throbber
Article
`
`pubs.acs.org/IECR
`
`Integrated Selective Catalytic Reduction−Diesel Particulate Filter
`Aftertreatment: Insights into Pressure Drop, NOx Conversion, and
`Passive Soot Oxidation Behavior
`Kenneth G. Rappé*
`
`Pacific Northwest National Laboratory, P.O. Box 999, Richland, Washington 99354, United States
`
`ABSTRACT: Integrating urea-selective catalytic reduction (SCR) and diesel particulate filter (DPF) technologies into a single
`device has the potential to reduce the complexity of current diesel aftertreatment strategies. Fundamental studies were performed
`to shed light on the pressure drop and reaction behavior of integrated SCR and DPF systems. Details of SCR washcoat amount
`and location were investigated for effect on pressure drop during soot filtration. The SCR catalyst primarily impacted depth
`filtration of soot, promoted by increased catalyst located within the upstream portion of the porous filter wall. This effect is
`believed to be related to the nature of the porous filter substrate and pore network and changing of the rate at which pores plug
`in the presence of catalyst. SCR catalyst on the wall of the inlet filter channel also had an effect on the pressure rise during cake
`filtration of soot. NOx reduction efficiency measurements were performed to determine the nature and magnitude of the effect of
`soot on SCR performance. The effect of soot on the SCR performance is primarily attributed to the contribution of passive soot
`oxidation, and the propensity for soot oxidation to shift the NO2/NOx fraction relative to 0.5. SCR performance at NO2/NOx <
`0.5 is adversely affected by the presence of soot oxidation by increasing the SCR dependency on standard (NO only) SCR
`reactions; conversely, at NO2/NOx > 0.5, the SCR performance is positively impacted by a decreased dependency on NO2-only
`SCR reactions. Temperature-programmed oxidation studies were performed to evaluate the impact of SCR on passive soot
`oxidation. SCR adversely impacts soot oxidation performance via NO2 diffusive effects, decreasing NO2 concentration in the inlet
`channel. This impact can be minimized or recovered at higher NO2 concentration and NO2/NOx fractions >0.5.
`
`1. INTRODUCTION
`The diesel engine is currently the primary engine employed in
`the long- and short-haul commercial trucking industry because
`of its attractive wear characteristics, increased engine durability,
`and ability to deliver power efficiently under high load
`conditions. Manufacturers are planning much more widespread
`usage than historically employed, motivated by reduced fossil
`fuel consumption, reduced CO2 emissions, and the ability to
`operate on biologically derived fuels. However,
`for such a
`paradigm shift to take place, there are hurdles to overcome.
`Diesel close-coupled injection, compression, and ignition result
`in a heterogeneous two-phase burn that takes place largely at
`fuel droplet edges where the fuel and the air are mixing. The
`result of this is a near-stoichiometric burn that is very hot and
`comparatively long in duration, leading to elevated levels of
`NOx and PM formation. Engine controls and advanced fuel
`injection and combustion strategies have made significant
`strides in lowering diesel engine emission levels. However,
`persistently high particulate matter (PM) and nitrogen oxides
`(NOx) emissions dictate that exhaust aftertreatment will be a
`necessity.
`Exhaust aftertreatment reduces pollutants from the engine
`exhaust by treating the exhaust gas in a thermally and
`chemically controlled environment. Widespread introduction
`of advanced diesel exhaust aftertreatment
`is necessary for
`simultaneous compliance with NOx and particulate matter
`emission standards. Aftertreatment may lessen the trade-offs
`involved in controlling both NOx and PM emissions while
`minimizing losses in fuel efficiency.
`
`Several existing emission control technologies have proven
`effective at controlling emissions individually. A diesel oxidation
`catalyst (DOC) effectively oxidizes NO, CO, and hydrocarbon
`(HC) emissions, and the diesel particulate filter effectively
`removes particulate matter from the exhaust stream. The wall-
`flow diesel particulate filter
`is a honeycomb monolithic
`structure made from porous ceramic that has channels
`alternately plugged at the ends so that exhaust gas is forced
`through the porous channel wall. As engine exhaust is passed
`through a DPF, particulates are retained on the upstream
`portion of the filter wall and accumulate over time. Soot
`filtration first occurs by depth filtration, which is the collection
`of soot within the porous microstructure of the filter wall. This
`is typically accompanied by a sharp increase in the pressure
`drop across the filter. However, filtration will usually quickly
`transition to cake filtration, which is when soot is no longer
`accumulating within the wall microstructure, but rather, on the
`upstream surface of the filter wall. In this instance, the soot cake
`present on the upstream filter wall serves as the filter, and this is
`typically accompanied by a smaller increase in pressure drop
`across the filter. Together, these lead to an increase in pressure
`drop across DPF and, thus, an increase in the back pressure on
`the engine that adversely affects the engine operation and fuel
`consumption.
`
`Received:
`July 16, 2014
`Revised: October 9, 2014
`Accepted: October 20, 2014
`Published: October 20, 2014
`
`© 2014 American Chemical Society
`
`17547
`
`dx.doi.org/10.1021/ie502832f | Ind. Eng. Chem. Res. 2014, 53, 17547−17557
`
`BASF-2015.001
`
`

`
`Industrial & Engineering Chemistry Research
`
`Article
`
`HNCO
`
`(1)
`
`(2)
`
`Selective catalytic reduction (SCR) has been developed as an
`effective technology at reducing NOx emissions under lean
`burn conditions. SCR catalysts selectively reduce NOx to N2
`and H2O through the use of NH3 as the reductant. NH3 is
`supplied via a urea injection system and subsequent thermal
`decomposition to isocyanic acid and 1 equiv of NH3, as shown
`in eq 1. A second equivalent of NH3 is produced from the
`surface-catalyzed decomposition of isocyanic acid, as shown in
`eq 2.
`+
`→
`−
`−
`NH
`NH CO NH
`3
`2
`2
`+
`→
`+
`CO
`NH
`HNCO H O
`2
`3
`2
`The SCR process begins with reversible ammonia chem-
`isorption on catalytic Brønsted acid sites, where it becomes
`−SCR reactions
`available for SCR reaction. The main NH3
`promoted by Fe- and Cu-exchanged zeolites are shown in eqs
`3−5 and consist of the standard SCR reaction (NO only), the
`fast SCR reaction (equimolar NO and NO2), and the NO2-only
`SCR reaction, respectively. Zeolite-based catalysts (such as
`vanadium systems) exhibit maximum NOx reduction activity at
`NO2/NOx = 0.5 as a result of the prevalence of the fast SCR
`reaction. The standard SCR reaction participates at NO2/NOx
`< 0.5 because of the relative abundance of NO; similarly, the
`NO2-only SCR reaction participates at NO2/NOx > 0.5 because
`of the relative abundance of NO2.
`+
`+ →
`+
`4NH
`4NO O
`4N
`2
`3
`2
`+
`+
`→
`+
`NO NO
`2NH
`2N
`3
`2
`2
`+
`→
`+
`3NO
`3.5N
`6H O
`4NH
`3
`2
`2
`2
`Current focus in the U.S. has been on development of Fe-
`and Cu-based zeolite formulations as effective SCR catalysts for
`meeting NOx emission regulations.1,2 Early zeolite catalysts
`favored Fe-exchanged versions (Fe−ZSM-5) with medium-
`sized pores (∼5.5 Å) for heavy-duty diesel (HDD) applications
`for their improved activity above 350 °C and superior thermal
`stability as compared with Cu-exchanged versions, which had
`previously been confined to light-duty diesel (LDD)
`applications. Fe zeolites exhibit less sensitivity to NO2/NOx >
`0.5 because of the efficiency of the NO2-only SCR reaction on
`these catalysts;3,4 however,
`recent advancements
`in Cu-
`exchanged small pore zeolites have made them more much
`attractive for both HDD and LDD applications.5 Specifically,
`the small pore Cu-exchanged chabazite-type zeolites Cu-SSZ-
`13 and Cu-SAPO-34 have recently exhibited good activity over
`a broad temperature range, good thermal stability, and good
`resistance to hydrocarbon deactivation.6−10 These Cu zeolites
`are shown to exhibit less sensitivity to NO2/NOx < 0.511 and
`have also demonstrated improved SCR selectivity to N2 versus
`undesirable products, including N2O.7
`Integration of catalytic SCR and DPF technologies is a
`relatively new area in exhaust aftertreatment systems. Integrated
`reduction and soot
`SCR/DPF technology combines NOx
`filtration in a single 2-way device. Its development is motivated
`by emission compliance in a manner that reduces aftertreat-
`ment
`system volume and cost and increases packaging
`flexibility. As such,
`it is currently being pursued for both
`LDD and HDD diesel applications.
`Figure 1 shows a comparison of wall-flow and flow-through
`substrates as well as a schematic of an integrated wall-flow 2-
`−SCR
`way device.12 The technology consists of coating an NH3
`catalyst within the porous wall of a wall-flow filter (i.e., DPF).
`
`6H O
`2
`3H O
`2
`
`(3)
`
`(4)
`
`(5)
`
`Figure 1. Schematic comparison of a wall-flow DPF, flow-through
`SCR and integrated wall-flow 2-way DPF/SCR. Reproduced with
`permission from ref 12. Copyright 2011 Sage Publications.
`
`An integrated SCR/DPF device combines soot filtration with
`an SCR catalyst, achieving simultaneous soot filtration and NOx
`reduction within a single device.13
`For current conventional aftertreatment strategies, placement
`of SCR and DPF functionalities in the exhaust stream is
`governed by SCR activity and soot management. Regeneration
`of filtered soot requires both heat and oxidant; major oxidants
`in a diesel application are O2 and NO2, with NO2 oxidizing soot
`at lower temperature. When exhaust gas temperatures do not
`reach levels necessary for soot oxidation, active regeneration is
`necessary; this is facilitated by the injection of fuel in front of
`the DOC and subsequent oxidation, heating the downstream
`DPF via the ensuing exotherm to temperature necessary for
`soot oxidation. For HDD applications, exhaust temperature is
`comparatively high, and oxidation of soot with NO2 (i.e.,
`passive soot oxidation) is highly desirable. This motivates
`placement of the SCR catalyst downstream of the DPF to avoid
`depletion of NO2 prior to the DPF; however, SCR catalyst
`performance is adversely affected by the upstream DPF because
`of thermal inertia; Koltsakis and co-workers predicted as high as
`10% reduced NOx reduction efficiency due to an upstream
`DPF.14 For LDD applications,
`in which the exhaust gas
`temperature is comparatively low, resulting in low passive soot
`oxidation potential, the SCR catalyst is placed upstream of the
`DPF for improved activity and cold-start performance.
`Assuming adequate soot filtration efficiency is achievable
`with the wall-flow substrate employed, the requirements for
`successful SCR/DPF integration are maximizing NOx reduction
`efficiency while demonstrating acceptable pressure drop across
`the filter. Maximizing NOx reduction efficiency will be a
`function of total SCR catalyst washcoat volume and optimum
`catalyst coating techniques for maximum SCR efficiency.
`Demonstrating acceptable pressure drop across the filter will
`be a function of filter characteristics (e.g., total porosity),
`catalyst washcoat, and soot management. Thus, high porosity
`filter materials development
`is a technology facilitator for
`integrated SCR/DPF systems for integration of larger catalyst
`volumes while retaining an acceptable pressure drop.15−18 The
`washcoating technique is also a technology facilitator for similar
`reasons and, thus,
`is an area of continued optimization by
`catalyst manufacturers.
`focus for
`This leaves soot management as an area of
`successful deployment of integrated SCR/DPF technology. For
`LDD applications that must rely predominantly on active soot
`
`17548
`
`dx.doi.org/10.1021/ie502832f | Ind. Eng. Chem. Res. 2014, 53, 17547−17557
`
`BASF-2015.002
`
`

`
`Industrial & Engineering Chemistry Research
`
`regeneration, soot loading and time between regenerations
`must be closely controlled because of the exotherm from soot
`burn and possible adverse effects on SCR catalyst deactivation.
`In contrast,
`for HDD applications that rely on significant
`passive soot oxidation, SCR/DPF integration is more complex.
`Passive soot oxidation is NO2-dependent, and current emission
`control strategies employ oxidation functionality in DPFs for
`enhanced NO2 production and improved passive soot oxidation
`capacity. Integrated SCR/DPF systems will likely not possess
`oxidation functionality because of
`its adverse impact on
`reductant usage (i.e., NH3 oxidation).
`In addition, SCR
`processes will consume NO2 competitively with soot oxidation.
`Thus, the incorporation of SCR/DPF for HDD applications
`must be developed and controlled in such a manner that
`optimizes NO2 availability for passive soot oxidation to avoid
`(or minimize) the fuel penalty associated with the increased
`active regeneration frequency.
`The motivation for integration of DPF and SCR function-
`alities into a single device resides in both performance and
`reduced aftertreatment volume. The integration of DPF and
`SCR functionalities into a single device has the potential to
`reduce the number of aftertreatment “bricks” required, leading
`to reduced volume, reduced cost, and increased simplicity. With
`regard to performance, development of advanced wall-flow
`DPF substrates have allowed engine manufacturers to focus
`their attention on reducing the amount of NOx emitted from
`tailpipes. Integration of DPF and SCR provides a pathway for
`NOx emission reduction with greater flexibility. By integrating
`SCR and DPF functionalities, the adverse effect of upstream
`DPF thermal
`inertia on SCR catalyst performance can be
`largely mitigated,
`thereby improving the cold-start NOx
`performance.14 Increased SCR catalyst washcoat volumes are
`feasible through different configurations, which would provide
`increased NOx conversion capability that could facilitate fuel
`savings by allowing engines to be calibrated to higher engine-
`out NOx levels; however, there is uncertainty surrounding the
`effect of SCR catalyst on passive soot oxidation in integrated
`SCR/DPF solutions. The adverse impact of upstream SCR on
`DPF passive soot oxidation is well documented; the extent to
`which this can be minimized is a topic of significant current
`interest which this paper discusses.
`A number of early works have been performed to evaluate
`the feasibility of simultaneous NOx and PM aftertreatment with
`integrated SCR/DPF systems.19−23 These early works success-
`fully demonstrated the feasibility of combined NOx and PM
`reduction and identified key barriers to its effective deployment.
`Some of these key technology hurdles include
`(cid:129) Increased back pressure resulting from high SCR
`washcoat volumes
`(cid:129) Necessity of
`improved washcoating technology for
`maximizing SCR catalyst volume and gas contact
`(cid:129) Soot oxidation challenges
`Washcoat optimization is a topic to which many SCR/DPF
`investigators have alluded;24 however, little has been mentioned
`of the details of the SCR catalyst integration within the wall-
`flow filter. This work attempts to provide insight into the
`impact of SCR catalyst location on soot-loading characteristics
`of the filter. The intent is to improve the level of understanding
`for maximizing SCR catalyst washcoat volumes while
`minimizing the back pressure that results from soot collection.
`The impact of soot on SCR reactor performance has been
`reported by many with conflicting results in some cases. This
`
`Article
`
`work will report on the impact of soot on NOx conversion
`performance to develop a more fundamental understanding of
`the nature of
`the interaction.
`In addition, as previously
`mentioned, the adverse effect of SCR reaction on passive
`soot oxidation (with NO2) is a recognized barrier to successful
`SCR/DPF integration for HDD. However, detailed under-
`standing of the nature of this interaction and the primary
`reaction mechanisms that govern system performance are
`lacking. This work will report on the impact of SCR reaction on
`passive soot oxidation in an attempt
`to develop a more
`fundamental understanding of the primary reactive drivers that
`govern soot oxidation feasibility in the integrated system.
`
`2. EXPERIMENTAL METHODS
`High porosity cordierite wall-flow filters (i.e., DPFs; 25 mm o.d.
`× 40 mm long) were coated with SCR catalyst for the purpose
`of this investigation. The SCR catalyst employed was a Cu-
`chabazite catalyst loaded by an unnamed industrial partner to
`varying catalyst density. Upon receipt, samples were degreened
`under lean conditions at 10% O2, 5% H2O, balance N2.
`Soot loading and reactive interrogation of the SCR/DPF
`samples occurred separately in iterative fashion. The samples
`were loaded with soot employing the exhaust from a 2003 VW
`Jetta TDI following a DOC placed upstream of the filters.
`During soot loading, the exhaust was continually measured with
`a smoke meter to determine post-DOC particulate density. A
`slip stream of the exhaust was pulled through the SCR/DPF
`sample with the use of a vacuum pump, with the slip stream
`subsequently conditioned and measured for flow rate. Flow rate
`was electronically controlled to target 55 000 GHSV during
`loading; measured particulate density and flow rate were
`combined to target a desired total soot loading, typically 4 g/L.
`Pressure drop was measured at ports just upstream and
`downstream of
`the SCR/DPF sample holding apparatus.
`Subsequent to soot loading, the samples were removed and
`weighed at elevated temperature for redundant measurement of
`soot mass uptake. Estimated soot loading from smoke meter
`and mass measurements were in good agreement with typically
`less than 10% variance.
`Reactive measurements were performed on a dedicated test
`bench designed to accommodate the soot-loaded filter and
`housing. The simulated exhaust composition, consisting of 9%
`O2, 9% CO2, 8% H2O, 500−1500 ppm of NOx (NO and NO2),
`and balance N2, was blended together to simulate post-DOC
`heavy-duty lean burn diesel exhaust. For measurements
`including SCR reaction, NH3 was included at 500−1500 ppm.
`Dry gases were supplied by electronically controlled MKS
`mass flow controllers. Air, N2, and CO2 were combined as dry
`gases and preheated to the desired temperature with an
`electronically controlled tube furnace. Simultaneous
`to
`preheating, the feed stream was humidified to the desired
`level with an electronically controlled water vaporizer employ-
`ing a N2 sweep flow. Following preheating, NO, NO2, and NH3
`were supplied separately as dry gas feeds to the exhaust flow in
`desired quantities. The full simulated exhaust was then fed to
`the reactor assembly consisting of a quartz tube containing the
`SCR/DPF sample held in place by high temperature alumina
`paper. An electronically controlled tube furnace controlled the
`reactor temperature to the desired level. Temperature measure-
`ments were made employing Watlow type K thermocouples
`placed just upstream and downstream of the DPF/SCR sample.
`Pressure drop measurement was made via ports located just
`upstream and downstream of the sample holder. Gas analysis
`
`17549
`
`dx.doi.org/10.1021/ie502832f | Ind. Eng. Chem. Res. 2014, 53, 17547−17557
`
`BASF-2015.003
`
`

`
`Article
`
`alternating ends on the filter, channels in opposing corners of
`each subfigure in Figure 2 are common to a respective end of
`the filter; direction of exhaust flow through the samples is
`reversed by simply rotating the filter axially with respect to flow.
`With the SCR/DPF sample loaded to 60 g/L catalyst density
`(B), SCR catalyst is deposited solely within the filter wall
`microstructure and not on the channel walls. Catalyst is located
`heavily to one side of the filter wall and penetrating, on average,
`40−60% across the width of the wall. In the SCR/DPF sample
`loaded to 90 g/L catalyst density (C), SCR catalyst is similarly
`deposited predominantly within the filter microstructure and
`also heavily to one side of the filter wall. Periodic occurrences
`of the catalyst penetrating the full width of the porous filter wall
`are evident, and with little or no catalyst present on either set of
`the channel walls. Comparatively in the SCR/DPF sample
`loaded to 150 g/L catalyst density (D), there are notable
`differences to the previous two samples. First, catalyst of
`varying amounts is present on one set of channel walls, outside
`of
`the wall microstructure. And second,
`the SCR catalyst
`consistently penetrates across the full width of the filter wall;
`thus, catalyst is in intimate contact with both filter walls. It is
`worth noting that the SEM imaging was conducted at four
`locations within the coated samples, with the images shown
`above selected randomly from the samples analyzed. The
`results discussed in the text with regard to catalyst location
`were consistently reflected in all samples analyzed.
`Figure 3 shows the corresponding results of Hg porosimetry
`analysis of the DPF and SCR/DPF samples, presented as
`
`Industrial & Engineering Chemistry Research
`
`was accomplished with a Nicolet 6700 FTIR employing a metal
`gas cell heated to 190 °C. Sample gas transfer lines to the FTIR
`were heated to 200 °C to avoid NH4NO3 formation. FTIR gas
`analysis was conducted at 100 Torr through the use of a
`vacuum pump located downstream from the FTIR gas cell and
`an electronically controlled MKS pressure controller. Redun-
`dant NOx measurement was made via a heated California
`Analytics chemiluminescent NOx meter. Control of the entire
`assembly, including mass flow controllers, all heated zones, and
`data acquisition, was facilitated by a custom designed Visual
`Basic control system.
`Bench-scale reactive measurements were performed in
`pseudosteady state fashion by employing slow thermal ramping,
`∼2 °C/min, from ∼200 to 550 °C. Reactive measurements
`included SCR performance for NOx conversion efficiency and
`temperature-programmed oxidation (TPO) for soot oxidation
`characterization.
`Scanning electron microscopy (SEM) was performed with a
`JEOL 5900 scanning electron microscope. The sample was cut
`perpendicular to the cordierite channels with a wet diamond
`wafer saw, potted in LR white resin, and spun at 2000 rpm for
`10 min in a standard centrifuge and subsequently evacuated to
`remove gases from the filter and resin. Following polymer-
`ization and hardening of the resin, the sample was cross-
`sectioned perpendicular
`to the channel axis, ground and
`polished with a series of diamond grits (25, 9, 3, and 1 μm),
`and final polished with 0.05 μm γ-alumina.
`Pore characteristics (porosity, pore size distribution) were
`measured via Hg intrusion employing a Micromeritics
`AutoPore IV 9500 series mercury porosimeter.
`
`3. RESULTS AND DISCUSSION
`SEM imaging and porosity characteristics were compared with
`soot collection characteristics in an attempt to determine the
`combined effect of catalyst loading and location on filtration
`behavior. Figure 2 shows the SEM imaging of the uncoated
`DPF substrate (A) and SCR/DPF samples coated with ∼60
`(B), ∼90 (C), and ∼150 g/L (D) SCR catalyst loading density.
`In each of the subfigures, using A as a reference, black is
`background or void space, (off-)white is cordierite substrate,
`and gray is SCR catalyst. Since channels are plugged at
`
`Figure 3. Hg porosimetry analysis of DPF (no catalyst) and SCR/DPF
`samples loaded to 60, 90, and 150 g/L SCR catalyst.
`
`incremental pore volume (mL/g) versus pore diameter (mm).
`The uncoated DPF measured 63% porosity. The addition of 60
`g/L of SCR catalyst decreased the total porosity to 56%, with
`similar peak pore diameters for both the DPF and 60 g/L SCR/
`DPF sample. This, coupled with SEM imaging, provides good
`indication that the 60 g/L of SCR catalyst was depositing solely
`within the filter wall microstructure. Similar peak pore
`diameters support the observation that a large portion of the
`porous filter wall remained largely void of catalyst.
`With an additional 30 g/L catalyst (i.e., 90 g/L total SCR
`catalyst), the total porosity decreased to 52% with a slight shift
`in the peak pore diameter to a smaller maximum. The shift in
`the peak pore diameter suggests that the less porous filter wall
`is void of catalyst versus the 60 g/L sample. These results
`
`17550
`
`dx.doi.org/10.1021/ie502832f | Ind. Eng. Chem. Res. 2014, 53, 17547−17557
`
`Figure 2. SEM imaging of clean DPF (A) and SCR/DPF samples
`coated with 60 (B), 90 (C), and 150 (D) g/L SCR catalyst loading.
`
`BASF-2015.004
`
`

`
`Industrial & Engineering Chemistry Research
`
`Article
`
`Figure 4. Soot-loading characteristics of 90 (left) and 150 g/L (right) SCR/DPF samples configured such that catalyst was present predominantly
`on the upstream portion of the filter microstructure and on the inlet channel wall (for the 150 g/L sample).
`
`Figure 5. Soot-loading characteristics of 90 g/L (left) and 150 g/L (right) SCR/DPF samples configured such that catalyst was present
`predominantly on the downstream portion of the filter microstructure and on the outlet channel wall (150 g/L sample).
`
`coupled with SEM indicate that 90 g/L of SCR catalyst
`continued to deposit predominantly within the filter wall
`microstructure.
`With an additional 60 g/L catalyst (i.e., 150 g/L total SCR
`catalyst), the total porosity did not decrease significantly but,
`rather, remained fairly constant at 51%, with no change in the
`peak pore diameter (versus 90 g/L). The SCR catalyst filled a
`small number of very large pores >20 mm, leading to a slight
`asymmetric narrowing of the pore distribution. One difference
`from the previous two samples is that a significant amount of
`catalyst did not penetrate the wall microstructure but, rather,
`remained coated on the filter channel.
`In summary, total porosity and peak pore characteristics,
`coupled with SEM imaging, support the notion that up to 90 g/
`L SCR catalyst was loaded into the filter wall microstructure,
`penetrating the width of the porous wall to varying degrees but
`loaded predominantly heavy to one side of the filter wall.
`Comparatively, the results suggest that >90 g/L SCR catalyst
`was largely deposited as a coating on the filter channel, with the
`catalyst deposited typically across the full width of the porous
`filter wall.
`3.1. Soot-Loading Behavior. The 90 and 150 g/L SCR/
`DPF samples were loaded with soot to determine the effect of
`catalyst location on the soot loading characteristics of the
`samples. Figures 4 and 5 show on-engine backpressure (i.e.,
`pressure drop) resulting from the 90 g/L sample (left) and the
`150 g/L sample (right) as they are loaded with soot. In Figure
`4, samples are configured with the catalyst on the upstream
`
`portion of the filter (in close proximity to the collected soot).
`Figure 5 shows analogous backpressure measurements with the
`samples configured in the opposite direction (i.e., catalyst on
`the downstream portion of the filter). The data presented here
`were very reproducible, with subsequent pressure drop (versus
`collected soot) traces in close agreement.
`The pressure drop of the clean SCR/DPF samples was
`comparatively low: 0.35 and 0.44 kPa for 90 and 150 g/L SCR
`catalyst loading, respectively, under the conditions tested. Thus,
`collected soot affected filter permeability much more
`significantly than the catalyst washcoat; however, both the
`catalyst concentration and location impacted the magnitude of
`pressure drop resulting from the collection of soot. The most
`prominent effect occurred during depth filtration of soot; this is
`not surprising, given the knowledge that the catalyst is largely
`located within the wall microstructure. With the catalyst on the
`upstream portion of
`the filter (Figure 4),
`the system
`demonstrated 11.5 and 20.5 kPa back pressure for the 90 and
`150 g/L SCR catalyst densities,
`respectively, under
`the
`conditions tested when it transitioned from depth to cake
`filtration mode. Comparatively, with the catalyst on the
`downstream portion of
`the filter (Figure 5),
`the system
`demonstrated 3.2 and 7.4 kPa back pressure for the 90 and 150
`g/L SCR catalyst densities, respectively, under the conditions
`tested when it transitioned from depth to cake filtration mode.
`The pressure drop during depth filtration is affected by the
`amount of SCR catalyst located on the upstream portion of the
`filter wall microstructure. The magnitude of pressure drop does
`
`17551
`
`dx.doi.org/10.1021/ie502832f | Ind. Eng. Chem. Res. 2014, 53, 17547−17557
`
`BASF-2015.005
`
`

`
`Industrial & Engineering Chemistry Research
`
`Article
`
`Figure 6. SCR/DPF NOx reduction efficiency at 35 000 GHSV, NH3/NOx = 1, NO2 = 250 ppm, and varying NO2/NOx fraction; NO2/NOx = 0.5
`and less shown on the left, NO2/NOx = 0.5 and greater shown on the right.
`
`not correlate solely with the mass of soot collected, but appears
`to also be a function of SCR catalyst amount and location. Two
`feasible contributing factors to this would be (1) relative
`density, and thus permeability, of soot deposits, or (2) relative
`impact on the flow field in the porous network. Evaluating the
`first in more detail,
`if back pressure is governed (solely or
`predominantly) by the density of soot deposits collected during
`depth filtration, then one would expect a closer quantitative
`relationship between the total mass of soot collected, the
`amount of SCR catalyst present on the upstream portion of the
`filter wall, and the magnitude of resulting back pressure.
`Although there appears to be a very general trend, it is not
`consistent; neither is it quantitative. For example,
`in both
`figures, the 90 and 150 g/L catalyst samples filtered similar
`masses of soot at the transition from depth to cake filtration but
`exhibited distinctively different back pressure as a result.
`These results suggest that SCR catalyst on the upstream
`portion of the filter wall directly affects the impact of soot on
`the porous network flow field. This brings to light the concept
`of pore throats. Pore throats are constriction points in the
`three-dimensional channels of wall-flow monolith pores, and
`they play a significant role in permeability.25 They are physically
`manifested a number of ways in the wall microstructure,
`ranging from simple constriction points along a pore channel to
`a directional change in the flow field passing through the
`porous network. Soot collected in depth filtration primarily
`impedes exhaust flow at pore throats26 and results in increased
`pressure drop during depth filtration; the extent is governed by
`the nature and density of those pore throats. Because the
`difference in clean pressure drop between the samples (0.35
`versus 0.44 kPa) was small compared with the effect of soot, the
`density of pore throats is not expected to be significantly
`different. Thus, it appears that the nature of a large fraction of
`pore throats (and their surrounding volumes) is affected by the
`presence of SCR catalyst. The result is an increased extent of
`plugging as a result depth filtration of soot in regions of
`elevated SCR catalyst concentration.
`Comparing the 150 g/L catalyst sample with the 90 g/L
`sample, the samples collected similar masses of soot during
`depth filtration in both configurations; however, the difference
`in back pressure between the samples was significantly different,
`depending on flow configuration. With the catalyst on the
`downstream portion of the filter wall, the 150 g/L sample
`
`measured 4.2 kPa greater back pressure versus the 90 g/L
`sample. This can be attributed to more SCR catalyst located
`across the full width of the filter wall for the 150 g/L sample
`and in close contact with collected soot near the upstream
`filter-channel wall. However, with the catalyst on the upstream
`portion of the filter wall, the 150 g/L sample exhibited 9.0 kPa
`greater back pressure compared with the 90 g/L sample. This
`can largely be attributed to significantly more catalyst present
`on the inlet channel wall for the 150 g/L sample versus the 90
`g/L sample. This suggests that catalyst on the inlet channel wall
`has further impact on the extent of plugging during depth
`filtration and is hypothesized to be a magnification of the pore
`throat effect in the proximity of the upstream channel-wall
`interface.
`Once the system transitioned to cake filtration, additional
`discrepancies are apparent. Without catalyst on the inlet filter
`channel walls (Figure 4, left, and Figure 5, left and right), cake
`filtration behavior was similar in all three samples, exhibiting a
`pressure-drop rise in the range of 1.5−1.7 kPa/(g/L soot)
`under the conditions tested. However, with significant SCR
`catalyst on the inlet filter channel walls (Figure 4, right), cake
`filtration exhibited a pressure-drop increase of 3.4 kPa/(g/L
`soot). This is double the mass-specific effect of cake-filtered
`soot on pressure drop relative to the case with little or no
`catalyst present on the inlet channel.
`It
`is evident
`that
`permeability of the filter during cake filtration is adversely
`affected by the presence of catalyst on the inlet channel wall.
`Contributing to this effect is believed to be a combination of
`reduced open volume in the inlet channel and effective
`filtration area on the upstream channel wall surface, combined
`with the effect of the catalyst on flow dynamics at the wall and
`the

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket