13.07.2015 Views

Stagnant Mobile Phase Mass Transfer in ... - ResearchGate

Stagnant Mobile Phase Mass Transfer in ... - ResearchGate

Stagnant Mobile Phase Mass Transfer in ... - ResearchGate

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

7656 J. Phys. Chem. B, Vol. 103, No. 36, 1999 Tallarek et al.TABLE 1: Some Characteristics of the Spherical-Shaped, Rigid Porous Particlesproperties: material YMC ODS-A S50 POROS 50 R2 SOURCE 30Qmean diameter, d p [µm] 50.2 49.6 30.7standard deviation [µm] 10.6 9.5 0.7 ( 0.9996 for the best fit of the raw data to eq 5and a reproducibility of the results with<strong>in</strong> 5%.All calculations were performed us<strong>in</strong>g IDL (Interactive DataLanguage, Research Systems Inc., Boulder, CO). To accountfor the contribution of <strong>in</strong>traparticle longitud<strong>in</strong>al relaxation overtime (κ 2 - κ 1 ) ) (∆ - κ 1 ) to the slope of the A <strong>in</strong>tra (∆) vs∆curve (cf. Equation 5), comb<strong>in</strong>ed T 1 /T 1 2 H relaxation timemeasurements were made with a standard <strong>in</strong>version-recoverytechnique (T 1 ), <strong>in</strong>corporat<strong>in</strong>g a proper echo-tra<strong>in</strong> (T 2 ). Thesemeasurements did not reveal the existence of discrete T 1distributions for the <strong>in</strong>traparticle and the <strong>in</strong>terparticle fluidmolecules. Although the surface-to-volume ratios of the respectivepore spaces are different, it suggests that surface relaxationis not strong enough to impart a significant difference <strong>in</strong>longitud<strong>in</strong>al relaxation behavior. It is probably caused by thelow-field NMR measurements ( 1 H 20.35 MHz), comb<strong>in</strong>ed withthe extremely low amount of paramagnetic impurities present<strong>in</strong> these chromatographic media and the respective chemicalmodification of the particles (<strong>in</strong>ternal and external) surfaces.In the case of the porous silica particles (the only <strong>in</strong>organicoxide-based particles used <strong>in</strong> this work), for example, the Fe 3+contam<strong>in</strong>ation of the parent silica is well below 10 ppm. Froma chromatographic po<strong>in</strong>t of view, surface trace metal impuritieswould give strong Lewis acid adsorption sites which adverselyaffect performance. 77 Further, the orig<strong>in</strong>ally polar, silanolicsurface has been chemically derivatized by a bonded alkyl layer(high coverage of C 18 cha<strong>in</strong>s), <strong>in</strong>clud<strong>in</strong>g a C 1 -endcapp<strong>in</strong>g ofresidual silanols. It is a classical chromatographic support usedfor many reversed-phase separations, most applicable for polarto moderately nonpolar samples. Thus, return<strong>in</strong>g to the NMRstandpo<strong>in</strong>t, surface (and bulk) trace metal activity is very small,and the respective sites become sterically almost <strong>in</strong>accessible.Results and DiscussionWith respect to the actual fluid dynamics encountered <strong>in</strong> theexperiments, Figure 1a shows a representative s<strong>in</strong>gle-fluid phaseaveraged propagator distribution, Ph(R,∆), recorded <strong>in</strong> the axial(flow) direction, at a volumetric flow rate (F v ) of 8.0 mL/m<strong>in</strong>and an observation time (∆) of 25 ms. This quantitativedisplacement probability distribution illustrates both the potentialsteady-state character (regard<strong>in</strong>g <strong>in</strong>traparticle diffusion) andtransient nature (with respect to the stagnant mobile phase masstransfer k<strong>in</strong>etics) of these PFG-NMR measurements. For


<strong>Mass</strong> <strong>Transfer</strong> <strong>in</strong> Chromatographic Media J. Phys. Chem. B, Vol. 103, No. 36, 1999 7657Figure 2. Intraparticle self-diffusion coefficient of water <strong>in</strong> two ofthe chromatographic media as a function of the observation (diffusion)time; F v ) 8.0 mL/m<strong>in</strong>.Figure 1. Axial displacement probability distributions of the fluidmolecules (water) <strong>in</strong> a packed bed of porous particles (YMC ODS-AS50). 4.6 × 150 mm PEEK column; volumetric flow rate F v of 8.0mL/m<strong>in</strong>. The stagnant, purely diffusive fluid is found centered at zeronet displacement. Observation times, ∆ ) 25 ms and ∆ ) 420 ms.example, under the experimental conditions (Figure 1a), we f<strong>in</strong>d24.6% of the water molecules <strong>in</strong> the packed bed be<strong>in</strong>g leftunexchanged <strong>in</strong> the deep stagnant pools <strong>in</strong>side the porous C18silica particles. By a systematic variation of the observation time,this actually transient character of the measurement allows torecord the stagnant mobile phase mass transfer k<strong>in</strong>etics towarda complete exchange between stagnant and mov<strong>in</strong>g fluid <strong>in</strong> thepacked bed. 54 As is demonstrated <strong>in</strong> Figure 1b, at ∆ ) 420 msonly less than 0.5% of the stagnant water rema<strong>in</strong>s unexchanged.On the basis of tortuosity factor, the average size and sphericalshape of the particles it should be f<strong>in</strong>ished after some characteristicexchange time, ∆ e , assum<strong>in</strong>g that the fluid moleculeshave to travel a net distance of d p /2 (from the spheres center toits external surface) to leave the porous particles completely 1∆ e ) τ <strong>in</strong>tra2 D m[ d p2 ] 2 2with (〈R <strong>in</strong>tra 〉) ∆)∆e ) d p2We will see how this dimensional consideration matches theexperimental data. On the other hand, the root-mean-squared(rms) translational displacement of the yet unexchanged <strong>in</strong>traparticlefluid molecules (as long as ∆ < ∆ e , cf. Figure 1a)probed <strong>in</strong> the direction of the applied magnetic field gradientis given byFor the particle dimensions of the chromatographic media,mobile phase diffusivities and observation times selected forthis study, the correspond<strong>in</strong>g diffusive displacements of the(6)2〈R <strong>in</strong>tra 〉 ) (2 D <strong>in</strong>tra ∆) 1/2 (7)stagnant fluid molecules are much larger than the range oftypical <strong>in</strong>homogeneities <strong>in</strong> either <strong>in</strong>traparticle pore network. Inthis respect, the experiment then provides a steady-state porediffusion measurement which allows to extract an effective<strong>in</strong>traparticle diffusion coefficient, D <strong>in</strong>tra . Furthermore, these datanecessarily have to be consistent with the characteristic decayrates of the aforementioned, diffusion-limited exchange k<strong>in</strong>etics.These two series of data <strong>in</strong> fact provide us with complementary<strong>in</strong>formation, not only <strong>in</strong> view of the stagnant (diffusioncontrolled)boundary layer exist<strong>in</strong>g at the particles externalsurface under lam<strong>in</strong>ar flow conditions. 5 Consequently, theobta<strong>in</strong>ed results are now discussed separately under the lightof these aspects.Intraparticle, Effective Self-Diffusion Coefficients. Thes<strong>in</strong>gle fluid phases used <strong>in</strong> this study consist of water, acetonitrile,methanol, 2-propanol, and methylene chloride. Thesenonsorbed molecules are rather small compared to the poreconstrictions encountered <strong>in</strong>side the porous particles. Thus,hydrodynamic friction is negligible and the directly measured<strong>in</strong>traparticle rms translational displacements (and D <strong>in</strong>tra ) at longenough times approach the values given by the topology, i.e.,the <strong>in</strong>terconnectiveness of the pore network, by means of eq 3(with K -1 H ≈ 1). Because only s<strong>in</strong>gle-fluid phases are used,the calculated values of D <strong>in</strong>tra represent the correspond<strong>in</strong>g selfdiffusioncoefficients. 78Because of different mean pore sizes, pore size distributions,and pore <strong>in</strong>terconnectivities <strong>in</strong> the particles (Table 1), theassociated transport heterogeneities <strong>in</strong> the respective porenetworks are expected to be characterized by different correlationtimes and length scales. Figure 2, however, demonstratesthat, with<strong>in</strong> the range of observation times ∆ encountered <strong>in</strong>the measurements, D <strong>in</strong>tra <strong>in</strong> fact is found <strong>in</strong>dependent of thisparameter and <strong>in</strong> the tortuosity-limited regime of the respectivepore network, correspond<strong>in</strong>g to a steady-state pore diffusion <strong>in</strong>effeither case (i.e., such that D <strong>in</strong>tra ≡ D <strong>in</strong>tra ) D m /τ <strong>in</strong>tra ). Otherwise,D <strong>in</strong>tra would depend on ∆ and still decrease toward thetortuosity asymptote. The atta<strong>in</strong>ment of a steady-state porediffusion, together with its verification here by the data shown<strong>in</strong> Figure 2 is also important with respect to the average particlediameters (Table 1). For example, when larger pores start tocontribute to the <strong>in</strong>traparticle porosity, the pore network becomes<strong>in</strong>homogeneous on an <strong>in</strong>creas<strong>in</strong>g scale that may approach thesize of the whole particle. On the other hand, as long as theasymptotic (tortuosity limited) diffusion regime is reached, thetortuosity factor of a network consist<strong>in</strong>g of 100% small poresis the same as with 100% large pores. 42


7658 J. Phys. Chem. B, Vol. 103, No. 36, 1999 Tallarek et al.TABLE 2: Intraparticle Tortuosity Factors, τ <strong>in</strong>tra ) D m /D <strong>in</strong>trafluid: material YMC ODS-A S50 POROS 50 R2 SOURCE 30Q D m [10 -5 cm 2 /s]water 2.24 1.47 2.08 2.15acetonitrile 2.29 1.56 2.16 3.83methanol 2.23 1.51 2.11 2.232-propanol 2.15 1.43 2.03 0.58methylenechloride 2.21 1.53 2.09 3.07The <strong>in</strong>traparticle tortuosities measured by PFG-NMR aresummarized <strong>in</strong> Table 2, and the discussion necessarily focuseson the topology of the three pore networks. The most strik<strong>in</strong>gfeature of these data is the tortuosity factor of the gigaporousparticles (POROS 50 R2) which, for all the mobile phasesconsidered <strong>in</strong> this work, is significantly lower than that of theother two supports. In contrast to conventional media, thesePOROS beads have two discrete classes of pores <strong>in</strong>terconnected<strong>in</strong> a well-correlated manner. 23 The bidisperse particles are madeby <strong>in</strong>ter-adher<strong>in</strong>g PS-DVB microspheres which are fused <strong>in</strong>tothe f<strong>in</strong>al, cont<strong>in</strong>uous particle structure by us<strong>in</strong>g several stepsof cluster<strong>in</strong>g. This process creates a first set of 6000-8000 Åpores which transect the whole particle (throughpores) and asecond, more abundant set of smaller pores with 500-1500 Å<strong>in</strong> diameter. Because of the hierarchical design of these particles,the spac<strong>in</strong>g between throughpores is seldom greater than 1 µm,thereby reduc<strong>in</strong>g diffusion path lengths <strong>in</strong> the <strong>in</strong>terconnect<strong>in</strong>gsmaller pores to below that distance. 23 Thus, the actual correlationof <strong>in</strong>terconnectivity between the large and small porescreates a gigapore network (d pore /d p > 10 -2 ) that spans the wholeparticle and m<strong>in</strong>imizes diffusion distances <strong>in</strong> the macroporousdoma<strong>in</strong> (d pore > 500 Å). On the basis of the lower networktortuosities, the stagnant mobile phase mass transfer rates with<strong>in</strong>such hierarchical bidisperse networks will be superior comparedto bidisperse structures <strong>in</strong> which the pore sizes are assigned atrandom. 79,80 The transport behavior of such nonrandom bidispersepore networks <strong>in</strong> fact is expected to be closer to that ofthe <strong>in</strong>dividual giga- and macropore (or macro- and mesopore)networks arrayed <strong>in</strong> parallel rather than to the behavior of thecorrespond<strong>in</strong>g random network with bimodal pore size distribution.81 Although the correlation of <strong>in</strong>terconnectivity is now wellrecognizedas be<strong>in</strong>g of prime importance <strong>in</strong> the design of(bidisperse) porous particles, 79-83 surpris<strong>in</strong>gly little experimentaldata (like D <strong>in</strong>tra ) are available which directly characterize thetopology of these hierarchically structured pore networkscompared to less correlated ones.In contrast to the 50-µm PS-DVB particles (POROS 50 R2),the pore network of the 30-µm beads (SOURCE 30Q) ischaracterized by a relatively broad, unimodal pore size distribution,with the <strong>in</strong>terconnection of the larger and smaller poresbe<strong>in</strong>g rather uncorrelated, i.e., comparatively random. Thus, thelarge and small pores do not form <strong>in</strong>terconnected, dist<strong>in</strong>ct porenetworks, and the tortuosity factor then <strong>in</strong>creases due to theshield<strong>in</strong>g of large pores by small ones (Table 2). 84 F<strong>in</strong>ally, theyet slightly higher <strong>in</strong>traparticle tortuosity factors found for themesoporous silica-based particles (YMC ODS-A S-50) couldbe caused by the presence of a larger number of dead-end pores.However, the exact topological differences of the latter two porenetworks are hardly known with sufficient accuracy to allowsuch small dist<strong>in</strong>ctions to be made mean<strong>in</strong>gfully. Neverthelessand as a more general trend <strong>in</strong> view of the arguments followed<strong>in</strong> this work, the tortuosity factors observed for the nonhierarchical<strong>in</strong>traparticle pore networks are significantly higher thanthose found for the hierarchical bidisperse particles; however,the values are still close to 2. This value has been predicted foran isotropic porous medium, 31 recall<strong>in</strong>g the basic arguments ofCarman 85 <strong>in</strong> context of the parallel-pore model. Other argumentsTABLE 3: Cubic Lattices of Identical Spheres at DensePack<strong>in</strong>gcubic latticecoord<strong>in</strong>ationnumber<strong>in</strong>terstitialporositytortuosityfactorSC 6 0.476 1.38BCC 8 0.320 1.47FCC 12 0.259 1.62suggest a tortuosity factor of 3 for a particle network consist<strong>in</strong>gof randomly oriented, non<strong>in</strong>tersect<strong>in</strong>g cyl<strong>in</strong>drical pores whereeach straight capillary covers the length of the whole particle. 86With respect to the remarkably lower value of τ <strong>in</strong>tra , thehierarchical bidisperse particles can be viewed as consist<strong>in</strong>g ofloosely packed agglomerates be<strong>in</strong>g responsible for the severaltypes of pores f<strong>in</strong>ally present <strong>in</strong> the matrix. 23 Thus, the actualnetwork with<strong>in</strong> a s<strong>in</strong>gle support particle resembles itself, to somedegree at least, the overall (<strong>in</strong>tra- and <strong>in</strong>terparticulate) pore spacetypically encountered <strong>in</strong> packed beds of porous particles. Us<strong>in</strong>gthis similarity for a moment, Table 3 summarizes someproperties of cubic, i.e., regular lattices of identical, impenetrablespheres at dense pack<strong>in</strong>g so that the hard spheres are justtouch<strong>in</strong>g each other. 87,88 An <strong>in</strong>crease <strong>in</strong> the distance betweenneighbor<strong>in</strong>g spheres (or correspond<strong>in</strong>g decrease, while spheresstart to overlap) causes also an <strong>in</strong>crease (decrease) <strong>in</strong> the<strong>in</strong>terstitial porosity and <strong>in</strong> the effective, translational diffusivityof small molecules entra<strong>in</strong>ed <strong>in</strong> the respective pore network. 89Further, any k<strong>in</strong>d of positional disorder <strong>in</strong>troduced <strong>in</strong>to theparticle arrangement (at a constant porosity) leads to an<strong>in</strong>creased lengthscale of heterogeneities and a higher tortuosityfactor.Given the fact that a conservative estimate for the <strong>in</strong>traparticleporosity of the hierarchically structured particles (POROS 50R2) is 0.5, 23 comparison with Table 3 and the tortuosity factorsgiven <strong>in</strong> Table 2 suggests that neither regular (cubic) lattice atdense pack<strong>in</strong>g is appropriate to describe well this pore network.On the basis of a pure value identification, the simple cubicarrangement comes closest. Not surpris<strong>in</strong>gly, it is the one withthe smallest coord<strong>in</strong>ation number. These observations are <strong>in</strong>agreement with recent f<strong>in</strong>d<strong>in</strong>gs that the pore structure of theseparticles rather behaves like an <strong>in</strong>homogeneous assemblage ofloosely packed microspheres. 90 This situation is quite differentfrom the one typically encountered <strong>in</strong> random close-packed bedsof chromatographic columns. Here, the <strong>in</strong>terparticle porosity isusually found to vary between 0.36 and 0.4. More extremevalues are possible, depend<strong>in</strong>g on pack<strong>in</strong>g and operationalprocedures, as well as on particle characteristics. 91Remarkably low <strong>in</strong>traparticle tortuosity factors have also beenpostulated for hierarchically structured, spherical, porous zirconiaparticles us<strong>in</strong>g PFG-NMR methods. 92,93 Even though theauthors did not report the net effective, <strong>in</strong>traparticle diffusioncoefficients, the trend <strong>in</strong> their diffusivity data at least <strong>in</strong>dicatesvalues of τ <strong>in</strong>tra between 1.5 and 2 (i.e., below the isotropic porousmedium value). This behavior has been attributed to the veryopen pack<strong>in</strong>g of the colloids with<strong>in</strong> the particles. 92 Highertortuosity factors, yet close to 2, have been reported for prote<strong>in</strong>diffusion with<strong>in</strong> macroporous size-exclusion chromatographicmedia. 43,94 By restrict<strong>in</strong>g the diffus<strong>in</strong>g species to <strong>in</strong>dividual


<strong>Mass</strong> <strong>Transfer</strong> <strong>in</strong> Chromatographic Media J. Phys. Chem. B, Vol. 103, No. 36, 1999 7659particles us<strong>in</strong>g exclud<strong>in</strong>g solvent <strong>in</strong> the <strong>in</strong>terparticle void space,<strong>in</strong>traparticle labeled-prote<strong>in</strong> (tracer) diffusion could be studiedselectively by fluor<strong>in</strong>e PFG-NMR.All these data seem to <strong>in</strong>dicate a general trend for the actualhierarchy <strong>in</strong> pore structure and the associated tortuosity factor.However, they also call for further studies demonstrat<strong>in</strong>gsystematically its dependence on the process<strong>in</strong>g conditions andthereby achieved correlation (or randomness) of <strong>in</strong>terconnectionbetween differently sized pores. These experimental data willhelp <strong>in</strong> the ref<strong>in</strong>ement of chromatographic separations 95 and <strong>in</strong>the development of a numerical approach (and model validation<strong>in</strong> general) for predict<strong>in</strong>g the <strong>in</strong>terrelationship of structure andtransport <strong>in</strong> porous adsorbent particles. 96Another aspect closely related to the <strong>in</strong>traparticle diffusivityof the hierarchically structured particles (POROS 50 R2) meritsattention. On the basis of dimensional considerations andmorphological observations concern<strong>in</strong>g the gigaporous (throughpore)network, it has been proposed that these hierarchicallybidisperse particles can be operated under perfusion conditions,i.e., <strong>in</strong> such a mode that both <strong>in</strong>traparticle forced-convectionand diffusion contribute to the fluid transport <strong>in</strong> the gigapores,while only diffusion is active <strong>in</strong> the smaller macropores. 23Accord<strong>in</strong>g to Liapis and McCoy, 97 the term “perfusion chromatography”refers to any chromatographic system <strong>in</strong> whichthe <strong>in</strong>traparticle fluid velocity is nonzero. The effect of<strong>in</strong>traparticle convection is an enhancement of the overall rateof mass transfer <strong>in</strong> the particle, thereby improv<strong>in</strong>g the efficiencyof diffusion-limited chromatographic operations. It has beenpo<strong>in</strong>ted out conceptually that its contribution can be wellrepresentedby a convection-augmented effective, <strong>in</strong>traparticlediffusivity. 23,98 The first experimental direct confirmation of netflow through gigaporous particles has been reported recently<strong>in</strong> a test apparatus that isolates s<strong>in</strong>gle particles. 90 Figure 3ahowever shows that with<strong>in</strong> the range of hydrodynamic conditionsencountered <strong>in</strong> our current work, D <strong>in</strong>tra <strong>in</strong> all the chromatographicmedia rema<strong>in</strong>s constant <strong>in</strong> the purely diffusive,tortuosity-limited regime (D <strong>in</strong>tra < D m and τ <strong>in</strong>tra > 1). As canbe concluded from Figure 3b which records the accompany<strong>in</strong>gpressure drop over the 4.6 × 150 mm columns and also froma comparison with the results reported by Pfeiffer et al., 90 theactual pressure gradient across <strong>in</strong>dividual particles <strong>in</strong> the packedbed is probably too small to act as a driv<strong>in</strong>g force for anymeasurable <strong>in</strong>traparticle forced-convection assist<strong>in</strong>g or evendom<strong>in</strong>at<strong>in</strong>g fluid transport. Thus, compared to the other twosupports, the low tortuosity factor recorded for the hierarchicallybidisperse particle structure resembles diffusion control <strong>in</strong> a yetsuperior pore network, without the need to postulate, at thisstage at least, the operation of a perfusive mechanism.Next, we turn to the stagnant mobile phase mass transferk<strong>in</strong>etics which is associated with the exchange of fluid moleculesbetween the <strong>in</strong>vestigated purely diffusive <strong>in</strong>traparticle porenetwork (already with a preknowledge of the respective valuesof D <strong>in</strong>tra ) and the diffusive-convective <strong>in</strong>terparticle void space,how we can acquire and analytically describe it, and to thecomplementary <strong>in</strong>formation it provides.Nonsteady-State Diffusion <strong>in</strong>to/out of Spherical Particles.On the basis of the net (dynamic) displacements of the fluidmolecules over time ∆, the propagator formalism provides aquantitative characterization of the stagnant and mov<strong>in</strong>g fluid<strong>in</strong> the packed bed us<strong>in</strong>g the PFG-NMR technique. Averagedpropagator distributions recorded at <strong>in</strong>creas<strong>in</strong>g observation timesthus directly monitor the progressive diffusion out of thespherical particles. 60 Those fluid molecules which leave thespheres and participate <strong>in</strong> the <strong>in</strong>terparticle forced convectionFigure 3. Intraparticle self-diffusion coefficients of water <strong>in</strong> (a) thechromatographic media and (b) column pressure drop as a function ofthe volumetric flow rate; ∆ ) 60 ms.with<strong>in</strong> (<strong>in</strong>creased) time ∆ will no longer contribute to theensemble of purely diffusive fluid molecules which is foundcentered at zero net displacement. At long enough observationtimes the stagnant fluid fraction has disappeared and exchangeis complete (cf. Figure 1). Consequently, this situation resemblesthe general case of an unsteady-state diffusion of a solute <strong>in</strong>to/out of spherical, porous particles <strong>in</strong> which the spheres areemptied or, vice versa, filled by diffusion, and it can be treatedanalytically <strong>in</strong> the same manner. Thus, when we then considerthe case <strong>in</strong> which the diffusion (no chemical reaction yet) fromthe bulk of the particle of radius r p to its surface is radial, i.e.,the <strong>in</strong>itial and surface conditions are such that the concentrationC of a solute only depends on r and time ∆, the diffusionequation for a constant diffusion coefficient takes the form 99With the substitution u ) Cr and assum<strong>in</strong>g an <strong>in</strong>itialconcentration C i (r) <strong>in</strong> the sphere and surface concentrationC s (∆), the equations for u become 99with∂C∂∆<strong>in</strong>tra[ ) D ∂2 C∂r + 2 ∂C(8)2 r ∂r]∂u∂∆ ) D ∂ 2 u<strong>in</strong>tra for 0 e r < r∂r 2 p (9)u ) 0 when r ) 0 and ∆ > 0u ) C s (∆)‚r p when r ) r p and ∆ > 0u ) C i (r)‚r when 0 e r < r p and ∆ ) 0


7660 J. Phys. Chem. B, Vol. 103, No. 36, 1999 Tallarek et al.These are the equations for diffusion <strong>in</strong> a slab of thickness r p ,with its ends at r ) 0 and r ) r p kept at zero and C s (∆)r prespectively, and with the <strong>in</strong>itial distribution C i (r)r. Thus, theproblem of radial diffusion <strong>in</strong> (uniform) spheres can be deducedfrom the solution of the correspond<strong>in</strong>g l<strong>in</strong>ear problem. Equation9 is of the same form as the analogous differential equationcharacteriz<strong>in</strong>g heat transfer for unsteady-state heat conduction<strong>in</strong> plane sheets (and spheres). For that case, Carslaw andJaeger 100 have derived an analytical solution which is the productof a trigonometric and an exponential series. In terms of thedifferential equation (eq 9) and the boundary conditions thatthe sphere <strong>in</strong>itially is at a uniform concentration (C i (r) )constant ) C 0 ) and that the concentration at the surface of thesphere is ma<strong>in</strong>ta<strong>in</strong>ed steady at zero (C s (∆) ) constant ) 0),which causes the sphere to be emptied by diffusion, the averageamount of diffus<strong>in</strong>g substance still rema<strong>in</strong><strong>in</strong>g <strong>in</strong> the sphere atany time, A <strong>in</strong>tra (∆), is given by 100A <strong>in</strong>tra (∆)A <strong>in</strong>tra (0) ) 6 π 2 ∑ n)1∞ 1[ D n 2exp -n2 π 2 <strong>in</strong>tra ∆] 2r p(10)The correspond<strong>in</strong>g solution for small exchange times (andsmall values of the dimensionless parameter ζ ) (D <strong>in</strong>tra ∆)/r p 2 )isA <strong>in</strong>tra (∆)[ D A <strong>in</strong>tra (0) ) 1-6 <strong>in</strong>tra ∆ ]1/2 3D <strong>in</strong>tra ∆+ -2 2πr p r p12[ D <strong>in</strong>tra ∆ ]1/2 ∞ nr p2 ∑ ierfc =r n)1p D <strong>in</strong>tra1-6[ ζ + 3ζ (11)π]1/2∆At large observation (exchange) times ∆, on the other hand,eq 10 approximates to a simple exponential decay functionA <strong>in</strong>tra (∆)A <strong>in</strong>tra (0) ) 6 π 2exp[-π2 ζ] (12)Figure 4a shows the dependence on the observation time ofthe amount of stagnant water entra<strong>in</strong>ed <strong>in</strong> the respective particlepack<strong>in</strong>gs. At first glance, all the mass transfer data are welldescribedby eq 10 and values of the constant B ) D <strong>in</strong>tra π 2 /r p2obta<strong>in</strong>ed from the best fit of the data to this equation aresummarized <strong>in</strong> Table 4, together with the correspond<strong>in</strong>g valuesobta<strong>in</strong>ed by use of eqs 11 and 12 at short and long times,respectively. Evidently, the experimental approach followed <strong>in</strong>this work us<strong>in</strong>g s<strong>in</strong>gle-fluid phases, while monitor<strong>in</strong>g a (fictitious)diffusional empty<strong>in</strong>g of the porous particles (Figure 4a),avoids problems that could arise due to a nonconstant (i.e.,concentration-dependent) diffusion coefficient dur<strong>in</strong>g transientdiffusion experiments. 101A closer look at the data however reveals slight, butcharacteristic deviations from the best fit to eq 10 which<strong>in</strong>herently assumes, apart from a constant diffusion coefficient,a uniform diameter of the spherical particles. In this respect,both the particle shape and a size distribution will have apronounced effect on the experimental mass transfer curves. 102Compared to the predictions of eq 10, smaller-than-averageparticles will cause a higher <strong>in</strong>itial rate, while the larger-thanaverageparticles will be responsible for lower decay rates atlong times, especially when A <strong>in</strong>tra (∆)/A <strong>in</strong>tra (0) becomes smallerthan about 0.2. This is exactly what we observed for the supportswith the larger particle size distribution (Figure 4b and TableFigure 4. (a) Unsteady-state diffusion and total <strong>in</strong>traparticle fluidfraction unremoved as a function of the observation time. <strong>Mobile</strong> phase,pure water. F v ) 4.8 mL/m<strong>in</strong> (30 µm particles) and 8.0 mL/m<strong>in</strong> (50µm particles). (b) The effect of the f<strong>in</strong>ite particle size distributionbecomes evident, best fit of the data to eqs 10 and 14.4). In view of the fact that the monodisperse polymer particles(SOURCE 30 Q) have a remarkably narrower particle sizedistribution than the other two supports (cf. Table 1), 65 the effectwas <strong>in</strong>vestigated further. In pass<strong>in</strong>g we note that the <strong>in</strong>fluenceof the particle shape is not addressed separately and that a perfectspherical shape of the particles is assumed, a fairly reasonableassumption for these modern supports as can be shown byelectron microscopy of the media.Introduc<strong>in</strong>g a particle size distribution function f(r p ), wheref(r p )dr p is the fraction of particles hav<strong>in</strong>g a radius <strong>in</strong> the range(r p + dr p ), the appropriate expression for describ<strong>in</strong>g diffusionout of the spherical particles isA <strong>in</strong>tra (∆)A <strong>in</strong>tra (0) ) 6 ∑ ∞ 1∫ ∞f(rp )π 2 n)1 n 2 r p )0exp[ D -n2 π 2 <strong>in</strong>tra ∆] dr p (13)2r pThe particle size distributions of the supports have beenmeasured and particle sizes were found to be normally (Gaussian)distributed. When µ and σ, respectively, denote the meanand standard deviation of the particle radius, eq 13 thenbecomes 102A <strong>in</strong>tra (∆)A <strong>in</strong>tra (0) ) 6∑∞ 1∫ ∞σπ 2 n)12π n 2 r p )0 ×exp[ ( r - p - µσ2)2D - n 2 π 2 <strong>in</strong>tra ∆] dr p (14)2r p


<strong>Mass</strong> <strong>Transfer</strong> <strong>in</strong> Chromatographic Media J. Phys. Chem. B, Vol. 103, No. 36, 1999 7661TABLE 4: Exchange K<strong>in</strong>etics B ) D <strong>in</strong>tra π 2 /r p2 and D <strong>in</strong>tra for Watermaterialparameter YMC ODS-A S50 POROS 50 R2 SOURCE 30QB (eq 10) 11.51 19.11 35.48B (eq 11) a 11.94 19.54 eB (eq 12) b 8.62 16.47 35.53D <strong>in</strong>tra [10 -5 cm 2 /s], eq 10 c 0.73 1.21 0.81D <strong>in</strong>tra [10 -5 cm 2 /s], eq 14 d 0.80 1.30 0.83aFor A <strong>in</strong>tra(∆)/A <strong>in</strong>tra(0) > 0.45. b For 0 < A <strong>in</strong>tra(∆)/A <strong>in</strong>tra(0) < 0.15. c Based on the mean particle radius alone. d Us<strong>in</strong>g the mean particle radius andits standard deviation (cf. Table 1). e The <strong>in</strong>itial part of the mass transfer data is miss<strong>in</strong>g due to the smaller particle size and the lower limit of theobservation times.TABLE 5: D <strong>in</strong>tra from Exchange K<strong>in</strong>etics and Steady-State Pore DiffusionWith the <strong>in</strong>formation about µ and σ, obta<strong>in</strong>ed by means of<strong>in</strong>dependent measurements, eq 14 was evaluated numericallyand an <strong>in</strong>deed excellent agreement with the experimental datacould be established (Figure 4b). The values of D <strong>in</strong>tra obta<strong>in</strong>edby this procedure (eq 14) are also given <strong>in</strong> Table 4.Thus, by tak<strong>in</strong>g proper account of the supports particle sizedistribution (and the shape), a very satisfactory quantitativedescription of the experimental mass transfer data may beobta<strong>in</strong>ed from the diffusion equation, and for the moderateparticle size distributions encountered <strong>in</strong> this work, already eq10 provides a good work<strong>in</strong>g def<strong>in</strong>ition of the <strong>in</strong>traparticlediffusivity. With the monodisperse particles hardly any error is<strong>in</strong>troduced <strong>in</strong> the determ<strong>in</strong>ation of D <strong>in</strong>tra by assum<strong>in</strong>g a uniformparticle size (eq 10). For the other two supports, the effect ofthe broader size distribution leads to an upward correction ofD <strong>in</strong>tra of the order of 8%. These particle size distributions arestill representative for today’s high-performance pack<strong>in</strong>gs wherethe distributions tend to be no wider than about 25% aroundthe mean. 61 The results are <strong>in</strong> agreement with theoreticalanalyses on this topic, 103,104 which <strong>in</strong>dicate that the particle sizedistribution has a significant <strong>in</strong>fluence on chromatographicperformance only for much broader or asymmetric distributions.The same series of measurements and calculations reportedfor water (Figure 4 and Table 4) were also made us<strong>in</strong>gacetonitrile as the mobile phase, with similar results. Thesevalues of D <strong>in</strong>tra for water and acetonitrile (eq 14), together withthe correspond<strong>in</strong>g data obta<strong>in</strong>ed from the steady-state porediffusion measurements (Table 2) are summarized <strong>in</strong> Table 5.It is now <strong>in</strong>structive to compare the agreement between thesetwo sets of D <strong>in</strong>tra values because they necessarily have to beconsistent <strong>in</strong> terms of the <strong>in</strong>herently associated fluid dynamics,network topology and particle dimensions.Comparison between Steady-State and Transient Measurements.There exists a fair qualitative agreement betweenthese two approaches, but it becomes evident that the <strong>in</strong>traparticlediffusivities based on the exchange k<strong>in</strong>etics are systematicallysmaller (between 10 and 20%), even though these datahave been corrected for the effects of the particle size distribution.In turn, the particles shape or size distribution does not<strong>in</strong>fluence the values of D <strong>in</strong>tra which are calculated from thesteady-state pore diffusion measurement, because the rmstranslational displacement of the stagnant fluid over time ∆ isYMC ODS-A S50 a POROS 50 R2 a SOURCE 30Q aparameter H 2O b CH 3CN b H 2O b CH 3CN b H 2O b CH 3CN bD <strong>in</strong>tra [10 -5 cm 2 /s], eq 14 0.80 1.45 1.30 2.18 0.83 1.48D <strong>in</strong>tra [10 -5 cm 2 /s], Table 2 0.96 1.67 1.46 2.46 1.03 1.77δ film [µm] 2.2 1.9 1.7 1.7 1.6 1.4δ film/d p 0.044 0.038 0.034 0.034 0.052 0.046Pe c 491 275 491 275 177 99aMaterial. b Fluid. c Pe ) particle Peclet number ) 4F vd p/(ɛ <strong>in</strong>terπd c2 D m), with ɛ <strong>in</strong>ter, the <strong>in</strong>terparticle porosity (assumed 0.38) and d c, the columndiameter (0.46 cm).here directly measured (cf. Figure 1 and eq 7). Thus, the lattervalues of D <strong>in</strong>tra represent the time-weighted average of the fluidmolecules <strong>in</strong> stagnant regions of the packed bed. On the otherhand, the values of D <strong>in</strong>tra extracted from the mass transfer curvesare based on the predeterm<strong>in</strong>ation of the stagnant fluid fractionsover times ∆, which represent the volume-weighted average offluid molecules <strong>in</strong> (any) stagnant regions of the packed bed,and D <strong>in</strong>tra is then obta<strong>in</strong>ed from the decay rate of these curves(us<strong>in</strong>g eq 10 or 14). Included <strong>in</strong> these data, A <strong>in</strong>tra (∆)/A <strong>in</strong>tra (0) vs∆ (Figure 4), is the contribution of the stagnant boundary layerwhich exists under lam<strong>in</strong>ar flow conditions at the particlesexternal surface. 5,13 When the fluid molecules diffuse out ofthe particles and reach the spheres boundaries they first movethrough this stagnant film, where the transport normal to thesurface is dom<strong>in</strong>ated by diffusion, 4 before they enter the flowstreaml<strong>in</strong>es. The effect of the boundary layer, an idealizedstagnant, spherically symmetric film, is to slightly enlarge eachparticle. 9Consider<strong>in</strong>g the effect of the stagnant boundary layer on thetime-weighted average value of D <strong>in</strong>tra , any <strong>in</strong>fluence is due tothe higher value of the diffusivity with<strong>in</strong> this film (D film ≈ D m )compared to the situation <strong>in</strong> the <strong>in</strong>traparticle pore space (D <strong>in</strong>tra) D m /τ <strong>in</strong>tra ). Follow<strong>in</strong>g a fluid molecule from the center of thespherical particle to its external surface, a characteristic time∆ <strong>in</strong>tra of 325 ms is calculated based on eq 6 for travell<strong>in</strong>g thisnet distance of r p (us<strong>in</strong>g water and the porous C18 silica particlesas an example, d p ) 50 µm and τ <strong>in</strong>tra ) 2.24). Assum<strong>in</strong>g a filmthickness of 2.0 µm (0.04d p ), the time ∆ film it takes the fluidmolecules to pass this stagnant boundary layer (mov<strong>in</strong>g normalto the particle surface) is less than 1 ms (D m ) 2.15 × 10 -5cm 2 /s). When each stagnant fluid molecule has enough time tosample both the <strong>in</strong>traparticle pore space and the film region,the time-weighted average of the diffusion coefficients D stag ofthe stagnant fluid can be expressed as∑D n ∆ nn (D <strong>in</strong>tra ∆ <strong>in</strong>tra ) + (D film ∆ film )D stag ) )= D <strong>in</strong>tra (15)∆∑∆ <strong>in</strong>tra + ∆ filmnnUs<strong>in</strong>g for D n and ∆ n the values calculated <strong>in</strong> the aboveexample, we f<strong>in</strong>d D stag ) 1.0038D <strong>in</strong>tra . Unfortunately, this


7662 J. Phys. Chem. B, Vol. 103, No. 36, 1999 Tallarek et al.situation corresponds to observation times where the stagnantmobile phase mass transfer becomes already complete, because<strong>in</strong> terms of probability, fluid molecules orig<strong>in</strong>ally located at thecenter of the particle must also be able to reach the stagnantboundary layer so that eq 15 can be applied (i.e., ∆ ≈ ∆ <strong>in</strong>tra )∆ e , eq 6). But even at the shorter observation times, where theeffective, <strong>in</strong>traparticle self-diffusion coefficients are typicallymeasured (cf. Figure 2), a similar dimensional considerationshows that D stag ≈ D <strong>in</strong>tra to with<strong>in</strong> 1.5%. Thus, the <strong>in</strong>fluence ofthe slightly different diffusivity experienced by the fluidmolecules <strong>in</strong> the stagnant film envelop<strong>in</strong>g the spheres on thedeterm<strong>in</strong>ed values of the steady-state <strong>in</strong>traparticle diffusivities(given <strong>in</strong> Table 2 <strong>in</strong> form of the respective tortuosity factors)can be neglected.The situation is different for the reported transient measurements.The volumetric effect of the stagnant boundary layermanifests itself <strong>in</strong> an additional (diffusional) mass transferresistance external to the spheres. Thus, the actually stagnant<strong>in</strong>terstitial fluid <strong>in</strong> the packed bed at a given Peclet numberdictates an effective <strong>in</strong>terparticle porosity available for flow. 5On the other hand, the mean particle radius enters <strong>in</strong> eqs 10and 14 as a squared “constant”. From an <strong>in</strong>traparticle viewpo<strong>in</strong>t,this immediately suggests the <strong>in</strong>troduction of a hydrodynamicallyeffective particle radius (<strong>in</strong>clud<strong>in</strong>g the film thickness δ film )to account for its contribution to the recorded mass transferk<strong>in</strong>etics. It certa<strong>in</strong>ly represents an oversimplified picture of the(considerably streaml<strong>in</strong>ed) stagnant fluid at the free surface ofthe particles and <strong>in</strong> the cusp regions around their contact po<strong>in</strong>ts,but nevertheless provides a tractable start<strong>in</strong>g po<strong>in</strong>t with<strong>in</strong> thisstudy to cover the average effect.For the calculation of the data reported <strong>in</strong> Tables 4 and 5,the mean particle radius was used as determ<strong>in</strong>ed from theparticle size distribution measurement. Now, the accurate valuesof the steady-state pore diffusivities act as a boarder l<strong>in</strong>e forD <strong>in</strong>tra and the systematically lower diffusivities obta<strong>in</strong>ed fromthe exchange k<strong>in</strong>etics are adjusted to these numbers by def<strong>in</strong><strong>in</strong>gthe hydrodynamically effective particle radius, r p * ) r p + δ film ,i.e., the steady-state diffusivities (Table 2) are used as <strong>in</strong>putparameters <strong>in</strong> eq 14 and the (“effective”) mean particle radius,with µ ) r p *, is fitted. The values of δ film obta<strong>in</strong>ed by thisprocedure are listed <strong>in</strong> Table 5. The measurements were madeat a volumetric flow rate, F v , of 8.0 mL/m<strong>in</strong> <strong>in</strong> case of the 50µm particles and at one of 4.8 mL/m<strong>in</strong> with the smaller 30 µmparticles due to a higher column pressure drop <strong>in</strong> this case. Thecorrespond<strong>in</strong>g (particle) Peclet numbers of each experiment arealso given <strong>in</strong> Table 5. Depend<strong>in</strong>g on the actual hydrodynamicconditions, we f<strong>in</strong>d a film thickness for the different fluid phasesand support particles rang<strong>in</strong>g between 1.4 and 2.2 µm (i.e., upto 0.05d p ). These data are now <strong>in</strong>spected <strong>in</strong> view of theirdimension, but they are too limited to allow conclusions aboutthe dependence of the stagnant boundary layer thickness onmolecular diffusivity (H 2 O and CH 3 CN) and Peclet number.Compared to the 50 µm particles, the results obta<strong>in</strong>ed for the30 µm spheres (Table 5, δ film /d p ), however, <strong>in</strong>dicate that theboundary layer is thicker at lower Peclet numbers.The flow rate dependence (above Pe ≈ 50) of its thicknessis an important criterion for the existence of a stagnant boundaryat the <strong>in</strong>terface. To confirm this observation, we conductedexperiments at different flow rates and carefully recorded thecomplete mass transfer k<strong>in</strong>etics with purely diffusive particles.This ensures that the <strong>in</strong>traparticle diffusivity itself will be<strong>in</strong>dependent of Pe (cf. Figure 3a) and any significant differences<strong>in</strong> the observed exchange k<strong>in</strong>etics may then be traced back toa change <strong>in</strong> the hydrodynamically effective particle diameter.Figure 5 shows the results for acetonitrile <strong>in</strong> the column packedwith the 50 µm C18 silica particles. The parameters obta<strong>in</strong>edFigure 5. Influence of the flow rate on the diffusion-limited exchangek<strong>in</strong>etics. <strong>Mobile</strong> phase, pure acetonitrile. Column packed with porousC18 silica particles (YMC ODS-A S50). Best fit of the data to eq 14(solid l<strong>in</strong>e).with eqs 10 and 14 are summarized <strong>in</strong> Table 6. As may be seen,the flow rate probably has an effect and the value of D <strong>in</strong>traapparently <strong>in</strong>creases steadily when the mean particle radius isma<strong>in</strong>ta<strong>in</strong>ed strictly constant at its value obta<strong>in</strong>ed from theparticle size distribution measurement.The same k<strong>in</strong>d of observation has already been made by Boydet al. 9 <strong>in</strong> their classical work on the k<strong>in</strong>etics of ion exchangeadsorption processes <strong>in</strong> a shallow-bed apparatus. For theirdiffusion-limited stagnant mobile phase mass transfer k<strong>in</strong>eticsthese authors found a clear dependence on the flow rate of the“constant” B ) D <strong>in</strong>tra π 2 /r p 2 <strong>in</strong> eq 10. They concluded that theparticles were made effectively slightly larger by a surround<strong>in</strong>gliquid film which should be an <strong>in</strong>verse function of the flowvelocity.Thus, by keep<strong>in</strong>g now <strong>in</strong> our calculations the acetonitrile<strong>in</strong>traparticle diffusivity constant at its value known from thesteady-state pore diffusion measurement (D <strong>in</strong>tra ) 1.67 × 10 -5cm 2 /s), the <strong>in</strong>crease <strong>in</strong> Peclet number is accompanied by a steadydecrease of the film thickness from 3.1 to 1.9 µm (r p * from28.2 to 27.0 µm, Table 6). Due to the dimensions <strong>in</strong>herently<strong>in</strong>volved (50 and 30 µm particles), the absolute effects areexpectedly small. We now seek for dimensional estimates ofthe boundary layer and its dependence on the hydrodynamicsavailable from literature data.In an attempt to describe the <strong>in</strong>terplay of convective anddiffusional mix<strong>in</strong>g of unsorbed solutes <strong>in</strong> the <strong>in</strong>terstitial spaceof packed beds, Horváth and L<strong>in</strong> 105 argued that eddy dispersionoccurs only outside the stagnant film, i.e., <strong>in</strong> the liquid whichis actually mov<strong>in</strong>g between the particles. To evaluate thethickness of this film they came up with the “free surface model”of Pfeffer and Happel 106,107 who developed it <strong>in</strong> their analyticalstudy of heat and mass transfer to a bed of spherical particlesat low Reynolds numbers. These authors found that at highPeclet (Pe) numbers the Sherwood (Sh) number depends on PebySh ≡ k film d pD m) ΩPe 1/3 (16)Here, k film is the fluid film mass transfer coefficient and thevalue of Ω is a function of the <strong>in</strong>terstitial porosity only, chang<strong>in</strong>gfrom 3.6 to 3.2 when ɛ <strong>in</strong>ter varies from 0.36 to 0.42. Fairagreement with this dependence has been obta<strong>in</strong>ed by Wilsonand Geankoplis 108 on the basis of their electrochemical measurements(for 0.0016 < Re < 55).


<strong>Mass</strong> <strong>Transfer</strong> <strong>in</strong> Chromatographic Media J. Phys. Chem. B, Vol. 103, No. 36, 1999 7663TABLE 6: Flow Rate Dependence of <strong>Mass</strong> <strong>Transfer</strong> K<strong>in</strong>etics aF v [ml/m<strong>in</strong>] Pe D <strong>in</strong>tra [10 -5 cm 2 /s] b r p* ) r p + δ film [µm] c δ film/d p B d δ film [µm] e1.6 55 1.34 28.2 0.062 19.2 3.74.1 141 1.42 27.3 0.044 20.9 2.78.0 275 1.45 27.0 0.038 21.3 2.2a<strong>Mobile</strong> phase: acetonitrile. Pack<strong>in</strong>g material: 50 µm C18 silica (YMC ODS-A S50). b Calculated with eq 14, assum<strong>in</strong>g a constant meanparticle radius, r p ) 25.1 µm (Table 1). c Obta<strong>in</strong>ed from eq 14, us<strong>in</strong>g a constant diffusivity, D <strong>in</strong>tra ) 1.67 × 10 -5 cm 2 /s (Table 2). d B ) D <strong>in</strong>traπ 2 /r p2 ,eq 10. e Based on eq 17, with Ω ) 3.6.s<strong>in</strong>gle decay curve characteristic of Fickian diffusion. Itdemonstrates that <strong>in</strong> this case simple geometrical aspects(particle shape, mean diameter, particle size distribution, andthe tortuosity factor) are the only important parameters whichcharacterize the stagnant mobile phase mass transfer resistance.In this respect, the results <strong>in</strong>dicate aga<strong>in</strong> the importance of theactual correlation (or randomness) of pore <strong>in</strong>terconnectivity <strong>in</strong>the particles.ConclusionsFigure 6. Total fraction of the stagnant fluid rema<strong>in</strong><strong>in</strong>g unexchanged<strong>in</strong> the packed beds as a function of the observation time, normalizedby the characteristic exchange time (∆ e). <strong>Mobile</strong> phases, water andacetonitrile. F v ) 4.8 mL/m<strong>in</strong> (30 µm particles) and 8.0 mL/m<strong>in</strong> (50µm particles).The film thickness was then obta<strong>in</strong>ed by relat<strong>in</strong>g it to themass transfer coefficient via the Nernst diffusion layer 105δ film ) D m)(17)k film ΩPe 1/3The stagnant boundary layer thickness expected on the basisof eq 17 (strictly applicable for Pe > 50) is given <strong>in</strong> Table 6for comparison. Although the absolute values of the filmthickness obta<strong>in</strong>ed from our measurements are found to divergefrom the calculated ones by up to 20%, they still fall <strong>in</strong>to areasonable order of magnitude and follow qualitatively thepredictions of eq 17. It should be po<strong>in</strong>ted out that at this stageand with the available data such a comparison should help toga<strong>in</strong> dimensional <strong>in</strong>sight, rather than provid<strong>in</strong>g quantitativedescription or agreement.Thus, by <strong>in</strong>troduc<strong>in</strong>g a flow rate dependent, effective particleradius, which is similar to the def<strong>in</strong>ition of an effective (velocitydependent) <strong>in</strong>terparticle porosity, the PFG-NMR approachappears promis<strong>in</strong>g to study the (relative) hydrodynamic behaviorof the <strong>in</strong>terstitial stagnant fluid volume <strong>in</strong> packed beds of porousparticles. (N.B. For the sake of completeness, it has been po<strong>in</strong>tedout correctly that the film mass transfer coefficient used byHorváth and L<strong>in</strong> 105 was orig<strong>in</strong>ally derived under conditions ofa concentration boundary layer profile. 109 The correct filmthickness thus would be the concentration boundary layerthickness which will be less than the Nernst diffusion layerthickness.)F<strong>in</strong>ally, we close by return<strong>in</strong>g to the topological aspect ofthe <strong>in</strong>traparticle pore network. All mass transfer data recordedfor the three pack<strong>in</strong>g materials are plotted <strong>in</strong> Figure 6 aga<strong>in</strong>stthe observation time which now has been normalized by thecharacteristic exchange time def<strong>in</strong>ed by eq 6. In the calculationfor ∆ e the <strong>in</strong>traparticle tortuosity factors obta<strong>in</strong>ed from thesteady-state pore diffusion measurement (Table 2) have beenused. The exchange k<strong>in</strong>etics obta<strong>in</strong>ed for the small, nonsorbedfluid molecules <strong>in</strong> the different pore networks collapse onto ad pThe PFG-NMR approach followed <strong>in</strong> this work allows torecord two series of data with complementary character regard<strong>in</strong>gthe diffusion and exchange k<strong>in</strong>etics <strong>in</strong> chromatographicmedia. The steady-state pore diffusion measurement gives aneffective, <strong>in</strong>traparticle self-diffusion coefficient from which therespective tortuosity factor can be calculated. The resultsdemonstrate the profound effect of the actual correlation of<strong>in</strong>terconnectivity between different pore sizes and therebyachieved hierarchy <strong>in</strong> pore structure on the associated transportbehavior. It has a strong impact on the chromatographicperformance of the support. These data are <strong>in</strong>dependent of theactual pack<strong>in</strong>g quality of particles <strong>in</strong> the column, their shape orsize distribution, and external contributions due to the stagnantboundary layer.The observational time scale of the PFG-NMR method canalso be used to record the complete stagnant mobile phase masstransfer k<strong>in</strong>etics <strong>in</strong> the packed bed of porous particles. Thesedata are <strong>in</strong>fluenced by the respective particle shape and sizedistribution, and by the hydrodynamics of the stagnant boundarylayer (consequently, also by the pack<strong>in</strong>g density), which presentsdiffusional mass transfer resistance external to the particles. Theeffects of the particle shape and size distribution can beaccounted for by <strong>in</strong>dependent measurements and the diffusionequation provides an excellent quantitative description of theexchange k<strong>in</strong>etics <strong>in</strong>clud<strong>in</strong>g this <strong>in</strong>formation. Slight corrections( 1) encountered <strong>in</strong> this work. Further, the k<strong>in</strong>etics ofadsorption/desorption could be derived more accurately if wecorrect properly for the <strong>in</strong>traparticle diffusivity contribution.Thus, it will be extremely useful to carry out similar determ<strong>in</strong>ationson a variety of other relevant chromatographic and ionexchange media.Acknowledgment. One of the authors (U. T.) gratefullyacknowledges the award of a Marie Curie Fellowship under


7664 J. Phys. Chem. B, Vol. 103, No. 36, 1999 Tallarek et al.the Tra<strong>in</strong><strong>in</strong>g and Mobility of Researchers (TMR) Program fromthe European Union (Contract ERBFMBI-CT98-3437).References and Notes(1) Gidd<strong>in</strong>gs, J. C. Dynamics of Chromatography; Marcel Dekker: NewYork, 1965.(2) Weber, S. G.; Carr, P. W. The Theory of the Dynamics of LiquidChromatography. InHigh Performance Liquid Chromatography; Brown,P. R., Hartwick, R. A., Eds.; Wiley: New York, 1989; Chapter 1.(3) Dullien, F. A. L. Porous Media: Fluid Transport and PoreStructure; Academic Press: New York, 1992.(4) K<strong>in</strong>g, C. V. J. Am. Chem. Soc. 1935, 57, 828.(5) Gottschlich, C. F. AIChE J. 1963, 9, 88.(6) Deans, H. A.; Lapidus, L. AIChE J. 1960, 6, 656.(7) Ergun, S. Chem. Eng. Prog. 1952, 48, 227.(8) Mart<strong>in</strong>, A. J. P.; Synge, R. L. M. Biochem. J. 1941, 35, 1358.(9) Boyd, G. E.; Adamson, A. W.; Meyers, L. S., Jr. J. Am. Chem.Soc. 1947, 69, 2836.(10) Boyd, G. E.; Meyers, L. S., Jr.; Adamson, A. W. J. Am. Chem.Soc. 1947, 69, 2849.(11) Kl<strong>in</strong>kenberg, A.; Sjenitzer, F. Chem. Eng. Sci. 1956, 5, 258.(12) van Deemter, J. J.; Zuiderweg, F. J.; Kl<strong>in</strong>kenberg, A. Chem. Eng.Sci. 1956, 5, 271.(13) Rosen, J. B. Ind. Eng. Chem. 1954, 46, 1590.(14) Done, J. N.; Knox, J. H. J. Chromatogr. Sci. 1972, 10, 606.(15) Huber, J. F. K. Ber. Bunsen-Ges. Phys. Chem. 1973, 77, 179.(16) Horváth, Cs.; L<strong>in</strong>, H.-J. J. Chromatogr. 1978, 149, 43.(17) Unger, K. K.; Jilge, G.; K<strong>in</strong>kel, J. N.; Hearn, M. T. W. J.Chromatogr. 1986, 359, 61.(18) Kirkland, J. J. Anal. Chem. 1992, 64, 1239.(19) Kalghatgi, K.; Fellegvári, I.; Horváth, Cs. J. Chromatogr. 1992,604, 47.(20) Hjertén, S.; Liao, J.-I.; Zhang, R. J. Chromatogr. 1989, 473, 273.(21) Petro, M.; Svec, F.; Gitsov, I.; Fréchet, J. M. J. Anal. Chem. 1996,68, 315.(22) Reeder, D. H.; Clausen, A. M.; Annen, M. J.; Flick<strong>in</strong>ger, M. C.;Carr, P. W.; McCormick, A. V. J. Colloid Interface Sci. 1996, 184, 328.(23) Afeyan, N. B.; Gordon, N. F.; Mazsaroff, I.; Varady, L.; Fulton,S. P.; Yang, Y. B.; Regnier, F. E. J. Chromatogr. 1990, 519, 1.(24) Frey, D. D.; Schwe<strong>in</strong>heim, E.; Horváth, Cs. Biotechnol. Prog. 1993,9, 273.(25) Nir, A.; Pismen, L. M. Chem. Eng. Sci. 1977, 32, 35.(26) Rodrigues, A. E.; Lopes, J. C.; Lu, Z. P.; Loureiro, J. M.; Dias, M.M. J. Chromatogr. 1992, 590, 93.(27) Heeter, G. A.; Liapis, A. I. J. Chromatogr. A 1997, 761, 35.(28) Davis, R. H.; Stone, H. A. Chem. Eng. Sci. 1993, 48, 3993.(29) Satterfield, C. N.; Colton, C. K.; Pitcher, W. H., Jr. AIChE J. 1973,19, 628.(30) Limbach, K. W.; Wei, J. AIChE J. 1990, 36, 242.(31) Epste<strong>in</strong>, N. Chem. Eng. Sci. 1989, 44, 777.(32) Yau, W. W.; Kirkland, J. J.; Bly, D. D. Modern Size-ExclusionLiquid Chromatography; John Wiley & Sons: New York, 1979; Chapter2.(33) Gidd<strong>in</strong>gs, J. C.; Kucera, E.; Russel, C. P.; Myers, M. N. J. Phys.Chem. 1968, 72, 4397.(34) Satterfield, C. N. <strong>Mass</strong> <strong>Transfer</strong> <strong>in</strong> Heterogeneous Catalysis; MITPress: Cambridge, MA, 1970.(35) Burganos, V. N.; Sotirchos, S. V. AIChE J. 1987, 33, 1678.(36) Carberry, J. J.; Bretton, R. H. AIChE J. 1958, 4, 367.(37) Aris, R. Chem. Eng. Sci. 1959, 11, 194.(38) Koch, D. L.; Brady, J. F. J. Fluid Mech. 1985, 154, 399.(39) Deans, H. A. Soc. Pet. Eng. J. 1963, 3, 49.(40) Coats, K. H.; Smith, B. D. Soc. Pet. Eng. J. 1964, 4, 73.(41) Sahimi, M.; Tsotsis, T. T. J. Catal. 1985, 96, 552.(42) Hollewand, M. P.; Gladden, L. F. Chem. Eng. Sci. 1992, 47, 1761.(43) Coffman, J. L.; Lightfoot, E. N.; Root, T. W. J. Phys. Chem. B1997, 101, 2218.(44) Li, J.; Cantwell, F. F. J. Chromatogr. A 1996, 726, 37.(45) Li, J.; Litw<strong>in</strong>son, L.; Cantwell, F. F. J. Chromatogr. A 1996, 726,25.(46) Villermaux, J. J. Chromatogr. Sci. 1974, 12, 822.(47) Haynes, H. W., Jr. Catal. ReV.-Sci. Eng. 1988, 30, 563.(48) D<strong>in</strong>g, A.; Candela, D. Phys. ReV. E1996, 54, 656.(49) Tallarek, U.; Albert, K.; Bayer, E.; Guiochon, G. AIChE J. 1996,42, 3041.(50) Lebon, L.; Oger, L.; Leblond, J.; Hul<strong>in</strong>, J. P.; Martys, N. S.;Schwartz, L. M. Phys. Fluids 1996, 8, 293.(51) Seymour, J. D.; Callaghan, P. T. AIChE J. 1997, 43, 2096.(52) Tessier, J. J.; Packer, K. J.; Thovert, J.-F.; Adler, P. M. AIChE J.1997, 43, 1653.(53) Am<strong>in</strong>, M. H. G.; Gibbs, S. J.; Chorley, R. J.; Richards, K. S.;Carpenter, T. A.; Hall, L. D. Proc. R. Soc. A 1997, 453, 489.(54) Tallarek, U.; van Dusschoten, D.; Van As, H.; Bayer, E.; Guiochon,G. J. Phys. Chem. B 1998, 102, 3486.(55) Tessier, J. J.; Packer, K. J. Phys. Fluids 1998, 10, 75.(56) Stapf, S.; Packer, K. J.; Graham, R. G.; Thovert, J.-F.; Adler, P.M. Phys. ReV. E1998, 58, 6206.(57) Tallarek, U.; Bayer, E.; Guiochon, G. J. Am. Chem. Soc. 1998,120, 1494.(58) Tallarek, U.; Bayer, E.; van Dusschoten, D.; Scheenen, T.; VanAs, H.; Guiochon, G.; Neue, U. D. AIChE J. 1998, 44, 1962.(59) Hollewand, M. P.; Gladden, L. F. Chem. Eng. Sci. 1995, 50, 309.(60) Tallarek, U.; van Dusschoten, D.; Van As, H.; Guiochon, G.; Bayer,E. Angew. Chem. 1998, 110, 1983; Angew. Chem., Int. Ed. Engl. 1998, 37,1882.(61) Neue, U. D. HPLC Columns: Theory, Technology, and Practice;Wiley-VCH: New York, 1997.(62) Poole, C. F.; Poole, S. K. Chromatography Today; Elsevier:Amsterdam, 1993.(63) Schenck, J. F. Med. Phys. 1996, 23, 815.(64) Regnier, F. E. Nature 1991, 350, 634.(65) Ugelstad, J.; Soderberg, L.; Berge, A.; Bergstrom, J. Nature 1983,303, 95.(66) Stejskal, E. O.; Tanner, J. E. J. Chem. Phys. 1965, 42, 288.(67) Stejskal, E. O. J. Chem. Phys. 1965, 43, 3597.(68) Callaghan, P. T. Pr<strong>in</strong>ciples of Nuclear Magnetic ResonanceMicroscopy; Clarendon Press: Oxford, 1993; Chapter 6.3.(69) Kärger, J.; He<strong>in</strong>k, W. J. Magn. Reson. 1983, 51, 1.(70) Callaghan, P. T.; Coy, A.; MacGowan, D.; Packer, K. J.; Zelaya,F. O. Nature 1991, 351, 467.(71) Callaghan, P. T.; MacGowan, D.; Packer, K. J.; Zelaya, F. O. J.Magn. Reson. 1990, 90, 177.(72) Cotts, R. M. Nature 1991, 351, 443.(73) Zimmerman, J. R.; Britt<strong>in</strong>, W. E. J. Phys. Chem. 1957, 61, 1328.(74) Brownste<strong>in</strong>, K. R.; Tarr, C. E. Phys. ReV. A1979, 19, 2446.(75) Stilbs, P. Prog. Nucl. Magn. Reson. Spectrosc. 1987, 19, 1.(76) Kärger, J.; Pfeifer, H.; He<strong>in</strong>k, W. AdV. Magn. Reson. 1988, 12, 1.(77) Nawrocki, J. Chromatographia 1991, 31, 177.(78) Albright, J. G.; Mills, R. J. Phys. Chem. 1965, 69, 3120.(79) Reyes, S. C.; Iglesia, E. J. Catal. 1991, 129, 457.(80) Hollewand, M. P.; Gladden, L. F. Chem. Eng. Sci. 1992, 47, 2757.(81) Nicholson, D.; Petrou, J. K.; Petropoulos, J. H. Chem. Eng. Sci.1988, 43, 1385.(82) Petropoulos, J. H.; Liapis, A. I.; Kolliopoulos, N. P.; Petrou, J. K.;Kanellopoulos, N. K. Bioseparation 1990, 1, 69.(83) Petropoulos, J. H.; Petrou, J. K.; Liapis, A. I. Ind. Eng. Chem. Res.1991, 30, 1281.(84) Sharratt, P. N.; Mann, R. Chem. Eng. Sci. 1987, 42, 1565.(85) Carman, P. C. Trans. Inst. Chem. Eng. (London) 1937, 15, 150.(86) Feng, C.; Stewart, W. E. Ind. Eng. Chem. Fundam. 1973, 12, 143.(87) Venema, P.; Struis, R. P. W. J.; Leyte, J. C.; Bedeaux, D. J. ColloidInterface Sci. 1991, 141, 360.(88) Blees, M. H.; Leyte, J. C. J. Colloid Interface Sci. 1994, 166, 118.(89) Dunn, K.-J.; Bergman, D. J. J. Chem. Phys. 1995, 102, 3041.(90) Pfeiffer, J. F.; Chen, J. C.; Hsu, J. T. AIChE J. 1996, 42, 932.(91) Guiochon, G.; Golshan-Shirazi, S.; Katti, A. M. Fundamentals ofPreparatiVe and Nonl<strong>in</strong>ear Chromatography; Academic Press: Boston, MA,1994.(92) Lorenzano-Porras, C. F.; Carr, P. W.; McCormick, A. V. J. ColloidInterface Sci. 1994, 164, 1.(93) Lorenzano-Porras, C. F.; Annen, M. J.; Flick<strong>in</strong>ger, M. C.; Carr, P.W.; McCormick, A. V. J. Colloid Interface Sci. 1995, 170, 299.(94) Gibbs, S. J.; Lightfoot, E. N.; Root, T. W. J. Phys. Chem. 1992,96, 7458.(95) Lightfoot, E. N.; Athalye, A. M.; Coffman, J. L.; Roper, D. K.;Root, T. W. J. Chromatogr. A 1995, 707, 45.(96) Rigby, S. P.; Gladden, L. F. Chem. Eng. Sci. 1996, 51, 2263.(97) Liapis, A. I.; McCoy, M. A. J. Chromatogr. 1992, 599, 87.(98) Rodrigues, A. E.; Ahn, B.; Zoulalian, A. AIChE J. 1982, 28, 541.(99) Crank, J. The Mathematics of Diffusion; Clarendon Press: Oxford,1956; Chapter 6.(100) Carslaw, H. S.; Jaeger, J. C. Conduction of Heat <strong>in</strong> Solids;Clarendon Press: Oxford, 1959; Chapter 9.3.(101) Westr<strong>in</strong>, B. A.; Zacchi, G. Chem. Eng. Sci. 1991, 46, 1911.(102) Ruthven, D. M.; Loughl<strong>in</strong>, K. F. Chem. Eng. Sci. 1971, 26, 577.(103) Rasmuson, A. Chem. Eng. Sci. 1985, 40, 621.(104) L<strong>in</strong>, Y. S.; Ma, Y. H. AIChE J. 1990, 36, 1569.(105) Horváth, Cs.; L<strong>in</strong>, H.-J. J. Chromatogr. 1976, 126, 401.(106) Pfeffer, R.; Happel, J. AIChE J. 1964, 10, 605.(107) Pfeffer, R. Ind. Eng. Chem. Fundam. 1964, 3, 380.(108) Wilson, E. J.; Geankoplis, C. J. Ind. Eng. Chem. Fundam. 1966,5, 9.(109) Arnold, F. H.; Blanch, H. W.; Wilke, C. R. J. Chromatogr. 1985,330, 159.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!