National Academies Press: OpenBook

A Research Agenda for Transforming Separation Science (2019)

Chapter: Appendix C: Examples of Characterization

« Previous: Appendix B: Meeting and Webinar Agendas
Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×

C

Examples of Characterization

CHARACTERIZING STRUCTURES FROM THE ATOMIC SCALE TO THE NANOSCALE

One example of how the state-of-the-science approaches described in this report can be used to gain detailed insight into structures from the atomic scale to the nanoscale is the exploitation of scattering to access information on interatomic correlations. Atom–atom correlations are obtained by a Fourier transformation of scattering data to obtain pair distribution functions, wherein peak positions represent correlation distances and peak intensities are related to coordination numbers (Egami and Billinge, 2003). Obtaining adequate peak resolution to understand chemical speciation requires the use of a spallation neutron source or a synchrotron.

With this method, one can characterize adsorbed gases in nanoporous membrane materials (Gallis et al., 2017) and study interpenetrating networks of relevance to carbon dioxide sequestration (Mulfort et al., 2010). High-energy x-ray scattering (energy > 60 keV) patterns have been used to determine thermodynamic stability constants for uranyl chloride complexation (see Figure C-1). The advantage of this approach over more standard techniques is that the results provide a direct link between the targeted ion’s coordination structure and thermodynamic free energies (Soderholm et al., 2011). Further development and use of this approach can provide modelers with the structural information necessary to understand the thermodynamic energy changes that drive mass transfer.

Extended x-ray absorption fine structure (EXAFS) spectroscopy has become a tool of choice to elucidate the local structure and bonding of target species. An exemplary study to identify an ionic liquid-based solvent to extract cobalt (Co) from samarium (Sm) used EXAFS to determine that the extraction occurs mainly via an ion pair rather than an ion-exchange mechanism (Sobekova Foltova et al., 2019). The competing roles of chloride, thiocyanate, and nitrate anions for complexation with Sm and Co were determined by comparing the coordination environments of these metals after extraction into the ionic liquid. The results provide direction for further development of an environmental-friendly approach to the recycling and recovery of high-purity metals from wasted permanent magnets. A similar study, coupling small-angle neutron scattering (SANS) with EXAFS, demonstrated the important role of extended aggregates and solvent structuring in the performance of a task-specific ionic liquid designed for the targeted extraction of trivalent lanthanides from spent nuclear-fuel waste (Abney et al., 2017).

CHARACTERIZING STRUCTURE, DYNAMICS, AND FIELDS AT INTERFACES

Experimental techniques that probe the speciation of surfaces and their adsorbates have been largely restricted to high-vacuum environments and cryogenic temperatures. Questions have been continually raised about the applicability of the experimental results to operando systems. Complicating the difficulties for liquid–solid and liquid–liquid processes is that changes in the targeted moiety’s speciation occur across a buried phase boundary, a problem not limited to separation science but relevant to a wide variety of chemical and physical systems from biological membranes to catalysis to geochemical transport. The result of the widespread need for experimental probes of interfaces under operando environments has been the recent development of a wide variety of techniques that yield information on the structure of interfaces on the atomic scale.

The opportunities for the separations community are just now being conceptualized, with recent papers reporting interesting and underappreciated techniques for studying the structure of interfaces that might eventually provide a mechanism for controlling a separation outcome. For example, studies have demonstrated an unexpected two-step process for PtCl62- adsorption at an air–water interface (Uysal et al., 2017). A recent study hints at the power of a multimodal approach to surface studies; it combines x-ray surface techniques with sum frequency generation to reveal the surface-water structure in a system related to solvent extraction (Rock et al., 2018). The combination of x-ray and neutron

Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×
Image
FIGURE C-1 Pair distribution functions obtained from a series of uranyl solutions as a function of chloride ion concentration. Analyses of these patterns provide stability constants for uranyl chloride complexation, which is linked directly with structure. SOURCE: Adapted from Soderholm et al., 2011. Reprinted with permission; copyright 2011, Journal of Physical Chemistry A.

reflectivity has been shown to be a powerful probe of a separation interface. Using that approach, researchers identified the mixing of water and extractant molecules within an interphase region in an oil and water solvent extraction system (Scoppola et al., 2015).

INTERFACES WITH SOLIDS

Accessing atomic-level detail of binding at sites at a solid interface under operando conditions is often complicated by multiscale roughness or porosity, the presence of multiple reactive site types, and reactive behavior convoluted with transport processes, such as advection and diffusion. Similar challenges are encountered in other disciplines, such as heterogeneous catalysis, electrochemistry, and geochemistry. Those disciplines have helped to advance characterization techniques that overcome such obstacles in situ.

For example, researchers have made substantial progress in probing the multiscale structure of solid surfaces down to the atomic scale directly in a fluid medium of interest. Dynamics can also be probed. For example, three-dimensional crystal truncation rod x-ray scattering techniques with supporting ab initio molecular dynamics simulations based on density functional theory have disentangled relationships between time-averaged interfacial atomic structure and dynamic processes, such as ligand exchange in complex aqueous electrolytes at a given pH (McBriarty et al., 2017, 2018). In situ scanning probe methods, such as atomic force microscopy (AFM), can image molecular dynamics at video-rate speeds. For example, AFM with a high scan rate is opening a window into such nanoscale processes as nucleation and growth, molecular translocation, and self-assembly (Kodera et al., 2010; Uchihashi and Scheuring, 2018). Another example is the use of nuclear magnetic resonance (NMR) and molecular in silico simulations using coarse-grained ligand docking to determine which region of a protein binds to multimodal ligand chromatographic beads (Chung et al., 2010).

Under operando conditions, the average interfacial structure might depend on the interplay of relatively static atoms at the solid surface with relatively dynamic molecules in an overlying fluid. Time-averaged structure captured by x-ray scattering can now be scrutinized and interpreted in terms of fast molecular dynamic processes, such as water-ligand exchange on metal oxide surfaces using ab initio molecular dynamic simulations (see Figure C-2).

In situ x-ray scattering techniques have been used to probe interface structure under equilibrium conditions. Recently, those techniques were extended into the more challenging case of far-from-equilibrium conditions, revealing how interfacial atomic structure responds to changes in electrostatic potential and adsorbed ion distributions (McBriarty et al., 2018). Conceptually, such measurements can be extended to steady-state flow conditions to examine more complex operando conditions.

Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×
Image
FIGURE C-2 Example of combining synchrotron x-ray scattering techniques, which can now provide information on the time-averaged three-dimensional atomic structure at solid–liquid interfaces (left), with ab initio molecular dynamics simulations of time-dependent processes such as water exchange (right), to produce a comprehensive atomic-structure model for interfaces (middle). SOURCE: Adapted from McBriarty et al., 2017. Reprinted with permission; copyright 2017, American Chemical Society.

LIQUID–LIQUID INTERFACES

Opportunities regarding the liquid–liquid interface (LLI), and in particular probing and understanding the interface structure and dynamics, have arisen. The opportunities include understanding the extent of heterogeneity within the interface and mechanism of transinterface mass transport (Zhou et al., 2017). In principle, understanding of the LLI could be used to design interface structure and properties. The LLI plays three roles in liquid–liquid extraction systems: mass transport, dispersion behavior, and soft-matter interactions. Thus, understanding interfacial structure could be used for kinetic control of selectivity and control of drop formation and coalescence. Understanding of the LLI could affect the effectiveness and utility of liquid–liquid extraction, supported liquid membranes, emulsion liquid membranes, polymer-inclusion membranes, membrane electrodes, membrane-assisted solvent extraction, two-phase aqueous extraction, and extraction chromatography.

X-ray techniques similar to those mentioned above for solid–fluid interfaces, such as x-ray reflectivity and x-ray fluorescence near-total reflection, that are available at synchrotron light sources have allowed a much more detailed inspection of the arrangement of extractant molecules at LLIs (Bu et al., 2014). Optical techniques have also proved valuable as probes to reveal the nature of species and supramolecular ordering at the LLI. In one example, an apparatus called a centrifugal liquid membrane has facilitated analysis by reducing the bulk-phase thicknesses to less than 100 μm (Watarai, 2014). When highly absorbing species, such as porphyrins, are used as extractant molecules, subtle interfacial speciation effects can be observed. In that way, ultraviolet-visible absorbance and fluorescence spectroscopies have provided detailed kinetic and mechanistic information not accessible with x-ray reflectance techniques.

Determining interfacial structure has been difficult. Fluorescence methods have detected single molecules at the LLI and measured lateral diffusion under different conditions. Other informative techniques include Raman spectroscopy, second harmonic generation, vibrational sum frequency generation, and optical chirality methods (Kocsis et al., 2018). When combined with molecular simulations, those techniques can reveal significant detail regarding interfacial structure. For example, recent work has shown that after adsorption, PtCl62- complexes partially retain their first and second hydration spheres, and it is possible to identify three types of water molecules around them on the basis of their orientational structures and hydrogen bonding strengths. Those results have important implications for relating interfacial water structure and hydration enthalpy to the general understanding of specific ion effects (Rock et al., 2018).

A critical component for understanding phase transfer during a liquid–liquid separation is the target ion’s speciation, including any change in ligation when it is bound by the extractant molecule. A recent study answered some questions. It reported the extraction of erbium (Er) from an acidic aqueous phase by using dihexadecyl phosphate (DHDP) dissolved in dodecane as an extractant. The system was used as a model system. Coupling several related x-ray surface techniques with EXAFS spectroscopy and using temperature to arrest the separation process revealed a counterintuitive interfacial arrangement of extractants in an intermediate state: an inverted bilayer of extractants with headgroups complexing and shielding the Er3+ ions from the organic phase. That

Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×

unexpected novel molecular arrangement might provide insights into both the mechanism of charged ion transfer and reverse micelle formation in the organic phase (see Figure C-3) (Bu et al., 2014).

CONFINEMENT BETWEEN INTERFACES

As in other fields, confinement between interfaces can be important in separation processes. For example, many of the proposals to use nanoporous materials to separate ethanol from water in biofuel purification rely on the fact that ethanol responds to confinement entirely differently from water (Wang et al., 2014).

The theory behind the interfacial forces under confinement are well established. However, techniques that can isolate and measure their magnitudes for a given pair of interacting oriented material surfaces are few and are restricted to flat or nearly flat interacting surfaces. Scanning probe microscopies and spectroscopies constitute one proven path forward (Zhang et al., 2017, 2018). For example, dynamic force spectroscopy (DFS) has recently demonstrated the self-organization of confined solvent molecules during the approach of solid surfaces of a defined crystallographic structure and twist angle. The self-organization was related to the net interaction force (see Figure C-4) (Zhang et al., 2017).

Image
FIGURE C-3 Left, an electron-density profile (red line) obtained by using x-ray reflectivity data from an interface of water and dodecane+extractant, in conjunction with molecular dynamics simulations to construct a model of the ordered arrangement of amphiphilic DHDP molecules. Right, the addition of Er3+ to the aqueous phase results in a change in the experimentally determined electron-density profile (blue line). Modeling the profile revealed an inverted bilayer composed of a 3:1 ratio of extractant to Er3+. SOURCE: Adapted from Bu et al., 2014. Reprinted with permission; copyright 2014, Journal of Physical Chemistry B.
Image
FIGURE C-4 Example of combining DFS with molecular simulations to probe interfacial forces and the structure of nanoconfined water between oriented single crystal surfaces. Oriented, single-crystal AFM tips (left) were used to measure the total interaction force with an oriented opposing substrate as a function of azimuthal angle (middle) and then compared directly with potentials of mean force computed by using molecular dynamics simulations (right). SOURCE: Zhang et al., 2017. Reprinted with permission; copyright 2017, Nature Communications.

Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×

CHARACTERIZING STRUCTURES FROM THE NANOSCALE TO THE MICROSCALE

Understanding reverse micelles and other aggregates requires extending structural studies from the atomic to the nanoscale and mesoscale. Besides observing cloud points optically (Quina et al., 1999), small-angle x-ray scattering (SAXS) and SANS have been used to characterize these and other reverse-micelle systems more fully (Jensen et al., 2007; Hollamby et al., 2010; Ellis et al., 2013). SAXS and SANS can also be used to characterize aggregate structures, including membrane and polymer pores and channels, as well as the role of reverse micelles in separation efficacy. Researchers have combined x-ray absorption spectroscopy (to understand the role of metal coordination) with SANS data to observe the role of the counterion on ionic liquid-based solvent microstructure separation (Abney et al., 2017). The results show important mechanistic parallels between ionic liquids and classic solvents and suggest opportunities for process improvements. Application of dynamic light scattering (DLS) combined with SAX and SANS has proved a powerful option for analyzing spatial and temporal fluctuations in polymer gels (Shibayama, 1998) but remains a topic of opportunity for sample characterization. Questions concerning uptake and fouling are particularly amenable to a combination of SAXS and SANS data: the complementary scattering could be used effectively to probe the roles of membrane and analyte structuring.

NEW OPPORTUNITIES FOR CHARACTERIZING STRUCTURAL DYNAMICS

Most structural characterization related to separation systems is time-averaged, so quantitative information on system dynamics and its potential effect on chemical separations is scarce. For liquid–liquid systems, dynamic properties are of low enough energy that they could affect selectivity or capacity. For example, control or limiting of the splitting of the organic phase into two distinct phases (third-phase formation) in liquid–liquid extraction could depend on structural dynamics.

NMR has provided information on dynamic processes and, although the technique is limited to slower time scales and the structural information is indirect, its potential effect is apparent (Witherspoon et al., 2018). DLS could be extended into a temporal regime that touches on complex solution structuring. Emerging opportunities are presented by the recent developments in x-ray photon correlation spectroscopy (XPCS) (Shpyrko, 2014), a technique that uses the coherent x-ray beams only now becoming available at synchrotron sources. Newly developing capabilities are enabling access to information on the nanometer and microsecond scales that are relevant to separations (see Figure C-5).

Image
FIGURE C-5 New opportunities in XPCS. Red lines show current and near-term upper limits due to detector speed. Black line shows typical upper limit due to signal level. The pink region of new XPCS opportunity covers time and length scales relevant to ion and micelle dynamics (green lines). The blue region shows new opportunities for molecular dynamics (MD) simulations as the limit moves to longer time and length scales and overlaps with XPCS studies. SOURCE: Stephenson, 2018. Reprinted with permission; copyright 2018, Brian Stephenson.

Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×

REFERENCES

Abney, C. W., C. Do, H. Luo, J. Wright, L. He, and S. Dai. 2017. Controlling the intermediate structure of an ionic liquid for f-block element separations. The Journal of Physical Chemistry Letters 8(9):2049–2054. doi: 10.1021/acs.jpclett.7b00755.

Bu, W., H. Yu, G. Luo, M. K. Bera, B. Hou, A. W. Schuman, B. Lin, M. Meron, I. Kuzmenko, M. R. Antonio, L. Soderholm, and M. L. Schlossman. 2014. Observation of a rare earth ion–extractant complex arrested at the oil–water interface during solvent extraction. The Journal of Physical Chemistry B 118(36):10662–10674. doi: 10.1021/jp505661e.

Chung, W. K., A. S. Freed, M. A. Holstein, S. A. McCallum, and S. M. Cramer. 2010. Evaluation of protein adsorption and preferred binding regions in multimodal chromatography using NMR. Proceedings of the National Academy of Sciences 107(39):16811–16816. doi: 10.1073/pnas.1002347107.

Egami, T., and S. J. L. Billinge. 2003. Underneath the Bragg peaks. Materials Today 6(6):57. doi: 10.1016/S1369-7021(03)00635-7.

Ellis, R. J., Y. Meridiano, R. Chiarizia, L. Berthon, J. Muller, L. Couston, and M. R. Antonio. 2013. Periodic behavior of lanthanide coordination within reverse micelles. Chemistry – A European Journal 19(8):2663–2675. doi: 10.1002/chem.201202880.

Hollamby, M. J., J. Eastoe, A. Chemelli, O. Glatter, S. Rogers, R. K. Heenan, and I. Grillo. 2010. Separation and purification of nanoparticles in a single step. Langmuir 26(10):6989–6994. doi: 10.1021/la904225k.

Jensen, M. P., T. Yaita, and R. Chiarizia. 2007. Reverse-micelle formation in the partitioning of trivalent f-element cations by biphasic systems containing a tetraalkyldiglycolamide. Langmuir 23(9):4765–4774. doi: 10.1021/la0631926.

Kocsis, I., M. Sorci, H. Vanselous, S. Murail, S. E. Sanders, E. Licsandru, Y.-M. Legrand, A. van der Lee, M. Baaden, P. B. Petersen, G. Belfort, and M. Barboiu. 2018. Oriented chiral water wires in artificial transmembrane channels. Science Advances 4(3):eaao5603. doi: 10.1126/sciadv. aao5603.

Kodera, N., D. Yamamoto, R. Ishikawa, and T. Ando. 2010. Video imaging of walking myosin V by high-speed atomic force microscopy. Nature 468(7320):72-76. doi: 10.1038/nature09450.

McBriarty, M.E., J.E. Stubbs, P.J. Eng, and K.M. Rosso. 2018. Potential-specific structure at the hematite-electrolyte interface. Advanced Functional Materials 28:1705618. doi: 10.1002/adfm.201705618.

McBriarty, M. E., G. F. von Rudorff, J. E. Stubbs, P. J. Eng, J. Blumberger, and K. M. Rosso. 2017. Dynamic stabilization of metal oxide–water interfaces. Journal of the American Chemical Society 139(7):2581–2584. doi: 10.1021/jacs.6b13096.

Mulfort, K.L., O.K. Farha, C.D Malliakas, M.G Kanatzidis, and J.T Hupp. 2010. An interpenetrated framework material with hysteretic CO2 uptake. Chemistry – A European Journal 16(1):276–281. doi: 10.1002/chem.200902104.

Quina, Frank H., and Willie L. Hinze. 1999. Surfactant-Mediated Cloud Point Extractions: An Environmentally Benign Alternative Separation Approach. Industrial & Engineering Chemistry Research 38(11):4150-4168. doi: 10.1021/ie980389n.

Rock, W., B. Qiao, T. Zhou, A. E. Clark, and A. Uysal. 2018. Heavy anionic complex creates a unique water structure at a soft charged interface. The Journal of Physical Chemistry C 122(51):29228–29236. doi: 10.1021/acs. jpcc.8b08419.

Sava Gallis, D. F., I. Ermanoski, J. A. Greathouse, K. W. Chapman, and T. M. Nenoff. 2017. Iodine gas adsorption in nanoporous materials: A combined experiment–modeling study. Industrial & Engineering Chemistry Research 56(8):2331–2338. doi: 10.1021/acs.iecr.6b04189.

Scoppola, E., E. Watkins, G. Li Destri, L. Porcar, R. A. Campbell, O. Konovalov, G. Fragneto, and O. Diat. 2015. Structure of a liquid/liquid interface during solvent extraction combining X-ray and neutron reflectivity measurements. Physical Chemistry Chemical Physics 17(23):15093–15097. doi: 10.1039/C5CP01809A.

Shibayama, M. 1998. Spatial inhomogeneity and dynamic fluctuations of polymer gels. Macromolecular Chemistry and Physics 199(1):1–30. doi: 10.1002/(SICI)1521-3935(19980101)199:1<1::AID-MACP1>3.0.CO;2-M.

Shpyrko, O. G. 2014. X-ray photon correlation spectroscopy. Journal of Synchrotron Radiation 21(5):1057–1064. doi: 10.1107/s1600577514018232.

Sobekova Foltova, S., T. Vander Hoogerstraete, D. Banerjee, and K. Binnemans. 2019. Samarium/Cobalt separation by solvent extraction with undiluted quaternary ammonium ionic liquids. Separation and Purification Technology 210:209–218. doi: 10.1016/j.seppur.2018.07.069.

Soderholm, L., S. Skanthakumar, and R. E. Wilson. 2011. Structural correspondence between uranyl chloride complexes in solution and their stability constants. The Journal of Physical Chemistry A 115(19):4959–4967. doi: 10.1021/jp111551t.

Stephenson, G. B. 2018. Opportunities in separations science using advanced synchrotron x-ray methods. Presented at the 3rd Data-gathering Meeting of the Separations Committee Meeting, Washington, DC, August 22-23, 2018.

Uchihashi, T. and S. Scheuring. 2018. Applications of high-speed atomic force microscopy to real-time visualization of dynamic biomolecular processes. Biochimica et Biophysica Acta 1862(2):229-240. doi: 10.1016/j.bbagen. 2017.07.010.

Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×

Uysal, A., W. Rock, B. Qiao, W. Bu, and B. Lin. 2017. Two-step adsorption of PtCl62– complexes at a charged Langmuir monolayer: Role of hydration and ion correlations. The Journal of Physical Chemistry C 121(45):25377–25383. doi: 10.1021/acs.jpcc.7b09350.

Wang, C.-H., P. Bai, J. I. Siepmann, and A. E. Clark. 2014. Deconstructing hydrogen-bond networks in confined nanoporous materials: Implications for alcohol–water separation. The Journal of Physical Chemistry C 118(34):19723–19732. doi: 10.1021/jp502867v.

Watarai, H. Interfacial Molecular Aggregation in Solvent Extraction Systems. In Supramolecular Aspects of Solvent Extraction, Ion Extraction and Solvent Extraction: A Series of Advances; CRC Press: Boca Raton, 2014; Vol. 21; pp 159-195.

Witherspoon, V. J., J. Xu, and J. A. Reimer. 2018. Solid-state NMR investigations of carbon dioxide gas in metal–organic frameworks: Insights into molecular motion and adsorptive behavior. Chemical Reviews 118(20):10033–10048. doi: 10.1021/acs.chemrev.7b00695.

Zhang X., Z. Shen, J. Liu, S.N. Kerisit, M.E. Bowden, M. L. Sushko, J. J. De Yoreo, K. M. Rosso. 2017. Direction-specific forces underlying zinc oxide crystal growth by oriented attachment. Nature Communications 8:835.

Zhang, X., Y. He, J. Liu, M. E. Bowden, L. Kovarik, S. X. Mao, C. Wang, J. J. De Yoreo, and K. M. Rosso. 2018. Accessing crystal–crystal interaction forces with oriented nanocrystal atomic force microscopy probes. Nature Protocols 13(9):2005–2030. doi: 10.1038/s41596-018-0027-4.

Zhou, T., A. McCue, Y. Ghadar, I. Bakó, and A. E. Clark. 2017. Structural and dynamic heterogeneity of capillary wave fronts at aqueous interfaces. The Journal of Physical Chemistry B 121(38):9052–9062. doi: 10.1021/acs. jpcb.7b07406.

Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×

This page intentionally left blank.

Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×
Page 91
Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×
Page 92
Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×
Page 93
Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×
Page 94
Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×
Page 95
Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×
Page 96
Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×
Page 97
Suggested Citation:"Appendix C: Examples of Characterization." National Academies of Sciences, Engineering, and Medicine. 2019. A Research Agenda for Transforming Separation Science. Washington, DC: The National Academies Press. doi: 10.17226/25421.
×
Page 98
A Research Agenda for Transforming Separation Science Get This Book
×
Buy Paperback | $58.00 Buy Ebook | $46.99
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

Separation science plays a critical role in maintaining our standard of living and quality of life. Many industrial processes and general necessities such as chemicals, medicines, clean water, safe food, and energy sources rely on chemical separations. However, the process of chemical separations is often overlooked during product development and this has led to inefficiency, unnecessary waste, and lack of consensus among chemists and engineers. A reevaluation of system design, establishment of standards, and an increased focus on the advancement of separation science are imperative in supporting increased efficiency, continued U.S. manufacturing competitiveness, and public welfare.

A Research Agenda for Transforming Separation Science explores developments in the industry since the 1987 National Academies report, Separation and Purification: Critical Needs and Opportunities. Many needs stated in the original report remain today, in addition to a variety of new challenges due to improved detection limits, advances in medicine, and a recent emphasis on sustainability and environmental stewardship. This report examines emerging chemical separation technologies, relevant developments in intersecting disciplines, and gaps in existing research, and provides recommendations for the application of improved separation science technologies and processes. This research serves as a foundation for transforming separation science, which could reduce global energy use, improve human and environmental health, and advance more efficient practices in various industries.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    Switch between the Original Pages, where you can read the report as it appeared in print, and Text Pages for the web version, where you can highlight and search the text.

    « Back Next »
  6. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  7. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  8. ×

    View our suggested citation for this chapter.

    « Back Next »
  9. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!