phys. Fluids
18
,
073101
(
2006
).
35.
R.
Krishna
, “
Problems and pitfalls in the use of the fick formulation for intraparticle diffusion
,”
Chem. Eng. Sci.
48
,
845
861
(
1993
).
36.
S.
Kjelstrup
and
D.
Bedeaux
,
Non-Equilibrium Thermodynamics of Heterogeneous Systems
(
WSPC
,
2020
).
37.
S.
Govender
and
H. B.
Friedrich
, “
Monoliths: A review of the basics, preparation methods and their relevance to oxidation
,”
Catalysts
7
,
62
(
2017
).
38.
J. A.
Moulijn
and
F.
Kapteijn
, “
Monolithic reactors in catalysis: Excellent control
,”
Curr. Opin. Chem. Eng.
2
,
346
353
(
2013
), a part of Special Issue: Energy and Environmental Engineering/Reaction Engineering and Catalysis.
39.
R.
Zwanzig
, “
Diffusion past an entropy barrier
,”
J. Phys. Chem.
96
,
3926
(
1992
).
40.
D.
Reguera
and
J. M.
Rubi
, “
Kinetic equations for diffusion in the presence of entropic barriers
,”
Phys. Rev. E
64
,
061106
(
2001
).
41.
P.
Kalinay
and
J. K. P.
Percus
, “
Projection of two-dimensional diffusion in a narrow channel onto the longitudinal dimension
,”
J. Chem. Phys.
122
,
204701
(
2005
).
42.
P.
Kalinay
and
J. K.
Percus
, “
Approximations of the generalized Fick-Jacobs equation
,”
Phys. Rev. E
78
,
021103
(
2008
).
43.
S.
Martens
,
G.
Schmid
,
L.
Schimansky-Geier
, and
P.
Hänggi
, “
Entropic particle transport: Higher-order corrections to the Fick-Jacobs diffusion equation
,”
Phys. Rev. E
83
,
051135
(
2011
).
44.
G.
Chacón-Acosta
,
I.
Pineda
, and
L.
Dagdug
, “
Diffusion in narrow channels on curved manifolds
,”
J. Chem. Phys.
139
,
214115
(
2013
).
45.
P.
Malgaretti
and
J.
Harting
, “
Closed formula for transport across constrictions
,”
Entropy
25
,
470
(
2023
).
46.
D.
Reguera
,
G.
Schmid
,
P. S.
Burada
,
J. M.
Rubi
,
P.
Reimann
, and
P.
Hänggi
, “
Entropic transport: Kinetics, scaling, and control mechanisms
,”
Phys. Rev. Lett.
96
,
130603
(
2006
).
47.
D.
Reguera
,
A.
Luque
,
P. S.
Burada
,
G.
Schmid
,
J. M.
Rubi
, and
P.
Hänggi
, “
Entropic splitter for particle separation
,”
Phys. Rev. Lett.
108
,
020604
(
2012
).
48.
U.
Marini Bettolo Marconi
,
P.
Malgaretti
, and
I.
Pagonabarraga
, “
Tracer diffusion of hard-sphere binary mixtures under nano-confinement
,”
J. Chem. Phys.
143
,
184501
(
2015
).
49.
P.
Malgaretti
,
I.
Pagonabarraga
, and
J.
Rubi
, “
Rectification and non-Gaussian diffusion in heterogeneous media
,”
Entropy
18
,
394
(
2016
).
50.
A.
Puertas
,
P.
Malgaretti
, and
I.
Pagonabarraga
, “
Active microrheology in corrugated channels
,”
J. Chem. Phys.
149
,
174908
(
2018
).
51.
V.
Bianco
and
P.
Malgaretti
, “
Non-monotonous polymer translocation time across corrugated channels: Comparison between Fick-Jacobs approximation and numerical simulations
,”
J. Chem. Phys.
145
,
114904
(
2016
).
52.
E.
Locatelli
,
V.
Bianco
,
C.
Valeriani
, and
P.
Malgaretti
,
Phys. Rev. Lett.
131
,
048101
(
2023
).
53.
P.
Malgaretti
,
I.
Pagonabarraga
, and
J. M.
Rubi
, “
Entropic electrokinetics: Recirculation, particle separation, and negative mobility
,”
Phys. Rev. Lett.
113
,
128301
(
2014
).
54.
M.
Chinappi
and
P.
Malgaretti
, “
Charge polarization, local electroneutrality breakdown and eddy formation due to electroosmosis in varying-section channels
,”
Soft Matter
14
,
9083
(
2018
).
55.
P.
Malgaretti
,
M.
Janssen
,
I.
Pagonabarraga
, and
J. M.
Rubi
, “
Driving an electrolyte through a corrugated nanopore
,”
J. Chem. Phys.
151
,
084902
(
2019
).
56.
M.
Sandoval
and
L.
Dagdug
, “
Effective diffusion of confined active brownian swimmers
,”
Phys. Rev. E
90
,
062711
(
2014
).
57.
P.
Kalinay
, “
Transverse dichotomic ratchet in a two-dimensional corrugated channel
,”
Phys. Rev. E
106
,
044126
(
2022
).
58.
G. C.
Antunes
,
P.
Malgaretti
,
J.
Harting
, and
S.
Dietrich
, “
Pumping and mixing in active pores
,”
Phys. Rev. Lett.
129
,
188003
(
2022
).
59.
G. C.
Antunes
,
P.
Malgaretti
, and
J.
Harting
, “
Turning catalytically active pores into active pumps
,”
J. Chem. Phys.
159
,
134903
(
2023
).
60.
H.
Schlichting
,
Boundary Layer Theory
(
McGraw-Hill
,
1979
).
61.
M. S.
Cramer
, “
Numerical estimates for the bulk viscosity of ideal gases
,”
Phys. Fluids
24
,
066102
(
2012
).
62.
L. M.
Brekhovskikh
and
V.
Goncharov
,
Mechanics of Continua and Wave Dynamics
,
Springer Series on Wave Phenomena
(
Springer
,
1994
).
63.
K.
Kadoya
,
N.
Matsunaga
, and
A.
Nagashima
, “
Viscosity and thermal conductivity of dry air in the gaseous phase
,”
J. Phys. Chem. Ref. Data
14
,
947
970
(
1985
).
64.
C. R.
Wilke
and
C. Y.
Lee
, “
Estimation of diffusion coefficients for gases and vapors
,”
Ind. Eng. Chem.
47
,
1253
1257
(
1955
).
65.
M.
Reinke
,
J.
Mantzaras
,
R.
Schaeren
,
R.
Bombach
,
A.
Inauen
, and
S.
Schenker
, “
High-pressure catalytic combustion of methane over platinum: In situ experiments and detailed numerical predictions
,”
Combust. Flame
136
,
217
240
(
2004
).
66.
C.
Papp
and
H.-P.
Steinrück
, “
In situ high-resolution x-ray photoelectron spectroscopy—Fundamental insights in surface reactions
,”
Surf. Sci. Rep.
68
,
446
487
(
2013
).
67.
S.
Ghavam
,
M.
Vahdati
,
I. A. G.
Wilson
, and
P.
Styring
, “
Sustainable ammonia production processes
,”
Front. Energy Res.
9
,
580808
(
2021
).
68.
G.
Xu
,
R.
Liu
, and
J.
Wang
, “
Electrochemical synthesis of ammonia using a cell with a Nafion membrane and SmFe0.7Cu0.3−xNixO3 (x = 0–0.3) cathode at atmospheric pressure and lower temperature
,”
Sci. China, Ser. B: Chem.
52
,
1171
1175
(
2009
).
69.
S.
Colin
,
P.
Lalonde
, and
R.
Caen
, “
Validation of a second-order slip flow model in rectangular microchannels
,”
Heat Transfer Eng.
25
,
23
30
(
2004
), cited by: 249; All Open Access, Bronze Open Access, Green Open Access.
70.
S.
Jennings
, “
The mean free path in air
,”
J. Aerosol Sci.
19
,
159
166
(
1988
).
71.
P.
Balcombe
,
K.
Anderson
,
J.
Speirs
,
N.
Brandon
, and
A.
Hawkes
, “
The natural gas supply chain: The importance of methane and carbon dioxide emissions
,”
ACS Sustain. Chem. Eng.
5
,
3
20
(
2017
).
72.
J. A.
Raub
,
M.
Mathieu-Nolf
,
N. B.
Hampson
, and
S. R.
Thom
, “
Carbon monoxide poisoning—A public health perspective
,”
Toxicology
145
,
1
14
(
2000
).
73.
M.
Zoccali
,
P. Q.
Tranchida
, and
L.
Mondello
, “
Fast gas chromatography-mass spectrometry: A review of the last decade
,”
TrAC, Trends Anal. Chem.
118
,
444
452
(
2019
).
74.
J.
de Zeeuw
,
J.
Peene
,
H.-G.
Jansen
, and
X.
Lou
, “
A simple way to speed up separations by GC-MS using short 0.53 mm columns and vacuum outlet conditions
,”
J. High Resolut. Chromatogr.
23
,
677
680
(
2000
).
75.
J.
Englhard
,
Y.
Cao
,
S.
Bochmann
,
M. K. S.
Barr
,
S.
Cadot
,
E. A.
Quadrelli
, and
J.
Bachmann
, “
Stabilizing an ultrathin MoS2 layer during electrocatalytic hydrogen evolution with a crystalline SnO2 underlayer
,”
RSC Adv.
11
,
17985
17992
(
2021
).
76.
C. V.
Ruiz M
,
M.
Terlinden
,
M.
Engelhardt
,
G.
Magnabosco
,
G.
Papastavrou
,
N.
Vogel
,
M.
Thommes
, and
J.
Bachmann
, “
Preparation and surface functionalization of a tunable porous system featuring stacked spheres in cylindrical pores
,”
Adv. Mater. Interfaces
10
,
2300436
(
2023
).
77.
M.
Yarema
,
M.
Wörle
,
M. D.
Rossell
,
R.
Erni
,
R.
Caputo
,
L.
Protesescu
,
K. V.
Kravchyk
,
D. N.
Dirin
,
K.
Lienau
,
F.
von Rohr
,
A.
Schilling
,
M.
Nachtegaal
, and
M. V.
Kovalenko
, “
Monodisperse colloidal gallium nanoparticles: Synthesis, low temperature crystallization, surface plasmon resonance and Li-ion storage
,”
J. Am. Chem. Soc.
136
,
12422
12430
(
2014
).
78.
M. A.
Rahim
,
J.
Tang
,
A. J.
Christofferson
,
P. V.
Kumar
,
N.
Meftahi
,
F.
Centurion
,
Z.
Cao
,
J.
Tang
,
M.
Baharfar
,
M.
Mayyas
,
F.-M.
Allioux
,
P.
Koshy
,
T.
Daeneke
,
C. F.
McConville
,
R. B.
Kaner
,
S. P.
Russo
, and
K.
Kalantar-Zadeh
, “
Low-temperature liquid platinum catalyst
,”
Nat. Chem.
14
,
935
941
(
2022
).
79.
C. M.
Bender
and
S. A.
Orszag
,
Advanced Mathematical Methods for Scientists and Engineers I
, 1st ed. (
Springer
,
New York
,
1978
).