You are on page 1of 11

Journal of Aerosol Science 114 (2017) 233–243

Contents lists available at ScienceDirect

Journal of Aerosol Science


journal homepage: www.elsevier.com/locate/jaerosci

An electrospray in a gaseous crossflow MARK



Elyar Allaf-Akbari, Nasser Ashgriz
Department of Mechanical and Industrial Engineering, University of Toronto, Toronto, Ontario, CanadaM5S 3G8

AR TI CLE I NF O AB S T R A CT

Keywords: An experimental study on the dynamics of an electrospray (ES) injected into a test chamber with
Electrospray a gaseous crossflow is presented. The study is relevant to the ion introduction into mass spec-
Crossflow trometers that use ES as the ion source. Bending and trajectory of the ES plume is determined for
Plume a range of air crossflows and for different ES injection locations with respect to the crossflow
Cone jet
centerline. It is shown that the ES injection conditions correlate with the intensity detections of a
Hysteresis
Mass spectrometry
mass spectrometer. The effect of crossflow on the cone jet and ES performance are also discussed.
A new hysteresis phenomenon for the electrospray in crossflow is detected.

1. Introduction

Electrosprays have a broad range of applications, such as in MEMS and microfluidics (Chiarot, Sullivan, & Ben Mrad, 2011),
emulsion production (Barrero & Loscertales, 2007), fuel injection (Deng, Klemic, Li, Reed, & Gomez, 2007; Kyritsis, Coriton, Faure,
Subir, & Gomez, 2004), etc. (Bailey, 1988). Their most important application, however, is to generate ions for mass spectrometry
(MS) (Cole, 1997; Fenn, Mann, Meng, Wong, & Whitehouse, 1989; Pramanik, Ganguly, & Gross, 2002). Therefore, there are numerous
investigations on various aspects of their operation, such as the operating modes. Cloupeau and Prunet-Foch (1990), Jaworek and
Krupa (1999), and Sultan, Ashgriz, Guildenbecher, and Sojka (2011), have reviewed different operating modes of ES. Some of these
modes are: dripping, spindle, cone jet, and multi-jet. In the cone jet mode, a conical liquid meniscus is formed at the tip of ES nozzle,
emitting a charged jet from its apex (Smith, 1986; Taylor, 1964; Zeleny, 1917). This jet later breaks into fine charged droplets, which
form a conical plume due to the repulsion among the droplets, i.e., space charge. Valaskovic, Murphy, and Lee (2004) and Nguyen
et al. (2014) tried to improve the ES performance in the cone jet mode by monitoring the ES plume, and cone, respectively. Park, Kim,
and Kim (2004) studied the effect of different solution flow rates and guard plates, i.e., different equipotential fields, on the plume
angle and on the onset voltage of cone jet mode. Wang, Tan, Go, and Chang (2012) investigated the effect of surface tension and ionic
strength of the spraying solution on the plume angle and jet breakup length. Dynamics of droplets and the shape of the plume in
stagnant gaseous environment have been simulated Wilhelm, Mädler, and Pratsinis (2003), and Grifoll and Rosell-Llompart (2012).
For detailed discussion on the modelling of ES plumes, the reader is referred to Arumugham-Achari, Grifoll, and Rosell-Llompart
(2015) and the references therein.
Studies on the formation and dynamics of electrosprays in a gaseous flow are limited. Gañán-Calvo, Lopez-Herrera, and Riesco-
Chueca (2006) mixed the flow focusing technique with electrospraying and achieved smaller droplets and more stable spraying, in
comparison to individual flow focusing or ES. They also discussed the jet instability and size, both theoretically and experimentally
(Gañán-Calvo & Montanero, 2009; Gañán-Calvo, 2007). Another study is provided by Sultan (2013), who considered the effect of a
nebulizing or a co-axial airflow on spindle, cone jet, and unstable modes of an ES. Spray characteristics, such as size and velocity
distributions, were determined using a PDPA system for both no-air and with-air flow conditions. Savtchenko, Ashgriz, Jolliffe,


Corresponding author.
E-mail address: ashgriz@mie.utoronto.ca (N. Ashgriz).

http://dx.doi.org/10.1016/j.jaerosci.2017.09.021
Received 19 October 2016; Received in revised form 11 September 2017; Accepted 19 September 2017
Available online 22 September 2017
0021-8502/ © 2017 Elsevier Ltd. All rights reserved.
E. Allaf-Akbari, N. Ashgriz Journal of Aerosol Science 114 (2017) 233–243

Cousins, and Gamble (2014) numerically compared the traveling path and the evaporation of droplets generated by an ES with a co-
axial air flow. Wang et al. (2011) studied the effect of high speed nebulizer gas, varying from subsonic to supersonic, on electrosonic
spray ionization (ESSI) using ESSI-MS. They also simulated the flow pattern of the nebulizing air with respect to the ES plume and
theoretically discussed the droplet acceleration and breakup conditions. Fomina, Masyukevich, Sviridovich, and Gall (2014) tested a
polycapillary ES source with a carrier air in order to produce and transport charged droplets at high liquid flow rates, which can be
applied for concentration detection of toxic material in the air. Holbrook, Hindle, and Longest (2015) used an ES with a crossflow in
order to improve the respiratory drug delivery to infants.
To the best of our knowledge, there have been no investigation on the interaction between an ES and a gaseous crossflow.
However, many researchers have focused on the behavior of liquid spray and jet in crossflow (JICF), mostly related to combustion
problems. For a review on JICF, the reader is referred to Eslamian, Amighi, and Ashgriz (2014), Mashayek and Ashgriz (2011) and the
references therein. Ghosh and Hunt (1998) studied the interaction of the weak to moderate crossflows with a spray, considering the
induced airflow by the spray. The induced air can change the crossflow pattern, especially in the region close to the nozzle tip. Based
on the crossflow velocity, they developed different theoretical models for the behavior of the spray in crossflow. Deshpande, Gao, and
Trujillo (2011) simulated the hollow cone sprays in crossflow. Similar to Ghosh and Hunt (1998), they discussed the spray behavior
with respect to the distance from an emitter.
ESs have three main differences with the uncharged jets and sprays reported in the literature: (i) In an ES, the liquid flow rate and
droplet sizes are usually much smaller than regular sprays. (ii) The spray profiles and planar size distributions are not alike, because
the atomization processes are completely different. (iii) In addition to the air drag, the external electric field and space charge affect
the motion of electrosprayed charged droplets. Therefore, the results of the regular sprays in crossflows cannot directly be used in ES
in crossflows.
The focus of this study is on the ES behavior in an air crossflow inside a test section. Similar to JICF, we have adopted ESICF for
“ES In CrossFlow”. Our test section is essentially a chamber with an air inlet and an outlet for the crossflow gas. The crossflow is used
to carry the generated ions towards an analyzer. The images of ES plume at different crossflows are shown and characterized. Then,
the observation of plume bending is matched and justified with MS intensity results. The effects of crossflow and voltage on the
Taylor cone are discussed and different operating modes of ES are reported. The importance of the nozzle tip wetting on the per-
formance of ES is examined by changing the sequence of voltage, liquid, and crossflow inputs. Lastly, a hysteresis effect in ESICF is
reported.

2. Experimental setup and methods

An experimental setup was built to visualize the ES performance inside a test section. Fig. 1 shows the test section, which has four
parts: (I) a gas inlet with sudden expansion; (II) a chamber; (III) a gas outlet with a tapered contraction, and (IV) an inlet for ES
sprayer. A copper plate was attached to the bottom of the chamber and used as the ground electrode. The ES nozzle operates inside
this chamber and the air carries the generated ions to the mass spectrometer through the outlet. A similar design was used in MS,
which provided better ion detections (Jolliffe, Javahery, Cousins, & Savtchenko, 2010). The advantages of this ion source over the
sources with co-axial gas, are in coupling the ES with the sampling orifice of a spectrometer and a better ion desolvation. In the
sources with co-axial flow, the location of ES with respect to the sampling orifice affects the signal to noise ratio, sensitivity, and

Fig. 1. (a) The test (desolvation) chamber, and (b) a typical airflow pattern inside the chamber with sudden expansion (similar to Forrester & Evans, 1997).

234
E. Allaf-Akbari, N. Ashgriz Journal of Aerosol Science 114 (2017) 233–243

Table 1
Crossflow air categories and the corresponding flow rates, velocities and Reynolds numbers.

Crossflow Flow Rate [Cfh] a


Pressureb [kPa] Flow Rate [Scfh] a
Flow Rate [Slpm] Velocityc [m/ s] Reynoldsc

Low 2–10 0 2–10 0.95–4.72 3.8−18.9 576–2882


Medium 10–20 0 10–20 4.72–9.44 18.9–37.9 2882–5764
High >20 >0 >20 >9.44 >37.9 >5764

a
Cfh: cubic feet per hour, equivalent to 0.472 L/min or Lpm, S: standard.
b
Room pressure: 99.32 kPa.
c
Calculated based on standard values and inlet diameter.

stability. These sources usually need additional adjustments (Cole, 1997; Manisali, Chen, & Schneider, 2006). Besides, the generated
droplets do not completely break into single ions, because of insufficient droplet flight time. The present desolvation chamber, due to
its geometry, does not need additional adjustments and keeps the ions focused at its outlet. It also provides enough time for ion
desolvation, before entering the analyzer (Jolliffe et al., 2010).
Two chamber dimensions with square cross sections were constructed for our tests: 16 × 16 × 19mm3 for plume study (Sections
3.1 and 3.2) and 10 × 10 × 19mm3 for cone jet study (Sections 3.3 and 3.4). The gas inlet diameter of chamber was 2.3 mm and its
outlet diameter was 2.4 mm with a 45° tapered contraction. Dry and filtered compressed air was used as the carrier gas, and its flow
rate was controlled with rotameters and a pressure gauge. The flow measurement error was less than 5%. A summary of the crossflow
conditions is provided in Table 1.
OD and ID of the ES nozzle are 150 µm and 300 µm, respectively. The nozzle was located 5 mm from the left wall of chamber, i.e.,
the air inlet. A syringe pump (Series 74900 Cole Parmer) operating between 0.01 and 210 ml/h with 0.01 ml/h precision, supplied
the liquid to a capillary tube. A high voltage DC power supply (Series ER, Glassman High Voltage Inc.) applied the required potential
to the ES emitter, up to +10 kV. A high voltage probe with the conversion factor of 1000 was used to measure the applied voltage
with less than 2% error. A 75/25 volumetric solution of methanol/water was selected as the operating liquid and its flow rate was set
to 1–2 µL/min. The density, viscosity, and surface tension of the operating liquid are 843.6 kg/m3 , 0.00136 Pa. s , and 0.042 N/m ,
respectively. All of the operating modes, including the cone jet mode, could be captured with this solution.
A 13.5 Megapixels DSLR camera (D300 Nikon) with two different imaging systems was used to capture the ES plume and the cone
jet images, separately. At least two sets of 6 images at 6 fps were captured for each condition. The second set was shot with a time
delay, e.g., 10 min, to check the stability of the results with time. For the plume photography, a macro lens (AF Micro 60 mm Nikkor)
coupled with 64 mm long extension tubes were used. A laser sheet was generated with passing a green laser beam (532 nm-50mw
Intelite) through a cylindrical lens. In this arrangement, the laser plane and the camera view were perpendicular. A black metal plate
was placed behind the test section as the background of plume images. For the imaging of the nozzle tip, a stereo microscope (SMZ-10
Nikon) and a camera flash (S13900 Nikon) with the flash duration of ~1 µs were used. The flash was located on the opposite side of
the camera, at 30° to 60° angle with respect to the viewing axis. The camera was operating at 2–2.5 µs shutter speed. The image
processing was performed with ImageJ (Abràmoff, Magalhães, & Ram, 2004), for background subtraction, brightness/contrast ad-
justment, and measurements.

3. Results and discussion

In this section, we present and discuss the interaction between the air crossflow and the ES. Fig. 1b shows a typical flow pattern
inside a chamber with sudden expansion. In our working Reynolds numbers, i.e., greater than 500, the recirculation zone spreads in
the whole length of the chamber; therefore, there are two recirculation regions at the top and the bottom of the two-dimensional flow
domain. The chamber-long recirculation zone is in agreement with the existing predictions of recirculation length in sudden ex-
pansion (Back & Roschke, 1972; Forrester & Evans, 1997).

3.1. Effect of Carrier gas flow rate on plume bending and angles

In the first set of experiments, the tip of the ES nozzle is located at the chamber centerline. In this case, the applied voltage and the
liquid flow rate are kept at 3 kV and 1.167 µL/min, respectively. Fig. 2 shows the ES plume generated at stable cone jet mode and low
crossflows (i.e. less than 10 Scfh). With no airflow, a symmetric conical plume is generated (Fig. 2a). With a crossflow, both the
leeward and the windward boundaries of the plume bend toward the flow (Fig. 2b-f). This means that the induced airflow by the
spray, which depends on the liquid flow rate and droplet sizes (Gañán-Calvo, Lasheras, Dávila, & Barrero, 1994; Ghosh & Hunt, 1998),
is not strong with respect to the crossflow. This ESICF has certain characteristics, which are shown in Fig. 3 and are discussed below:

3.1.1. Jet angle


The emitting jet from the Taylor cone may have an angle with respect to the vertical axis. The measured angle, α1, is essentially
the angle of the vertical axis with the straight line that connects the emission point to the plume formation location. When there is no
crossflow, the jet is vertical and the jet angle is zero. The jet angle increases with increasing the crossflow velocity, as shown with a
dashed line in Fig. 3b. This angle increases from 13° at 2 Scfh to 20° and 31° at 6 Scfh and 10 Scfh, respectively.

235
E. Allaf-Akbari, N. Ashgriz Journal of Aerosol Science 114 (2017) 233–243

Fig. 2. The ES operating at the stable cone jet mode with different air crossflows inside the desolvation chamber. The tip of the nozzle is located at the chamber's
centerline (the solid and dashed lines show the chamber's inlet diameter and the centerline, respectively).

Fig. 3. Top: illustrations of jet bending and plume angles; Bottom: characteristics of the ES plume in different crossflows inside the chamber (nozzle tip is located at the
centerline).

3.1.2. Initial plume angle


The droplets generated by the jet disperse shortly after the breakup, forming a plume with an initial plume angle of α2. The solid
line in Fig. 3b shows that the initial plume angle decreases with increasing the crossflow, from 44° at 2 Scfh to 24° at 10 Scfh. This
reduction is mainly due to a higher effect of the crossflow on the windward plume boundary than the leeward boundary. Thus, the
windward side bends more toward the flow, reducing the initial plume angle.

3.1.3. Plume bending angle


In addition to the initial plume angle, we have defined a bending angle for the plume, α3. This angle is defined as the angle

236
E. Allaf-Akbari, N. Ashgriz Journal of Aerosol Science 114 (2017) 233–243

between the centerline of the plume and the vertical axis. Centerline of the plume is drawn at half angle of the initial plume angle.
According to the dotted line in Fig. 3b, this bending angle also increases with the flow, changing from 18° at 2 Scfh to 44° at 10 Scfh.
The plume bending angle is slightly higher than the jet angle at each crossflow (~13° for flows between 4 and 10 Scfh). This is
because the generated droplets lose their initial inertia faster than the emitting jet.

3.1.4. Dispersed plume angle


Because the flow field inside the chamber has a recirculation zone below the main crossflow, and because a downward external
electric field is applied in the present configuration, the plume disperses asymmetrically. The windward trajectory passes the central
zone earlier, and bends back sharply toward the ground electrode. On the other hand, the leeward trajectory travels more in the
central region. With increasing the air, the leeward trajectory moves even further in the flow direction; therefore, the plume opens up
and spreads more. In order to characterize this plume dispersion in the vessel, a dispersed plume angle, α4, is defined. This is the
angle between the downstream windward and leeward boundaries of the plume. The dispersed plume angle (shown with a dash-dot
line in Fig. 3b) is much larger than the initial plume angle, which indicates a significant dispersion. The angle increases from 61° at 2
Scfh to 82° at 10 Scfh.
The mentioned angles are the averages from a set of six images. The measurement discrepancies are less than 4° and 2° for the
plume and the bending angles, respectively. At around 10 Scfh, the plume became slightly dynamic but still stable (dynamic cone jet).
The electrosprayed droplets disperse, evaporate and undergo breakups. Therefore, the plume image fades away and cannot be
observed further downstream of the injection point. At 3 kV and with increasing the air flow above 10 Scfh, the plume becomes
unstable.

3.2. Effect of carrier gas flow rate on the intensity measured by mass spectrometer

The main objective of introducing an ES plume in crossflow is to carry the ions toward the entrance of a mass spectrometer. Since
the chamber has a complicated internal flow, the injection point of the ES makes a large difference in the droplet trajectories. In the
previous Section 3.1, the ES nozzle was located at the crossflow centerline. In such a case, droplets quickly pass through a high
velocity crossflow and most of them enter the recirculation zone. Such droplets, i.e., ions, do not leave the chamber and collide with
the walls. In order to prevent this ion loss, the ES capillary is lifted upwards by 2.75 mm. The applied voltage and the liquid flow rate
are 3.4 kV and 1 µL/min, respectively. In this part, the results of this configuration are compared with MS detection rates and then
with the previous case.
Fig. 4 shows the ES plume images and the intensity detections of a mass spectrometer at different carrier airflows for the case of ES
tip located above the centerline. The MS tests performed and provided by IONICS Inc. (Private communications, 2015), using a
Reserpine solution with the molality of 10−7 and the flow rate of 1 µL/min. At no crossflow, a symmetric plume is formed (not shown).
By increasing the crossflow, the plume is bent in the flow direction. Up to around 8 Scfh, i.e., 15.1 m/s, the increase of the carrier air
enhances the detection rate. This observation is compatible with the visual results, indicating that low air flows cannot bend the
plume sufficiently; therefore, many ions collide with the walls. With increasing the airflow, the plume is dragged closer to the
centerline and more ions can exit the vessel. In a range of airflows, 8–12 Scfh, the detected intensity is at the peak. This peak is
detected when the generated droplets in the stable plume are carried by the crossflow and reach the spectrometer's inlet. Above this
optimum flow rate, the detection rate decreases moderately. As the flow is increased to 14 Scfh, i.e., 26.5 m/s, the droplets are
dispersed rapidly, making it difficult to see the plume. At this crossflow, the Reynolds number is ~4100, indicating the turbulent
condition inside the chamber, which enhances the dispersion and ion loss.

Fig. 4. The ES operating at the stable cone jet mode with different air crossflows inside the desolvation chamber in comparison to the detected intensity of a mass
spectrometer (tip was located 2.75 mm above the chamber's centerline).

237
E. Allaf-Akbari, N. Ashgriz Journal of Aerosol Science 114 (2017) 233–243

Fig. 5. The schematic of a charge droplet trajectory in the crossflow of the vessel (yellow and red arrows illustrate the applied drag and external electric forces,
respectively).

3.2.1. Effect of tip to centerline distance


The breakup of the electrosprayed jets takes place at some axial distance from the tip. When the nozzle is located well above the
centerline, the droplets are also generated above the centerline. As shown in Fig. 5, when the droplets are above the centerline, they
experience more time and flight distance in the main crossflow. As a result, once the air flow is sufficient, they receive a sufficient
momentum from the air in the horizontal direction and are carried by the air. Besides, in the off-centered case, the droplets become
more dispersed and are expected to have lower inertia when they reach the centerline. Therefore, the drag force can dominate the
motion of the droplets and bend both the windward and leeward boundaries toward the air path.
The generated charged droplets experience a high electric force when they are at the vicinity of the needle tip. This force is
induced by both the columbic repulsion among the droplets and the external electric field. The droplets are very close to each other
initially and repel each other in the opposite directions, i.e., in and against the crossflow direction. For the case of tip on the chamber
centerline, the electric force on the droplets is still significant in comparison to the drag force of the central zone. This electric force
can push the ions off the central zone and bend the plume downwards.
Park et al. (2004) showed that when the induced electric fields at the tip are similar for two ESs, the applied voltages do not
influence the operation mode. The applied electric field to the tip determines jet and droplet emission velocity, when the other
2V
parameters are similar. The electric field at the nozzle tip for the needle-to-plate configuration is calculated with E0 = r ln (4L 0 / r ) ;
c Gr c
where V0 , rc , and LGr are the applied voltage, outer diameter of the capillary, and the distance of capillary from the ground plate,
respectively (Jones & Thong, 1971). Using this equation, the difference in the electric fields between the centered and off-centered
cases is less than 10%. Therefore, the difference in the initial droplet velocity between these two cases should not have a vital role in
the observed bending of plumes.
Another important attribute of locating the ES above the centerline is that only the crossflows higher than 14 Scfh disrupt the
plume generation. This value was 10 Scfh for the nozzle located at the centerline. In the off-centered case, the tip experiences a lower
airflow and the ES is more stable than the centered positioning. From this observation, we noticed that higher crossflows destabilize
the Taylor cone and, therefore, the ES.

3.3. Cone jet in crossflow

A cone jet in the stagnant air environment is typically a symmetric cone, formed on a capillary, with a single liquid jet emitting
from the tip of the cone. Fig. 6a shows this case for a cone jet at the liquid flow rate of 2 µL/min and applied voltage of 3.05 kV.
The crossflow bends the jet and deforms the cone jet. Fig. 6b and c show that at relatively lower crossflows (4 and 8 Scfh), the
cone is bent and its apex is shifted toward the right. In these conditions, the cone jet is stable, meaning that it is steady and stationary.
The cone boundary is curved and its apex is pointier. This observation can be explained based on a study by Pantano, Gañán-Calvo,
and Barrero (1994) on the shape of a Taylor cone with respect to the pressure difference between the liquid and its surrounding air.
They noted that at higher relative air pressures the cone boundary becomes curved toward the axis and at lower relative pressures it
becomes curved away from the axis (Fernández de la Mora, 2007). In Fig. 6b and c, the cone shape looks similar to the lower liquid
pressure case. Similar results are also reported by Gubarenko, Chiarot, Ben Mrad, and Sullivan (2008). They used the expanded
Laplace-Young equation (Landau, Lifshitz, Sykes, Bell, & Dill, 1961) and investigated the stress equilibrium at the liquid/air interface,
in order to determine the curvature of the meniscus: ( pliquid − pair −γ ∇ . n + Tnn = 0), where the terms from the left to right are liquid
pressure, air pressure, pressure due to the surface tension, and the pressure due to the electric field. According to this equation, the
fluid pressure and the external electric tension tend to push the interface outwards and make its shape convex. On the other hand, the
air pressure and surface tension push the liquid inwards and tend to form a concave meniscus. In our tests, increasing the crossflow
causes the air pressure to increase in the chamber and around the cone. The concave boundaries at 4 and 8 Scfh are attributed to the
higher air pressure inside the chamber relative to the no crossflow case.
When the crossflow is increased to 12 Scfh, i.e., 22.7 m/s, the crossflow velocity becomes so large that the Taylor cone becomes

238
E. Allaf-Akbari, N. Ashgriz Journal of Aerosol Science 114 (2017) 233–243

Fig. 6. Cone jet in different crossflows - applied voltages: [a-f] 3.05 kV; [g-m] 3.35 kV.

unstable. Fig. 6d to f show this unstable behavior in three instants. The shape and the size of the meniscus changes constantly and the
cone jet is unstable. Unstable cone jet is referred to a mode that jets emit at different spots, one at a time, and intermittently. During
our tests, we noticed that better plumes and jets can be captured with increasing the voltage. This is attributed to a higher outward
electric pressure on the meniscus, which can compensate for the increase in the local air pressure (see the stress equilibrium
equation). We increased the voltage from 3.05 to 3.35 kV at 16 Scfh (air flow velocity of 30.3 m/s). A stable cone jet was established
at this condition, as shown in Fig. 6g. It is also observed that the size of the cone reduces with the voltage. Further increase of the
crossflow to 20 Scfh (air velocity of 37.9 m/s) makes the cone jet dynamic, and although it is formed steadily, its location and
bending are not constant (see Fig. 6h-i). We refer to this mode as dynamic cone jet.
Fig. 6j shows the ES operating at high crossflow of ~48 Scfh (~91 m/s) and 3.35 kV. In this condition, the electrospraying mode is
unstable cone jet and the cone appears much rarely. The emission of the jet occurs only when the fluid can form an instantaneous
cone. In these instants, the jet has a short time to emit, before the airflow suppresses the instantaneous balance between the electric
stress, surface tension, and the pressure difference at the liquid/air interface. With further increase of the crossflow to ~77 Scfh
(~145 m/s), Fig. 6k, the liquid at the tip is washed, sheared, and dispersed by the air. Most of the liquid is pushed inside the capillary
by the air and sometimes cloud-like pulsations emit from the tip.
Up to this point, the applied voltage to the nozzle was kept below 4 kV. In another set of tests, we investigated the behavior of ES
at ~48 Scfh (~91 m/s), 2 µL/min and relatively high voltages, 4.4–5.2 kV. A stable cone jet could not be established with these
conditions. The highest frequency and multiple number of jet emissions were captured at 4.6–4.8 kV, where the ES was operating in
either the unstable cone jet, or the unstable multi-jet mode, Fig. 7. We expect the highest spectrometer's detection rate in this voltage
range. This prediction is simply because the emission of jets is at its maximum level, which should lead to the highest ion generation.

Fig. 7. A sample schematic of the ES performance at ~48 Scfh (~91 m/s) and 4.6 kV.

239
E. Allaf-Akbari, N. Ashgriz Journal of Aerosol Science 114 (2017) 233–243

3.4. Effect of input sequence on the ES performance

The experimental procedures indicated that the sequence of applying voltage, solution flow, and air crossflow can significantly
influence the ES operation and its characteristics, such as jet emission angle. Three different input sequences are considered for a case
in which the capillary tip is located at the centerline of the chamber (tip to ground electrode distance was 5 mm). The three sequences
are: (1) Solution-Voltage-Crossflow, (2) Voltage-Solution-Crossflow, and (3) Solution-Crossflow-Voltage. In all of the cases discussed
earlier, first, the solution flow was set; then, the voltage was modified to capture the cone jet mode, and finally, the crossflow air was
initiated (this is Sequence 1). The second and the third sequences are discussed here.
For Sequence 2, first the voltage was raised to 2.81 kV. At this voltage, the tip became clear of any fluid, since any liquid
remaining on the tip from the previous tests was ejected. Then, the solution was introduced at 2 µL/min. Once a stable jet was
formed, the air crossflow was added. Without any crossflow, an asymmetric steady cone was formed at the right half of the nozzle tip
and a bent jet emitted toward the right. This cone was much smaller than the ones in a typical cone jet, in which a symmetric cone
covers all the tip surface. As the cross flow was increased, the ES underwent the following changes: At relatively low air crossflows,
the ES bends more in the direction of the flow. With increasing the air flow, the ES becomes more dynamic and the jet emission moves
from one spot to another spot on the capillary, until it eventually becomes unstable at higher crossflows. At even higher crossflows,
no jet emission was observed. These events are qualitatively similar to the first sequence.
Chiarot, Gubarenko, Ben Mrad, and Sullivan (2009) have reported similar studies on the operation of an ES with applying a
voltage before a solution (Sequence 2 without a crossflow). However, they noted a steady and symmetrical cone jet. The initial
asymmetry of the cone jet in our tests is related to the local wetting of the tip. Since, in practice, it may be difficult to prevent local tip
wetting at all times, Sequence 1 is preferred to Sequence 2, in order to have a more repeatable results.
For Sequence 3, first a 1 µL/min solution was introduced, then the air was set to a constant low crossflow of ~4.7 Scfh (~9 m/s),
and finally, the voltage was increased to 3 kV. Below this voltage, a cone was not observed. At 3 kV, the meniscus elongates in the
axial direction, then a jet or a ligament is ejected, and then the meniscus returns to a cusp shape (similar to pulsating mode).
Increasing the voltage up to 3.5 kV, pushed the ES to the unstable or droplet ejection mode, where the liquid can eject in bulks or
droplets. Returning to 3.38 kV resulted in a pulsating mode with more variation in ejections. Further efforts to achieve a cone jet were
not successful. Note that we could have a stable cone jet at similar conditions with Sequence 1. In this sequence, when the crossflow
air was shut off and the voltage was decreased to 3.25 kV, then the ES operated in the stable cone jet mode. The inability to form a
stable cone jet with Sequence 3, is believed to be due to the existing perturbations on the liquid surface caused by the crossflow. The
liquid surface is perturbed by the crossflow, preventing the equilibrium at the interface and the formation of a jet. However, in
Sequence 1, the liquid at the tip is initially stationary, forming a steady cone jet. Therefore, when the crossflow is added, the jet can
still operate at medium crossflows. At higher crossflows, even a stable cone jet becomes perturbed, and the stress equilibrium at the
cone surface is disrupted.

3.5. Electrohydrodynamic (EHD) hysteresis in ESICF

The experimental results show an EHD hysteresis effect for an ESICF, namely, the results are different for when the voltage is
increased than that when the voltage is decreased. Fig. 8a-d and e-h respectively illustrate the ES performance at 20 Scfh (37.8 m/s),
with increasing the voltage up to 3.4 kV, and then decreasing it down to the same potential. The four panels of each case show
different instants of the ESICF in the unstable cone jet mode. The decreasing voltage resulted in more frequent plume appearances
with longer stable durations, as compared to the increasing voltage. This behavior was observed repeatedly at different input con-
ditions. This is believed to be due to the variation of the boundary conditions, namely, the cone shape, the space charge of droplets,
and the wetting at the capillary tip. This EHD hysteresis is caused by both the electric field and the air crossflow. The operation of ES

Fig. 8. ES performance and plume formation with increasing and decreasing the applied voltage; air crossflow: 20 Scfh (37.8 m/s).

240
E. Allaf-Akbari, N. Ashgriz Journal of Aerosol Science 114 (2017) 233–243

with EHD hysteresis, Fig. 8e-h, is expected to result in a better ion detection in a mass spectrometer, as the plume is generated more
frequently (in comparison to the other unstable cone jet modes).
This hysteresis is different than the spray current hysteresis phenomenon, between the cone jet and the pulsating modes, as
reported by Smith (1986), Chen, Pui, and Kaufman (1995), and Noymer and Garel (2000). One should note that in the stable cone jet
mode, the operation of ES and Taylor cone is stable, which is not the case in the unstable cone jet mode in a crossflow. Marginean,
Parvin, Heffernan, and Vertes (2004) referred to standing waves as the reason of a periodic liquid ejection in the pulsating mode. In
another paper, Marginean, Kelly, Page, Tang, and Smith (2007) stated that the hysteresis between the cone jet and the pulsating
modes can be attributed to the stationary behavior of the liquid at the meniscus. In other words, in order to ES fall back into the
pulsating mode, a mechanical fluctuation must initiate the instability to the liquid surface. Therefore, as long as this fluctuation does
not arise, the standing waves are not generated and the stable cone jet can operate below the onset voltage. Having this in mind, the
plume and Taylor cone in our case are highly affected by the crossflow and are not stationary; therefore, our observed phenomenon is
not the same as that discussed in the literature. Detailed investigations on the current hysteresis and this new observation are beyond
the scope of this paper.
In general, the final voltage and the air crossflow are not the only parameters that predict the exact behavior of an ES. The
capillary tip wetting is another important parameter that determines the ES behavior at medium-high crossflows. The tip wetting
determines the shape and fluid dynamics of the meniscus, which leads to the variation of boundary conditions. This effect is a serious
obstacle for repeatability, quantification, and prediction of the exact results. Understanding the physics of the tip wetting and its
relation with operational parameters, such as airflow and voltage, will lead to new developments in ESICF, and its existing and
possible applications. The analysis of the transient behavior of an ESICF using an equilibrium model, similar to Chiarot et al. (2009)
and Gubarenko et al. (2008), would help in better understating of this observation. It is also required to study the other types of
capillaries and nozzles under the effect of crossflow. For the time being, monitoring the ES, similar to Valaskovic et al. (2004), and
Nguyen et al. (2014), can be used to tackle the instability of ES plume and cone jet in the medium-high crossflows. Lastly, the stable
behavior of ES in low-medium crossflows, should be targeted as the favorable operating condition for the design of ion sources.

4. Conclusion

• The behavior of an ES in a gaseous crossflow inside a chamber is investigated. The chamber deign is mainly used for introducing
the electrosprayed ions into a mass spectrometer. The crossflow influences the cone shape, jet emission, and plume structure and
trajectory. The injection location of the ES can result in large changes in the ES plume behavior and structure:
i. When the tip of the ES nozzle is placed at the centerline of the crossflow, a stable plume is found for crossflows up to 18.9 m/s
(10 Scfh) inlet velocities. In this configuration, the plume initially bends in the main flow direction. However, once it pene-
trates further and passes the central zone, it bends back toward the ground plate, located at the bottom of the chamber.
Therefore, the generated ions may not exit the chamber. At medium crossflow inlet velocities, i.e., 18.9–37.8 m/s, the plume
becomes unstable. The applied voltage to the ES has to be increased to obtain a stable ES.
ii. When the tip of the ES nozzle is placed above the centerline (by 2.75 mm), a stable plume is found for crossflows up to 26.5 m/s
(14 Scfh) inlet velocities. For crossflow velocities in the range of 15.1–22.7 m/s (8–12 Scfh), the plume bends sufficiently to
direct the droplets towards the exit of the chamber. However, for inlet velocities below this range, the plume does not bend
enough and the ions mostly collide with the ground plate. Above this range, the crossflow becomes highly turbulent, leading to
dispersion and ion loss. These findings matched well with an actual intensity measurement with a mass analyzer.
• The operating modes of an ESICF are identified as stable, dynamic, and unstable cone jet or multi jet. The general behavior of an
ES in the stable cone jet mode with a crossflow can be described as follows:
i. As a crossflow is imposed on an ES cone jet, it deforms the Taylor cone and bends the emitting jet. The crossflow results in a
higher air pressure in the chamber and at the cone surface, therefore, making the cone boundary curved and its apex pointier.
Above a certain crossflow rate, the cone jet becomes dynamic. In the dynamic cone jet mode, the cone jet is stable, but it
continuously deforms and moves. With further increase in the cross flow, the cone jet becomes unstable. Increasing the voltage
may help the restoration of the stable cone. In the unstable cone jet mode, the cone repeatedly forms and disappears and the jet
emission occurs intermittently.
ii. At high crossflows, e.g., 90 m/s, the ES operates at the unstable cone jet mode, since the liquid at the meniscus is highly
perturbed and pushed inside the capillary tube. With relatively high voltages, instantaneous jet emissions occur in these
crossflows. At some specific voltages, multiple instantaneous jets emit from the meniscus with relatively high frequencies.
• The sequence of applying the voltage, starting the liquid solution flow, and introducing the crossflow, influences the performance
of the ES. The most stable and predictable plume was obtained by “Solution-Voltage-Crossflow” sequence. The ES performance
with “Voltage-Solution-Crossflow” sequence depends on the tip wetting and boundary conditions. In the “Crossflow-Solution-
Voltage” sequence, unlike the previous ones, the cone jet mode was not observed. This is attributed to the perturbations on the
liquid surface induced by the crossflow, which prevents the formation of the Taylor cone.
• An EHD hysteresis phenomenon was observed in the medium-high crossflow range, i.e., 18.9–37.8 m/s, when the ES operated in
the unstable cone jet mode. The plume generation in this mode became more stable when the applied voltage was reduced to a
target voltage from a higher value (in comparison to a case when the voltage was increased to the target voltage from a lower
value). This hysteresis is attributed to the wetting of the ES nozzle tip coupled with the boundary conditions of the cone (due to
voltage changes). The temporary variation of inputs can change the boundary conditions at the tip and, therefore, the perfor-
mance of the ESICF.

241
E. Allaf-Akbari, N. Ashgriz Journal of Aerosol Science 114 (2017) 233–243

Acknowledgment

This study was financially supported by IONICS Inc. (Grant number 495678) and the Natural Sciences and Engineering Research
Council of Canada (Grant number 495677). We also highly acknowledge Drs. Reza Javahery, Lisa Cousins, Hui Qiao, and Serguei
Savitchenko for the constructive discussions and the provided information.

References

Abràmoff, M. D., Magalhães, P. J., & Ram, S. J. (2004). Image processing with ImageJ. Biophotonics International, 11(7), 36–42.
Arumugham-Achari, A. K., Grifoll, J., & Rosell-Llompart, J. (2015). A comprehensive framework for the numerical simulation of evaporating electrosprays. Aerosol
Science and Technology, 49(6), 436–448. http://dx.doi.org/10.1080/02786826.2015.1039639.
Back, L. H., & Roschke, E. J. (1972). Shear-layer flow regimes and wave instabilities and reattachment lengths downstream of an abrupt circular channel expansion.
Journal of Applied Mechanics, 39(3), 677–681.
Bailey, A. G. (1988). Electrostatic spraying of liquids. New York: Wiley.
Barrero, A., & Loscertales, I. G. (2007). Micro- and nanoparticles via capillary flows. Annual Review of Fluid Mechanics, 39(1), 89–106. http://dx.doi.org/10.1146/
annurev.fluid.39.050905.110245.
Chen, D. R., Pui, D. Y. H., & Kaufman, S. L. (1995). Electrospraying of conducting liquids for monodisperse aerosol generation in the 4 nm to 1.8 μm diameter range.
Journal of Aerosol Science, 26(6), 963–977. http://dx.doi.org/10.1016/0021-8502(95)00027-A.
Chiarot, P. R., Gubarenko, S. I., Ben Mrad, R., & Sullivan, P. E. (2009). On the pulsed and transitional behavior of an electrified fluid interface. Journal of Fluids
Engineering, 131(9), 91202. http://dx.doi.org/10.1115/1.3203203.
Chiarot, P. R., Sullivan, P., & Ben Mrad, R. (2011). An overview of electrospray applications in MEMS and microfluidic systems. Journal of Microelectromechanical
Systems, 20(6), 1241–1249. http://dx.doi.org/10.1109/JMEMS.2011.2168810.
Cloupeau, M., & Prunet-Foch, B. (1990). Electrostatic spraying of liquids: Main functioning modes. Journal of Electrostatics, 25(2), 165–184. http://dx.doi.org/10.
1016/0304-3886(90)90025-Q.
Cole, R. B. (1997). Electrospray ionization mass spectrometry: Fundamentals, instrumentation and applications. New York: John Wiley & Sons Inc.
Deng, W., Klemic, J. F., Li, X., Reed, M. A., & Gomez, A. (2007). Liquid fuel microcombustor using microfabricated multiplexed electrospray sources. Proceedings of the
Combustion Institute, 31, 2239–2246. http://dx.doi.org/10.1016/j.proci.2006.08.080.
Deshpande, S., Gao, J., & Trujillo, M. F. (2011). Characteristics of hollow cone sprays in crossflow. Atomization and Sprays, 21(4), 349–361.
Eslamian, M., Amighi, A., & Ashgriz, N. (2014). Atomization of liquid jet in high-pressure and high-temperature subsonic crossflow. AIAA Journal, 52(7), 1374–1385.
http://dx.doi.org/10.2514/1.J052548.
Fenn, J. B., Mann, M., Meng, C. K., Wong, S. F., & Whitehouse, C. M. (1989). Electrospray ionization for mass spectrometry of large biomolecules. Science, 246(4926),
64–71. http://dx.doi.org/10.1126/science.2675315.
Fernández de la Mora, J. (2007). The fluid dynamics of Taylor cones. Annual Review of Fluid Mechanics, 39(1), 217–243. http://dx.doi.org/10.1146/annurev.fluid.39.
050905.110159.
Fomina, N. S., Masyukevich, S. V., Sviridovich, E. N., & Gall, N. R. (2014). A polycapillary electrospray source for generating charged droplet flows. Instruments and
Experimental Techniques, 57(2), 226–231. http://dx.doi.org/10.1134/S0020441214010205.
Forrester, S. E., & Evans, G. M. (1997). Computational modelling study of the hydrodynamics in a sudden expansion tapered contraction reactor geometry. Chemical
Engineering Science, 52(21), 3773–3785.
Gañán-Calvo, A. M. (2007). Electro-flow focusing: The high-conductivity low-viscosity limit. Physical Review Letters, 98, 134503. http://dx.doi.org/10.1103/
PhysRevLett.98.134503.
Gañán-Calvo, A. M., Lasheras, J. C., Dávila, J., & Barrero, A. (1994). The electrostatic spray emitted from an electrified conical meniscus. Journal of Aerosol Science,
25(6), 1121–1142. http://dx.doi.org/10.1016/0021-8502(94)90205-4.
Gañán-Calvo, A. M., Lopez-Herrera, J. M., & Riesco-Chueca, P. (2006). The combination of electrospray and flow focusing. Journal of Fluid Mechanics, 566, 421–445.
http://dx.doi.org/10.1017/S0022112006002102.
Gañán-Calvo, A. M., & Montanero, J. M. (2009). Revision of capillary cone-jet physics: Electrospray and flow focusing. Physical Review E - Statistical, Nonlinear, and Soft
Matter Physics, 79, 66305. http://dx.doi.org/10.1103/PhysRevE.79.066305.
Ghosh, S., & Hunt, J. C. R. (1998). Spray jets in a cross-flow. Journal of Fluid Mechanics, 365, 109–136. http://dx.doi.org/10.1017/S0022112098001190.
Grifoll, J., & Rosell-Llompart, J. (2012). Efficient Lagrangian simulation of electrospray droplets dynamics. Journal of Aerosol Science, 47, 78–93. http://dx.doi.org/10.
1016/j.jaerosci.2012.01.001.
Gubarenko, S. I., Chiarot, P., Ben Mrad, R., & Sullivan, P. E. (2008). Plane model of fluid interface rupture in an electric field. Physics of Fluids, 20(4), 43601. http://dx.
doi.org/10.1063/1.2891311.
Holbrook, L., Hindle, M., & Longest, P. W. (2015). Generating charged pharmaceutical aerosols intended to improve targeted drug delivery in ventilated infants.
Journal of Aerosol Science, 88, 35–47. http://dx.doi.org/10.1016/j.jaerosci.2015.05.015.
Jaworek, A., & Krupa, A. (1999). Classification of the modes of EHD spraying. Journal of Aerosol Science, 30(7), 873–893. http://dx.doi.org/10.1016/S0021-8502(98)
00787-3.
Jolliffe, C., Javahery, G., Cousins, L., & Savtchenko, S. (2010). Ion source vessel and methods.
Jones, A. R., & Thong, K. C. (1971). The production of charged monodisperse fuel droplets by electrical dispersion. Journal of Physics D: Applied Physics, 4(8),
1159–1166. http://dx.doi.org/10.1088/0022-3727/4/8/316.
Kyritsis, D. C., Coriton, B., Faure, F., Subir, R., & Gomez, A. (2004). Optimization of a catalytic combustor using electrosprayed liquid hydrocarbons for mesoscale
power generation. Combustion and Flames, 139, 77–89. http://dx.doi.org/10.1016/j.combustflame.2004.06.010.
Landau, L. D., Lifshitz, E. M., Sykes, J. B., Bell, J. S., & Dill, E. H. (1961). Electrodynamics of Continuous Media.
Manisali, I., Chen, D. D. Y., & Schneider, B. B. (2006). Electrospray ionization source geometry for mass spectrometry: Past, present, and future. TrAC - Trends in
Analytical Chemistry, 25(3), 243–256. http://dx.doi.org/10.1016/j.trac.2005.07.007.
Marginean, I., Kelly, R. T., Page, J. S., Tang, K., & Smith, R. D. (2007). Electrospray characteristic curves: In pursuit of improved performance in the nanoflow regime.
Analytical Chemistry, 79(21), 8030–8036. http://dx.doi.org/10.1021/ac0707986.
Marginean, I., Parvin, L., Heffernan, L., & Vertes, A. (2004). Flexing the electrified meniscus: The birth of a jet in electrosprays. Analytical Chemistry, 76(14),
4202–4207. http://dx.doi.org/10.1021/ac049817r.
Mashayek, A., & Ashgriz, N. (2011). Atomization of a liquid jet in a crossflow. In N. Ashgriz (Ed.). Handbook of atomization and sprays (pp. 657–683). US: Springer.
Nguyen, T. K., Nguyen, V. D., Seong, B., Hoang, N., Park, J., & Byun, D. (2014). Control and improvement of jet stability by monitoring liquid meniscus in electrospray
and electrohydrodynamic jet. Journal of Aerosol Science, 71, 29–39. http://dx.doi.org/10.1016/j.jaerosci.2014.01.004.
Noymer, P. D., & Garel, M. (2000). Stability and atomization characteristics of electrohydrodynamic jets in the cone-jet and multi-jet modes. Journal of Aerosol Science,
31(10), 1165–1172. http://dx.doi.org/10.1016/S0021-8502(00)00019-7.
Pantano, C., Gañán-Calvo, A. M., & Barrero, A. (1994). Zeroth-order, electrohydrostatic solution for electrospraying in cone-jet mode. Journal of Aerosol Science, 25(6),
1065–1077.
Park, H., Kim, K., & Kim, S. (2004). Effects of a guard plate on the characteristics of an electrospray in the cone-jet mode. Journal of Aerosol Science, 35(11), 1295–1312.
http://dx.doi.org/10.1016/j.jaerosci.2004.05.012.
Pramanik, B. N., Ganguly, A. K., & Gross, M. L. (2002). Applied electrospray mass spectrometry: Practical spectroscopy series VOL 32. New york - Basel: Marcel Dekker.

242
E. Allaf-Akbari, N. Ashgriz Journal of Aerosol Science 114 (2017) 233–243

Private communications (2015). IONICS Mass Spectrometry Inc., Bolton, Ontario, Canada.
Savtchenko, S., Ashgriz, N., Jolliffe, C., Cousins, L., & Gamble, H. (2014). Effect of swirling desolvation gas flow in an atmospheric pressure ion source. Journal of the
American Society for Mass Spectrometry, 25(9), 1549–1556. http://dx.doi.org/10.1007/s13361-014-0933-9.
Smith, D. P. H. (1986). The electrohydrodynamic atomization of liquids. IEEE Transactions on Industry Applications, 3, 527–535. http://dx.doi.org/10.1109/TIA.1986.
4504754.
Sultan, F. (2013). Characterization of an electrospray with co-flowing gas. University of Toronto.
Sultan, F., Ashgriz, N., Guildenbecher, D. R., & Sojka, P. E. (2011). Electrosprays. In N. Ashgriz (Ed.). Handbook of atomization and sprays (pp. 727–753). US: Springer.
Taylor, G. (1964). Disintegration of water drops in an electric field. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences,
280(1382), 383–397. http://dx.doi.org/10.1098/rspa.1964.0151.
Valaskovic, G. A., Murphy, J. P., & Lee, M. S. (2004). Automated orthogonal control system for electrospray ionization. Journal of the American Society for Mass
Spectrometry, 15(8), 1201–1215. http://dx.doi.org/10.1016/j.jasms.2004.04.033.
Wang, R., Allmendinger, P., Zhu, L., Gröhn, A. J., Wegner, K., Frankevich, V., & Zenobi, R. (2011). The role of nebulizer gas flow in electrosonic spray ionization
(ESSI). Journal of the American Society for Mass Spectrometry, 22(7), 1234–1241. http://dx.doi.org/10.1007/s13361-011-0124-x.
Wang, Y., Tan, M. K., Go, D. B., & Chang, H.-C. (2012). Electrospray cone-jet breakup and droplet production for electrolyte solutions. EPL (Europhysics Letters), 99(6),
64003. http://dx.doi.org/10.1209/0295-5075/99/64003.
Wilhelm, O., Mädler, L., & Pratsinis, S. E. (2003). Electrospray evaporation and deposition. Journal of Aerosol Science, 34(7), 815–836. http://dx.doi.org/10.1016/
S0021-8502(03)00034-X.
Zeleny, J. (1917). Instability of electrified liquid surfaces. Physical Review, 10, 1–6. http://dx.doi.org/10.1103/PhysRev.10.1http://journals.aps.org/pr/pdf.

243

You might also like