746
Views
8
CrossRef citations to date
0
Altmetric
Original Articles

Knudsen transition effects on the thermophoretic properties of fractal-like aggregates: Implications for thermophoretic sampling of high-pressure flames

&
Pages 1262-1274 | Received 17 Apr 2017, Accepted 01 Jun 2017, Published online: 01 Aug 2017

ABSTRACT

Aggregated particles nucleated and grown by Brownian encounters in atmospheric pressure gaseous flames, usually have primary particles (of radius R1) smaller than the prevailing gas molecule mean-free-path, lg. This simplifies their drift behavior in a strong temperature gradient, i.e., thermophoresis (TP), as has been exploited in the popular technique of TP/TEM soot sampling. Indeed, thermophoretic sampling has become the effective “calibration standard” because it also provides unambiguous aggregate morphology information and is independent of optical properties needed to interpret alternate “non-invasive” methods. However, we show here that at pressures of current engineering interest (e.g., 30–50 bar) the Knudsen number (Kn1 lg/R1) is O(1) and the sensitivity of aggregate TP-behavior to pressure, morphology and, especially, aggregate size is altered considerably. While the recent gas kinetic theory results of Young reveal that the thermal force on each spherule should diminish as one leaves the free-molecule limit, large fractal-like aggregates receive the benefit of inter-particle “momentum shielding,” and we predict that larger aggregates drift faster than smaller ones in the same temperature gradient. We show that aggregate TP-diffusivity approximately scales with a Kn1-dependent power, k, of the spherule number N, where the exponent k is as large as 0.44 for aggregates characterized by “dimension” Df = 1.8, kp/kg = 1000 and Kn1 below O(1). However, if the particle thermal conductivity far exceeds that of the carrier gas, the reduction in thermal force overwhelms the inter-particle momentum shielding below about Kn1 = 0.7, with the aggregate TP-diffusivity becoming inadequate for TP-sampling near Kn1 = 0.2. Based on these results we conclude that to infer accurate high-pressure mainstream aggregate size distributions and volume fractions via TP-sampling, it is necessary to correct observed TEM-information for the expected “over-representation” of large aggregates. These appreciable corrections (up to 40% for the number-mean aggregate size and approximately one-decade for the associated spherule volume fraction, φ, at pressure near 50 bar), are shown to be straightforward to implement for sufficiently large mainstream and approximately log-normal aggregate populations.

Copyright © 2017 American Association for Aerosol Research

1. Introduction, motivation, and objectives

While admittedly an “immersion” method requiring post-sampling TEM-grid image analysis, the experimental technique known as “thermophoretic (TP) sampling” (to determine the nature and quantity of particulate matter in combustion gases, first implemented by Dobbins and Megaridis [Citation1987]), has become the “gold-standard” in near-atmospheric- (or sub-atmospheric) pressure combustion studies. Its convenience for characterizing aggregated carbonaceous or inorganic “soots” has been based on the remarkable insensitivity of the TP-dominated capture efficiency to differences in aggregate size-, structure- (morphology) and even the intrinsic thermal conductivity of its constituent spherules (Rosner et al. Citation1991). Thus, if TP is indeed the dominant deposition mechanism (cf. Brownian diffusion or inertial impaction), relative abundances seen on the target often provide a sufficiently accurate picture of what must have existed in the local mainstream—independent of the often uncertain optical properties of such precipitates.

The principal reason for the abovementioned insensitivity has been the fact that the constituent spherules (often with radii of order only 10–20 nm) found themselves in the “free-molecule” domain of gas/solid momentum- and energy-transfer, i.e., the prevailing gas molecule mean free path far exceeded the individual spherule radii. Moreover, even large aggregates (containing thousands of primary particles) are often “open” enough to facilitate gas molecule “penetration,” i.e., their “internal” surface area is said to be quite “accessible” to the much smaller probing gas molecules (Rosner and Tandon Citation1994).

Our present theoretical extensions (Section 5) are motivated by current practical interest in the TP/TEM-sampling of high-pressure flames (Leschowski et al. Citation2014; Vargas and Gulder Citation2016), frequently well above 30 bar. Additionally, we anticipate the need for more realistic mathematical modeling of soot production and transport in practical high pressure combustion environments. This requires that we gain an understanding of the thermophoretic behavior of aggregates under gas-dynamic conditions in which the constituent spherules find themselves in the so-called Knudsen transition regime (Kn1 < 1)Footnote1 —if not the near-continuum regime (with Kn1 ≪ 1).

In this article, we develop/report/illustrate a rational yet tractable method (Section 5) to predict the orientation-averaged TP-properties of large fractal-like aggregates (FA) over the entire expected range of Knudsen numbers and thermal conductivity ratios: kp/kg. As will be seen, our numerical results now suggest that the abovementioned remarkable insensitivities of the dimensionless thermophoretic diffusivity (when Kn1 ≫ 1) are NOT expected to carry over to the more complex Knudsen transition regime in which primary particles of large fractal-like aggregates will find themselves in sampled, “sooting” high-pressure flames. Our present predictions will be seen to provide the basis for a rational/tractable “deconvolution” procedure (Section 6, , and the Appendix) capable of systematically inferring local mainstream aggregate size distributions and corresponding total mainstream aggregate volume fractions from what is experimentally observed on TEM-grids immersed in combustion products at pressures even up to ca. 200 bar—probably the present upper limit considering our underlying assumption (Sections 2 and 5.8) of the relative dominance of thermophoretic transport (over Brownian diffusion).

2. Present assumptions and simplifications

To proceed with our prediction of the effects of high-pressure on the thermophoretic properties of aggregated precipitate particles in combustion gases, we now explicitly introduce the following assumptions and simplifications:

A1

The underlying molecular accommodation coefficients which determine gas/solid thermal creep and momentum and energy transfer, already near their theoretical upper limits for most gas/macroscopic surface interactions (Li and Wang Citation2005) even below 1 bar, will not be significantly altered when the carrier gas molecular volume fraction remains small enough to fall within the ideal gas domain.

A2

In the large N (N ≫ 1) aggregate size range being investigated here (nominally gyration radii exceeding ca. 0.2 micrometers and comprised of “nano-spherules” with diameters of the order of 30 nm, Brownian motion of the constituent primary particles will not appreciably affect the thermal force, even in the isolated sphere free-molecule (Waldmann-) limit: Kn1 ≫ 1 ( Li and Wang Citation2004).

A3

Estimates of the prevailing thermal force on each primary particle (at each Knudsen number Kn1 and thermal conductivity ratio) will be based on the Young correlation (Young Citation2011) for an isolated sphere–recently demonstrated to be in better agreement with available experimental data than earlier correlations (e.g., that due to Talbot et al. Citation1980), especially for thermally conductive particles within the Knudsen transition range [where “higher-order kinetic theory effects” (i.e., non [Navier–Stokes–Fourier] theory) become important.

A4

To approximately account for thermal shielding (Section 5.1) we further assume that the temperature gradient experienced by each such spherule within the FA is simply reduced from the maximum (imposed) value by the local factor (3/2)·{1 + [keff/(2 kg,eff)]}1, where keff is the effective thermal conductivity of the local spherule suspension and kg,eff corrects the prevailing gas thermal conductivity for Knudsen sublayer effects in the vicinity of each primary spherule (Rosner and Papadopoulos Citation1996). For this purpose the orientation- averaged FA will be considered to be a radially non-uniform granular “porous medium”–as in the earlier mobility (inverse drag) calculations of Tandon and Rosner Citation(1995).

A5

To estimate the more significant effects of multi-spherule momentum shielding (i.e., drag reduction) we simply exploit recent drag (mobility) results (Thajudeen et al. Citation2015; Melas et al. Citation2014) which relate the Knudsen transition regime drag on the FA to the drag on an equivalent (“adjusted”) solid sphere (Dahneke Citation1973a,b,c) whose radius is calculated from both the orientation-averaged projected area of the FA, written ≪PA≫, and the continuum-limit mobility radius—taken to be a structure-dependent fraction of the gyration radius of the FA (Sorensen Citation2011).

A6

To make use of available “isolated” single-sphere results (Young Citation2011) and for simplicity, we assume that the spherules comprising the aggregate are only “weakly necked” (i.e., have nearly point contacts with their [often two-] nearest neighbors) and are all of nearly the same material and radius, R1. (Thus, we currently preclude partially “sintered” [or “re-structured”] aggregates and/or “mixed” aggregates with unusually large primary particle “polydispersity.”)

This combination of plausible assumptions and simplifications will be shown to lead to a tractable “quadrature”-method (Sections 5.2 and 5.3) for predicting the significant amount by which the thermophoretic diffusivity of an N-particle FA (with N in the range 102–106) should exceed the corresponding thermophoretic diffusivity of an isolated primary particle in the same environment. Of particular interest will be our predictions of the dimensionless TP-diffusivity: ≪αT·DpgN, for both Diffusion Limited Cluster Aggregates (DLCAs) and Reaction Limited Cluster Aggregates (RLCAs), with N-values in the frequently observed range: 102–106, kp/kg-values from, say, 102–103, and pressures so large that the prevailing Knudsen number Kn1 is of order unity or less (even when the primary spherule radius R1 is only ca. 10–20 nm).

Summarizing, our present focus is on the following question: At high pressures of current combustion interest what correction factors must be applied to TP/TEM sampling data to account for now-expected systematic differences in TP-dominated aggregate collection efficiency? Or, put another way: What quantitative correction factors must be applied to relative abundances observed on the TEM grid to predict actual relative abundances in the local mainstream? Provided the collection mechanism is, in fact, TP-dominated (for a convenient sufficient condition, see Section 5.8), our present tractable methods/preliminary results should now enable systematic corrections (Section 6 and the Appendix) for aggregate/spherule size (N,R1), FA-morphology (e.g., DLCAs or RLCAs), kp/kg-values, at sampling pressures of current engineering interest. The “correction” to the apparent total volume fraction of aggregates is found (Appendix) to rise to the level of a 10-fold systematic overestimate for either mainstream DLCAs or RLCAs at sampling pressures near 50 bar.

3. Relevance of recent work on the thermophoretic properties of isolated spherical particles in non-isothermal gases

In comparing Young's theoretical analysis (Young Citation2011) of the thermophoretic diffusivity of an isolated spherical particle in a monatomic gas temperature gradient with that of Talbot et al. (Citation1980) it becomes clear that while both results essentially possess the same Kn1-asymptotes (i.e., the same values in the free-molecule limit [Kn1 ≫ 1] and the [Navier-Stokes-Fourier] continuum limit [Kn1 ≪ 1]) they differ quite significantly in the Knudsen transition region (e.g., for Kn1-values between, say, 1 and 0.01), especially when the particle-to-gas thermal conductivity ratio: kp/kg is large, i.e., >>1. Indeed, as shown in , when kp/kg is larger than about 35, as it often is for combustion-generated particles (especially ordinary carbonaceous “soot”), Young's method (which includes non-Navier-Stokes-Fourier effects in the gas phaseFootnote2 ) actually predicts a domain of small and ultimately negative (reverse-) thermophoresis before changing sign again and approaching from below the now-familiar (Epstein-) continuum limit value of the dimensionless TP-diffusivity {approx. (3/4)· [1 + (kp/2kg)] −1}.

Figure 1. Dimensionless thermophoretic diffusivity of an isolated spherical particle of thermal conductivity kp in a non-isothermal ideal gas of thermal conductivity kg; alternative predictions of expected Knudsen transition behavior (Young Citation2011; Talbot et al. Citation1980).

Figure 1. Dimensionless thermophoretic diffusivity of an isolated spherical particle of thermal conductivity kp in a non-isothermal ideal gas of thermal conductivity kg; alternative predictions of expected Knudsen transition behavior (Young Citation2011; Talbot et al. Citation1980).

For Kn1 = O(1), Young (Citation2011) has critically reviewed available experimental data on the N = 1 thermal force reported by no less than 16 international groups over the time interval: 1952–1995, as well as N = 1 TP-drift velocity data reported by 9 groups between 1946 and 2009 (including microgravity experiments). On this basis Young (Citation2011) has shown that his present non-monotonic interpolation recommendation (motivated in part by higher-order Boltzmann equation solutions, plotted in for the dimensionless TP-drift velocity, and adopted in Sections 5 and 6 for our present aggregate drift velocity predictions) is more accurate than the previously-used Talbot et al. (Citation1980) monotonic interpolation—especially for the usually encountered case of thermally conductive spherules: kp/kg ≫ 1.

In what follows, we therefore make use of Young's correlationFootnote3 for the net thermal force—in effect, this enables our estimate of the dimensionless TP-“factor” (αT)1, defined here in such a way that the thermal force per spherule is:[1]

If this force acts on each of the N spherules in an aggregate, and if we invoke what is now known about the orientation-averaged mobility of FAs (Section 4) then we can infer the expected drift velocity of a FA in a gas of spatially non-uniform temperature. In Section 5 we use this approach to develop results for the expected orientation-averaged TP-diffusivity ≪αT·DpN defined by the drift-velocity relation:[2]

To a first approximation (Rosner Citation2000; Appendix), TP-dominated particle deposition rates will be directly proportional to this scalar coefficient ≪αT·DpN, with units m2/s. In Section 5 we develop our general method and use this method to estimate/ plot present values of the corresponding dimensionless aggregate TP-diffusivity, i.e., ≪αT·DpgN, in terms of Kn1 and N for both DLCAs (Df = 1.8, k0 = 1.3) and RLCA (Df = 2.1, k0 = 0.94) with, say, kp/kg = 1000. To open the door to a rational “deconvolution” procedure, in Section 6 we show how these predicted Knudsen transition effects would systematically distort the TP-sampled aggregate size distribution function, pdfw(N), providing a numerical illustration for the case of TP-sampling from a 2000 K flame at a pressure of 50 bar. (Corresponding systematic over-estimates of the total aggregate volume fraction are dealt with in the Appendix). We conclude (Section 7) with a brief discussion of several immediate theoretical and experimental implications of these predictions, especially timely in view of current international interest in thermophoretic sampling from high-pressure flames (Leschowski et al. Citation2014; Vargas and Gulder Citation2016).

4. Recent work on the mobility properties of large fractal aggregates

As summarized by Sorensen (Citation2011), useful scaling laws have been developed to facilitate calculations of the orientation-averaged drag on large FAs, both in the continuum limit (Kn1 ≪ 1), and in the Knudsen transition regime (Kn1 = O(1)).

Continuum results are conveniently stated in terms of the equivalent “mobility radius” of a FA, found to be expressible in the form:[3] i.e., due to “momentum shielding”, the low Reynolds number orientation-averaged drag on an FA containing N spherules is less than N-times the drag on a single isolated spherule in the same stream by the above mentioned dimensionless factor ηmom, which is a function of N for a given FA structure (i.e., k0, Df)Footnote4 .

To account for the drag reduction associated with gas rarefaction use is made of the Cunningham slip correction function (Allen and Raabe Citation1982) for an isolated solid sphere but using (in the definition of Kn) an effective aggregate radius which depends both upon the orientation-averaged projected area, ≪PA≫, of the FA and its continuum mobility radius, i.e.:[4]

This method, having roots in the small-chain aggregate “ASM”-work of Dahneke (Citation1973a,Citationb,Citationc), has been shown to be successful for large FAs in the more recent computational work of Zhang et al. (Citation2012); Thajudeen et al. (Citation2015) and Melas et al (Citation2014).

By balancing the total thermal force (Section 3) on an N-spherule FA against the transition regime total orientation-averaged aggregate drag, we are now in a position to predict N-dependent dimensionless aggregate TP-diffusivities when the prevailing gas pressures are high enough for the constituent spherules to enter the interesting Knudsen transition regime. Especially for large values of kp/kg (of practical interest) and FAs with large N, this method now leads to several previously unanticipated trends—as pointed out/explored in Sections 5 and 6.

5. Present predictions of ≪αT·DpgN in the Knudsen transition regime

5.1. Total thermal force on a fractal-like aggregate

We now imagine that the (orientation-averaged-) fractal aggregate has a radial structure defined by the solid-fraction function:[5] except for a small “random-(nearly-) close packed” core (see Tandon and Rosner Citation1995 for details). The total thermal force on such an aggregate can then be formally written (Equation (Equation6)):[6] where we have neglected the small changes in Kn1 within the FA interior and also explicitly factored out [grad T], the local temperature gradient in the carrier gas if the spherules comprising the aggregate were not presentFootnote5 .

5.2. FA drift velocity via force balance

Equating this thermal force (Equation (Equation6)) to the total drag on the aggregate (Section 4), and making use of Equations (Equation3) and (Equation4) of Section 4, we are led to the following interesting separable representation for the aggregate dimensionless thermophoretic diffusivity, defined by Equation (Equation2), i.e.8:[7]

Here we identify (αTDpg)1 as the Young-predicted dimensionless thermophoretic diffusivity of an isolated spherical particle at the prevailing Kn1 and kp/kg, () along with two non-dimensional multi-spherule “shielding factors” — one for momentum defined by:[8]

Introducing assumption A4 (Section 2) we are led to the following quadrature representation for the thermal shielding contribution:[9] a factor which is seen to be expressible as the ratio of two volume-averages over the radially symmetric (orientation-averaged) FA structure.

Because, for DLCAs, <φ> is so small (, which also summarizes our Df, k0 choices for the DLCA and RLCA parameters, as well as the correlations we have used for ηPA and ηmom) the internal effective thermal conductivity keff, say evaluated at <φ>, is not much larger than kg,eff and we expect (and find9, Section 5.3) that Sh will be close to unity. Even for very large RLCAs, inter-spherule thermal shielding remains weak, e.g., for N = 10,000, kp/kg = 103, Kn1 = 1 and Df = 2.1 we find (via Equation (Equation9)) that Sh = 0.987. Because <φ> is often ≪1, a useful rough first approximation to the thermal shielding factor Sh for situations with kp/kg≫1 is found to be: 1–3 <φ> (where use has been made of the classical Maxwell relation for the conductivity of a dilute suspension of spherical conductors). For DLCAs with N = 1000 the “quadrature” value of Sh (Equation (Equation9))Footnote6 for kp/kg = 1000 is in fact 0.99 (cf. 1–3·(0.0051) = 0.985 with <φ> = 0.0051).

Table 1. Statistical parameters and normalized correlations chosen to characterize two important classes of fractal (-like) aggregates (FAs).

5.3. Numerical predictions for ≪αT·DpgN ; dependence on Kn1, N, and fractal-like structure

For an important particular class of FAs, i.e., DLCAs with kp/kg = 1000, we depict in the result of applying the factor Smom· Sh (defined by Equation (Equation7)) to the N = 1 Young-values of (αT·Dpg)1 displayed in . To display the now appreciable N-dependence, the predicted Kn1-dependence is shown () for each of the distinct N-values: 10n (2 ≤ n ≤ 6). Especially noteworthy is the now-expected increase in TP-diffusivity (for any N) as one leaves the free-molecule (Waldmann) limit—prior to the ultimate dramatic descent attributed to the Young-monomer contribution () as Kn1 approaches values near 0.2. Peak aggregate TP diffusivities (favorable for the dominance of the TP-deposition mechanism over that of Brownian diffusion [Section 5.8], hence simplifying the deconvolution process) should be expected near Kn1 = 0.7—which, for R1 = 20 nm and T = 2000 K, corresponds to a combustor pressure of about 42 bar.

Figure 2. Predicted (Equation (Equation7)) dimensionless (orientation-averaged) thermophoretic diffusivity of a fractal aggregate (Df = 1.8, k0 = 1.3) containing N-spherules (Kn1-transition behavior for N = 1 based on ; Young Citation2011).

Figure 2. Predicted (Equation (Equation7[7] )) dimensionless (orientation-averaged) thermophoretic diffusivity of a fractal aggregate (Df = 1.8, k0 = 1.3) containing N-spherules (Kn1-transition behavior for N = 1 based on Figure 1; Young Citation2011).

Regarding the treatment of more “compact” aggregate structures, we note that similar calculations for RLCAs (with Df = 2.1, k0 = 0.94) lead to the results summarized in . While qualitatively similar to , significantly, the N-dependence of ≪αT·DpgN for Kn1 = O(1) is somewhat greater than that shown in . This is quantified in Section 5.7, where we extract/report useful correlations for the power-law dependence of ≪αT·DpgN on spherule number N. In Section 6 we demonstrate how these Kn1-dependent exponents can be used to correct future experimentally reported TP/TEM high pressure data for the expected “falsification” of both pdf(N) and the corresponding mainstream aggregate volume fraction.

Figure 3. Predicted (Equation (Equation7)) dimensionless (orientation-averaged) thermophoretic diffusivity of a fractal aggregate (Df = 2.1, k0 = 0.94) containing N-spherules (Kn1 transition behavior for N = 1 based on ; Young Citation2011).

Figure 3. Predicted (Equation (Equation7[7] )) dimensionless (orientation-averaged) thermophoretic diffusivity of a fractal aggregate (Df = 2.1, k0 = 0.94) containing N-spherules (Kn1 transition behavior for N = 1 based on Figure 1; Young Citation2011).

5.4. Predicted behavior of ≪αT·DpgN in the free-molecule and continuum limits

It is clear from the discussion above that the principal Kn1- dependencies are contained in the factors [αT·Dpg]1 () and the momentum shielding factor Smom defined by Equation (Equation8).The latter contains the nonlinear function Cslip(Kn1) but for the present asymptotic limits the simple approximation mentioned by Sorensen (Citation2011), i.e., Cslip(Kn1) ≈ 1+ 1.612 Kn1 will suffice.

5.4.1. Free molecule limit: Kn11

When Kn1 ≫ 1 we find that, independent of the continuum limit function ηmom, Smom approaches (ηPA)−1, which for DLCAs is 1.67 for N = 104 and for N ≫ 1 approaches the limiting value 2.09 (Meakin et al. Citation1989). These multipliers would then be applied to the dimensionless TP-diffusivity: 0.5385 for N = 1 (Waldmann value with diffuse molecular “reflection”). Our current predictions in this limit appear to be somewhat larger than those predicted by Mackowski (Citation2006), who used a Monte Carlo-numerical scheme (obtaining (his ) ca. 1.1 for Smom at N = 104). In any case, even when using TP/TEM sampling from near- or sub-atmospheric pressure flames, modest systematic capture-fraction corrections appear to be warranted if the FAs are sufficiently large. It is interesting to also note that in the (less-likely) limiting case of specular molecular reflection, Zurita-Gotor (Citation2006) has shown that, independent of N, all aggregates of any structure should exhibit ¾ for the dimensionless TP-diffusivity, corresponding to the complete absence of inter-spherule shielding, i.e., Smom·Sh = 1, when Kn1 ≫ 1.

Figure 4. Predicted “falsification” of observed fractal-like aggregate distributions due to expected N-dependent thermophoretic diffusivities (). Comparison between mainstream- and predicted wall- log-normal distributions for a p = 50 bar, T = 2000 K mainstream DLCA aggregate population characterized by = 100, Df = 1.8 and σg = 3; see the Appendix for the corresponding over-estimate of the inferred mainstream total aggregate volume fraction.

Figure 4. Predicted “falsification” of observed fractal-like aggregate distributions due to expected N-dependent thermophoretic diffusivities (Figure 2). Comparison between mainstream- and predicted wall- log-normal distributions for a p = 50 bar, T = 2000 K mainstream DLCA aggregate population characterized by = 100, Df = 1.8 and σg = 3; see the Appendix for the corresponding over-estimate of the inferred mainstream total aggregate volume fraction.

5.4.2. Continuum limit: Kn1 ≪ 1

When Kn1 ≪ 1 we find that Smom approaches 1/ηmom irrespective of ηPA, with correlations for ηmom summarized in for both DLCAs and RLCAs. Moreover, the continuum limits of (αT·Dpg)1 for both Young (Citation2011) and Talbot et al. (Citation1980) become indistinguishable (). However, as quantified in Section 5.5, all of the “low Kn1” results presently collected in and will be “inaccessible” if the primary spherules in the aggregates are in the frequently encountered 10–20 nanometer range. This is because the attainment of such small Kn1 values would require pressures inconsistent with our restriction to the ideal gas law. Only if dealing with aggregates containing spherules which are of the order of micrometers in radii could this regime be accessed without violating our ideal gas constraint.

5.5. Restrictions associated with departures from gas ideality

The ideal gas kinetic theory expression for the gas mean-free-path can be re-written in terms of the molecular volume fraction: (φ)molec = ng· [(π) (σ)3]/6, where ng = p/[kB· T] for an ideal gas. If the ideal gas -law fails when (φ)molec exceeds, say, 5%, then we find that the lg cannot be less than about 2.4· (σ)g,eff and therefore Kn1 cannot be less than about 2.4· [σ/R1]. Because (σ)g,eff is near 0.4 nm (and only weakly T-dependent) and R1 is typically near 15 nm, on this basis, Kn1,min will be about 6.3 × 10−2, almost independent of T. Thus, because of our present “Ideal Gas constraint” (see A1), we should not expect the “low Kn1 branches,” formally displayed in and to be “accessible.” However, we note that the precipitous fall in aggregate TP-diffusivity (after its notable rise well above fm-limit-values!) should be observable for combustion-generated “soot” at sufficiently high pressures w/o violating our near-ideal gas law “self-consistency” condition.

5.6. Relations between Smom and ≪D≫N ; Sh and (α)T,N

We remark here that what we have called the momentum shielding factor, Smom, bears a close relation to the Brownian diffusivity ratio [≪D≫N]/D1 which was calculated via our earlier “non-uniform porous sphere” model of a fractal aggregate (Tandon and Rosner Citation1995). Indeed, we find that the Einstein mobility equation and the FA relations in Section 4 imply:[10]

In this respect Smom, shown by our numerical results to be significantly greater than unity (Section 5.3), quantifies the amount by which N ≪D≫N exceeds the N = 1 diffusivity D1.

Somewhat analogously, the thermal shielding factor Sh introduced above is closely related to the ratio of dimensionless TP-factors: ≪αTN/(αT)1, i.e.,[11]

Recalling the close relation between αT and the net thermal force (Equation (Equation1)), this relation shows that the factor Sh quantifies the expected systematic difference between the actual FA TP-factor ≪αTN and N times the dimensionless TP factor for a single, isolated spherule, (αT)1.

Because Sh is near unity9 the prevailing value of ≪αTN will be about N-times the prevailing value of αT,1. It is also useful to point out that in the free-molecule limit the Waldmann-value of αT,1 can be shown to be expressible (via ideal gas kinetic theory) as:[12] which is itself ≫1 when R1 ≫ σg,HS,eff —as in the ideal gas-applications of present interest.

5.7. Power-law correlations for the N-dependence of ≪αTD/νgN expected at high pressures

One of our objectives (Section 6) is to calculate and illustrate the expected “distortion” of sampled aggregate size distributions (ASDs) and inferred aggregate mainstream volume fractions that results when the FA deposition rate acquires a significant dependence on aggregate size N—all other factors remaining constant. If, as is often the case, measured ASDs are well fit by a log-normal representation then a particularly convenient representation of the N-dependence of ≪αT·DpgN is the power-law form, i.e., an Nk dependence, where because Kn1 = O(1), k > 0.

From the dimensionless TP diffusivity results displayed in (for Df = 1.8, k0 = 1.3, kp/kg = 1000) we find that the effective exponent k increases as we move away from the free-molecule (Kn1 ≫ 1) limit, taking on the values: k(100) = 0.145 k(10) = 0.344 and k(1) = 0.432, k(0.2) = 0.442. A reasonable curve-fit to this behavior is:[13a]

A similar curve-fit to our Df = 2.1 results (depicted in ) is:[13b]

These values will be shown to greatly facilitate the interpretation of future TP/TEM sampling data from high pressure flames. Our methods will be illustrated in Section 6 and the Appendix.

5.8. Sufficient condition for the dominance of the TP-deposition mechanism

Equation (Equation11) also facilitates the development/use of a simple sufficient condition for the dominance of the TP-deposition mechanism. The respective laws governing deposition rates by thermophoresis or Brownian convective-diffusion (Rosner Citation2000) suggest that the following simple condition on ≪αTN will be sufficient to ensure TP-dominance:[14]

The diffusivity ratio Sc on the RHS of Equation (Equation14) can be estimated via the gas kinematic viscosity and the use of the slip-corrected Stokes-Einstein relation:[15]

Whether ≪αTN actually meets condition (Equation (Equation12)) can be verified because ≪αTN will be simply the ordinate of or 3 multiplied by the prevailing Sc-value calculated via Equation (Equation15).

6. Predicted “falsification” of observed aggregate size distributions (and corresponding mainstream particle volume fractions)

In this section, we calculate and illustrate the “distortion” of the sampled aggregate size distribution (ASD) that results when the FA deposition rate has a Nk power-law dependence on aggregate size N—all other factors remaining constant. This dependence will be seen to be particularly convenient if the measured ASDs are well fit by a log-normal (LN-) representation. From the DLCA dimensionless TP diffusivity results displayed in we find that the effective exponent k increases as we move away from the free-molecule (Kn1 ≫ 1) limit, with the k(Kn1) correlations for DLCA and RLCA given by Equations (Equation13a) and (Equation13b), respectively.

If, for sufficiently large N, the aggregate number flux to the cooler TEM grid is proportional to Nk then it is easy to show that if the mainstream normalized log normal (LN) aggregate size distribution (ASD) is written pdf(N), i.e.,[16] then the normalized pdfw(N) will also be log-normal and can be compactly written:[17] whereand we have exploited the fact that for ≫ 1 we can, in effect, replace the lower limit N = 1 by zero. A useful property of LN distribution functions is that any moment Mk can be expressed in closed-form via Ng and σg:[18]

Using these relations, as well as the LN-based relation between the pdf spread parameter and the pdf's moments: M2, M1 and M0, it is possible to now prove that, irrespective of the magnitude of the exponent k, the wall ASD will have the same spread as the mainstream pdf, i.e.,[19]

However, as now expected, Ng,w will be larger than Ng,∞ and the enlargement factor is found to be:[20]

Because the mainstream and TEM grid ASD-spreads are equal this also implies that the number-average aggregate size on the TEM grid will be systematically larger than by this same factor: .

While high-pressure experimental results for aggregate size distributions via TP-sampling have not yet appeared in the combustion literature, we can illustrate this expected systematic ”ASD-falsification” with the help of a simple numerical special case. Consider that we are sampling from a LN- population of FAs with Df = 1.8, and σg = 3, with = 100 comprised of, say, 12 nm radius spherules at p = 50 bar, 2000 K. Under these conditions we find that the gas mfp is about 11.7 nm so that Kn1 based on R1 is 0.975, i.e., quite close to unity. At this Knudsen number the exponent k is found to be about 0.43 and the corresponding pdfs are shown plotted in , along with each value, with the sampled being 168. The corresponding values of parameter k are found to be 0.39 and 0.34 for Knudsen number Kn1 of 5 and 10, respectively, resulting in sampled of 160 and 151. Such corrections become particularly significant, because properly implemented TP/TEM sampling, being independent of material optical properties and morphological uncertainties, has assumed the role of a calibration standard in most experimental studies of flame-generated particulate matter. As shown in the Appendix the consequences for inferred mainstream total aggregate volume fraction, which involves an integral over all N, are even more dramatic—rising to the level of one order of magnitude under these same conditions.

As for the canonical “inverse” problem of de-convoluting observed pdfw(N) data, when is sufficiently large (see, also, Section 7) the implications are now clear. To reconstruct the corresponding mainstream pdf(N) our TP-diffusivity predictions () show that one can assume the same pdf spread, but both Ng,∞ and will be smaller than their TP- sample counterparts by the factor , where k is evaluated from either and or Equations (Equation13a) and (Equation13b) at the prevailing Knudsen number based on the relevant gas mfp and .

Over the range of viable TP-sampling conditions considered here, we also expect that our principal conclusions (Sections 5 and 6) will not be significantly altered if our aggregate TP-diffusivity calculations are simply based on the number-mean R1-value, i.e., . This expectation is based on the facts that: (a) recently observed spherule radius spread-parameters (Vargas and Gulder Citation2016; Leschowski et al. Citation2014) appear to be quite modest (apparently less than 1.5 (with 1.0 corresponding to strict R1-uniformity), and (b) the condition for TP-dominance (Section 5.8) will likely rule out future Kn1-values which are small enough to cause the spherule size-dependence of the thermal force on conductive spherulesFootnote7 to become significant.

7. Conclusions, implications for TP/TEM sampling in high-pressure flames, and future work

Quite clearly, when flame conditions are such that the individual spherules found within aggregates are large enough to no longer find themselves in the “free-molecule” regime, the indicated dependencies of ≪αT·DpgN on aggregate size (N), structure (Df, k0) and thermal conductivity (via kp/kg) are found to be no longer “modest”! We provisionally conclude that, especially because kp/kg is usually large, the quantitative interpretation of TP/TEM sampling data from high-pressure flames will normally require a non-trivial “deconvolution” with the help of ≪αT·DpgN -results of the type shown for the first time in and , i.e., large aggregates would be appreciably “over-represented” on TEM grids. This is now explicitly illustrated in Section 6, with the expectation that suitable high-pressure experimental data on aggregate size distributions will soon become available. We find (Section 5.7) that over useful domains of Kn1 and N, power-law representations of our numerical results are adequate, greatly facilitating systematic “corrections” to “falsified” TP-data both for back-calculating pdf(N) (Section 6) and the corresponding mainstream aggregate volume fraction, (φ)p, ∞ (Appendix).

Despite the fact that for intermediate Kn1-values the nonlinearity of the presently accepted functions Cslip(Kn1), ηPA, ηmom and Sh will normally preclude simple exponent power-law relations, when -values are sufficiently large, our present TP-diffusivity prediction methods are sufficiently tractable to enable a straightforward data-reduction procedure. However, we find (Section 6) that over useful domains of Kn1 and N, power-law representations of our numerical results are adequate, greatly facilitating such systematic “corrections” to “falsified” TP-data, e.g., the Kn1 dependence of k can be represented by Equation (13). (The N-dependence of ≪αT·DpgN appears to be even greater [ca. N0.57] for RLCAs when Kn1 = O(1)). As shown in Section 6, if aggregate size distributions are well-fit by log-normal relations then these power-law representations make correcting “raw” TEM-observed abundances for the now-expected systematic differences in TP capture fraction rather straightforward.

As additional theoretical, numerical or experimental information becomes available (e.g., about the thermophoretic properties of isolated spherules, or the orientation-averaged mobility properties of spherule aggregates) it should be possible to incorporate these improvements in our present analysis. Of course, experimental confirmation of some or all of our present qualitative and quantitative predictions ( and ) would build confidence in our underlying assumptions (Section 2). While our present rational approximation methods rely on the largeness of the mean spherule number in the mainstream population, it remains to be seen how small N can actually be before the associated systematic errors become unacceptable. Since there is also likely considerable interest in situations dominated by “small” aggregates (e.g., ≤O(10)) systematic “small N-1”-corrections to our present results will undoubtedly be of future value. In any case, our present analysis will, hopefully, motivate further advances in high-pressure thermophoretic sampling and perhaps have presently unexpected applications. In this regard, we note that the Kn1-transition regime seems to be well-suited for exploiting thermophoresis for aggregate size separation—perhaps in combination with inertial effects. Thermophoresis in this Kn1-domain may also prove to be practical for aggregate separations based on spherule thermal conductivity (Rosner and Arias-Zugasti Citation2011). In closing, we also conjecture that analogous inter-particle shielding effects may cause a similar size-dependence for the electrophoresis of large fractal-like aggregate in liquid solutions of biological interest—when one leaves the thin Debye sheath limit.

Added in Press

The dimensionless TP-diffusivity results shown plotted in for DLCAs were based on the ηmom(N;Df = 1.8) correlation displayed in (Sorensen Citation2011). However, our recent extensions (Rosner and Tandon Citation2017, Aerosol Sci. and Tech. submitted), carried out to demonstrate the role of “momentum shielding” for the rival aggregate transport mechanisms of isothermal convective-diffusion, or inertial impaction, have revealed that our “porous medium” approach (see, also, Footnote 4) leads to somewhat higher ηmom -factors—amounting to ca. 1.6-1.8 fold, depending upon N. On this basis we now believe that the dimensionless TP-diffusivity predictions displayed in should be systematically reduced by about this modest factor near Kn1 = O(1). In this same follow-up paper our “porous sphere” model has also been generalized to independently validate the ‘Adjusted Sphere Method’ in the broad domain of N and Kn1-values for which KnN ≪ 1, i.e., the “outer” flow is near-continuum. Finally, we wish to add that our present Kn1-transition results for large aggregate thermophoresis will also be applicable to the TEM-analysis of thermally precipitated aggregated smokes sampled under near-atmospheric pressure conditions, especially at more modest (not flame-) temperatures, and when the spherule radii may reach ca. 50 nm. Examples of this type will evidently occur in sampling “pyrolysis smokes,” including those from orbiting spacecraft environments (Meyer et al. Citation2015; Mulholland et al. Citation2015).

Nomenclature

=

molecular mean thermal speed, [8kBT/(πmg)]1/2

D=

Brownian diffusivity

Df=

fractal dimension ()

Cslip=

Cunningham slip factor (Kneff-dependent)

FT,1=

thermal force on a spherule

fB(κ)=

Brinkman function (Tandon and Rosner Citation1995; their Equation (3.1-Equation2))

k=

Exponent in relation ≪αTDp/ν≫ ∼ Nk (Section 5.7, Equation (13))

kB=

Boltzmann constant

k0=

pre-factor (relating N to (Rgyr/R1)Df)

kg=

gas thermal conductivity (Fourier)

kp=

particle (spherule) thermal conductivity

Kn=

Knudsen number (lg/R1)

l=

molecular mean-free-path (in carrier gas), ≈ 2νg/

m=

mass of one gas molecule

N=

number of spherules in fractal aggregate (FA)

Np=

aggregate number density; Equation (EquationA1)

p=

pressure (units:1 bar(105 Pa) unless otherwise specified)

R1=

spherule radius

Rgyr=

aggregate gyration radius, R1·(N/k0)1/Df

Rmob=

mobility radius of FA containing N spherules; Equation (Equation3)

r=

radial position within FA (from FA center-of-mass)

Sc=

Schmidt number for particle diffusion in gas, νg/Dp,N

Smom=

momentum shielding factor for aggregate (Equations (Equation7) and (Equation8))

Sh=

energy (thermal) shielding factor for a FA (Equations (Equation7) and (Equation9))

T=

absolute temperature (Kelvins)

U=

thermophoretic drift velocity; Equation (Equation2)

V=

total included volume of one FA (based on Rmax)

Greek

αT=

thermophoretic factor (dimensionless); Equation (Equation1))

β=

prefactor relating N to (Rmax/R1)Df

χ=

permeability of local granular medium

φ=

local solid fraction within (orientation-avgd) FA; Equation (Equation5)

ηPA=

normalized, orientation-avgd proj. area of FA; Equation (Equation4) ≡ ≪PA≫/[N·(π R12)]

ηmom=

normalized drag of FA (cf. N- individual spheres); Equation (Equation3))

μg=

Newtonian viscosity of carrier gas

κ =

Rmax1/2

ν=

momentum diffusivity (kinematic viscosity)

σ=

effective hard sphere diameter of molecule

σg=

geometric mean spread parameter in LN pdf

Φ=

dimensionless thermal force, N = 1 (Young Citation2011)

ψ=

dimensionless drag on spherule (Young Citation2011)

Subscripts

c=

continuum

eff=

effective

fm=

free-molecule

g=

pertaining to the carrier gas

g=

pertaining to geometric mean

h=

pertaining to heat (energy) transfer

HS=

effective “hard-sphere” value

N=

pertaining to a FA with N spherules

p=

pertaining to the (solid-like) particles

min=

minimum value (consistent with A1)

mob=

mobility (drift velocity per unit force)

molec=

pertaining to the molecular gas

mom=

pertaining to momentum transfer

1=

pertaining to N = 1 (single, isolated spherule)

=

in the gas far from the aggregate center of mass

Acronyms and abbreviations

ASD=

aggregate size distribution

ASM=

adjusted sphere model (Dahneke Citation1973a,b,c)

cf=

short for confer meaning compare

DLCA=

diffusion-limited cluster aggregate ()

FA=

fractal (-like) aggregate

LN=

log-normal (Equation (Equation15))

mfp=

mean free path in carrier gas

PA=

projected area (orientation-averaged)

PDE=

partial differential equation

pdf=

probability density function

RLCA=

reaction-limited cluster aggregate ()

TEM=

transmission electron microscope

TP=

thermophoretic

Operators

grad()=

spatial gradient operator

|| grad T ||=

magnitude of the local vector grad T

O( )=

order-of-magnitude

< >=

volume-averaged

≪ ≫=

orientation-averaged

=

average value of ( )

Acknowledgments

This article is dedicated to the memory of Prof. Richard A. Dobbins (Brown University, Division of Engineering), who (with his then graduate student C. M. Megaridis) put the experimental technique of “Thermophoretic Sampling” on the combustion-generated particle “map” nearly 3 decades ago (Langmuir, vol 3 (1987) loc cit). It is also our pleasure to acknowledge the helpful comments of our current and former Yale/SEAS colleagues: J. Fernandez de la Mora, A. Gomez, A. G. Konstandopoulos, D. W. Mackowski, M. Arias-Zugasti, and M. Zurita-Gotor, as well as C.M. Megaridis (U IL-Chicago, Mechanical Engineering.), C. Sorensen (KSU Physics), R. L. McGraw (Brookhaven National Laboratory, Aerosol Chemistry), and G. W. Mulholland (University of Maryland, Chemical Engineering).

Funding

This work was partially supported by the Industrial Affiliates of the Yale/ChE Sol Reaction Engineering Research Group.

Notes

1 For present purposes, a useful preliminary estimate of the relevant mean-free-path (units: nm) will be lg(p,T)= 66 · [p(bar)]−1 · [(T(Kelvins)/300)1.15. (This dimensional form results from the ideal gas kinetic theory expression for mfp, together with a numerical estimate of the [weakly temperature-dependent] effective hard-sphere diameter of molecular nitrogen, i.e., (σ)N2,HS,eff.)

2 Associated with “higher order” gas kinetic theory—i.e., beyond the Enskog–Chapman level of ideal gas kinetic theory.

3 In the notation of Young (Citation2011), our (αT·Dpg )1 is Young's: Φ/ψ

4 As done in Tandon and Rosner (Citation1995), by viewing the (orientation-averaged) FA as a quasi-spherical “porous” object with a radially variable permeability, it is possible to relate (η)mom (N;Df,k0) to the closed-form Brinkman function fB(κ) (Equation 3.1-2 in Tandon and Rosner, loc cit) associated with the Re ≪ 1 non-dimensional drag on a sphere of uniform permeability. On this basis we find that a reasonable, independent estimate of (η)mom can be obtained from the relation:where (κ)effRmax/(χeff)1/2 and (χ)eff = (χ)Happel evaluated at the volume-averaged solids fraction: <φ>( ). See, also: “Added in Press.”

5 If the relative change in gas temperature over one molecular mean-free path is indeed small then transport coefficients defined by “linear” flux laws of the type used here for the aggregate TP-diffusivity: ≪ (αT·D)≫N in Equation (Equation2), and calculated as we have done in Section 5 (subject to the assumptions stated in Section 2), will not have a dependence on ||grad T|| itself. Moreover, when the condition: lg·||grad T||/T ≪ 1 is met then the dimensionless scalar coefficient transport coefficient ≪ (αT·D)/νg)≫ N , calculated and plotted in and , contains not only the entire dependence of the TP-drift velocity, UT, on pressure and spherule radius R1 (via Kn1), but also its dependence on the thermal conductivity ratio: kp/kg (with the help of ) and the aggregate's fractal-like structure via its parameters: Df and k0. As a numerical example, available data on the pressure-dependence of laminar flame speeds (say, for pre-mixed CH4/air) suggest that even at ca. 50 bar the product: lg·||grad T||/T would be of the order of only 10−4, i.e., ca.1/100th %. Thus, the internal structure of such gaseous flames would remain amenable to a “pseudo-continuum” analysis, including the use of flux-laws compatible with the principles of linear irreversible thermodynamics. Moreover, based on currently available experimental data for individual spherules, and fractal-like assemblies of such spherules (Sections 2 and 3), we provide in and the first available quantitative estimates for the relevant aggregate thermophoretic transport coefficients applicable in the Knudsen transition regime. Implications for investigators currently sampling aggregates from high pressure “sooting flames” are considered in Section 6.

6 Volume integrals needed for our estimates of the thermal shielding function, Sh (Equation (Equation9)) incorporated in and were calculated numerically via the radial integral of {(2/3)[ 1 + (keff/2kg,eff)]}−1 · [φ(r;Df,k0)] · 4(π) r2 dr where [φ(r) is given by Equation (Equation5), 0 ≤ rRmax, and V = (4/3)(π)Rmax3. For this purpose keff was taken to be the Maxwell-effective conductivity of the local suspension of uniform size conductive spherules, and both keff and kg,eff were reduced to account for non-zero Kn1 via the methods summarized in Rosner and Papadopoulos (Citation1996). This quadrature approximation to Sh (see Assumption A4 and Equation (Equation9)), which circumvents solution of the heat diffusion PDE within the (orientation-averaged) aggregate, is expected to be adequate for large fractal-like aggregates because <φ> ≪1 (), causing Sh to remain close to unity. This expectation has been verified a posteriori for the DLCAs and RLCAs explicitly considered here.

7 While beyond the scope of the present manuscript, our analysis of the FA TP diffusivity via both the (R1-dependent) individual particle thermal force (Young Citation2011)) and the drag on an assembly of N particles (Section 4) opens the door to quantitative estimates of the consequences of “spherule polydispersity.” Indeed, available high pressure TP-sampling data (Leschowski et al. Citation2014; Vargas and Gülder Citation2016) enable estimates of pdfw(R1) irrespective of N.

References

  • Allen, M. D., and Raabe, O. G. (1982). Re-evaluation of Millikan's Oil Drop Data for the Motion of Small Particles in Air. J. Aerosol Sci. 10:537–547.
  • Dahneke, B. E. (1973a). Slip Correction Factors for Nonspherical Bodies—I Introduction and Continuum flow. J. Aerosol Sci., 4:139–145.
  • Dahneke, B. E. (1973b). Slip Correction Factors for Nonspherical Bodies—II Free Molecule Flow. J. Aerosol Sci., 4:147–161.
  • Dahneke, B. E. (1973c). Slip Correction Factors for Nonspherical Bodies—III the Form of the General Law. J. Aerosol Sci., 4:162.
  • Dobbins, R. A., and Megaridis, C. M. (1987). Morphology of Flame-Generated Soot as Determined by Thermophoretic Sampling. Langmuir, 3:254–259.
  • Gomez, A., and Rosner, D. E. (1993). Thermophoretic Effects on Particles in Counterflow Laminar Diffusion Flames. Comb. Sci. Tech. 89:335–362.
  • Leschowski, M., Dreier, T., and Shultz, C. (2014). An Automated Thermophoretic Soot Sampling Device for Laboratory-Scale High-Pressure Flames. Rev. Sci. Instrum., 85, Paper 045103.
  • Li, Z., and Wang, H. (2005). Gas-Nanoparticle Scattering: A Molecular View of Momentum Accommodation Function. Phys. Rev. Lett., 95:014502.
  • Li, Z., and Wang, H. (2004). Thermophoretic Force and Velocity of Nanoparticles in the Free Molecule Regime. Phys. Rev. E, 70:021205.
  • Mackowski, D. W. (2006). Monte Carlo Simulation of Hydrodynamic Drag and Thermophoresis of Fractal Aggregates of Spheres in the Free-Molecule Flow Regime. J. Aerosol Sci., 37:242–259.
  • McEnally, C. S., Koylu, U. O., Pfefferle, L. D., and Rosner, D. E. (1997). Soot Volume Fraction and Temperature Measurements in Laminar Nonpremixed Flames Using Thermocouples. Combust. Flame, 109:701–720.
  • Meakin, P., Donn, B., and Mulholland, G. W. (1989). Collisions Between Point Masses and Fractal Aggregates. Langmuir, 5:510–518.
  • Melas, A. D., Isella, L., Konstandopoulos, A. G., and Drossinos, Y. (2014). Friction Coefficient and Mobility Radius of Fractal-Like Aggregates in the Transition Regime. Aerosol Sci. Technol., 48:1320–1331.
  • Meyer, M. E., Mulholland, G. W., Bryg, V., Urban, D. L., Yuan, Z., Ruff, G. A., Cleary, T., and Yang, J. (2015). Smoke Characterization and Feasibility of the Moment Method for Spacecraft Fire Detection. Aerosol Sci. Tech., 49:299–309.
  • Mulholland, G. W., Meyer, M. E., Urban, D. L., Yuan, Z., Ruff, G. A., Bryg, V., Cleary, T., and Yang, J. (2015). Pyrolysis Smoke Generated Under Low-Gravity Conditions. Aerosol Sci. Tech., 49:310–321.
  • Rosner, D. E. (2000). Transport Processes in Chemically Reacting Flow Systems, DOVER, Mineola, New York.
  • Rosner, D. E., and Arias-Zugasti, M. (2011). Novel Features of Aerosol Coagulation in Non-Isothermal Environments. Ind. Eng. Chem.-Res., 50:8932–8940.
  • Rosner, D. E., Mackowski, D. W., and Garcia-Ybarra, P. (1991). Size- and Structure-Insensitivity of the Thermophoretic Transport of Aggregated “Soot” Particles in Gases. Comb. Sci. Tech., 80(1-3):87–101.
  • Rosner, D. E., and Papadopoulos, D. H. (1996). Jump, Slip, and Creep Boundary Conditions at Nonequilibrium Gas/Solid Interfaces. Ind. Eng. Chem., 35:3210–3222.
  • Rosner, D. E., and Tandon, P. (1994). Prediction and Correlation of Accessible Area of Large Multi-Particle Aggregates. AIChE J., 40:1167–1182.
  • Rosner, D. E., and Tandon, P. (2017). Aggregation and Rarefaction Effects on Particle Mass Deposition Rates by Convective-Diffusion, Thermophoresis or Inertial Impaction; Consequences of ‘Momentum Shielding’. Aerosol Sci. Tech., in review.
  • Sorensen, C. M. (2011). The Mobility of Fractal Aggregates: A Review. Aerosol Sci. Technol., 45:765–779.
  • Talbot, L., Cheng, R. K., Schefer, R. W., and Willis, D. R. (1980). Thermophoresis of Particles in a Heated Boundary Layer. J. Fluid Mechs., 80:737–758.
  • Tandon, P., and Rosner, D. E. (1995). Translational Brownian Diffusion Coefficient of Large (multi-particle) Suspended Aggregates. Ind. Eng. Chem., 34:3265–3277.
  • Tandon, P., Terrell, J. P., Fu, X., and Rovelstad, A. (2003). Estimation of Particle Volume Fraction, Mass Fraction and Number Density in Thermophoretic Deposition Systems. Int. J. Heat Mass Transfer, 46:3201–3209.
  • Thajudeen, T., Jeon, S., and Hogan, C. J. (2015). The Mobilities of Flame Synthesized Aggregates/Agglomerates in the Transition Regime. J. Aerosol Sci., 80:45–57.
  • Vargas, A. M., and Gülder, Ö. L. (2016). A Multi-Probe Thermophoretic Soot Sampling System for High-Pressure Diffusion Flames. Rev. Sci. Instrum., 87(5):055101.
  • Vargas, A. M., and Gülder, Ö. L. (2016). Pressure Dependence of Primary Soot Particle Size Determined Using Thermophoretic Sampling in Laminar Methane-Air Diffusion Flames. Proceedings of Combustion Institute, in press, https://doi.org/10.1016/jproci2016.05.023.
  • Young, J. B. (2011). Thermophoresis of a Spherical Particle: Reassessment, Clarification, and New Analysis. Aerosol Sci. Tech., 45:927–948.
  • Zhang, C., Thajudeen, T., Larriba, C., Schwartzentruber, T. E., and Hogan, C. J. (2012). Determination of the Scalar Friction Factor for Nonspherical Particles and Aggregates Across the Entire Knudsen Number Range by Direct Simulation Monte Carlo (DSMC). Aerosol Sci. Tech., 46:1065–1078.
  • Zurita-Gotor, M. (2006). Size- and Structure-Independence of the Thermophoretic Transport of an Aerosol Particle for Specular Boundary Conditions in the Free Molecule Regime. J. Aerosol Sci., 37:283–291.

Appendix: Effect of predicted aggregate size-dependent thermophoretic velocity on inferred mainstream soot volume fractions

Underlying earlier “non-optical” estimates of absolute “soot” spherule volume fractions using TP-dominated flame sampling data (McEnally et al. Citation1997; Gomez and Rosner Citation1993; Tandon et al 2003) is the assumption of an aggregate size-insensitive TP-velocity, i.e., equivalent (Sections 5.7 and 6) to setting the exponent k = 0. However, as emphasized above (Section 5), we now find that this simplifying assumption fails badly when the sampling pressure is high enough to reduce the operative spherule Knudsen numbers (based on ) from ≫1 to O(1). Because the soot volume fraction is in fact calculated from the absolute total aggregate deposition rate on the cooler solid target, we turn our attention here to the theoretical correction factor associated with the calculable dimensionless deposition rate ratio—again assuming that the normalized mainstream aggregate size distribution (ASD), written pdf(N), is approximately “log-normal” (as in Section 6).

If m1 is the mass of one spherule and the total aggregate number density in the local mainstream being sampled then the total TP-dominated deposition rate can be approximated by:[A1]

But, suppose that most aggregates in the population contain many spherules, i.e., Ng ≫ 1 and, following Equation (Equation7), we can write:[A2]

Here Equation (EquationA2) may be regarded as a local power-law fit to the behavior of ≪αT*DpN/(αTD)1 near N = Ng,∞.

Equations (EquationA1) and (EquationA2) lead to the closed-form result:[A3]

Dividing by the corresponding expression when k = 0 then leads to the predicted deposition rate ratio, i.e.,[A4]

For example, when Kn1 = 1 and k = 0.43 (for Df = 1.8, k0 = 1.3) we find that if Ng,∞ = 100 and σg = 3 this particular ratio is as large as ca. 1.9—larger than the previously calculated -ratio (Section 6). But as calculated above is itself larger than the corresponding spherule deposition rate by the factor [Smom*Sh]N = Ng,∞, which we estimate (using and ) to be ca. 4.4. Thus, the overall enhancement in spherule deposition rate to the cold target is the product of 1.9 and 4.4, giving ca. 8.4—or nearly a one-decade correction.

On this basis we expect that the systematic corrections required to infer realistic mainstream soot volume fractions based on high-pressure TP/TEM flame sampling (often used as the calibration standard in this field) can rise to the level of necessary reduction factors by ca. one -decade. We note here that at the pressure levels of current combustion turbine interest (up to ca. 50 bar) this important conclusion would not be significantly altered if we had retained the earlier Talbot et al. (Citation1980) thermal force law (for individual spherules)—as a reference to will confirm.

This calculation can be repeated for more compact fractal-like aggregates (e.g., with Df = 2.1, k0 = 0.94, as in ) using the important deposition rate ratio (Equation (EquationA4)) in combination with and . The overall result is again a correction of ca. 10-fold. On this basis we conclude that apparent soot volume fractions calculated from total spherule counts on TEM grids at flame pressures near 50 bar using previously available methods would be systematically high by as much as one decade. This is but one dramatic consequence of the fact that such aggregates will no longer find themselves in the more familiar “free-molecule” regime, i.e., their spherules can take advantage of increased “momentum shielding” (Rosner and Tandon Citation2017).

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.