Fluorometric Measurement and Modeling of Droplet Temperature

Dec 5, 2013 - The Evolution of Electrospray Generated Droplets is Not Affected by Ionization Mode ... Journal of The American Society for Mass Spectro...
0 downloads 0 Views 621KB Size
Subscriber access provided by DUESSELDORF LIBRARIES

Article

Fluorometric Measurement and Modeling of Droplet Temperature Changes in an Electrospray Plume Stephen C. Gibson, Charles S. Feigerle, and Kelsey D. Cook Anal. Chem., Just Accepted Manuscript • Publication Date (Web): 05 Dec 2013 Downloaded from http://pubs.acs.org on December 7, 2013

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Analytical Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

Fluorometric Measurement and Modeling of Droplet Temperature Changes in an Electrospray Plume Stephen C. Gibson1 Charles S. Feigerle1 Kelsey D. Cook1,2*

1Department

of Chemistry, University of Tennessee, 552 Buehler Hall, 1420 Circle Dr.,

Knoxville, TN 37996-1600, United States 2National

Science Foundation, 4201 Wilson Blvd., Arlington, VA 22230, United States

1 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract The evolution of droplet temperatures in an electrospray plume was measured via ratiometric fluorescence. Under typical operating conditions, droplet temperatures decrease ~30 K over the first 5.0 mm along the spray axis, followed by a slight (~2-3 K) rewarming. Experimental axial profiles (Z-axis) were fit using diffusion-controlled and surface-controlled evaporation models. Both models fit the experimental data well for the cooling portion of the spray (Pearson correlation coefficient R ≥ 0.994), but the surfacecontrolled model required unrealistic droplet radius values to obtain a good fit. In lateral profiles at a given Z near the emitter tip, temperatures are lower (by 3.0 - 10 K) in the periphery than on the spray axis. This behavior is consistent with the expected enrichment of the spray periphery with smaller droplets. At longer axial distances, lateral profiles were relatively flat. Droplet temperature as a function of axial displacement fell more rapidly at lower liquid flow rates, possibly attributable to changes in droplet size and/or velocity with flow rate. Introduction To understand the relationship between ions formed in electrospray (ES) mass spectrometry and those present in bulk solution, it is necessary to consider spray-induced physical and chemical changes.1,2,3,4,5 For example, non-covalent interactions may be affected by pH and polarity changes that occur in evaporating ES droplets.6,7,8,9 In earlier studies, Zhao and Cook used polarity- and pH-sensitive fluorophores, respectively, to characterize solvent fractionation10 and electrochemically-induced pH changes11 in the ES plume via laser-induced fluorescence (LIF). Similar studies have been reported by Wang and Zenobi12 and by Antoine et al. ,13 who established a relationship between pH changes and the charge state distributions of peptide anions observed in the ES mass spectrum. Parks et al. also used LIF to study spray-induced conformational changes of cytochrome c.14 Temperature is another parameter that can affect solute chemistry either directly (e.g., affecting the conformation of biopolymers15) or indirectly (e.g., by affecting solvent evaporation and fractionation processes). Droplet cooling is recognized to have an effect on the kinetics and equilibria of reactants in the spray.8 Tang and Kebarle16 modeled 2 ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

electrospray-induced evaporative cooling of the droplets, but to our knowledge, there have been no experimental determinations of temperature profiles in an ES plume. Temperature can be measured using fluorophores with temperature-dependent emission wavelengths and/or intensities. For spray measurements, there is some advantage to relying on temperature-dependent changes in emission wavelengths17,18,19 because intensity changes also reflect changes in concentration as droplets evaporate and disperse. However, temperature-induced shifts in emission wavelengths are usually small. For example, in the present experiment Rhodamine B (RhB) showed only a 1.5 nm red shift in the emission maximum when heated from 283 K to 319 K. In contrast, the emission intensity of RhB decreases by 2.3% per in the same temperature range.20 This temperature sensitivity can best be exploited by using a temperature-insensitive internal standard and a ratiometric, two-color LIF thermometry approach.21 Rhodamine 110 (Rh110) is well characterized21,22 and relatively temperature insensitive (0.13% per K).21 It has been employed as a standard in the analysis of solution thermal gradients21 and in spatiallyresolved temperature measurements at the micrometer scale.23 In this study, the fluorescent temperature probe RhB is used with Rh110 as an internal standard to ratiometrically determine droplet temperatures at various positions along an electrospray plume. Further, the temperature profiles are fitted to evaporative models to provide insight into droplet characteristics and dynamics. Experimental Spectrometer Fluorescence spectra were acquired using a Raman/fluorescence spectrometer (HORIBA Jobin Yvon, Edison, NJ; model T64000) equipped with a 488 nm edge filter (to remove the Rayleigh line), a 600 groove/mm grating, and a 1024 x 256 open electrode Synapse CCD array detector. Data were acquired in the single spectrograph mode over the wavelength range 510 to 580 nm using the macrostage in a 180° “backscattering” geometry. The 488 nm line of an argon ion laser (Coherent Innova 200, Palo Alto, CA) was used for excitation.21 The laser cross section at the focal point was ~0.2 mm and the depth

3 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

of field was ~1.3 ± 0.2 mm. The laser power at the spray was typically ~25 mW (assessed with a model 3803 power meter; New Focus, Santa Clara, CA). Reagents RhB and Rh110 were used as received (Eastman Kodak Company, Rochester, NY; item #4453 and 11927, respectively). A solution containing 5.0 μM RhB and 0.50 μM Rh110 in methanol was prepared from separate ~0.5 mM dye stock solutions which were prepared fresh every 3-4 days to avoid degradation artifacts. The 10:1 mixture ratio was used in order to obtain similar emission intensities for the two dyes while avoiding detector saturation. Calibration The fluorescence emission intensity ratio of RhB (567 nm) to Rh110 (523 nm) was used as a measure of spray temperature. Calibration (see Fig. S-1 and S-2) was performed by placing an aliquot of the mixed dye solution in a thermostatted 1-cm glass cuvette (Starna, Atascadero, CA) positioned on the macrostage of the spectrometer in such a way as to maximize the RhB fluorescence signal. This resulted in the laser being focused near the front face of the cuvette. Temperature was maintained with a circulating water bath (Haake, Berlin, Germany; model F 4391). The temperature of the solution in the cuvette was monitored using a linear thermistor probe (Omega, Stamford, CT; OL-700 Series; calibrated at 273 and 373 K) and a DMM/scanner (Keithley, Cleveland, OH; model 199). For these calibration measurements, an OD = 2 neutral density filter was placed in the laser beam path before the sample to avoid saturation of the detector. Ten one-second acquisitions were signal averaged to improve the signal-to-noise ratio (S/N). Between each set of acquisitions, the laser shutter was closed to minimize photodegradation of the solution. Electrospray Apparatus The custom-built electrospray apparatus was similar to that previously described.10,11,24,25 Briefly, the apparatus consists of a 250 μm inner diameter (i.d.) x 500 μm outer diameter (o.d.) stainless steel inner capillary for sample introduction positioned 4 ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

coaxially within a 690 μm i.d. x 3.33 mm o.d. stainless steel capillary used to introduce ~327 K N2 nebulizing gas (Airgas Co., Knoxville, TN; industrial grade, 99.95%) at 1.6 L/min. The emitter end of the inner capillary was tapered with a diamond-coated file and lathe.26 This tapered tip extended ~1.5 mm beyond the end of the outer capillary. The two capillaries were supported by a stainless steel tee fitting (Valco, Houston, TX), which was heated to ~329 K and insulated with glass wool. The capillaries extended 1 cm beyond the tee and were not insulated. The temperature of the tee was measured with a thermocouple epoxied to the surface of the tee and monitored with a digital thermometer (Omega, Stamford, CT; HH-51 thermometer with an SA2C-K thermocouple). 4.0 kV was applied to the tee using a high voltage power supply (Gamma High Voltage Research, Ormand Beach, FL; model RR50-1.2SR/DDPM). The spray apparatus was electrically isolated from and mounted to a two-dimensional manipulator (L.S. Starrett Co., Athol, MA) enabling adjustment of the position of the spray relative to the optical path. To enhance thermal insulation and to reduce the impact of changes in ambient conditions, the apparatus was housed in an 18.5 x 16.0 x 7.5 cm purged enclosure made of transparent 2-mm thick poly(methyl methacrylate). The spray was operated at least thirty minutes prior to taking data, providing methanol vapor and nitrogen gas sufficient to flush the purged enclosure ~20 times while displacing ambient air. The sample solution was introduced into the inner capillary via a syringe pump (Harvard Apparatus, Holliston, MA; model 11). A 2.54 x 2.54 cm2 stainless steel plate (Kimball Physics, Wilton, NH; SS-PL-C7x7-B) was oriented orthogonally to the spray axis ~15 mm downstream from the electrospray tip and served as a counterelectrode. It was connected to ground through a picoammeter (Keithley, Cleveland, OH; model 417) which was used to monitor spray current. Spray Profiles and Data Acquisition To generate profiles of the spray plume, the electrospray apparatus was translated parallel (Z-axis; axial profiles) or orthogonal (Y-axis; lateral profiles) to the (vertical, downward) spray axis while the laser remained fixed along the X-axis (see Fig. 1). Ten 10-

5 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

second acquisitions were averaged for each data point. Each data point was taken in triplicate; plotted error bars represent one standard deviation (σ) for these replicates. Spectra for axial profiles at Y = 0 mm were acquired in random sequence at 14 different Z-axis positions between 0.25 mm and 12 mm from the emitter tip. At Z > 12 mm, the signal intensities became erratic, probably due to interaction between the laser and deposited solid and/or liquid on the counterelectrode. Temperatures are therefore reported only for distances < 12 mm. Lateral profiles were obtained at four different Z-axis positions with each profile including up to seven different Y values taken in random sequence. In each lateral profile, the maximum |Y| value was that for which the S/N fell to ≤ 10. The number of points therefore varies among profiles at different Z. Mass Spectrometry Electrospray mass spectra of the dye mixture solution were acquired using a JEOL Model JMS-T100LC (AccuTOF) orthogonal time-of-flight mass spectrometer (Peabody, MA). The needle, orifice 1, and ion guide peak voltage were 2500, 30, and 800 V, respectively. The desolvating chamber temperature was 573 K. The sample was delivered by flow injection (Rheodyne, Rohnert Park, CA; model 7125) of a 10 μL aliquot at 50 μL/min using a syringe pump (Harvard Apparatus, Holliston, MA; model 55-2222). Modeling Calculations were performed using Microsoft Excel 2010 (Redmond, WA). Least squares fits were created via successive approximations by optimizing one parameter at a time. Although the pitfalls of this approach were recognized, arriving at the same optimized parameter values even when optimizing parameters in different sequences increased confidence of discovering the global minimum. Uncertainties for fitting parameters representing one standard deviation (σ) were determined from the experimental pooled standard deviation and the sum squared error (i.e., the squared difference between the experimental data and the model predictions), as described in

6 ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

Reference 27. Pearson correlation coefficients (R) were generated via OriginPro 8.1 (OriginLab, Northampton, MA). Results and Discussion Axial temperature profiles of ES plumes created using various solvent flow rates (F = 12.5, 25, and 50 μL/min) are shown in Fig. 2 (see Fig. S-3 for spectra used to construct the profile at F = 25 μL/min). There is a significant decrease in temperature from the spray tip to approximately Z = 3 - 5 mm followed by slight reheating. The initial cooling qualitatively resembles that modeled by Kebarle16 and is attributable to endothermic solvent evaporation. Reheating at larger Z may be attributed to a thermal competition whereby the evaporative cooling process (slower at lower temperatures) is overtaken by collisional warming by ambient background and/or heated nebulizing gas. At larger Z, the droplets are expected to be smaller, with correspondingly higher surface-to-volume ratio and lower heat capacity, potentially enhancing the warming effect of collisions. Collisional frequency may also be enhanced at larger Z by turbulence from gas reflection off the counterelectrode; this may contribute to droplet reheating. Further study (outside the scope of this work) is required to understand the mechanism of rewarming. Two differences are evident at lower F (Fig. 2): a slightly steeper initial temperature drop followed by greater reheating. Several factors likely contribute, relating to droplet size, velocity, and environment. Among the factors affecting the rate of solvent evaporation is the partial pressure of solvent vapor in the surrounding environment, designated P∞. Employment of the purged enclosure to contain the spray apparatus allows estimation of P∞ from the assumptions that 1) the total ambient pressure is roughly 760 Torr; 2) the mole ratio of carrier gas to solvent is approximately the same as the mole ratio of their flow rates; and 3) ambient air is swept from the box during the 30-minute pre-measurement equilibration period. P∞ calculated from these assumptions is approximately 4, 8 and 16 Torr at F = 12.5, 25 and 50 µL/min, respectively, considerably less than the saturated methanol vapor pressure of ~100 Torr at room temperature. Based solely on this trend, evaporation should be enhanced at lower F, consistent with the observed enhancement of initial cooling. 7 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Complex effects of F on droplet velocity and size28 may also contribute to the trends evident in Fig. 2. The liquid flow velocity (vF; calculated from F and the capillary i.d.) decreases from 1.7 to 0.42 cm/s as F decreases from 50.0 to 12.5 µL/min. These values are small compared with droplet velocities measured elsewhere by Doppler phase anemometry29 (~2-3 m/s), suggesting at most a small contribution, albeit in a direction consistent with the trends of Fig. 2. A more significant effect may derive from the interplay between F and the measured spray current (I). From the data in Table 1, charge density within the droplets (proportional to I/F) would appear to increase slightly from F = 12.5 to 25 µL/min, then fall as F is increased to 50 µL/min. The initial increase is unexpected and may result from the uncertainty in measuring small I; the decrease from 25 to 50 µL/min is consistent with predictions of de la Mora et al.28 Charge density would be expected to affect both the size of the droplets (smaller at higher I/F) and their acceleration by the extracting field and by collisions with the nebulizing gas28,30,31 (more acceleration for smaller and more highly charged droplets). There may be offsetting effects: smaller droplets evaporate more quickly, but their higher velocity gives them less time to evaporate before reaching a given position. Further study, including direct experimental measurement of droplet sizes and velocities (again, beyond the scope of this study), is necessary to unambiguously resolve these contributions, although some insight may be derived from modeling the cooling process (see below). Lateral profiles obtained at various axial positions (Z = 0.25, 1, 4, and 7 mm) are displayed in Fig. 3. On-axis (Y = 0 mm) temperatures in these profiles generally follow trends presented in Fig. 2. At Z = 0.25 and 1 mm, temperatures decrease sharply as the distance from the Z-axis increases. Smaller droplets are preferentially driven to the periphery of the electrospray due to space charge32 and simple dispersion effects. The higher surface-to-volume ratio of smaller droplets promotes faster evaporative cooling, consistent with lower temperatures off axis. Downstream lateral profiles (Z = 4 and 7 mm) are relatively flat, suggesting that thermal equilibration and/or reduced polydispersity has occurred by this point. These profiles extend to higher Y values than those nearer to the emitter; the dispersion of the spray plume provides measurable signals farther into the periphery at higher Z, although the error bars in Fig. 3 become large at high Z and Y. 8 ACS Paragon Plus Environment

Page 8 of 28

Page 9 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

Fig. 4 illustrates the impact of F on lateral profiles. These data are consistent with axial profile comparisons where higher F generally yields warmer temperatures at a given Z position. Data at Z = 4 mm (not shown) are analogous. Lateral profiles at Z = 0.25 and 1 mm (Fig. 4a and 4b, respectively) also reflect a physically broader spray plume at higher F (i.e., S/N > 10 at larger |Y|), consistent with the greater amount of material present in the plume. Modeling Evaporation of airborne droplets may be either diffusion-controlled (when the ratelimiting process is diffusion of vapor away from the droplet in the gas phase) or surfacecontrolled (for droplets too small to saturate the droplet’s surface vapor layer).33 The critical radius (rc) for transition between these modes is defined as: ∙

  ∙

(1)

where λ is the mean free path of vapor molecules in the gas around the droplet (~ 68 nm at 298.15 K and 760 Torr, using a 36.9-nm hard sphere diameter for methanol34) and α is the condensation coefficient (temperature-dependent probability of molecule capture upon collision with the droplet; α ~ 0.9 at 300 K for methanol35). The critical radius for methanol is thus about 0.10 μm under conditions used here. By contrast, droplet size measurements29,32,36,37,38 indicate that the initial droplet radius r0 should be on the order of 4.0 to 40 µm for conventional electrospray. From this simple consideration, evaporation should initially be diffusion-controlled, possibly transitioning to surface control later in the spray. Tang and Kebarle16 were the first to model cooling during droplet evaporation in an electrospray. They considered surface-controlled evaporation of droplets of r0 = 10.0 nm, and estimated that such a droplet would cool to a temperature 8 K below that of its surroundings (306 K to 298 K) in about 8 μs. This small radius would apply to the final stages of residue or offspring droplet evolution, nearing complete evaporation. Observation times corresponding to the distances presented in Fig. 2 occur much earlier in the droplet lifetime and cover timespans much longer than those modeled by Kebarle. For 9 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 28

example, assuming a constant velocity of 2 m/s (as in Reference 29), the Z-axis (0-12 mm) of Fig. 1 may be estimated to correspond to a timespan of 6 ms. Hence, it is of interest to ascertain whether the present data can be adequately described by the surface-controlled evaporation model used by Kebarle. Specifically, modeling of the evaporation process was undertaken to gain further insight into the cooling portion of the spray (Z = 0.25 to 5.00 mm in Fig. 2). As noted above, the re-warming at Z > 5.0 mm cannot be ascribed to simple evaporation and thus is beyond the scope of simple evaporation models. Droplet cooling was calculated using models adapted from Ref. 33 for both diffusion-controlled (DCM; Eqn. 2) and surface-controlled (SCM; Eqn. 3) evaporation:  !∙"∙# ̅ (  ' ) %&  



,      ∙  ∙ ∙  

, 12   

  ∙ ∙ 



 ∙-∙"∙# ̅ (   '  ∙%& )

 )* +  ,- .  /0 (

*

(

 )* +  *

3 4 .  

(2)

 /50 (3)

where δTd is the change in a droplet’s temperature over an incremental axial distance δZ for a droplet moving at velocity vd (δZ/vd = δt, the time increment used in Davies’ model33); the other terms are defined in Table 2. In applying these equations, δZ was selected to provide 1000 calculated points between the most closely spaced experimental data points (δZ = 0.25 μm). Eqn. 2 and 3 apply to pure liquids, but since solute concentrations used were on the order of micromolar, colligative effects were assumed to be negligible. The first measured experimental temperature at Z = 0.25 mm, Y = 0 mm was presumed accurate and provided the initial value of droplet temperature (Td,0), one boundary condition. If the droplet radius at this position is designated r0, rd at a given position (Z) (i.e. rd(Z)) may be calculated from Eqn. 4 for DCM evaporation33 or Eqn. 5 for SCM evaporation:16 8∙!∙"

 /   678  9  ∙% ∙ '  /12  7 

(  )

(



 )* + ∙  : *

∙-∙( ∙"  ∙ ∙%∙)

∙





(4) 5

where r0 = radius at Z = 0.25 mm, and other terms are defined in Table 2. 10 ACS Paragon Plus Environment

Page 11 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

For the temperature-dependent parameters ∇, l, and k, Davies33 recommends using the geometric mean of values at the surface and at ambient temperature. For the diffusion coefficient,∇(T), the needed values can be estimated from Eqn. 6:39 ) =.?=

    3  ∙ 9 :

(6)

)
5 mm.

16 ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

Further corroboration of the minimal impact of these artifacts comes from the observation that the measured fluorescence intensity ratios (and the derived temperature) at the plateaus in the axial profiles (Z = 3 – 6 mm in Fig. 2) are similar for all three flow rates despite differences in modeled droplet sizes and (presumably) dye and [H+] concentrations. Additionally, the largest measured temperature changes occur over regions (Z = 0.25 – 1.0 mm) where there is a relatively small change in droplet radii, providing additional confidence in calibrations. Conclusions The total temperature change predicted by these models is up to 30 K. This is much larger than the 8 K change calculated by Tang and Kebarle16 (who modeled offspring droplets of initial radius 10 nm using the SCM), but should provide a better estimate of the total temperature excursion to which ES solutes may be subjected. Both the SCM and DCM models fit the experimental data very well and yield reasonable values for P∞ and vd (or r02∙vd for the DCM). The value of r0 needed to fit these data with the SCM far exceeds the critical radius, rc, contrary to the conditions for evaporation to be surface-controlled. This suggests that evaporation in the region sampled in this experiment is diffusion-controlled. This conclusion would be further strengthened by independent experimental measurement of droplet velocity and radius – beyond the scope of the current study. Regardless of model, the amplitude of ΔT is clearly large enough to significantly impact some chemical equilibria if the kinetics are fast enough for changes on the timescale of the droplet lifetime. Acknowledgements S.C.G. would like to thank Eric Barrowclough for assistance in a preparation of a figure. Participation by K.D.C. while at the National Science Foundation was supported through the NSF Independent Research and Development program. Programming assistance from and discussions with Dr. Frank Vogt are gratefully acknowledged.

Supporting Information Available: This material is available free of charge via the Internet at http://pubs.acs.org. 17 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 28

Purged Enclosure Outer Capillary Inner Capillary Electrospray Plume Y (Lateral)

X (Laser)

Z (Axial)

488 nm Laser

Counterelectrode Vent

Figure 1: Graphic representation of the sampling system.

18 ACS Paragon Plus Environment

Page 19 of 28

300

F = 12.5 µL/min F = 25.0 µL/min F = 50.0 µL/min

295

Temperature (K)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

290 285 280 275 0

2

4

6

8

10

12

Axial (Z) Displacement (mm) Figure 2: Axial temperature profiles through the center of the spray (Y = 0) at indicated flow rates.

19 ACS Paragon Plus Environment

Analytical Chemistry

Z = 0.25 mm Z = 1.0 mm Z = 4.0 mm Z = 7.0 mm

295 290

Temperature (K)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

285 280 275 270

-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

Lateral (Y) Displacement (mm) Figure 3: Lateral temperature profiles at indicated axial (Z) positions each at 25 µL/min.

20 ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28

F = 12.5 µL/min F = 25.0 µL/min F = 50.0 µL/min

a

Temperature (K)

295

290

285

280 -0.3

-0.2

-0.1

0.0

0.1

0.2

0.3

Lateral (Y) Displacement (mm)

285

Temperature (K)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

-0.6

b

280

275

F = 12.5 µL/min F = 25.0 µL/min F = 50.0 µL/min

270

265 -0.4

-0.2

0.0

0.2

0.4

0.6

Lateral (Y) Displacement (mm) Figure 4: Lateral temperature profiles at indicated flow rates and a) Z = 0.25 mm and b) Z = 1 mm.

21 ACS Paragon Plus Environment

Analytical Chemistry

295

Temperature (K)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Experimental Data Diffusion-Controlled Model Surface-Controlled Model

290

285

280

275 0

2

4

6

8

10

12

Axial (Z) Displacement (mm) Figure 5: Comparison of experimental data and optimized model at F = 25 µL/min. Curves for the diffusion-controlled and surface-controlled models are indistinguishable.

22 ACS Paragon Plus Environment

Page 22 of 28

Page 23 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Analytical Chemistry

Table 1: Experimental and calculated parameter values for surface-controlled and diffusion-controlled models of droplet cooling. Parameter Solvent Flow Rate (F; μL/min)1 Emission Current (I; μA)1 Methanol Partial Pressure (P∞; Torr)2 Droplet Velocity (vd; m/s)3 Droplet Radius (rd(Z); μm) at Z = 04

Surface-Controlled Model (SCM) 12.5 25.0 50.0 0.055 0.12 0.19 12.9 ± 1.3 12.6 ± 0.9 20.0 ± 0.6 0.52 ± 0.05 0.60 ± 0.05 0.63 ± 0.03 11.9 ± 0.6 13.5 ± 0.8 15.5 ± 0.4

Diffusion-Controlled Model (DCM) 12.5 25.0 50.0 0.055 0.12 0.19 19.5 ± 1.2 15.7 ± 1.3 19.5 ± 0.6 2.0 2.0 2.0 6.7 ± 0.3 7.9 ± 0.3 8.8 ± 0.1

Droplet Radius (r0; μm) at Z = 0.252

11.7 ± 0.6

13.4 ± 0.8

15.4 ± 0.5

6.7 ± 0.3

7.8 ± 0.3

8.7 ± 0.1

Droplet Radius (rd(Z); μm) at Z = 54

10.9 ± 0.6

12.6 ± 0.8

14.5 ± 0.4

6.4 ± 0.3

7.5 ± 0.3

8. 5 ± 0.1

Droplet Radius (rd(Z); μm) at Z = 124 9.7 ± 0.6 11.6 ± 0.8 13.5 ± 0.4 6.0 ± 0.3 7.1 ± 0.3 8.2 ± 0.1 Droplet Temperature (Td; K) at Z = 04 302.0 ± 0.8 306.7 ± 0.8 305.3 ± 0.8 302.7 ± 0.8 306.7 ± 0.8 305.1 ± 0.8 Droplet Temperature (Td,0; K) at Z = 0.251 290.8 ± 0.8 294.7 ± 0.8 295.9 ± 0.8 290.8 ± 0.8 294.7 ± 0.8 295.9 ± 0.8 Droplet Temperature (Tf; K) at Z = 124 276.9 ± 0.8 275.7 ± 0.8 277.0 ± 0.8 276.9 ± 0.8 275.7 ± 0.8 277.0 ± 0.8 ΔT (K) (Z = 0.25 to 12 mm)4 13.9 19.0 18.9 13.9 19.0 18.9 ΔT (K) (Z = 0 to 12 mm)4 25 31 28 26 31 28 Pearson Correlation Coefficient (R)5 0.994 0.998 0.998 0.994 0.998 0.998 Pooled Standard Deviation5 0.76 0.83 0.35 0.76 0.83 0.35 1Experimental measurement 2Fitting parameter 3Fitting parameter for SCM; fixed value for DCM (see text) 4Calculated from model 5For fit of model to the experimental data between Z = 0.25 and 5.0 mm

23 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 28

Table 2. Parameters used in this work, in rough order of appearance. Parameter Definition F Liquid flow rate P∞ Partial vapor pressure of solvent in the surrounding gas vF Liquid flow velocity I Spray current rc Critical radius

λ α r0

Gas mean free path Condensation coefficient Droplet radius at Z = 0.25 mm

δTd

Change in droplet temperature Td in incremental distance δZ Incremental distance

δZ

vd δt rd(Z) ρ’ s’ ! M @̅

Rg M̅

Droplet velocity Evaporation time increment Droplet radius at axial position Z Solvent density Solvent specific heat Diffusion coefficient of solvent vapor in air Solvent molar mass Latent heat of solvent vaporization

Gas constant Mean speed of gaseous solvent

Value

Comments Fitting parameter

Experimentally measured. Radius at which transition between diffusion- and surfacecontrolled models occurs. 0.90

0.25 μm

For methanol at 300 K (Ref. 35). Fitting parameter; Expected values range from 4 - 40 µm (Refs. 29,32,36,37,38). Calculated using Eq. 2 or 3

Optimized to prevent round-off error while preserving sensitivity of parameters. Fitting parameter

δZ/vd Calculated using Eq. 4 or 5. 0.7914 g/mL 2.531 J/g·K 6∇ )* ∙ ∇ ) 32.04 g/mol 6@ )* ∙ @ )

8.314 J/mol·K 445.2 m/s

(Ref. 46) (Ref. 46) ∇ )* and ∇ ) are calculated using Eqn. 6 (Ref. 39). @ )* and @ ) are calculated via Eqn. 7. For methanol, A = 45.3, α = 0.3100 , β = 0.4241 (Ref. 42). The reduced temperature Tr = T/Tc, where the critical temperature Tc of methanol = 512.6 K. For methanol at 300 K; calculated from Ref. 46.

24 ACS Paragon Plus Environment

Page 25 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

Pd(Z)

,k’

Solvent vapor pressure at the droplet surface

log Pd = A (B/(C + Td)) Thermal conductivity of air Thermal conductivity of liquid methanol

T∞

Ambient temperature

Td

Droplet surface temperature

Tf

Final droplet temperature Time for temperature equilibration in a droplet Wavelength of maximum intensity or absorbance

teq

λmax

106.5 Torr

6, )* ∙ , ) 0.203 W/m·K

303 ± 1 K

Calculated via the Antoine equation. For methanol, A = 8.08097, B = 1582.27, C = 239.7 (Ref. 47). , )* and ,) are calculated using interpolation from Ref. 43. For methanol at 300 K; Calculated by interpolation from Ref. 43. Average temperature and standard deviation from measurements at six positions inside the purged enclosure. Td,0 = measured temperature at Z = 0.25, Y = 0 mm at each flow rate (See Table 1); Subsequent values calculated by adding δTd from Eqn. 2 or 3 to the preceding value. Model-derived temperature at Z = 12 mm. Calculated using Eq. 8.

25 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) (19) (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) (30) (31) (32)

Kebarle, P. J. Mass Spectrom 2000, 35, 804–817. Kiselev, P.; Rosell, J.; Fenn, J. B. Ind. Eng. Chem. Res. 1997, 36, 3081–3084. Zhou, S.; Cook, K. D. J. Am. Soc. Mass Spectrom. 2001, 12, 206–214. Fenn, J. B. Ion formation from charged droplets: roles of geometry, energy, and time. J. Am. Soc. Mass Spectrom. 1993, 4, 524–535. Pramanik, B. N.; Ganguly, A. K.; Gross, M. L. Applied Electrospray Mass Spectrometry; CRC Press, 2002. Mansoori, B. A.; Volmer, D. A.; Boyd, R. K. J. Rapid Commun. Mass Spectrom. 1997, 11, 1120–1130. Wang, H.; Agnes, G. R. Anal. Chem. 1999, 71, 3785–3792. Wang, H.; Agnes, G. R. Anal. Chem. 1999, 71, 4166–4172. Gatlin, C. L.; Turecek, F.; Vaisar, T. Anal. Chem. 1994, 66, 3950–3958. Zhou, S.; Cook, K. D. Anal. Chem. 2000, 72, 963–969. Zhou, S.; Prebyl, B. S.; Cook, K. D. 2002, 74, 4885–4888. Wang, R.; Zenobi, R. J. Am. Soc. Mass Spectrom. 2010, 21, 378–385. Girod, M.; Dagany, X.; Antoine, R.; Dugourd, P. Int. J. Mass Spectrom. 2011, 308, 41–48. Rodriguez-Cruz, S. E.; Khoury, J. T.; Parks, J. H. J. Am. Soc. Mass Spectrom. 2001, 12, 716–725. Grandori, R.; Matecko, I.; Müller, N. J. Mass Spectrom. 2002, 37, 191–196. Tang, L.; Kebarle, P. Anal. Chem. 1993, 65, 3654–3668. Schrum, K. F.; Williams, A. M.; Haerther, S. A.; Ben-Amotz, D. Anal. Chem. 1994, 66, 2788–2790. Lou, J.; Hatton, T. A.; Laibinis, P. E. Anal. Chem. 1997, 69, 1262–1264. Figueroa, I. D.; El Baraka, M.; Quiñones, E.; Rosario, O.; Deumié, M. A. Anal. Chem. 1998, 70, 3974–3977. Gallery, J.; Gouterman, M.; Callis, J.; Khalil, G.; McLachlan, B.; Bell, J. Rev. Sci. Instrum. 1994, 65, 712–720. Sakakibara, J.; Adrian, R. J. Exp. Fluids 1999, 26, 7–15. Coppeta, J.; Rogers, C. Exp. Fluids 1998, 25, 1–15. Kim, H. J.; Kihm, K. D.; Allen, J. S. Int. J. Heat Mass Tran. 2003, 46, 3967–3974. Zhou, S.; Edwards, A. G.; Cook, K. D.; Van Berkel, G. J. Anal. Chem. 1999, 71, 769–776. Davis, D.; Portelius, E.; Zhu, Y.; Feigerle, C.; Cook, K. D. Anal. Chem. 2005, 77, 8151– 8154. Chen, M.; Cook, K. D. Anal. Chem. 2007, 79, 2031–2036. Bevington, P. R. Data Reduction and Error Analysis for the Physical Sciences; McGrawHill: New York, 1969. De La Mora, J. F.; Loscertales, I. G. J. Fluid Mech. 1994, 260, 155–184. Nemes, P.; Marginean, I.; Vertes, A. Anal. Chem. 2007, 79, 3105–3116. Hinds, W. C. Aerosol Technology; 2nd ed.; John Wiler and Sons, Inc.: New York, 1999. Wang, R.; Allmendinger, P.; Zhu, L.; Gröhn, A. J.; Wegner, K.; Frankevich, V.; Zenobi, R. J. Am. Soc. Mass Spectrom. 2011, 22, 1234–1241. Olumee, Z.; Callahan, J. H.; Vertes, A. J. Phys. Chem. A 1998, 102, 9154–9160. 26 ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

(33) Davies, C. N. Evaporation of Airborne Droplets. In Fundamentals of Aerosol Science; Shaw, D. T., Ed.; Wiley-Interscience: New York, 1978; pp. 135–164. (34) Wilhelm, E.; Battino, R. Estimation of Lennard-Jones (6,12) J. Chem. Phys. 1971, 55, 4012–4017. (35) Mikami, S.; Kobayashi, K.; Ota, T.; Fujikawa, S.; Yano, T.; Ichijo, M. Exp. Therm. Fluid Sci. 2006, 30, 795–800. (36) Olumee, Z.; Callahan, J. H.; Vertes, A. Anal. Chem. 71, 4111–4113. (37) De Juan, L.; De la Mora, J. F. J. Colloid. Interf Sci. 1997, 186, 280–293. (38) Electrospray and MALDI Mass Spectrometry: Fundamentals, Instrumentation, Practicalities, and Biological Applications; Cole, R. B., Ed.; 2nd ed.; Wiley: Hoboken, NJ, 2010. (39) Poling, B. E.; Prausnitz, J. M.; O’Connell, J. P. The Properties of Gases and Liquids; McGraw-Hill Prof Med/Tech, 2001. (40) Winkelmann, J. Diffusion of methanol (1); air (2). In Data extract from of LandoltBörnstein IV/15A, Chapter 5.1.1: Diffusion of Gas/Vapor in Gas; Lechner, M. D., Ed.; Landolt-Börnstein - Group IV Physical Chemistry; Springer-Verlag Berlin Heidelberg, 2007; Vol. 15A. (41) Chen, N. H.; Othmer, D. F. J. Chem. Eng. Data 1962, 7, 37–41. (42) Enthalpies of Vaporization of Organic Compounds; Majer, V.; Svoboda, V., Eds.; 1st ed.; Blackwell Scientific Publications: Oxford, 1985. (43) Vargaftik, N. B.; Filippov, L. P.; Tarzimanov, A. A.; Totskii, E. E. Handbook of Thermal Conductivity of Liquids and Gases; CRC Press, 1993. (44) Chamarthy, P.; Garimella, S. V.; Wereley, S. T. International Journal of Heat and Mass Transfer 2010, 53, 3275–3283. (45) Novák, A. Collection of Czechoslovak Chemical Communications 1978, 43, 2869–2878. (46) CRC Handbook of Chemistry and Physics; Lide, D. R., Ed.; 86th ed.; CRC Press: Boca Raton, FL, 2005. (47) NIST Chemistry Web Book. Available at: http://webbook.nist.gov/chemistry/. Accessed 18 September, 2012.

27 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

For Table of Contents only:

28 ACS Paragon Plus Environment

Page 28 of 28