Advertisement

If you have an ACS member number, please enter it here so we can link this account to your membership. (optional)

ACS values your privacy. By submitting your information, you are gaining access to C&EN and subscribing to our weekly newsletter. We use the information you provide to make your reading experience better, and we will never sell your data to third party members.

ENJOY UNLIMITED ACCES TO C&EN

Analytical Chemistry

What could chemistry do with more expensive instruments?

Workshops conclude chemistry could benefit from NSF mid-scale instrumentation program, which funds instrument projects with price tags between $4 million and $70 million

by Celia Henry Arnaud
March 27, 2017 | A version of this story appeared in Volume 95, Issue 13

 

Close-up photo of part of a ultrahigh-vacuum scanning probe microscope.
Credit: Flickr/EMSL
At two NSF workshops, chemists concluded that developing new types of instruments and colocating existing ones could enable breakthrough research in chemistry’s “grand challenges.” Shown here is an ultrahigh-vacuum scanning probe microscope.

Chemistry and physics differ not only in the scales of the systems they study but also in the scales of the instruments they use for research. Physicists and astronomers work on some large projects involving multiple facilities and big instruments. For example, the Laser Interferometer Gravitational-Wave Observatory (LIGO) detected gravitational waves for the first time at two multi-kilometer-long facilities in the U.S. Meanwhile, much of chemistry research still happens with instruments that sit in individual researchers’ labs.

The National Science Foundation underwrites much of the research infrastructure in the U.S. at both the small and large scales. NSF’s Major Research Instrumentation Program will pay for instruments up to $4 million. That’s the program scientists apply to if they’re looking for instruments for individual labs or departments. In contrast, NSF’s Major Research Equipment and Facilities Construction funds much larger projects. It funded LIGO to the tune of about $1.1 billion.

NSF is starting to bridge the gap between the two with a focus on “midscale” instrumentation. Moves are afoot in the chemistry community to try to capture some of that funding. But first the community needs to figure out whether it can effectively use funding at such scales.

Grand challenges

At a workshop on mid-scale instrumentation sponsored by the National Science Foundation’s Division of Chemistry, participants came up with six “grand challenges” in chemistry that could be addressed by bringing together existing instruments at regional centers. Here are the six research areas with two examples of each challenge.

Structure and dynamics at interfaces
Charge-transfer processes
Interactions of biomolecules with solid surfaces

Highly parallel chemical synthesis and characterization
New catalysts
New chemistries for solar cells

Transient intermediates
Transition-state complexes
Surface-molecule complexes

Structure-function relationships in disordered and/or heterogeneous systems
Amorphous materials, such as polymers, glasses, and ceramics
Membrane-bound proteins

Multiscale dynamics of complex systems: integrating transport with reaction
Biogeochemical cycling
Atmospheric chemistry

New science arising from the characterization of heterogeneous mixtures
Identification of metabolites
Analysis of crude oil

Last fall, two workshops sponsored by NSF’s Division of Chemistry explored the needs for chemistry instrumentation in this midscale price range. At the Pittsburgh Conference on Analytical Chemistry & Applied Spectroscopy, held earlier this month in Chicago, workshop organizers gave the chemistry community a sneak peek at the reports from those workshops, which will come out soon.

The two workshops approached the issue from different angles. The first one concentrated on ways existing instrumentation could be combined at regional facilities to enable breakthrough research. The other focused on the need for new complex instruments and opportunities for developing them.

“Chemists aren’t used to thinking about instrumentation at this scale,” said Robert J. Hamers, a chemistry professor at the University of Wisconsin, Madison. He co-organized the workshop on regional facilities with Graham F. Peaslee of the University of Notre Dame and Sophia E. Hayes of Washington University in St. Louis.

“Most chemists tend to think in terms of individual projects where the instrumentation needs are more modest and where the instruments might reside largely in their own lab,” Hamers said. “The idea of combining several instruments—$500,000 to $1 million each—into a larger super-instrument in a way that they all interrogate the same sample is not something that most people think about, in large part because historically the resources have not been available.”

At the first workshop, participants asked how the chemistry community could use such resources if they became available. They identified six “grand challenges” for which colocating diverse instruments in regional facilities could advance science (see box on page above). Those challenges deal with systems ranging from the very small—the structure and dynamics of nanoscale interfaces—to the very large—molecular/chemical transport and reactions in complex systems such as the environment and the atmosphere.

Chemistry is ready for this.
Angela K. Wilson, director of NSF’s Division of Chemistry

For each challenge, workshop participants considered multiple questions. What is the best number of instruments to include? How should those instruments be integrated? Too many instruments poorly stitched together could result in a “Franken-instrument” that is too difficult to use, Hamers warned.

And simply having the instruments in the same facility might not be enough. They might need to be connected in such a way that a sample can be moved seamlessly among them. Or the instruments might even need to be set up so that they analyze a sample simultaneously.

Workshop participants identified the measurement of the structure and dynamics of interfaces between chemical systems as one grand challenge, Hamers said. A number of issues make such measurements difficult. Notably, the relevant behavior is typically confined to interfaces that are no more than a few nanometers thick, and sometimes even smaller. These interfaces sometimes are buried in hard-to-reach places. Often, only a small amount of material is available for analysis, and the phenomenon of interest might be happening at only a few sites in that sample.

Hamers further illustrated that grand challenge using the specific example of chemistry at electrode-electrolyte interfaces in batteries. These interfaces present a number of unanswered scientific questions: Do specific molecules in a complex solvent mixture preferentially adsorb at electrode interfaces? How do liquid-phase molecules orient near this interface? What are the mechanisms of ion solvation and desolvation? What are the compositions, structures, and mechanisms of formation of the layers that form between the electrode and the electrolyte?

To answer those questions, multiple instruments could be brought to bear in a regional facility, Hamers said. Of course, electrochemistry could be used to characterize electrode behavior. One or more optical probes, using methods such as sum frequency generation, second harmonic generation, and Raman or infrared spectroscopy, could provide structural information specific to the interface. An X-ray synchrotron beamline could be used for sensitive structural and elemental analysis. Both solid-state and high-resolution solution nuclear magnetic resonance (NMR) spectroscopy could provide information about molecular structure.

But building such regional facilities is only the beginning. Such centers must be organized to ensure their sustainability when the initial funding runs out, Hamers said. Ongoing operation will require steady funding for experienced staff and instrument maintenance. And that requirement means potential users must be ready to pay for use of the facility.

Whereas the first workshop considered combining existing instruments, the second one focused on research areas that could benefit from entirely new instruments. Paul W. Bohn of the University of Notre Dame and Marcos Dantus of Michigan State University organized the second workshop, which Bohn described during the symposium at Pittcon.

Before the workshop, meeting participants were asked to dream up ideas for new chemical instrumentation, Bohn said. During the workshop, participants considered the proposals in the context of NSF’s “10 Big Ideas for Future NSF Investments” and whittled the list down to three project concepts that they thought had the best chances of transforming chemical research. The participants estimated that each of these three concepts would cost between $12 million and $30 million over five years.

One proposal was to develop high-flux ion sources as tools for synthetic chemistry. Scientists could use such ion sources to accelerate organic synthesis in charged droplets and to tailor surfaces for new catalysts and biomaterials.

Participants identified “mapping the dark molecules of life” as another high-impact area. They defined dark molecules as the many proteins and metabolites whose structures and functions have yet to be identified. Bringing these molecules out of the dark will require mass spectrometers and NMR spectrometers with sensitivity and spectral resolution that is greater than what is currently available. In the case of NMR, it could compel the development of instruments able to detect nuclei that are currently difficult to observe, such as 17O, 25Mg, and 33S.

For the third concept, participants proposed an instrument they dubbed the “quantum explorer.” It would bridge the gap between experiment and computation and require the development of lasers producing nonclassical forms of light that can interact with matter in previously impossible ways. This quantum explorer could set up a feedback loop between experiment and theory that accelerates the development and understanding of new chemical reactions, the speakers said.

Neither workshop was intended to come up with detailed proposals for midscale instrumentation projects. Rather, they were intended to determine whether the chemistry community has the desire and need to undertake such projects, which, in the context of most chemistry research, are large. The workshops demonstrated that there are indeed needs that fall outside the bounds of current funding programs, the speakers noted.

At the symposium, Peaslee described how the projects proposed in both workshops could help NSF achieve some of its desired “broader impacts.” Such instruments and facilities are particularly good for developing scientific talent and for engaging a wider audience. High-end instrumentation attracts students and the broader public through outreach events at facilities, Peaslee said.

Angela K. Wilson, director of NSF’s Division of Chemistry, said the chemistry division would be pleased to see such projects come to fruition. But she cautioned that NSF’s funding situation right now is uncertain. In addition, she said, NSF’s budget has been flat for about 10 years. Any new initiatives could require taking money from somewhere else.

Although the chemistry community sees value in such large-scale, collaborative projects, it doesn’t want that funding to come at the expense of grants for individual investigators, Hamers said.

“Chemistry is ready for this,” Wilson said. “But we’re not going to do 10 of them.” If funding can be found, the division is likely to take a slow and steady approach. “We’d do one and then another one.” 

Article:

This article has been sent to the following recipient:

0 /1 FREE ARTICLES LEFT THIS MONTH Remaining
Chemistry matters. Join us to get the news you need.