ArticlePDF Available

Probing liquid surfaces under vacuum using SEM and ToF-SIMS

Authors:

Abstract and Figures

We report a newly developed self-contained interface for high-vapor pressure liquid surfaces to vacuum-based analytical instruments. It requires no wires or tubing connections to the outside of the instrument and uses a microfluidic channel with a 3 μm diameter window into the flowing fluid beneath it. This window supports the liquid against the vacuum by the liquid's surface tension and limits the high-density vapor region traversed by the probe beams to only a few microns. We demonstrate this microfluidic interface for in situ liquid surfaces in a time-of-flight secondary ion mass spectrometer (ToF-SIMS) and a scanning electron microscope (SEM) with chemical analysis.
Content may be subject to copyright.
Probing liquid surfaces under vacuum using SEM and ToF-SIMS
Li Yang,
a
Xiao-Ying Yu,*
b
Zihua Zhu,
c
Martin J. Iedema
c
and James P. Cowin*
a
Received 7th December 2010, Accepted 18th May 2011
DOI: 10.1039/c0lc00676a
We report a newly developed self-contained interface for high-vapor
pressure liquid surfaces to vacuum-based analytical instruments. It
requires no wires or tubing connections to the outside of the
instrument and uses a microfluidic channel with a 3 mm diameter
window into the flowing fluid beneath it. This window supports the
liquid against the vacuum by the liquid’s surface tension and limits
the high-density vapor region traversed by the probe beams to only
a few microns. We demonstrate this microfluidic interface for in situ
liquid surfaces in a time-of-flight secondary ion mass spectrometer
(ToF-SIMS) and a scanning electron microscope (SEM) with
chemical analysis.
Much chemistry takes place at the interface of liquid phases with
gases in environmental, industrial, and biological systems. The
surfaces of these aqueous phases and films (when <1 nm thick) have
unique kinetics and thermodynamics, distinct from the bulk.
1–3
However, many key surface analytical techniques are vacuum-based,
and cannot easily probe these high vapor-pressure interfaces. Our
goal is to permit studies of high vapor pressure liquids for electron
and vacuum-based ion/molecular based techniques. We first discuss
some existing approaches and their limitations.
The liquid jet interface was pioneered by Manfred Faubel
4
for
electron and X-ray analysis of liquids. It directs a 5–30 micron diam-
eter water jet into a cryopumped vacuum chamber.
5
It requires a very
specialized system, and water evaporation from the jet causes extreme
jet supercooling and freezing. The environmental SEM (ESEM),
originally developed by Danilatos
6
then commercialized via Electro-
scan and FEI, and a recent in situ X-ray photoelectron spectrometer
(XPS)
7,8
(not commercially available) both require specially built
instruments. They do look at in situ aqueous samples, but cannot
handle highly reactive gaseous environments, and will not work much
above about 25 Torr. The TEM via its tight geometry and close-
coupled side port has permitted several impressive in situ probes.
9
But
these are inherently tethered to specific instruments. Other techniques
including ‘‘WetStem’’ cells with liquid trapped between two thin silicon
nitride (SiN) films for SEM and STEM
10
and a similar recent TEM
interface (Hummingbird Scientific, Lacey, WA) are not self-contained,
nor are they true vacuum–liquid interfaces, even if the latter does make
flowing liquids directly available for studying TEM systems.
We engineered a unique liquid sample interface to provide
a completely portable and self-contained micro-environment, suitable
as a multimodal platform. That is, it can be potentially used in any
finely focused (<1 mm) analytical device, as well as optical micro-
scopes. Crucial to its design is the small hole size and the flowing
solution, both of which can reduce the effects of solvent evaporation
on sample temperature and concentration to manageable levels
(about 10 C drop and 20% change, respectively, see ESI†). These
effects would be very severe if there were no flow, or if one had the
360exposed geometry of a liquid jet.
Brivio and co-workers
11
took an innovative approach to exam-
ining liquids that shares some similarities to our device. They have
a bulk solution analyzed by passing through a microfluidic channel
which was capped with a 2 mm thick silicon film. Several holes
100–500 nm diameters were drilled via a focused ion beam (FIB) as
‘‘windows’’ into the flowing liquid beneath. A focused laser then
ablates the exposed liquid (and any ablatable surrounding material
that has seeped or diffused out of the hole) to be analyzed in a matrix-
assisted laser desorption ionization mass spectrometer (MALDI-
MS). It probes a true vacuum–liquid interface supported by its
surfacetension,aswedo.Howevertheirsisdesignedtoanalyzethe
bulk fluid, while ours is designed to study the liquid surface. Theirs is
aimed at being a specific MALDI interface, while ours aims to be
a generic interface. Moreover, our new device provides a continuous
flow using an electro-osmotic pump, whereas they use the pressure
difference between an air bubble and vacuum system to move the
liquid. As a result, our flow rates are easily controlled, while theirs
requires much more trial and error to achieve reasonable flows.
The design of the liquid vacuum module is shown in Fig. 1A,
mainly consisting of a PDMS microfluidic block with microchannel
on the top made by soft lithography (Sylgard 184, Dow Corning
Co., Midland, MI) (a), an electro-osmotic pump (Model number:
3000126, Dolomite) (b), a battery (Saft, Li-SOCl
2
)(c),andValco
PTFE connecting tubes (d). Although PDMS and PTFE are
permeable to water vapor, we calculate that only 3.8 10
8
and 2.0
10
7
g water diffuse through the tubing and PDMS block during
the 8 hours of run per fill (calculation in ESI†). These are small
compared with the 0.2 g of water in the device. The PDMS
a
Chemical and Materials Sciences Division, Pacific Northwest National
Laboratory, Richland, WA, 99354, USA. E-mail: jpcowin@charter.net
b
Atmospheric Sciences and Global Climate Change Division, Pacific
Northwest National Laboratory, Richland, WA, 99354, USA. E-mail:
xiaoying.yu@pnl.gov
c
Scientific Resources Division, W. R. Wiley Environmental Molecular
Science Laboratory, Pacific Northwest National Laboratory, Richland,
WA, 99354, USA
† Electronic supplementary information (ESI) available. See DOI:
10.1039/c0lc00676a
‡ Current address: Cowin In-Situ Science, L.L.C., Richland, WA 99354,
USA.
This journal is ªThe Royal Society of Chemistry 2011 Lab Chip, 2011, 11, 2481–2484 | 2481
Dynamic Article LinksC
<
Lab on a Chip
Cite this: Lab Chip, 2011, 11, 2481
www.rsc.org/loc COMMUNICATION
microfluidic block was coated with a gold film to further reduce gas
permeation
12,13
and prevent charging. The continuous liquid flow
was obtained by an electro-osmotic pump driven by a battery. A tee
on the line (a small viton tube (f) intersected with a short glass
capillary) was used during device filling to prevent overpressure
events. The flow channel (Fig. 1B) on the PDMS block was covered
by a silicon nitride (SiN) film,
14
which subsequently had a 2–3 mm
hole (Fig. 1C) drilled through by a focused ion beam (FIB, FEI
Helios, FIB/SEM).
Our microchip-based vacuum interface was deployed in an SEM
(FEI XL30) and a ToF-SIMS (IONTOF GmbH, M
unster, Ger-
many). The ToF-SIMS rasters a focused (0.25 mm) primary ion beam
over the sample. Ejected secondary ions are mass-analyzed to
form surface chemical maps with about 0.25 mm resolution. The
ToF-SIMS provides molecular information on the top nm of the
liquid. The SEM, via the energy dispersed X-rays generated by the
rastering electron beam, provides images and elemental maps for
elements from B and up. The depth resolution is 0.1 to 1 micrometre
for the elemental mapping. The aperture size needed to be large
enough to permit the analytical methods (SEM/EDX and ToF-
SIMS) to cleanly probe within the aperture, yet small enough to limit
mean-free path issues and fluid loss. Evaporating vapor will be dense
only around the distance of the aperture diameter away from the
device. For electron based SEM, the mean free path through a gas for
electron energies of 20 keV is millimetres,
15
much longer than the
aperture size. The ToF-SIMS uses 25 keV Bi ions as probes, while the
detected ions typically range from 0 to 10 eV. For simplicity, we
assume the probability of the ion mean free path is the same as that of
a water molecule. The mean free path of water molecules is 13.5 mm
at 24 Torr.
16
For the 3 mm aperture, the radius (R)is1.5mm, making
the ion collision approximately 0.11 (1.5/13.5), acceptably small.
The fluid was mechanically supported by its surface tension (s)
across the opening, against the pressure difference inside to outside.
This maximum pressure is P
max
¼2s/R.
17
For sof pure water
(0.073 N m
1
)andRof 1.5 10
6
m, P
max
is 97 000 Pa, just short of
1 bar. This can easily hold off the vapor pressure of water (about 0.03
bar), and the pressure needed to push the liquid through the channel
(about 0.1–0.2 bar). Flowing fluids limit concentration changes from
evaporation, temperature drop, and beam damage to the solution
compared to having a static fluid. In order to maintain higher linear
flow rates directly behind the aperture and yet prevent too high of
a pressure drop, the designed channel had two widths (80 mmwide,
1.97 mm long and 10 mmwide,30mm long). This makes for an
average linear flow rate at the aperture of 3.5 cm s
1
(at 208 nL min
1
fluid flow) at only 0.13 bar of pressure drop. The temperature drop is
about 17 K, according to our calculations
18
(see ESI†) by the time the
fluid crosses the 3 mm aperture.
Fig. 2A shows the SEM image of the microchannel with DI water
flowing through the channel. The hole in the SiN window is shown in
Fig. 2B, taken at a high energy (30 keV) electron beam. The liquid is
seen as a low contrast blurry region in the hole with a dark band near
the edge of the hole. The liquid presence is confirmed by the EDX
results. Fig. 2C shows the intensity of observed elements at different
spots (with the e beam parked), including the spot outside the channel
(S1), outside the hole but in the channel (S2), and in the hole (S3),
taken at a low energy beam (10 keV). The atomic percentages at the
different locations are summarized in Table 1. The lower energy
beam was chosen, as it barely penetrates the SiN film. Indeed the
much higher oxygen atom percent inside the hole demonstrates the
presence of the bare water liquid surface compared with the other two
locations. Similarly, the low atom percent of C, Si, and N for the
beam over the aperture is consistent with the bare liquid interface.
Images of holes with no water behind them (shown in the ESI†) have
very dark secondary electron and EDX signals.
This is the first time that highly volatile liquid surfaces have been
investigated using ToF-SIMS to the best of our knowledge. The main
chamber operating vacuum pressure in the experiments was 2.5 to 5.5
10
7
mbar. Fig. 3A shows the secondary ion images around the
aperture, with the channel unfilled. The primary ion beam was 25 keV
Bi
+
(beam size: 250 nm). The beam current is 1.0 pA instantaneously,
chopped at 20 kHz, with a beam width of 130 ns. Data shown in
Fig. 3A was taken for a total integration time of 65.5 s, over a scan area
of 10 mm
2
square, with 256 256 pixels. A clear hole can be found at
the center of the Si
+
image with a diameter about 2mm, consistent
Fig. 1 (A) Assembly of the liquid interface. The parts are: (a) channel,
(b) electro-osmotic pump, (c) battery, (d) 1 of 2 Teflon tube fluid reser-
voirs, (e) connecting wires to battery, (f) pressure relief tee (see text). (B)
Optical micrograph of channel. Water flow is from left to right, and the
narrow section is where the aperture is. (C) Optical micrograph of the 2
micron hole made by FIB.
2482 | Lab Chip, 2011, 11, 2481–2484 This journal is ªThe Royal Society of Chemistry 2011
with the SEM results. In addition, low H
,Na
+
and I
signals are
observed in the hole. When we flow DI water through the channel
(Fig. 3B), a hole is also found in the Si
+
image. The Na
+
signal origi-
nates from an area considerably larger than the hole (4mmdiameter),
and is not uniform around the hole. This Na signal appears to have
been brought to the aperture by the DI water, and may come from
fairly low trace impurities (the sensitivity of ToF-SIMS to Na is often
orders of magnitude higher than that for other species). Mobile ions on
surfaces have been observed in the presence of monolayer amounts of
adsorbed water.
19
There may be a partial monolayer of hydrated Na
impurities on the surface, at equilibrium with the DI water.
The high H
signal in the hole is observed as expected when water
is present. The H
signal comes from a region a little larger than the
hole seen in the Si
+
signal. This can be partly explained by a beam size
of around 250 nm. The H
signal may also be a little larger than the
actual hole size, as there is a signal originating from monolayer
amounts of surface water.
Fig. 3C shows the case where a 5 mM sodium iodide aqueous
solution is flowing through the channel, for the same conditions as
those of Fig. 3A and B. The Si
+
image shows a hole as expected. The
Na
+
signal has a very bright core, about the same size as the Si
+
hole.
The I
signal originates from a region about the size of the Si
+
hole.
This is compatible with what is expected for imaging the liquid
surface. The H
signal comes from a region a little larger than the
hole, as does the less-intense halo of the Na
+
,whichmayindicate
a small amount of hydrated Na
+
adsorbed on the surface of the SiN
film. Our results show the aqueous solution is exposed to the vacuum,
and its composition can be probed by the ToF-SIMS.
When an amino acid solution, 1% glutamic acid, was flown
through the channel, the anions H
and [M H]
were strongly seen
in the liquid exposed at the hole, as expected. The Na
+
signal showed
just trace amounts, also as expected (See Fig. 3D). Understanding
what biological molecules are exposed at the surface of a solution
could be important in many applications. This interface enables the
exploration of those systems.
The ToF-SIMS can directly be used to drill the aperture, instead of
using the FIB/SEM. It is done by turning up the ion beam current. This
reduces the handling of the whole device prior to usage and possible
contamination, compared to making the hole in advance with the FIB.
In this experiment, the channel contained flowing D
2
O. The focused
250 nm Bi
+
beam was rastered over a circular area with a diameter of
3mm. To get a rapid sputtering rate, a high average current was
required. This was achieved by lengthening the pulse width to 800 ms,
making the average current 730 times what was used for Fig. 3. In this
mode the mass resolution is much reduced, so only H
and D
could
be cleanly separated in the spectra. High mass resolution spectra
(narrow pulse width) show that H
2
peak is very weak (<1% of D
),
and thus the signal around 2.0 amu range can be regarded as pure D
.
We monitored the ToF-SIMS signal while the intense sputtering was
ongoing, yielding a depth profile. Fig. 4 shows that the Bi
+
beam can
drill a hole through the 100 nm silicon nitride layer in about 42 seconds.
The D
signal shows a dramatic jump (30 times in signal amplitude)
Fig. 2 SEM image of the microchannel (A) and the micron hole above
the channel (B). EDX spectra of different spots (including outside
channel, outside hole but in channel, and in hole) (C).
Table 1 Atomic percent of different spots (S1, S2 and S3 in Fig. 2)
Location
Atomic percent (%)
C O Si N Na Cl
Outside channel (S1) 38 6 14 41 1 0
Outside hole but in channel (S2) 36 15 12 37 1 0
In hole (S3) 9 86 2 3 0 0
Fig. 3 ToF-SIMS imaging (10 10 mm
2
) of the silicon nitride
membrane surface around the hole. Microchannel was (A) empty, (B)
holds flowing H
2
O, (C) holds a flowing 5 mM sodium iodide aqueous
solution, and (D) holds a flowing 1% glutamic acid solution.
This journal is ªThe Royal Society of Chemistry 2011 Lab Chip, 2011, 11, 2481–2484 | 2483
as soon as the Bi
+
beam pass through silicon nitride layer, and H
signal only shows a small jump, as low as 80%. At the initial punch-
through, probably only a small and irregular hole is formed, and the
size of the hole becomes larger and larger with additional doses of Bi
+
ions, and finally the hole size becomes about 3 mmsizeandtheD
and
H
signals become constant. Positive ion spectra (not shown here)
show similar behavior. Fig. 4 (inset) shows the optical image of a ToF-
SIMS-drilled hole after Bi
+
bombardment. It is nearly as round as that
made by FIB.
To conclude, we demonstrated that liquid surfaces can be studied
in situ by advanced vacuum based techniques such as SEM/EDX and
ToF-SIMS using the microfluidic interface assembly developed in our
group. The ToF-SIMS has been applied to study room temperature
aqueous surfaces for the first time. It is a significant technical
breakthrough to broaden the applications of vacuum based surface
techniques to study liquid surfaces of importance in various areas.
This development opens new avenues for understanding interfacial
phenomena occurring on liquid surfaces in the future.
Acknowledgements
We are grateful for help from Bruce Arey of the EMSL, in FIB
fabrication. Support from a Department of Energy (DOE) Division
of Chemical Sciences, Geosciences, and Biosciences (BES Chemical
Sciences grant, KC-0301020-16248) is gratefully acknowledged. The
research was performed in the W. R. Wiley Environmental Molec-
ular Sciences Laboratory (EMSL), a national scientific user facility
sponsored by the DOE’s Office of Biological and Environmental
Research (OBER) and located at the Pacific Northwest National
Laboratory. Pacific Northwest National Laboratory is operated for
DOE by Battelle.
References
1 N. Altimir, P. Kolari, J.-P. Tuovinen, T. Vesala, J. Back, T. Suni,
M. Kulmala and P. Hari, Biogeosciences, 2006, 3, 209–228.
2 A. L. Sumner, E. J. Menke, Y. Dubowski, J. T. Newberg,
R. M. Penner, J. C. Hemminger, L. M. Wingen, T. Brauers and
B. J. Finlayson-Pitts, Phys. Chem. Chem. Phys., 2004, 6, 604–613.
3 S. C. Ringwald and J. E. Permberton, Environ. Sci. Technol., 2000, 34,
259–265.
4 M. Faubel, S. Schlemmer and J. P. Toennies, Z. Phys. D: At., Mol.
Clusters, 1988, 10, 269–277.
5 B. Winter and M. Faubel, Chem. Rev., 2006, 106, 1176–1211.
6 G. D. Danilatos, Scanning, 1981, 4, 9–20.
7 J. Pantfoerder, S. Poellmann, J. F. Zhu, D. Borgmann, R. Denecke
and H.-P. Steinrueck, Rev. Sci. Instrum., 2005, 76, 014102–014109.
8 S. Ghosal, J. C. Hemminger, H. Bluhm, B. S. Mun,
E. L. D. Hebenstreit, G. Ketteler, D. F. Ogletree, F. G. Requejo
and M. Salmeron, Science, 2005, 307, 563–566.
9 F. Banhart, In situ Electron Microscopy at High Resolution, World
Scientific Publishing Co. Pte. Ltd., Singapore, 2008, pp. 229–230.
10 A. Bogner, G. Thollet, D. Basset, P.-H. Jouneau and C. Gauthier,
Micron, 2007, 38, 390–401.
11 M. N. R. Brivio, M. H. Tas, H. J. G. E. Goedbloed, W. Gardeneiers,
A. V. D. Verboom and D. N. R. Berg, Lab Chip, 2005, 5, 378–
381.
12 S. J. Metz, W. J. C. Ven van de, J. Potreck, M. H. V. Mulder and
M. Wessling, J. Membr. Sci., 2005, 251, 29–41.
13 G. C. Randall and P. S. Doyle, Proc. Natl. Acad. Sci. U. S. A., 2005,
102, 10813–10818.
14 J. C. Mcdonald and G. M. Whitesides, Acc. Chem. Res., 2002, 35,
491–499.
15 B. L. Thiel and M. Toth, J. Appl. Phys., 2005, 97, 051101–051118.
16 E. J. Davis, Atmos. Res., 2006, 82, 561–578.
17 M. C. Porter, in Handbook of Industrial Membrane Technology, Noyes
Publications, USA, 1990, p. 71.
18 J. Crank, The Mathematics of Diffusion, Clarendon Press, Oxford,
2nd edn, 1979, p. 102.
19 L. Xu and M. Salmeron, Langmuir, 1998, 14, 5841–5844.
Fig. 4 H
and D
signals vs. Bi
+
ion erosion time, while making a hole
into the channel in the ToF-SIMS. An optical image (10 10 mm
2
) of the
hole created is also shown (inset).
2484 | Lab Chip, 2011, 11, 2481–2484 This journal is ªThe Royal Society of Chemistry 2011
... The SALVI microreactor was composed of polydimethylsiloxane (PDMS) using soft lithography [42,43]. Briefly, the main reactions took place in a microchannel that was 200 μm wide and 300 μm deep, encapsulated in a PDMS block (Figure 1b). ...
... The devices were covered with aluminum foil (Figure 1b) and kept away from light for a series of aging times to simulate dark aging. For UV aging, the reactant solution was first injected into the SALVI microreactor channel [42,43] and the microreactor was set 10 cm below a UV illumination source (Oriel lamp model 6035, power supply model 6060, Franklin, MA, USA) for different periods of time [45]. The lamp current was 18 ± 5 mA. ...
... Normalization was performed using selected products' total ion counts. A ToF-SIMS V-100 instrument (IONTOF GmbH, Münster, Germany) with a 25 keV Bi 3 + ion beam was used to detect the interfacial products [42,43]. All results were acquired in replicates to ensure data quality and reproducibility. ...
... Our group has developed microfluidic cells for the multimodal spectroscopy and microscopy of liquids [34-36], called the system for analysis at the liquid-vacuum interface (SALVI). The electrochemical version of the SALVI cell, or the E-cell, contains three electrodes (working, counter-, or reference electrodes, or WE, CE, and RE), and is integrable to a suite of spectroscopy and imaging techniques for in situ and operando analysis [34][35][36]. ...
... Parts of these devices were fabricated using soft lithography. More details on SALVI fabrication are available in the Supplementary Information section and previous reports [34,36]. [34,36]. ...
... More details on SALVI fabrication are available in the Supplementary Information section and previous reports [34,36]. [34,36]. ...
Article
Full-text available
Understanding the corrosion of spent nuclear fuel is important for the development of long-term storage solutions. However, the risk of radiation contamination presents challenges for experimental analysis. Adapted from the system for analysis at the liquid–vacuum interface (SALVI), we developed a miniaturized uranium oxide (UO2)-attached working electrode (WE) to reduce contamination risk. To protect UO2 particles in a miniatured electrochemical cell, a thin layer of Nafion was formed on the surface. Atomic force microscopy (AFM) shows a dense layer of UO2 particles and indicates their participation in electrochemical reactions. Particles remain intact on the electrode surface with slight redistribution. X-ray photoelectron spectroscopy (XPS) reveals a difference in the distribution of U(IV), U(V), and U(VI) between pristine and corroded UO2 electrodes. The presence of U(V)/U(VI) on the corroded electrode surface demonstrates that electrochemically driven UO2 oxidation can be studied using these cells. Our observations of U(V) in the micro-electrode due to the selective semi-permeability of Nafion suggest that interfacial water plays a key role, potentially simulating a water-lean scenario in fuel storage conditions. This novel approach offers analytical reproducibility, design flexibility, a small footprint, and a low irradiation dose, while separating the α-effect. This approach provides a valuable microscale electrochemical platform for spent fuel corrosion studies with minimal radiological materials and the potential for diverse configurations.
... The SALVI microreactor was composed of polydimethylsiloxane (PDMS) using soft lithography [42,43]. Briefly, the main reactions took place in a microchannel that was 200 μm wide and 300 μm deep, encapsulated in a PDMS block (Figure 1b). ...
... The devices were covered with aluminum foil (Figure 1b) and kept away from light for a series of aging times to simulate dark aging. For UV aging, the reactant solution was first injected into the SALVI microreactor channel [42,43] and the microreactor was set 10 cm below a UV illumination source (Oriel lamp model 6035, power supply model 6060, Franklin, MA, USA) for different periods of time [45]. The lamp current was 18 ± 5 mA. ...
... Normalization was performed using selected products' total ion counts. A ToF-SIMS V-100 instrument (IONTOF GmbH, Münster, Germany) with a 25 keV Bi 3 + ion beam was used to detect the interfacial products [42,43]. All results were acquired in replicates to ensure data quality and reproducibility. ...
Article
Full-text available
In this work, the relative yields of aqueous secondary organic aerosols (aqSOAs) at the air–liquid (a–l) interface are investigated between photochemical and dark aging using in situ time-of-flight secondary ion mass spectrometry (ToF-SIMS). Our results show that dark aging is an important source of aqSOAs despite a lack of photochemical drivers. Photochemical reactions of glyoxal and hydroxyl radicals (•OH) produce oligomers and cluster ions at the aqueous surface. Interestingly, different oligomers and cluster ions form intensely in the dark at the a–l interface, contrary to the notion that oligomer formation mainly depends on light irradiation. Furthermore, cluster ions form readily during dark aging and have a higher water molecule adsorption ability. This finding is supported by the observation of more frequent organic water cluster ion formation. The relative yields of water clusters in the form of protonated and hydroxide ions are presented using van Krevelen diagrams to explore the underlying formation mechanisms of aqSOAs. Large protonated and hydroxide water clusters (e.g., (H2O)nH+, 17 < n ≤ 44) have reasonable yields during UV aging. In contrast, small protonated and hydroxide water clusters (e.g., (H2O)nH+, 1 ≤ n ≤ 17) form after several hours of dark aging. Moreover, cluster ions have higher yields in dark aging, indicating the overlooked influence of dark aging interfacial products on aerosol optical properties. Molecular dynamic simulation shows that cluster ions form stably in UV and dark aging. AqSOAs molecules produced from dark and photochemical aging can enhance UV absorption of the aqueous surface, promote cloud condensation nuclei (CCN) activities, and affect radiative forcing.
... Yu and coworkers developed a microfluidic device to overcome this challenge [71]. The device, termed SALVI, a system for analysis at the liquid-vacuum interface, is capable of maintaining liquids and living biological specimens with high vapor pressure in highvacuum instruments, such as a scanning electron microscope or a time-of-flight mass analyzer [71]. ...
... Yu and coworkers developed a microfluidic device to overcome this challenge [71]. The device, termed SALVI, a system for analysis at the liquid-vacuum interface, is capable of maintaining liquids and living biological specimens with high vapor pressure in highvacuum instruments, such as a scanning electron microscope or a time-of-flight mass analyzer [71]. The SALVI microfluidic device has been used to study plant-microbe interactions by mass spectrometry imaging [35,42,46,68,72]. ...
Article
Full-text available
Plant–microbe interactions are of rising interest in plant sustainability, biomass production, plant biology, and systems biology. These interactions have been a challenge to detect until recent advancements in mass spectrometry imaging. Plants and microbes interact in four main regions within the plant, the rhizosphere, endosphere, phyllosphere, and spermosphere. This mini review covers the challenges within investigations of plant and microbe interactions. We highlight the importance of sample preparation and comparisons among time-of-flight secondary ion mass spectroscopy (ToF-SIMS), matrix-assisted laser desorption/ionization (MALDI), laser desorption ionization (LDI/LDPI), and desorption electrospray ionization (DESI) techniques used for the analysis of these interactions. Using mass spectral imaging (MSI) to study plants and microbes offers advantages in understanding microbe and host interactions at the molecular level with single-cell and community communication information. More research utilizing MSI has emerged in the past several years. We first introduce the principles of major MSI techniques that have been employed in the research of microorganisms. An overview of proper sample preparation methods is offered as a prerequisite for successful MSI analysis. Traditionally, dried or cryogenically prepared, frozen samples have been used; however, they do not provide a true representation of the bacterial biofilms compared to living cell analysis and chemical imaging. New developments such as microfluidic devices that can be used under a vacuum are highly desirable for the application of MSI techniques, such as ToF-SIMS, because they have a subcellular spatial resolution to map and image plant and microbe interactions, including the potential to elucidate metabolic pathways and cell-to-cell interactions. Promising results due to recent MSI advancements in the past five years are selected and highlighted. The latest developments utilizing machine learning are captured as an important outlook for maximal output using MSI to study microorganisms.
... In situ liquid SIMS is enabled by the system for analysis at liquid vacuum interface (SALVI) [35,36], a microfluidic reactor compatible for vacuum instruments (Fig. 3c). Because ToF -SIMS is a mass spectral imaging technique, liquid SIMS can provide 1D, 2D, and 3D chemical information on the top few nm of the water surface. ...
Article
Formation of aqueous secondary organic aerosol (aqSOA) at the air–liquid interface recently has attracted a lot of attention in atmospheric chemistry. The discrepancies in mass distributions, aerosol oxidative capacity, liquid water content, hygroscopic growth of aerosols, and formation of clouds and fogs suggest that interfacial chemistry play a more important role than previously deemed. However, detailed mechanisms at the air–water interface remain unclear owing to the lack of comprehensive understanding that underpins complicated interfacial phenomena, which are not easily measurable from field campaigns and laboratory measurements or computational simulations. This review highlights relevant and recent technical advancement employed to study aqSOA encompassing spectroscopy and mass spectrometry. The current knowledge on the aqSOA processes is digested with an emphasis on recent research of interfacial aqSOA formation including laboratory studies and model simulations. Finally, future directions of the interfacial chemistry are recommended for field and laboratory studies as well as theoretical efforts to resolve interfacial challenges in atmospheric chemistry.
... Microfluidics is a viable approach to address the associated technical challenges. Our team developed a microfluidic-based platform for the multimodal spectroscopy and microscopy of liquids [11][12][13]. This vacuum-compatible platform is named System for Analysis at the Liquid-Vacuum Interface (SALVI). ...
Article
Full-text available
We developed a new approach to attach particles onto a conductive layer as a working electrode (WE) in a microfluidic electrochemical cell with three electrodes. Nafion, an efficient proton transfer molecule, is used to form a thin protection layer to secure particle electrodes. Spin coating is used to develop a thin and even layer of Nafion membrane. The effects of Nafion (5 wt% 20 wt%) and spinning rates were evaluated using multiple sets of replicates. The electrochemical performance of various devices was demonstrated. Additionally, the electrochemical performance of the devices is used to select and optimize fabrication conditions. The results show that a higher spinning rate and a lower Nafion concentration (5 wt%) induce a better performance, using cerium oxide (CeO2) particles as a testing model. The WE surfaces were characterized using atomic force microscopy (AFM), scanning electron microscopy-focused ion beam (SEM-FIB), time-of-flight secondary ion mass spectrometry (ToF-SIMS), and X-ray photoelectron spectroscopy (XPS). The comparison between the pristine and corroded WE surfaces shows that Nafion is redistributed after potential is applied. Our results verify that Nafion membrane offers a reliable means to secure particles onto electrodes. Furthermore, the electrochemical performance is reliable and reproducible. Thus, this approach provides a new way to study more complex and challenging particles, such as uranium oxide, in the future.
Article
Full-text available
Secondary Ion Mass Spectrometry (SIMS) is an outstanding technique for Mass Spectral Imaging (MSI) due to its notable advantages, including high sensitivity, selectivity, and high dynamic range. As a result, SIMS has been employed across many domains of science. In this review, we provide an in-depth overview of the fundamental principles underlying SIMS, followed by an account of the recent development of SIMS instruments. The review encompasses various applications of specific SIMS instruments, notably static SIMS with time-of-flight SIMS (ToF-SIMS) as a widely used platform and dynamic SIMS with Nano SIMS and large geometry SIMS as successful instruments. We particularly focus on SIMS utility in microanalysis and imaging of metals and alloys as materials of interest. Additionally, we discuss the challenges in big SIMS data analysis and give examples of machine leaning (ML) and Artificial Intelligence (AI) for effective MSI data analysis. Finally, we recommend the outlook of SIMS development. It is anticipated that in situ and operando SIMS has the potential to significantly enhance the investigation of metals and alloys by enabling real-time examinations of material surfaces and interfaces during dynamic transformations.
Article
Molecular structure conversion concomitant with mass transfer processes at the electrode-electrolyte interfaces plays a central role in energy electrochemistry. Mass spectrometry, as one of the most intuitive, sensitive techniques, provides the capability to collect transient intermediates and products and uncover reaction mechanisms and kinetics. In situ time-of-flight secondary ion electrochemical mass spectrometry with inherent high mass and spatiotemporal resolution has emerged as a promising strategy for investigating electrochemical processes at the electrode surface. This review illustrates the recent advancements in coupling time-of-flight secondary ion mass spectrometry and electrochemistry to visualize and quantify local dynamic electrochemical processes, identify solvated species distribution, and disclose hidden reaction pathways at the molecular level. Moreover, the key challenges in this field are further discussed to promote new applications and discoveries in operando studying the dynamic electrochemical interfaces of advanced energy systems.
Article
Full-text available
Scanning Electron Microscopy (SEM), while being amongst the most widespread analytical instrumentation, is not widely used to study nucleation and growth (NG) phenomena in liquids. This is, partially due to insufficient exposure of the electrochemical research community to its capabilities. Here, we report on a simple but versatile custom-made setup for liquid phase (LP) SEM to access chemically and electrochemically driven NG processes in liquids. In addition, we will reveal the experimental artifacts and limitations of the technique related to radiation damage of the liquids. Finally, we will discuss a few recent developments in beam damage-free LP SEM imaging in liquids.
Article
A quantitative description on dispersity of boehmite (γ-AlOOH) particles, a key component for waste slurry at Hanford sites, can provide useful knowledge for understanding various physicochemical nature of the waste. In situ liquid scanning electron microscopy (SEM) was used to evaluate the dispersity of particles in aqueous conditions using a microfluidic sample holder, System for Analysis at Liquid Vacuum Interface (SALVI). Secondary electron (SE) images and image analyses were performed to determine particle centroid locations and the distance to the nearest neighbour particle centroid, providing reliable rescaled interparticle distances as a function of ionic strength in acidic and basic conditions. Our finding of the particle dispersity is consistent with physical insights from corresponding particle interactions under physicochem-ical conditions, demonstrating delicate changes in dispersity of boehmite particles based on novel in situ liquid SEM imaging and analysis.
Article
Full-text available
of a paper presented at Microscopy and Microanalysis 2010 in Portland, Oregon, USA, August 1 – August 5, 2010.
Article
Progress in designing and constructing an atmospheric or environmental scanning electron microscope (SEM) is reported. The introduction of vacuum pumping between the objective and pressure limiting aperture has allowed the use of relatively large pressure limiting apertures, i.e., up to 57 μm for operation at atmospheric pressure or up to 400 μm for operation at saturation water vapour pressure and room temperature. The imaging obtained has been considerably improved by these developments. The first part of experimentation and analysis on the vacuum characteristics of the new system together with different detection configurations is also presented.
Article
Low vacuum scanning electron microscopy (SEM) is a high-resolution technique, with the ability to obtain secondary electron images of uncoated, nonconductive specimens. This feat is achieved by allowing a small pressure of gas in the specimen chamber. Gas molecules are ionized by primary electrons, as well as by those emitted from the specimen. These ions then assist in dissipating charge from the sample. However, the interactions between the ions, the specimen, and the secondary electrons give rise to contrast mechanisms that are unique to these instruments. This paper summarizes the central issues with charging and discusses how electrostatically stable, reproducible imaging conditions are achieved. Recent developments in understanding the physics of image formation are reviewed, with an emphasis on how local variations in electronic structure, dynamic charging processes, and interactions between ionized gas molecules and low-energy electrons at and near the sample surface give rise to useful contrast mechanisms. Many of the substances that can be examined in these instruments, including conductive polymers and liquids, possess charge carriers having intermediate mobilities, as compared to metals and most solid insulators. This can give rise to dynamic contrast mechanisms, and allow for characterization techniques for mapping electronic inhomogeneities in electronic materials and other dielectrics. Finally, a number of noteworthy application areas published in the literature are reviewed, concentrating on cases where interesting contrast has been reported, or where analysis in a conventional SEM would not be possible. In the former case, a critical analysis of the results will be given in light of the imaging theory put forth.
Article
Potassium mica was treated with different ionic solutions to replace the naturally occurring K+ on the surface by Ca2+, Mg2+, and H+ ions. The extent of the exchange was monitored by variable emergence angle X-ray photoelectron spectroscopy (XPS). Scanning polarization force microscopy (SPFM) was used to measure the mobility of the surface ions as a result of water adsorption when the mica is exposed to different humidity levels.
Article
Raman and FTIR spectroscopies are used to investigate the sorption mechanisms of benzene, toluene, and 2- and 4-picoline onto silica as models for volatile aromatic pollutant interactions with a soil constituent. Benzene and toluene vapor adsorption on silica occurs via weak π-system−hydrogen bonding with silanols on the silica surface. This weak interaction would likely result in low vadose zone retention, especially in wet conditions where water adsorption would successfully compete for surface sites. The vapor adsorption of 2- and 4-picoline (2- and 4-methylpyridine) is studied to model aza-arene environmental contaminants. These species adsorb to surface silanols by a more specific and stronger hydrogen-bonding mechanism involving the lone pair electrons on the N atom. The strength of these interactions is probably sufficient to result in their retention by dry or damp vadose zone soil, slowing their transport. This work illustrates the utility of these vibrational spectroscopic techniques in elucidating specific surface interactions of pollutants with mineral oxides and in helping to predict the fate of pollutants in the environment.
Article
A method to maintain a clean surface of a liquid in a high vacuum is described. Using a very thin and fast liquid jet it is not only possible to prevent freezing of the liquid but also to reduce the number of collisions between evaporating molecules to negligibly small values. Thus many of the standard, vacuum dependent, particle probing techniques for solid surfaces can be used for studies of rapidly vaporizing, high vapor pressure liquids. In a first molecular beam investigation we have used time-of-flight analysis to measure the velocity distribution of H2O molecules vaporizing from thin jets of pure liquid water. The experiments were carried out for liquid jet diameters between 50 and 5 m. In this range the expanding vapor is observed to undergo the transition to the collision-free molecular flow regime. From the measured velocity distributions the local surface temperature is determined to be less than 210 K. This appears to be the lowest temperature ever reported for supercooled liquid water.
Article
This Account summarizes techniques for fabrication and applications in biomedicine of microfluidic devices fabricated in poly(dimethylsiloxane) (PDMS). The methods and applications described focus on the exploitation of the physical and chemical properties of PDMS in the fabrication or actuation of the devices. Fabrication of channels in PDMS is simple, and it can be used to incorporate other materials and structures through encapsulation or sealing (both reversible and irreversible).
Article
In an effort to extend the pressure range for electron-based spectroscopies from ultrahigh vacuum into the so-called pressure gap region, we have built a new apparatus for in situ x-ray photoelectron spectroscopy up to 1 mbar . The principle of the experimental setup is based on a modified hemispherical electron energy analyzer, a modified twin anode x-ray source, and several differential-pumping stages between sample region and electron detection. The reaction gas is provided in situ either by background dosing or, as a new feature, by beam dosing, using a directed gas beam from a small tube. The latter allows for higher local pressures. The performance of the new setup is discussed, deriving normalization procedures from the analysis of the attenuation of the substrate photoemission intensity by the increasing gas phase pressure. In addition, the change of the work function due to changes in surface composition can be evaluated in situ by analyzing the binding energy shift of the gas phase core-level peaks. As a first study, measurements for the pressure dependence of CO adsorption on Pd(111) between 5×10<sup>-8</sup> and 1 mbar are presented.
Article
This paper reviews theory and measurements of transport processes between small particles and the surrounding gas. Evaporation and condensation coefficients and gas uptake coefficients are of particular interest. There has long been a great difference in coefficients reported by different experimentalists, and much of this disagreement is considered in this overview. A brief review of the kinetic theory of gases is provided to describe molecular transport to or from a surface when the mean free path of molecules, ℓ, is large compared with the particle dimensions, that is, when the Knudsen number is large. For a sphere of radius a the Knudsen number, Kn = ℓ / a. The condition Kn >> 1 is called the free molecule regime, and for Kn << 1 continuum theory applies. It is shown that accommodation coefficients cannot be determined by experiments operating in the continuum regime. At intermediate Knudsen numbers (the Knudsen regime) transport theory is more difficult, but results based on solution of the Boltzmann equation describing the evolution of the molecular velocity distribution are reviewed. With the advent of high-speed computers transport theory has been supplemented by molecular dynamics calculations, and these calculations are often at odds with experimental measurements of accommodation coefficients. Examples are provided.