• Non ci sono risultati.

Chemical Engineering Science

N/A
N/A
Protected

Academic year: 2021

Condividi "Chemical Engineering Science"

Copied!
14
0
0

Testo completo

(1)

Contents lists available atScienceDirect

Chemical Engineering Science

journal homepage:w w w . e l s e v i e r . c o m / l o c a t e / c e s

Hydrocarbon storage modeling for diesel oxidation catalysts

Chaitanya S. Sampara

a,

, Edward J. Bissett

b

, Dennis Assanis

a

aWalter E. Lay Automotive Laboratory, 1231 Beal Avenue, Ann Arbor, 48109 MI, USA

bChemical and Environmental Sciences Laboratory, General Motors R&D, 30500 Mound Road, Warren, 48090 MI, USA

A R T I C L E I N F O A B S T R A C T

Article history:

Received 16 January 2008

Received in revised form 1 May 2008 Accepted 14 June 2008

Available online 26 June 2008

Keywords:

Hydrocarbon adsorption–desorption kinetics Zeolites

Diesel oxidation catalysts Diesel engine cold-start emissions

A simple hydrocarbon (HC) storage model was proposed which represents the adsorption–desorption processes in zeolites incorporated in a diesel oxidation catalyst (DOC). Using four experiments in which the HC was stepped up and the reactor run until steady state, a Langmuir isotherm was generated that was sufficient to represent the equilibrium data, and the remaining rate constant capturing the adsorption time scale was obtained by fitting. The model thus developed was validated using reactor data which was obtained by stepping the inlet HC concentration to zero after outlet reached equilibrium.

A typical DOC which contained both a storage component (zeolite) and noble metal component (for oxidation) was studied using a full scale 1D reactor model which includes the storage kinetics developed here with the oxidation kinetics developed in our previous work [Sampara et al., 2008. Global kinetics for a commercial diesel oxidation catalyst with two exhaust hydrocarbons. Industrial & Engineering Chemistry Research 47, 311–322]. Simulations of a simplified warm-up process indicated that the zeolite storage component reduces the overall cold start HC emissions by at least a factor of two if the warm-up rate achieves 4565C min−1, a range commonly observed during start-up. Modeling results also showed that the HC oxidation for these reactors commonly starts at the rear end of the reactor due to reduced CO inhibition. The rates of the individual processes during cold start were analyzed in detail and compared with the rate of inlet temperature increase provided from the exhaust.

© 2008 Elsevier Ltd. All rights reserved.

1. Introduction

Reducing engine emissions resulting from cold start is a ma- jor impediment in meeting emissions standards. For late-model gasoline engines, nearly 60–70% of the total hydrocarbon (THC) emissions occur during cold start. In diesel engines, while the THC concentrations resulting from conventional combustion modes are low, advanced combustion strategies, such as pre-mixed compres- sion ignition (PCI), which are being used simultaneously to reduce soot and NOx, tend to produce significant amounts of THC and CO.

A diesel oxidation catalyst (DOC), which can oxidize all the THC and CO, is being proposed to meet the emissions standards in these scenarios. While such an oxidation catalyst is efficient in reducing THC and CO emissions after it is fully warmed-up, other strategies are being researched to address the cold start engine emissions.

Zeolites have been proven to be effective in storing cold start hydrocarbons. Catalyzed hydrocarbon traps, which contain a mix- ture of an adsorbent material, such as zeolite, and noble metals in the same washcoat, provide both trapping and oxidation functions.

Corresponding author. Tel.: +1 34 272 2091; fax: +1 734 764 4256.

E-mail address:csampara@umich.edu(C.S. Sampara).

0009-2509/$ - see front matter©2008 Elsevier Ltd. All rights reserved.

doi:10.1016/j.ces.2008.06.021

The use of zeolites as efficient hydrocarbon trap systems to re- duce cold-start hydrocarbon emissions has been well demonstrated for both gasoline (Mitsuishi et al., 1999; Yamamoto et al., 2000;

Nishizawa et al., 2000; Kanazawa, 2004) and diesel (Kamijo et al., 2001; Adams et al., 1996; Banno et al., 2004; Lafyatis et al., 1998) applications. For diesel applications zeolite is typically a part of the DOC which is used to oxidize CO and THC in the stream. This DOC will adsorb the hydrocarbons at low temperatures. As the temper- ature increases, hydrocarbons adsorbed on the zeolite desorb from the surface. However, if the noble metal becomes significantly ac- tive by this stage, the desorbed hydrocarbons are oxidized on the noble metals thus leading to near zero hydrocarbon emissions. CO is oxidized normally after catalyst light-off. Capturing THCs is more important since the CO oxidation occurs earlier compared to HC be- cause the CO reaction rate is faster (Sampara et al., 2008). In cases where the desorption of hydrocarbons takes place before catalyst light-off, we observe hydrocarbon slip, which leads to undesired hy- drocarbon emissions.

There are many different types of zeolites commercially available.

The use of a particular type of zeolite is application specific. Classi- fication of zeolite is commonly based on the Si/Al ratio. Two types of zeolites have been popular for automotive applications namely, Y- and-zeolites. These zeolites have 12 membered ring structures

(2)

and larger pore sizes compared to other forms of zeolites. To differ- entiate between these two forms, the Y-zeolites commonly have a lower Si/Al ratio (1–25) and a smaller pore size. The-zeolites have Si/Al ratios ranging between 10–100 and a larger pore size. There are obvious trade-offs between using these two types of zeolites.

While smaller pores or a low pore volume may lead to pore block- age by coke and limit the diffusional transport of feed and product molecules, larger pores lead to lower surface area and consequently reduced catalytic contribution (Weitkamp and Puppe, 1999). With respect to hydrocarbon adsorption, while Y-zeolite can adsorb straight chain hydrocarbons, -zeolite can adsorb both straight and ring type hydrocarbons (Corma, 1995). However, Y-zeolites are easier to produce and are known to have better performance characteristics with larger aliphatic hydrocarbons because of their higher bronsted acidity.

Equilibrium data for the adsorption–desorption reactions tak- ing place in such systems have been successfully represented with Langmuir isotherms (Goralski et al., 2000; Otto et al., 1991; Koltsakis and Stamatelos, 2000; Kryl et al., 2005). Others in the literature have also used slightly more complicated representations such as the Fruendlich isotherm to represent the equilibrium data (Watanabe et al., 2001). For diesel adsorption–desorption modeling Tanaka et al. (2005)used C3H6as their representative hydrocarbons. Kryl et al. (2005) used a combination of C3H6, toluene (C7H8) and n-decane (C10H22) as representative hydrocarbons in diesel exhaust.

They further assumed that C3H6does not adsorb on the zeolite sur- face in the presence of larger hydrocarbons. This is a good assump- tion in the case of 3D zeolites such as Y- or-zeolites (Czaplewski et al., 2002). Their model however indicates that toluene and n-decane are adsorbed onto two independent zeolite sites. This has not been otherwise observed in the literature. They also reported that both C10H22 and C7H8 have the same desorption activation energy and have nearly similar adsorption rate constants.

In the current paper, we develop a simple storage-release model for hydrocarbon storage on zeolites. We start by defining the experi- ments needed to generate adsorption and desorption rates, assuming that these are simple first-order processes on a single type of storage site, so as to obtain a Langmuir isotherm at equilibrium. Following the literature cited above, we idealize the HC content of the diesel exhaust with three representative hydrocarbons: propene (C3H6) as a partially oxidized HC, n-dodecane (C12H26) as a larger aliphatic HC, and toluene (C7H8) as an aromatic. C3H6is not expected to absorb, and preliminary testing showed no adsorption of toluene on our sam- ple. Therefore, transient experiments are conducted on a small scale reactor with n-dodecane (C12H26) representing adsorbable hydro- carbons. Total number of zeolite sites (Ntot), desorption activation energy (Edes) and the ratio of the adsorption and desorption pre- exponentials (Aads/Ades) are determined based on equilibrium data.

We then use simple optimization techniques coupled with simplified reactor codes to estimate one of the pre-exponentials while other is inferred by their ratio. Once the adsorption–desorption kinetic pa- rameters have been thus determined, we incorporate this rate into a full reactor model along with the oxidation rates developed in our previous work (Sampara et al., 2008) to understand the effect of var- ious rates of heat-up the catalyst undergoes during start-up. This allows us to assess the effectiveness of our catalyst in reducing HC emissions due to HC storage and to estimate the order of magnitude of the heat-up rate which gives improved performance due to the HC storage function.

2. Experimental

The hydrocarbon storage kinetics assumed here (i.e., Langmuir isotherm) allow us to generate the necessary adsorption–desorption rates with a minimum of four experiments for each hydrocarbon

species involved in the system. The derivation of equations which leads to this conclusion are discussed in the “Modeling” section. In this section we describe the experimental procedures, our choice of the HC species, temperature ranges considered and typical experi- mental results.

2.1. Test protocol

The experimental protocol should provide a means to infer the total number of moles of hydrocarbons that can be stored on a clean zeolite sample until equilibrium is reached. This equilibrium storage capacity is different from the total zeolite storage capacity and is a function of the hydrocarbon concentration and temperature of the system. The protocol should provide a means to validate the rate model thus developed, and should also include a procedure to clean the sample of any adsorbed hydrocarbons before attempting to run experiments on the same catalyst at a later stage.

We start all our experiments with a clean zeolite sample. Each experiment is performed in three phases. In the first and second phases the temperature is constant, and in the third phase the tem- perature increases linearly with time at about 10C s−1. The inlet HC concentration is held constant during the first phase and then dropped to zero in the second and third phases. Therefore, the first is an “adsorption phase”, in which the HCs adsorb on the zeolite surface until the outlet concentration reaches a constant value, that is, the equilibrium concentration for the given temperature. The second phase is then the “desorption phase” in which the negative concentration gradient between the surface and the gas phase HC concentrations cause the HCs to desorb from the zeolite surface.

Rather than wait for the HC concentration at the outlet of the reac- tor to approach all the way to zero at the given temperature, we end this phase when the outlet HC concentration shows a small gradient.

Finally in the third phase we perform a temperature programmed desorption (TPD) to ensure that all the HCs are removed from the zeolite surface. These experiments are illustrated in Figs. 1(a), (b) and 2.

We use the adsorption phase data to develop the adsorption–

desorption rate for the HC species under study. The desorption and the TPD phases are then used to validate the rate model. The TPD phase is also used to clean the zeolite surface of any remaining HCs before the next set of experiments could be performed.

2.2. Test matrix

With the test protocol defined, we specify the HC species which will represent the potentially adsorbable components (on zeolite) of diesel exhaust. We also establish the particular temperatures and HC concentrations used in the experiments.

Our choice of the hydrocarbons for this study is based on our previous work (Sampara et al., 2008), where we developed oxida- tion kinetics for various species in diesel oxidation catalysts. There we speciated our hydrocarbons as partially oxidized hydrocarbons, represented by C3H6, and unburnt fuel, represented by diesel fuel, and we assumed that each of these components was∼ 50% of the THCs on a molar C3basis. Here we further subdivide the unburnt fuel as a combination of n-dodecane (C12H26) and toluene (C7H8), again split equally for simplicity so that each of these components was 25% of the THCs.

The test matrix consisting of two HC concentrations and two tem- peratures is given inTable 1. From our earlier work we expected maximum inlet THCs for a DOC of around 2000 ppm on a C3 ba- sis, over the entire operating cycle. With our assumption that each potentially adsorbable representative HC (n-dodecane or toluene) might be 25% of the total, a reasonable concentration near the high end would be 340 ppm. The second concentration level chosen was

(3)

0 1000 2000 3000 4000 5000 0

10 20 30 40 50 60

time (s)

Dodecane concentration (ppm)

Inlet Outlet

0 500 1000 1500 2000

0 20 40 60 80 100 120 140 160 180

time (s)

Toluene concentration (ppm)

Inlet Outlet

Fig. 1. Typical adsorption–desorption experimental results for (a) dodecane and (b) toluene.

half of this. Note that these numbers will change when measured on an absolute scale, such as a C12(e.g., n-dodecane) basis. The two temperatures, 116C and 153C, were chosen as representative of where adsorption and desorption would be significant, respectively.

The lower temperature was also limited by the boiling points of n-dodecane and toluene.

2.3. Reactor set-up and analysis

An 8 g ft−3, 2:1 Pt: Pd/-Al2O3catalyst with zeolite was used for all the experiments. The monolith supported catalyst was hydro- thermally aged in a furnace at 650C for 16 hours to account for de-greening of the noble metal and dealumination of the zeolite.

0 1000 2000 3000 4000

100 120 140 160 180 200 220 240 260 280

time (s)

Temperature (°C)

Inlet Outlet

Fig. 2. Measured temperatures during the three phases of the experiment.

Table 1

Test matrix for generating a Langmuir isotherm

HC concentration (ppm, C3) Temperature (C)

340 116

170 116

340 153

170 153

A constant 2.2 L min−1flow of 10% H2O in air was fed to the fur- nace for the entire 16 hours aging period. The active site density (aj) for the noble-metal was found to be 0.483 mol-site m−3from CO chemisorption measurement, which corresponds to a dispersion of 26.1%. The description of the CO chemisorption method is given elsewhere (Sampara et al., 2007). We determined the zeolite loading capacity based on the Langmuir isotherm which is described in the next section.

The details of the reactor set-up are as follows:

• Monolith samples with thin washcoats (about 20m) were used to minimize diffusion resistance within the washcoat.

• Experiments were carried out in a 2 in. OD stainless steel tubular reactor containing a sample which is 1.5 in. in diameter and 1.5 in.

in length. Samples were held in place using a compressible ceramic paper wrap that prevented flow from bypassing the catalyst.

• The inlet gas contained 10% CO2, 8.7% H2O and the remainder N2. Water was vaporized at 400C using a length of coiled 1/4 in.

stainless steel tubing immersed in liquid Tin. No O2was included to avoid interference of oxidation.

• A Cole–Parmer 74900-10 syringe pump was used to inject either dodecane or toluene into the system. The hydrocarbons were in- jected directly into a heated N2 by-pass stream which helped vaporize and carry the fuel into the main flow. Since the mole- fraction of fuel after it enters the N2 by-pass stream is very low (ppm level), the saturation temperature corresponding to its par- tially pressure is substantially lower than the temperatures of our experiments. Hence we did not observe any condensation effects of the fuel. The fiberglass wick commonly used to supply liquid HCs for steady state experiments (Sampara et al., 2008) was re- jected for these transient experiments to decrease the response time for step changes in the HC concentration.

(4)

• Two inline heaters heated the inlet gas which in turn heated the catalyst. Reactor temperatures were monitored using using three Type K thermocouples: one at the inlet, one at the outlet, and one inside the catalyst at the approximate axial midpoint.

• After the inline heaters, the heated feed stream was mixed with the HC/N2mixture from the syringe pump. The complete inlet gas stream is then passed through a section of unheated silica mixing beads to attain the desired mixing and uniformity of the flow before entering the catalyst section.

• Two FTIRs were used to analyze the inlet and outlet gases simul- taneously. MKS MultiGas 2030 process stream FTIRs were used to analyze n-dodecane (C12H26), toluene (CH3-C6H5), H2O or CO2. No measurable quantities of aldehydes, alcohols or CH4were de- tected during the experiments.

• All the experiments were at the slightly elevated pressure of 1.15 atm to ensure proper flow through both the FTIRs.

• All experiments were carried out at a space velocity of 35, 281 h−1, which is the actual space velocity as seen by the catalyst sample after subtracting the flow going through the inlet FTIR.

2.4. Typical experimental results

Figs. 1(a) and (b) show typical results from the transient exper- iments conducted with dodecane and toluene. The temperatures measured by the upstream and downstream thermocouples are shown in Fig. 2. Note that during the adsorption and desorption phases the inlet temperature is held constant. After the desorption phase, the TPD continues until the HC concentration at the exit of the reactor is zero. For experimental purposes we roughly esti- mated 340 ppm on a C3basis, which converts to 100 ppm on a C12 (dodecane) basis and 160 ppm on a C7(toluene) basis.

We observe that n-dodecane exhibits storage behavior on the type of zeolite used for this study, and toluene does not. Therefore, the THCs in the exhaust are represented by C3H6, n-dodecane and toluene for purposes of modeling the HC oxidation and storage, with n-dodecane taken to be the only adsorbable hydrocarbon species on this zeolite.

3. Modeling

Let us denote the adsorbable hydrocarbon species, which for this study is n-dodecane, by DF1. The storage-release reaction can be written as

DF1+ ZeolDF1.Zeol (1)

Here Zeol refers to the zeolite sites which are not occupied by DF1, and DF1. Zeol represents the zeolite sites covered with hydrocar- bons. In the discussion, we will need to separately discuss the rate of the forward adsorption step, rads, and the reverse desorption step rdes. It is commonly assumed that the adsorption rate constant is non-activated, but the desorption rate constant has Arrhenius de- pendence with temperature. Therefore we propose a simple reaction rate model below.

rads= Aadscs,DF1(1−DF1) (2)

rdes= Adese−Edes/RTDF1 (3)

DF1 refers to the fraction of zeolite sites that are covered with HCs. It is important to note that the adsorption rate is expressed as a function of the concentration (mol m−3) rather than the mole- fraction of the HC species, DF1, to correctly capture the temperature dependence of this rate (c= p/RT).

From Eqs. (2) and (3), it is clear that we need three constants, namely, Aads, Ades and Edesto define the adsorption and desorp- tion rates. We also need the total zeolite storage capacity (Ntot) to completely define the species equations for DF1 in Appendix A and in Eq. (8). We develop these constants in the following discussion with the experimental data from the previous section. The ensuing discussion in this section is divided into three parts. First, we use adsorption phase data from the four experiments and perform equi- librium calculations to develop three of the four constants needed for the rates. Second, we perform transient calculations, where we use a simplified transient reactor code integrated with optimization routines to fit the adsorption phase data to generate the fourth con- stant. Finally, we use the desorption and TPD phases' data to validate the rate model thus developed.

3.1. Equilibrium calculations

At equilibrium, rads= rdes, so from Eqs. (2) and (3),

Aadscs,DF1(1−DF1)= Adese−Edes/RTDF1 (4) Re-arranging Eq. (4), and introducing the number of moles of DF1 stored (DF1= nDF1/Ntot),

1 neq,DF1= 1

Ntot + 1

K(T)∗ cs,DF1∗ Ntot

(Langmuir isotherm) (5)

where

K= Aads

Adese−Edes/RT (6)

In the left-hand side of Eq. (5), neq,DF1can also be measured directly from our experiments by integrating the total measured HCs into the catalyst minus the total measured HCs coming out of the catalyst until equilibrium. That is,

neq,DF1=

teq

0 (˙nDF1,in− ˙nDF1,out) dt (7)

where teqis the time at which the outlet HC concentration is in equi- librium with the inlet concentration. Eq. (5) is commonly referred to as the Langmuir isotherm. For a given temperature, it represents a line when 1/neq,DF1is plotted as a function of 1/cs,DF1. As there are no concentration gradients within the reactor at equilibrium, the inlet and outlet concentrations of n-dodecane indeed measure the concentration at the adsorber surface, cs,DF1.

At a fixed temperature (e.g., one of the temperatures in our test matrix,Table 1), two adsorption experiments at different HC con- centrations, run all the way to equilibrium, will generate the two points that determine the line of the Langmuir isotherm. The slope and intercept of this line then determines the two coefficients in Eq. (5), essentially Ntotand K at this temperature. Repeating this at the second temperature, not necessarily with the same two HC con- centrations, yields Ntotand K at this second temperature. Since Ntot is physically independent of temperature in the simplest case, the intercept of these two lines must be approximately the same, which we verify below for our data. We use K at two temperatures to gener- ate an Arrhenius plot (log(K) versus 1/T), which yields both the des- orption activation energy and the ratio of the two pre-exponentials.

3.1.1. Langmuir isotherm and Arrhenius plot

Before attempting to use the four sets of adsorption phase data to generate the isotherm, we performed two operations to decrease the effects of noise in the experimental data upon our subsequent calculations. First, after the target concentration was reached, we

(5)

Table 2

neq,DF1calculated for the experimental combinations of temperature and n-dodecane concentration

neq,DF1(moles) cs,DF1(C12basis) Temperature (C)

5.84× 10−4 3.5× 10−3 116

4.89× 10−4 1.8× 10−3 116

3.15× 10−4 2.9× 10−3 153

1.74× 10−4 1.4× 10−3 153

Concentrations were evaluated at the 1.15 atm of the experiments.

0 200 400 600 800

1000 1500 2000 2500 3000 3500 4000 4500 5000 5500 6000

1/cs, DF1 [mol/m3]−1 1/neq [mol−1]

116C 153C

y1 =1.05e3+1.90x y2 =1.05e3+6.48x

Fig. 3. Langmuir isotherms generated with the four concentration and two temper- ature combinations.

replaced the noisy inlet concentration with this constant target con- centration. Second, during the initial period in which the concentra- tions rose rapidly to the target inlet concentration, we smoothed all the experimental concentration data with a 12-point moving aver- age over the preceding 12 s of data.

The four sets of adsorption data, processed as above, were used to calculate the equilibrium molar capacities of n-dodecane, neq,DF1, from Eq. (7) and shown inTable 2. Besides being converted from mole fractions, the concentrations shown here only approximate the intended values fromTable 1. These values resulted in the two Langmuir isotherms at 116 and 153C in Fig. 3. Because Ntot is assumed independent of temperature, the least squares fit for these lines was constrained to have a common intercept, but the high de- gree of consistency with the data lends credibility to this assumption.

The results from this plot are used to calculate Ntot = 9.52×10−4mol for this sample of volume 45× 10−6m3. Therefore the surface site density for zeolite, which is used to solve the species equations de- scribed in the next section, is given by aze= 21.2 mol m−3. From the slopes of the two lines, K at the two temperatures of 116 and 153C is 553 and 162, respectively.Fig. 4shows the Arrhenius plot gener- ated by these two values of K, defined in Eq. (5). In the usual way, this yields Edes= 4.56 × 104J mol−1and Aads/Ades= 4.08 × 10−4. 3.2. Transient calculations

With Ntot, Edesand Aads/Adesalready known, estimating one of Aadsor Adesis enough to define all the constants. For our case, we

2.3 2.35 2.4 2.45 2.5 2.55 2.6

5 5.2 5.4 5.6 5.8 6 6.2 6.4 6.6

1/T

log (K)

data linear fit y = 5.49e+3x − 7.81

x 10−3

Fig. 4. Arrhenius plot of the ratio of rate constants.

chose to evaluate Aads by fitting model predictions based on the above rate expressions to the experimental data.

The general method to generate reaction rate constants by match- ing experiments in this manner involves solving an outer problem and a corresponding inner problem. The outer problem is the op- timization which minimizes the difference between the measured and calculated exit concentrations for all the tested conditions. The rate parameters are adjusted after every iteration to improve (de- crease) the objective function, which measures how well the model predictions are in agreement with the experimental measurements.

For each evaluation of the objective function we need solutions of a non-trivial inner problem where we predict the DF1 exit concen- tration as a function of time for each of the test conditions using the species conservation equations and the coverage equations. The objective function is then assembled as a weighted sum of the dif- ferences between measured and calculated exit DF1 concentrations.

3.2.1. Governing equations of the inner problem

For the inner problem which solves for the species exit concen- trations for given inlet conditions we used a simplified 1D reactor code. The basic assumptions of this model are as follows:

(1) The temperature field is specified by linearly interpolating the two experimentally temperatures. No energy equation is solved.

(2) The diffusion volume (used in the correlation determining the binary diffusion coefficient) of dodecane was assumed to be the same as that of diesel fuel as determined in our earlier work (Sampara et al., 2008).

(3) No pore diffusion effects within the washcoat layer.

(4) Transport properties of all species are calculated as though the bulk gas were N2.

In accepting the measured temperatures directly into the model in assumption 1, we are assuming mild temperature profiles within the reactor that can be well represented by a simple linear interpolant.

This is justified since the adsorption and desorption phases are es- sentially isothermal, there is no local heat release from oxidation because oxygen was absent in all experiments, and the heat losses

(6)

0 100 200 300 400 500 0

10 20 30 40 50 60 70 80 90 100 110

Time [s]

DF1 exit concentrations [ppm] inlet

exit expt exit model

0 200 400 600 800

0 5 10 15 20 25 30 35 40 45 50 55

Time [s]

DF1 exit concentrations [ppm] inlet

exit expt exit model

0 100 200 300 400 500 600 700 0

10 20 30 40 50 60 70 80 90 100

Time[s]

DF1 exitc oncentrations [ppm] inlet

exit expt exit model

0 100 200 300 400 500 600

0 5 10 15 20 25 30 35 40 45 50

Time [s]

DF1 exit concentrations [ppm] inlet

exit expt exit model

Fig. 5. Experimental results and model predictions during adsorption phase for the four test points at the optimized value of Aads= 13.5. All concentrations are ppm dodecane on a C12basis: (a) 100 ppm dodecane inlet at 116C; (b) 50 ppm dodecane inlet at 116C; (c) 90 ppm dodecane inlet at 153C; and (d) 45 ppm dodecane inlet at 153C.

from our reactor are small. For Assumption 2, dodecane, which is a long chain hydrocarbon, has nearly the same molecular weight as the molecular model for diesel fuel that we used in our previous work (170 versus 200). Since diffusion is largely dependent on the size of the molecule, we used the same diffusion volume. Other as- sumptions are fairly common in exhaust aftertreatment modeling and so we skip their discussion.

Based on these assumptions the equations for the inner problem are given as follows:

w A

dxg,DF1

dz = −km,DF1S(xg,DF1− xs,DF1)= aze(rdes− rads) (8) where

xs,DF1= cs,DF1/c and xg,DF1= cg,DF1/c (9)

and dDF1

dt = Aadscs,DF1(1−DF1)− Adese−Edes/RTDF1

= rads− rdes (10)

Note that azeis the surface site density of the adsorption–desorption reaction and was calculated by dividing Ntotby the volume of the re-

actor. The mass transfer coefficient is calculated based on the asymp- totic Sherwood number and the binary diffusitivity of individual trace HC within the mixture.

km,DF1= Sh

Dh(cDDF1,m) (11)

The binary diffusion coefficient for the trace species (hydrocarbon- DF1) is calculated based on the correlation given by Fuller et al.

(1966)and as written inPoling et al. (2002) is shown in Eq. (12) with the mixture approximated by N2.

cDDF1,m=3.85× 10−5T0.75

1/MDF1+ 1/MN2

[1/3DF1+1/3N2]2 (12) Here MDF1is 170 (g mol−1), andDF1, which is the diffusion volume of DF1, is taken as 80 fromSampara et al. (2008).

The species equations were scaled according to the procedure described in our earlier work (Sampara et al., 2008). The coupled ODE in time (coverage equation) and DAE in space (species equation) are solved using “ode15s” (MATLAB) which is called recursively to solve both the time and space problems. Some minor modifications are made to “ode15s” to improve the problem specific behavior.

(7)

0 500 1000 1500 2000 2500 0

20 40 60 80 100 120

Time [s]

DF1 exit concentration [ppm]

inlet exit expt exit model

0 500 1000 1500 2000 2500 3000 3500 0

10 20 30 40 50 60

Time [s]

DF1 exit concentration [ppm]

inlet expt exit model exit

Fig. 6. Validation using desorption and TPD data. All concentrations are ppm dode- cane on a C12basis: (a) 100 ppm dodecane inlet for adsorption phase at 116C; (b) 90 ppm dodecane inlet for adsorption phase at 153C.

3.2.2. Definition of objective function for the outer problem

The objective function for optimization defined below is based on the difference between the experimentally measured and model predicted DF1 exit concentration over the entire adsorption phase, i.e., until the outlet concentration reaches the equilibrium value. The summation over j refers to the four sets of experimental conditions listed inTable 2.





1 4

4 j=1

 teq 0 (xexpt_j

g,DF1(L)− xmodel_j

g,DF1 (L))2dt (13)

The optimization to generate the “best” value of Aadswas done us- ing MATLAB's “fmincon”, a constrained minimizer which uses local optimization methods.

3.2.3. Optimization results

The value of the objective function at the end of the optimization was 86.5. The final optimized value for Aadswas 13.5. The value of Adescalculated based on the ratio of the pre-exponentials estimated from the Arrhenius plot was 3.31× 104. The results from the opti- mization are shown inFigs. 5(a)–(d).

3.3. Validation

To validate the rate model developed in the previous sections, the model DF1 exit concentrations are compared with experimental DF1 exit concentrations over the entire desorption and TPD phases.

Since we use the adsorption phase to estimate the reaction rate for adsorption and equilibrium calculations to determine the desorption rate, using the desorption and TPD phases data gives us an indepen- dent means to check the desorption reaction rate. The two repre- sentative cases shown inFigs. 6(a) and (b) show quite reasonable representation of the both the adsorption and desorption phases of the data. As the validation comparison is not particularly hindered by the presence of experimental noise, these plots contain the fluc- tuations in inlet HC concentrations not present inFigs. 5(a)–(d).

4. Results and discussion

The intended function of a storage component such as zeolite in a DOC is to adsorb the hydrocarbons during early cold-start and then later release them after the noble metal is sufficiently warmed up to oxidize a substantial portion of the stored hydrocarbons. Ad- sorption and desorption are both occurring to some extent at all temperatures. However, since the desorption process is activated as compared to adsorption, desorption will become dominant as the reactor warms up, thereby releasing whatever hydrocarbons were stored at lower temperature when the desorption rate was small.

Our hope when introducing such storage devices is to oxidize the hydrocarbons immediately after there is net release from the ad- sorption sites, thereby minimizing early hydrocarbon emissions. The value of adsorption–desorption kinetics comes after coupling with oxidation kinetics to accurately predict hydrocarbon emissions dur- ing cold-start. In this section, we couple the adsorption/desorption kinetics developed in this work with our existing oxidation kinet- ics (Sampara et al., 2008) (also given in Appendix B) to assess the advantage of having a storage component within a DOC.

These kinetics are exercised with our “full adiabatic catalyst model” to assess the performance of a somewhat idealized but typi- cal storage+ oxidizer system. The basic governing equations used in the full model are given in Appendix A. The solution procedure for these equations is beyond the scope of this work and hence are not discussed here. For a more detailed discussion we refer the reader to our earlier work (Oh et al., 1993).

In our previous work (Sampara et al., 2008), we grouped the total hydrocarbons in the exhaust as diesel fuel, representing unburnt fuel component in the exhaust, and C3H6, representing partially oxidized hydrocarbons in the exhaust. When validating model predictions with experiments using diesel engine exhaust, reasonable agreement was obtained when the THC from the engine were divided roughly equally, on a molar basis, between diesel fuel and C3H6. For modeling purposes we had used C14.6H24.8to represent DF based onHeywood (1988). The molecular weight of this molecule is 200 g mol−1and its diffusion volume which is used in the calculation of the binary diffusion coefficient is taken as 80 based onSampara et al. (2008).

(8)

Table 3

Catalyst parameters for storage+ oxidation studies

L 0.2 m

A 1.62× 10−2m2

Cell density 400 cpsi

wsb 1.65× 10−4m

wwc 3× 10−5m

sb 1.72× 103kg m−3

wc 1.3× 103kg m−3

aj(Noble metal) 0.331 (mol-site m−3)

aj(Zeolite) 21.2 (mol-site m−3)

W 20 g s−1

The rate expressions inferred fromSampara et al. (2008)are given in Appendix B.

By comparison, the adsorption–desorption rate developed here is for n-dodecane (C12H26) and not diesel fuel. We however noted earlier in our discussion that the difference in molecular weight is small and that we had used the same diffusion volume for both the molecules. Recall that toluene was found not to adsorb on the par- ticular type of zeolite considered here. In other words, if the unburnt fuel is considered as some combination of long-chain and aromatic hydrocarbons, there should be a fraction of diesel fuel which does not adsorb on the zeolite. For simplicity we choose this fraction to be 50% (all stated percentages for the THCs in this discussion should be interpreted as C3on a molar basis). In summary, to simulate diesel engine exhaust, we divide the total hydrocarbons into three bins:

50% of the THC as C3H6(based onSampara et al., 2008), 25% of the THC as DF1, which is the adsorbable fraction of the unburnt fuel in the exhaust and, 25% of the THC as DF2, which is the non-adsorbable fraction of the unburnt fuel in the exhaust. Both DF1 and DF2 are assumed to have the same oxidation rates; they only differ in their behavior towards storage on zeolite.

We begin with the study of a representative DOC which serves to demonstrate the value of the reactor model with both storage and oxidation kinetics combined, and allows us to discern the specifics of the interaction between storage and oxidation that contribute to HC emissions during early warm-up. Of particular interest are the effects of different warm-up rates upon this representative catalyst. Finally, we show how different heat-up rates result in particular efficiencies of this catalyst to assess its performance as a HC storage device.

4.1. A representative storage+ oxidation catalyst

To make specific quantitative predictions with the reactor model, we choose a representative DOC and somewhat idealized set of in- let conditions describing exhaust warm-up. Catalyst dimensions are typical for a full scale DOC reactor in the exhaust of a 2L engine. The same catalyst specifications were used for all the parametric studies.

The dimensions of the catalyst are given inTable 3. To avoid attend- ing to the details of an actual transient driving cycle, we idealize the reactor inlet conditions during warm-up to a linearly increas- ing temperature with constant flow rate and species concentrations.

For this study, the flow rate is given inTable 3, and the typical inlet concentrations are given inTable 4. These are typical for a 2L en- gine during the cold start period of FTP operation. THC and CO emis- sions were chosen slightly on the higher side to critically evaluate the behavior of the DOC in effectively reducing these emissions. The temperatures are discussed below.

The transient nature of inlet gas during reactor warm-up is ide- alized as a steady ramp up to a constant. This allows us to capture the single greatest factor driving the early stages of reactor warm- up with a single parameter, the ramp rate. For each of the different ramp rates studied here, we started from an ambient temperature of 30C and linearly increased the inlet gas phase temperature to 240C, where it is subsequently held constant. The temperature of

Table 4

Species inlet concentrations for storage+ oxidation studies

Species Concentration

CO 1000 ppm

C3H6 300 ppm

DF1 (adsorbable fuel) 25 ppm (C14basis)

DF2 (non-adsorbable fuel) 25 mppm (C14basis)

H2 200 ppm

NO 200 ppm

NO2 100 ppm

O2 10%

H2O 8.7%

CO2 10%

0 200 400 600 800 1000 1200 1400

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

x 10−4

time (s)

DF1 concentratin [ppm]

Inlet = 25ppm

Without zeolite With zeolite

Fig. 7. DF1 emissions for a 10C min−1ramp rate.

240C was chosen because the oxidation reactions studied here are significant by this temperature (Sampara et al., 2008). After reach- ing 240C, we run the model at this temperature for 40 s to ensure that all reactions reached 100% conversion.

4.2. Effect of heat-up rate 4.2.1. 10C min−1

Initially, we consider a relatively slow ramp rate of 10C min−1 in order to easily see each stage of the adsorption–oxidation pro- cesses.Fig. 7shows the DF1 (adsorbable hydrocarbon) concentra- tion at the exit of the reactor plotted against time. For comparison we also show the exit DF1 concentration that would result in the absence of a storage device. We observe that all the DF1 is initially adsorbed on the zeolite until about 600 s, when the inlet gas tem- perature reaches 130C. After this, the catalyst starts net desorbing, leading to significant hydrocarbon emissions, until light-off of the oxidation reactions near 1100 s (214C inlet gas), followed by zero DF1 emissions at later times.

To explain the various processes which occur in this system, we propose a “rate plot” corresponding to this case (Fig. 8(a)). The rate plot contains the cumulative oxidation rate and the net release rate (difference between desorption and adsorption rates) integrated over the entire reactor (“oxidation” in rate plot= −L

0rDF1dz and “net

(9)

0 200 400 600 800 1000 1200 1400

−60

−40

−20 0 20 40 60

time (s)

rate

oxidation net release

0 0.2 0.4 0.6 0.8 1

80 100 120 140 160 180 200 220 240 260 280

z/L Gas Temperature - Tg (°C)

400 s 1200 s

1130 s

780 s 1104 s

1000 s

925 s

0 0.2 0.4 0.6 0.8 1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7

z/L θDF1

400 s

780 s

925 s 1000 s

1104 s

0 0.2 0.4 0.6 0.8 1

0 0.2 0.4 0.6 0.8 1 1.2

x 10−4

z/L

DF1 concentration [ppm]

780 s

400 s

925 s

1000 s

1104 s

1130 s 1200 s

Fig. 8. Profiles at various times when the inlet gas phase temperature is ramped up at 10C min−1: (a) Rate plot for 10C min−1 ramp rate; (b) gas phase temperature profiles; (c)DF1profiles; and (d) DF1 concentration profiles, cs,DF1.

release” in rate plot=L

0(rdes− rads) dz). These spatial integrals pro- vide a way to understand the gross behavior of the reactor at various times of its operation but do not provide any spatially resolved infor- mation. Referring these rates specifically to the DF1 concentration, the oxidation rate is shown here as negative since it depletes DF1 in the reactor. Similarly, net release produces DF1 in the reactor, so it is positive when desorption dominates and negative when adsorption dominates. To help our explanation ofFig. 8(a), we also plot axial profiles for the gas phase temperature, surface coverage of DF1 on zeolite and DF1 concentration profiles at specific times of interest as Figs. 8(b)–(d). For each time of interest, we explain the behavior of the rate plot by using these three auxiliary profile figures.

Details of the earliest phase, in which adsorption is dominant, are evident in the profiles at 400 s. Temperatures are too low for appreciable desorption or oxidation. The coverage profiles inFig. 8(c) and the DF1 concentration profiles inFig. 8(d) show that storage occurs first in the upstream portion of the reactor and that emissions of the adsorbable HCs represented by DF1 are zero.

At around 700 s, desorption, which is much more sensitive to temperature, begins to become significant as shown by the rapid increase in net release inFig. 8(a). By 780 s, there is overall more DF1 desorbing and exiting the reactor than adsorbing, since the net

release switches to positive values then. The coverage profiles in Fig. 8(c) and the DF1 concentration profiles inFig. 8(d) show that net desorption is occurring in about the first half of the reactor and net adsorption in the second half. By∼ 925 s, the net release inFig. 8(a) reaches its maximum, marking the point where the coverages have fallen so low that they can no longer sustain their current level of overall net release. By 1000 s, coverages are small throughout the reactor and still falling. HC oxidation rates have increased only slightly up to this point, and no significant exotherms are yet evident.

At 1104 s inFig. 8(a), HC oxidation has just become very signif- icant, and this time is very close to the sharp peaks in both rate curves. FromFig. 8(b), rapid oxidation has begun by this time, and all the profiles show that the reaction initiates at the downstream end of the reactor. The spatial gradients of both DF1 and CO (profile not shown) favor downstream light-off, while the small tempera- ture gradient favors light-off upstream. Detailed comparison of con- ditions in the front and rear of the reactor at this time shows that the temperature has a very minor effect, with the CO gradient be- ing the major factor. Because CO lights off slightly sooner than HCs, there is already a gradient in CO established by the time tempera- tures rise enough to initiate significant HC oxidation.The short sec- ond increase to a spike in net release near 1104 s is driven by the

(10)

0 200 400 600 800 1000 1200 1400

−60

−40

−20 0 20 40 60

time (s)

rate

oxidation net release

0 100 200 300 400 500 600 700

−200

−150

−100

−50 0 50 100 150 200

time (s)

rate

oxidation net release

0 100 200 300 400

−500

−400

−300

−200

−100 0 100 200 300 400 500

time (s)

rate

oxidation net release

0 50 100 150 200

−400

−300

−200

−100 0 100 200 300

time (s)

rate

oxidation net release

Fig. 9. Rate plots for varying ramp rates for inlet gas temperature: (a) 10C min−1; (b) 20C min−1; (c) 40C min−1; and (d) 90C min−1.

rapid oxidation of HCs. Specifically, while oxidation rapidly drives cs,DF1to zero at the exit, the adsorption rate, which is proportional to cs,DF1, is temporarily more strongly affected than the desorption rate, which is proportional toDF1× cs,DF1(z= L) = 0 at the tip of this spike. At later times, oxidation is dominant and prevents any more DF1 from exiting the reactor, whether it comes from the in- let gas or is desorbed from the remainder stored on the zeolite, as the reaction front moves upstream through the reactor (e.g., 1130 s or 1200 s).

We make the following observations based on the above discus- sion. Substantial amount of DF1 emissions are observed for slow heat-up rates of the catalyst, such as the one discussed above. The net release curve shows two peaks for such cases: first correspond- ing to significant desorption and the second corresponding to oxi- dation. Ideally we want to decrease the time between these peaks because this is the period during which stored HCs are released with- out being oxidized. Second, this case exhibits “wrong way” behavior similar to a case noted byOh and Cavendish (1982), where the local wall temperature at the downstream face rose above the inlet gas temperature. Oh et al. observed this effect during catalyst cool down when they stepped the gas phase inlet temperature down to room temperature with increased combustible species concentrations in the exhaust. This behavior is observed in our case due to lower CO inhibition at the downstream end of the reactor.

4.3. Varying heat-up rates

In the initial case studied above, the slow ramp rate of 10C min−1made it easier to distinguish the separate time intervals when adsorption, desorption, and oxidation were dominant. How- ever, this separation of these phases also make a HC storage device rather ineffective in controlling HC emissions since, fromFig. 7prior to light-off, most of the stored HCs desorbed before they could be oxidized. To improve emissions performance of the DOC, it is clear that the oxidation peak must be moved earlier relative to the net release peak. In the setting of idealized inlet conditions we consider here, this is accomplished by increasing the temperature ramp rate, which is illustrated inFigs. 9(a)–(d) where we plot the rate plots for ramp rates of 10, 20, 40 and 90C min−1. At the highest ramp rate, the oxidation spike is even pushed slightly before the desorption peak.

Fig. 10shows the DF1 concentration exiting the reactor for the various ramp rates considered here. Cases with faster ramp rates show lower DF1 emissions. There are at least two main factors to note that influence this conclusion. First, if we define a time to full light-off, for example as the time at which the exit DF1 concen- tration reaches zero, then cases with faster ramp rates will neces- sarily reach full light-off sooner and produce correspondingly less emissions. This effect is not related directly to HC storage. Second,

(11)

0 200 400 600 800 1000 1200 1400 0

0.2 0.4 0.6 0.8 1 1.2

x 10−4

time (s)

DF1 concentratin [ppm]

10°C/min 20°C/min

40°C/min

90°C/min

Inlet

Fig. 10. DF1 (adsorbable hydrocarbon) emissions for varying ramp rates.

even for times before full light-off, cases with higher ramp rates ox- idize a larger fraction of their (lower) total incoming HCs because there is less time between the start of net desorption and the start of significant oxidation. As a point of reference, Federal Test Proce- dure (FTP) cycles for diesel engines produce heat-up rates of around 45–65C min−1depending upon the size of the engine. Therefore, for the particular catalyst and conditions studied here, we would ex- pect good performance from the storage component of this DOC at practical warm-up rates. We also note that the catalyst used for this study had lower noble metal (8 g ft−3) as compared to conventional loadings commonly observed in a full-scale DOC. For standard noble metal loadings, we expect the oxidation activity to be faster and the benefits of cold start HC adsorption to be more significant. It is also important to realize that with the increase in the heat-up rate, the total amount of time available for adsorption also decreases, thus decreasing the coverage on the zeolite surface. For heat-up rates which give us low DF1 emissions, we observed that the coverage

DF1is very low (10%), thus implying lower utilization of the stor- age component.

To quantify how well increasing the heat-up rate improves the performance of this adsorber+ oxidizer system, and to understand the importance of the presence of the adsorber in a typical DOC, we define a parameteras follows:

=Cumulative DF1 emissions for an adsorber+ oxidizer system Cumulative DF1 emission for an oxidizer system

(14) Thus= 1 means that the adsorber is ineffective and has no effect on HC emissions. Lower values ofindicate increasing effectiveness.

A comparison ofvalues for various heat-up rates of the catalyst and a variety of typical total flow rates are plotted inFig. 11. For each of these cases the inlet concentrations fromTable 4were used for the predictions.

For any of these realistic flow rates, the figure expectedly shows poor performance at the lowest ramp rates and the best performance as the ramp rate approaches arbitrarily large values. Assuming  indeed is an appropriate measure of HC storage performance, the figures also show that the expected practical range of heat-up rates

0 25 50 75 100 125 150

0 10 20 30 40 50 60 70 80 90 100

Temperature ramp rate [°C/min]

 [%]

17500 h−1 21800 h−1 26200 h−1 30600 h−1

10°C/min

Typical heat up rates in diesel engines resulting from FTP

Fig. 11.comparison for varying heat-up rates and space velocities.

0 50 100 150

0 10 20 30 40 50 60 70 80 90 100

Temperature ramp rate [°C/min]

 [%]

500 ppm CO 1000 ppm CO 2000 ppm CO 3000 ppm CO

10°C/min

Fig. 12.comparison for varying heat-up rates and CO concentrations.

of 45–65C min−1 is generally within the favorable intermediate performance interval. That is, the ramp rate is sufficiently high to give improved performance, but not too large so that the performance is insensitive to changes in the rate. As a point of reference, the cases represented byFigs. 9(a)–(d) are also represented within the curve inFig. 11for the case where the flow rate is 20 g s−1(17500 h−1). As may also have been somewhat evident fromFigs. 9 and 10, increasing the heat-up rate beyond 40C min−1 at this flow rate would not give any substantial improvement in the HC storage performance as measured by.

Fig. 12 shows how the  varies with different heat-up rates and CO concentrations. Excluding very high concentrations of

(12)

CO∼ 3000 ppm, which are possible during PCI (pre-mixed com- pression ignition) operation in diesel engines, all other CO con- centrations show reasonably similar behavior in terms of their effect on . Similar studies on NO concentration showed smaller effect of NO ondue to the fact that the range and magnitude of NO concentration is much smaller than that of CO during typical operation.

5. Summary

This work has two purposes. The principal purpose is to de- velop simple reaction rate expressions for adsorption and desorp- tion of hydrocarbons on zeolites. Secondly, we want to exercise these adsorption kinetics with our previous DOC oxidation kinetics to make some general, useful observations on the combined adsorber+ oxidizer system. The conclusions for the first part of the work are as follows:

(1) The reaction kinetics for hydrocarbon adsorption–desorption on zeolite can be adequately described by first order adsorption and desorption on a single site, including a Langmuir isotherm to represent equilibrium.

(2) We represent the diesel exhaust HCs by a mixture of propy- lene (partially oxidized HCs from the engine), n-dodecane (aliphatic unburned fuel), and toluene (aromatic unburned fuel).

Only n-dodecane adsorbed significantly on the zeolite studied here.

(3) A minimum of four experiments are sufficient to generate the necessary kinetic constants. Elementary analysis and direct mea- surements during adsorption yield the total zeolite storage ca- pacity (Ntot), the activation energy for desorption (Edes) and the ratio of the pre-exponentials for adsorption and desorption. One of these pre-exponentials is then evaluated by fitting model pre- dictions to the experimental adsorption data using a simplified 1D reactor code integrated within optimization routines in Mat- lab.

(4) The resulting adsorption–desorption rate was validated with ad- ditional experimental data obtained during the later phases of the four tests.

The observations from exercising the adsorber+ oxidizer model are as follows:

(1) For the idealized warm-up conditions studied here, the HCs in the DOC light off from the downstream section of the catalyst.

This is primarily because the CO starts reacting earlier, creating a gradient of CO, which then produces decreased inhibition of the HC oxidation at the rear.

(2) The rate of heating of the inlet gas to the DOC plays an important role in determining overall system performance.

(3) A “rate plot” is developed from the model predictions to sep- arately reveal the rates of HC adsorption, desorption and ox- idation while they are interacting during a typical warm-up.

This plot clarifies the sequence of individual events that in- fluence the performance of the zeolite in helping reduce HC emissions.

(4) The zeolite studied here is reasonably effective in reducing exhaust hydrocarbon emissions. Specifically, for the exhaust conditions considered here (including realistic flow rates and inlet temperatures which increase at realistic rates of 45–65C min−1), storage on the zeolite reduced HC emissions during warm-up by at least a factor of two compared to cases with oxidation alone.

Notation

aj active site density for noble metal or zeolite depend- ing on the reaction j, mol-site m−3

aze storage site density, mol-site m−3

A face area, m2

Aads rate constant for adsorption, m3mol-site−1s−1 Ades rate constant for desorption, mol mol-site−1s−1 Ai pre-exponential for rate constant, m6mol−1

mol-site−1s−1

Aai pre-exponential for adsorption constant, mol−1m−3 c total molar concentration of gas, mol m−3

cpg molar specific heat of gas, J mol−1K−1

cs,i molar concentration of trace species at catalyst sur- face, mol m−3

Cps,sb specific heat of substrate, J kg−1K−1 Cps,wc specific heat of washcoat, J kg−1K−1 Dh hydraulic diameter of channel, m

Di,m binary diffusion coefficient of species i in the mix- ture, m2s−1

DF1 hydrocarbon component which can be both ad- sorbed and oxidized

DF2 hydrocarbon component which can be oxidized but not adsorbed

Ei activation energy for rate constant, J mol−1 Eai activation energy for adsorption constant, J mol−1 Edes activation energy for desorption constant, J mol−1 fsb solid fraction of the substrate

fwc solid fraction of the washcoat

h interphase heat transfer coefficient, J m−2s−1K−1

Hj enthalpy of reaction (< 0 for exothermic), J mol−1 km,i mass transfer coefficient for species i, mol m−2s−1 K ratio of rate constants for adsorption–desorption,

mol m3

L length of the reactor, m

Mi molecular weight of species i, kg mol−1

nDF1 number of moles of DF1 adsorbed on the zeolite surface, mol

˙nDF1 mole flow rate of DF1, mol s−1

neq,DF1 number of moles of DF1 adsorbed on the zeolite surface at equilibrium, mol

Ntot total number of moles of zeolite available for storage, mol

p system pressure, Pa

rj rate of reaction j, mol mol-site−1s−1

rads adsorption rate onto zeolite, mol mol-site−1s−1 rdes desorption rate from zeolite, mol mol-site−1s−1 R universal molar gas constant, J mol−1K−1 si,j stoichiometric coefficient of species i in reaction j Sh asymptotic Sherwood number

S surface area per reactor volume, m−1

T temperature, K

Tg temperature of bulk gas phase, K Ts temperature of solid phase, K w molar flow rate, mol s−1 W inlet mass flow rate, kg s−1 wsb substrate wall thickness, m wwc washcoat wall thickness, m

xg,i mole fraction of species i in bulk gas phase xs,i mole fraction of species i in gas at catalyst surface

z axial position, m

Riferimenti

Documenti correlati

This work was carried out in order to assess the geothermal potential of the water and sediment, and this will encourage the use of geothermal heat pump systems in densely

The investigated igneous complexes are those closer to the crustal anatectic end-member compositions and include intrusive bodies (Monte Capanne and Porto Azzurro monzogranites

Infine, è proprio a partire dal 1966 che i Beatles si misero alla ricerca dell’Io, della consapevolezza spirituale: gli strumenti furono trip fantastici con l’LSD, corsi di

Interesting to investigate in bacterial species that are implicated in the metabolism of dietary polyphenol (Dueñas et al., 2015) or in diary milk.. There is evidence for

Pathway analysis focused on these genes was carried on with KEGG tool (http://www.genome.jp/kegg/pathway.html) and is also reported in the Supplementary Tables 1 and

The focus of the research, however, did not lie on the description of Estuary English and its ongoing development; it rather tried to concentrate on the use of this supposed

Comparison with shell model and relativistic mean field calculations demonstrate that the observed effects arise from deformed prolate ground state configurations associated with

An integral neutronics experiment of the measurement of 239 Pu production rates in a water cooled natural uranium blanket mock-up of a fusion –fission hybrid reactor was