Presentation and Impact of Experimental Techniques in Chemistry

Jul 7, 2008 - probes: Table 2 contains the list of acronyms of some frequently ..... H.; van Leeuwen, H. P.; Lobinski, R. Pure Appl. Chem. 2000, 72,...
0 downloads 0 Views 570KB Size
In the Classroom

Advanced Chemistry Classroom and Laboratory 

Presentation and Impact of Experimental Techniques in Chemistry

edited by

  Joseph J. BelBruno Dartmouth College Hanover, NH  03755

Zbigniew Sojka* Laboratoire de Réactivité de Surface, UMR 7609-CNRS, Université Pierre et Marie Curie, 4, Place Jussieu, 75252 Paris-Cedex 05, France and Faculty of Chemistry, Jagiellonian University, ul. Ingardena 3, 30-060 Krakow, Poland; *[email protected] Michel Che Laboratoire de Réactivité de Surface, UMR 7609-CNRS, Université Pierre et Marie Curie, 4, Place Jussieu, 75252 Paris-Cedex 05, France

The large number of instrumental techniques available in laboratories is a remarkable feature of present day chemistry. The spectacular progress in all domains of chemistry would be hard to imagine without the application of all the classical methods of identification and characterization such as separation, imaging, diffraction, and spectroscopy, as well as ultra-sensitive chemical analysis at the level of the element, molecule, and phase. These analyses can now be performed with precision and accuracy unattainable even a few years ago. As a result microscopic (molecular) and macroscopic (molar) chemical information can be obtained by small-scale and often nondestructive experiments. Many of these instrumental techniques are now in routine use, thanks to the development of technology and computerization. Combining two or more techniques has been another more recent and major progress in chemistry (1). Several judicious combinations of methods have become popular because of their ability to solve problems far beyond those of each separate technique. The GC–MS tandem is one of the best examples, where the separation power of gas chromatography is integrated with the identification ability of modern mass spectrometry, providing a powerful approach to describe complex mixtures. This coupling has become a routine technique in chemical, environmental, biochemical, and forensic laboratories (2–5), successfully solving of crucial problems such as speciation and trace analysis (6, 7). The pace of progress in structure determination has significantly increased with the advent of the Fourier transform multidimensional spectroscopies. These still burgeoning methods are becoming available even to non-specialists and have revolutionized the way chemists think about structure determination. They also are increasingly used in student laboratory experiments (8–11). Recent chemistry curriculum reforms both in the United States (12) and Europe (13) have given more importance to the introduction of active-learning methods and toward developing a need-to-know mentality (14). The recommendations call for less emphasis on passive presentation and accumulation of information (teaching) and more on active development and communication of knowledge (“hands and minds on” learning). Students not only have to learn, comprehend, and apply factual material but also have to spend a large fraction of their time in lab courses where important laboratory skills are acquired. Obviously, students cannot gain an adequate appreciation of modern chemistry without using instrumental methods. The increasing number of techniques may generate some difficulties with the classification, description, and rational selec934

tion of the most important ones to be included in the curriculum. Crucial issues to be resolved are the level and scope (general and broad vs selective and advanced) and the approach (intuitive concept-based or more fundamental and physically rigorous) of introducing experimental techniques to students. It is probably easier to decide which subjects not to include, such as techniques still in their experimental stage or requiring large budget for acquisition and operation. The selection may be based on the need to comprehensively cover the principal aspects of chemical research (identification, structure determination, reactivity studies) or to introduce principal classes of techniques such as spectroscopy, diffraction, imaging, and so forth. There is a strong consensus that there is no need to teach every technique. Instead, it seems more appropriate to concentrate on fewer techniques and study those in-depth, after having introduced the instrumental techniques in a more general context (15). In the large panoply of currently available techniques, the problem is to find a rational and easy way to classify them in relation to their performance and usefulness. We suggest that an extended Probst-type diagram (16, 17), sometimes used in textbooks on experimental techniques in catalysis—one of the most multi-instrumental disciplines (18)—may serve as a suitable unifying tool of systematization. This article is the result of a literature search aiming at providing a unifying view of techniques along with their physical bases and at evaluating their impact, defined as the number of articles dealing with each, for the year 2000. Although there are numerous good books and review articles dealing with experimental techniques at different levels, there seems to be no information on their importance based on the frequency of their use in current research. Portfolio of Experimental Techniques: A Unifying Picture Spectroscopy and Electromagnetic Spectrum Spectroscopy is a broad discipline, embracing an extensive range of energetic, structural, and dynamic processes. Photons scattered, absorbed, or emitted upon interaction with matter become a versatile source of information, thus providing the basis for many fundamental spectroscopies. Figure 1 shows the electromagnetic spectrum along with the associated molecular motions and chemical events as well as the corresponding techniques, according to the particular region of wavelength involved.

Journal of Chemical Education  •  Vol. 85  No. 7  July 2008  •  www.JCE.DivCHED.org  •  © Division of Chemical Education 

In the Classroom

Effect

E/

Spectrum

(kJ mol-1)

O/ s-1

AC

CV 10

-5

NMR

nuclear spin precession 10

EPR

electron spin precession molecular rotations

10

molecular vibrations UV–vis

-3

10

1

10

3

10

5

10

10

9

microwave

10

-1

10

1

far infrared near infrared

10

valence electron excitation

7

1

10

radio

IR

UPS

M/ m

3

10

10

multicomponent

13

P T

10

7

10

9

10

vacuum ultraviolet 5

11

separable

mixture

e i



B

visible ultraviolet

Classification Diagram There are many instrumental techniques that can help to obtain thorough information on chemical systems. A great part of those techniques can be derived from a generic scheme known as the Probst diagram; an extended version is shown in Figure 2. The sphere in the center represents the sample to be analyzed, while the arrows going in and out indicate the various ways to probe the sample and the possible responses. There are essentially, eight types of probe of corpuscular, wave, or continuum nature used for characterization that can be applied either alone or in suitable combinations. Four types of probe (photons, electrons, neutrons, and ions of the corpuscular-wave duality) may interact with samples either in a quantum way at the microscopic atomic or molecular level or in a classical way (diffraction, interference, reflection) at the macro- or mesoscopic scales. The remaining four types of probe [electric, magnetic, thermal, and pressure (acoustic) fields] interact, with a few exceptions, usually in a classic macroscopic fashion with the samples (e.g., scanning calorimetry) or are used as ancillary means, such as the magnetic field in NMR or EPR spectroscopies. Each technique, classified according to their ability to absorb, emit, and scatter photons, electrons, neutrals, ions, and so forth, can then be associated with the combination of an arrow pointing in and an arrow pointing out of the sample (Figure 2).

10

n

15

hO

E

sample

17

hO

E

separation

Technique

of the mixture under a gradient of pressure, temperature, electric, or magnetic fields, such as retention, permeation, or diffusion, provide the physical basis for fractionation. All kinds of chromatography such as GC (21, 22), HPLC (23, 24), or capillary electrophoresis (25, 26) discussed in this Journal may serve here as good examples. As a result, multi-component mixtures can be analyzed in detail even from minute samples.

nonseparable

The interaction of a probe with a sample depends upon the nature of both the probe and the sample. For probes with nonzero charge and mass, the sample can be profoundly perturbed because of the absorption by matter. Furthermore, owing to small mean free paths in gases even at modest pressure such studies, being necessarily confined to ultra-high vacuum conditions, are hardly applicable to in situ investigations of chemical reactions. Fortunately, in some cases, these obstacles can be circumvented because of a proximal detection (at the atomic scale) as, for example, in scanning tunneling microscopy, where the electron flux between the probing tip and the investigated solid is measured. The techniques using photons, thermal, or electric fields can be applied in liquid or gas phase and the samples usually are only slightly perturbed, unless extremely high energies are used. They are widely employed in structural characterization, and for investigations into rate processes in a broad range of temperatures and pressures (19, 20). When dealing with mixtures, analytical techniques can be interfaced with various separation methods for increased performance. Specific transport properties of the various components

XRD XPS EXAFS

Mössbauer

core electron excitation

nuclear excitation

10

7

X-ray

10

11

10

19

n

B

i



e 10

9

H-ray

10

13

10



T P

21

cosmic ray

Figure 1. Electromagnetic spectrum along with the associated energy scale and molecular motions and chemical events, and the corresponding techniques according to the particular region of wavelength.

components

single component

FIgure 2. Modified Probst diagram showing the physical origin of experimental techniques. Each technique can be derived from the combination of in- and out-going arrows. The sample can be either single- or multi-component. In the latter case if it is a physical mixture, the analysis can be preceded by the separation into its components to improve the performance.

© Division of Chemical Education  •  www.JCE.DivCHED.org  •  Vol. 85  No. 7  July 2008  •  Journal of Chemical Education

935

In the Classroom Table 1. Techniques Viewed as Combinations of Eight Types of Probes (Photons, Electrons, Neutrals, Ions and Electric, Magnetic, Thermal and Acoustic Fields) Primary Probe/Field

Outgoing Probe/Field

Neutrals

Neutrals

INS, ND

Ions

FAMBS

Ions

Electrons

Photons

Electric Magnetic Field

Thermal Acoustic Field

ESD

TPD

SIMS, ISS, RBS, IMM

Electrons

Photons

LMMS

FIM

LEED, AES, EELS, ED, SEM, TEM

UPS, XPS, XPD CEMS

FEM, STM

XRE CL

IR, LRS, UV–vis, EXAFS, XRF, XRD, photoluminescence, elastic and quasi-elastic light scattering, Mössbauer,

PIXE, PAS

TE

NMR, EPR, ENDOR Electric Magnetic Field

CV, IS

Thermal Acoustic Field

PAS

DSC, DTA

Table 2. List of Frequently Used Acronyms Acronym

Full Name

Acronym

Full Name

AES

Auger electron spectroscopy

LEED

Low energy electron diffraction

AFM

Atomic force microscopy

MS, QMS

Mass spectroscopy, quadruple mass spectrometry

CL

Cathodoluminescence

NMR,

Nuclear magnetic resonance

CEMS

Conversion electron Mössbauer spectroscopy

MAS-NMR

Magic angle spinning NMR

CV

Cyclic voltammetry

PAS

Photo-acoustic spectroscopy

DRIFT

Diffuse reflectance infrared Fourier transform

PL

Photoluminescence

DSC

Differential scanning calorimetry

PD-MS

Plasma desorption mass spectrometry

DTA

Differential thermal analysis

PIXE

Proton induced X-ray emission

EELS

Electron energy loss spectroscopy

RS

Raman spectroscopy*

EI-MS

Electron impact mass spectrometry

RBS

Rutherford backscattering spectroscopy

ESI-MS

Electrospray mass ionization spectrometry

SAXS

Small angle X-ray scattering

ENDOR

Electron nuclear double resonance

SEM

Scanning electron microscopy

EPR/ESR

Electron paramagnetic (spin) resonance

SIMS

Secondary ion mass spectrometry

ESD

Electron stimulated desorption

STM

Scanning tunnelling microscopy

EXAFS

Extended X-ray absorption fine structure

TE

Thermionic emission

FABMS

Fast atom bombardment mass spectrometry

TEM

Transmission electron microscopy

FEM

Field emission microscopy

TPD

Temperature programmed desorption

FIM

Field ion microscopy

UPS

Ultraviolet photoelectron spectroscopy

GC

Gas chromatography

UV–vis

Ultraviolet–visible spectroscopy

IR, FTIR

Infrared, Fourier transform infrared

XRD, ED, ND

X-ray diffraction, electron diffraction, neutron

IMM

Ion microprobe microanalysis

INS

Inelastic neutron scattering

IS

Impedance spectroscopy

ISS

Ion scattering spectroscopy

XRE

X-ray emission

LAMMS

Laser microprobe mass spectrometry

XRF

X-ray fluorescence

diffraction XPS, ESCA

X-ray photoelectron spectroscopy, electron spectroscopy for chemical analysis

*Including a wide range of non-classical variants such as resonance Raman spectroscopy (RRS), surface enhanced Raman spectroscopy (SERS), and nonlinear spectroscopies for instance coherent Stokes Raman spectroscopy (CSRS), anti-Stokes Raman (CARS), stimulated Raman gain spectroscopy (SRGS) etc. Their concise description can be found, e.g., in Pure Appl. Chem., 1997, 69, 1451–1468.

936

Journal of Chemical Education  •  Vol. 85  No. 7  July 2008  •  www.JCE.DivCHED.org  •  © Division of Chemical Education 

In the Classroom

For example, one can irradiate a sample with X-rays and study how the latter are diffracted (X-ray diffraction, XRD), absorbed (spectroscopies such as X-ray absorption near edge structure, XANES, or extended X-ray absorption fine structure, EXAFS), or produce electrons emitted with different kinetic energies from the sample due to the photoelectric effect (X-ray photoelectron spectroscopy, XPS, also referred to as electron spectroscopy for chemical analysis, ESCA). Thus, the Probst diagram can help in making connections and drawing meaningful interactions and combinations among various techniques, and if necessary, it can be readily further extended to include new methods not implied by the present scheme explicitly. It is worth noting that a single combination of inward and outward arrows may lead to distinct spectroscopies depending on the property of the sample that is studied. This is the case of, say, microwave radiation that can be used to excite molecular rotations (rotational or microwave spectroscopy) or to induce, in presence of external magnetic field, flips of electron magnetic moments (EPR/ESR spectroscopy). In Table 1, the various tech-

niques are presented as a function of the incident and emergent probes: Table 2 contains the list of acronyms of some frequently used techniques. As seen in Table 1, most techniques are diagonal, that is, the incident and outgoing probes are of the same nature. Obviously, it is technically easier to construct instruments such as IR, UV– vis, or RS rather than XPS or NMR spectrometers that, apart from an electromagnetic radiation source, require high vacuum chambers and electron energy analyzers or highly homogeneous strong magnetic fields, respectively. Non-diagonal techniques are thus technically more involved and therefore traditionally usually less popular. However, owing to the recent progress in technology they are rapidly gaining in significance, as it can be exemplified by the variety of advanced mass spectrometry methods, for instance. For enhancement of its expediency, the classification diagram of techniques may be complemented by assessment of their impact to ascertain those of most value to graduating students and Ph.D. scientists.

500 400

Inorganic Chemistry

400

Journal of Physical Chemistry A and B

300

300

1200

Journal of Organic Chemistry

1000 800

GC

STM/AFM

SEM/TEM

SIMS

SAXS

LEED

EXAFS

300

MS, QMS

AES

XRD

Fluorescence

XPS, UPS, ESCA

UV–vis

Photoluminescence

IR

RS

Mössbauer

NMR, MAS-NMR

GC

STM/AFM

SEM/TEM

SIMS

SAXS

LEED

EXAFS

MS, QMS

AES

XRD

Fluorescence

XPS, UPS, ESCA

UV–vis

Photoluminescence

IR

RS

0 Mössbauer

0 EPR/ENDOR

100

NMR, MAS-NMR

100

EPR/ENDOR

200

200

Journal of Solid State Chemistry

200

600 400

100

200 0

GC

STM/AFM

SEM/TEM

SAXS

SIMS

LEED

EXAFS

MS, QMS

AES

XRD

Fluorescence

XPS, UPS, ESCA

UV–vis

Photoluminescence

RS

IR

Mössbauer

EPR/ENDOR

NMR, MAS-NMR

GC

STM/AFM

SEM/TEM

SAXS

SIMS

LEED

EXAFS

MS, QMS

AES

XRD

Fluorescence

XPS, UPS, ESCA

UV–vis

Photoluminescence

RS

IR

Mössbauer

EPR/ENDOR

NMR, MAS-NMR

0

Figure 3. Frequency of use (number of counts) of selected techniques for the year 2000, distributed by discipline.

© Division of Chemical Education  •  www.JCE.DivCHED.org  •  Vol. 85  No. 7  July 2008  •  Journal of Chemical Education

937

In the Classroom

Impact of Techniques in Chemistry Research To appreciate the importance of experimental techniques in contemporary research as a function of the various types of chemistry, we performed a literature survey of the use of twenty-five selected techniques in seven journals (Inorganic Chemistry, Journal of Organic Chemistry, Journal of Physical Chemistry A and B, Journal of Solid State Chemistry, Journal of Catalysis, and Journal of the American Chemical Society) covering the principal branches of chemistry. The Journal of Catalysis was selected as a reference field because of its multi-technique character. The techniques were selected on an ad hoc basis to encompass various spectroscopic (NMR, EPR/ENDOR, IR, RS, UV–vis, photoluminescence, XPS, UPS, AES, EXAFS, Mössbauer), diffraction (XRD, LEED), microscopic (SEM/ TEM, STM, AFM), and identification (GC, MS, SIMS) purposes. We are conscious that this selection is arbitrary, but for the sake of conciseness many widely used methods, such as liquid chromatography or light scattering techniques and others, could obviously not be included. Since many research articles dealing with one of the selected techniques do not mention it in their title or abstract and hence can not be extracted by an automated search, the survey was performed manually by carefully analyzing the experimental sections of all articles published in those journals in the year 2000. The results collected in Figure 3 give the corresponding histograms of the techniques in the leading journals representing four classic branches of chemistry. Figure 4 shows the results for a general journal such as Journal of the American Chemical Society and for a specialized one, Journal of Catalysis, since in both cases those techniques are extensively used. Inspection of Figure 3 indicates how the use of experimental techniques in the four branches of chemistry is uneven. The organic and solid state chemists are apparently highly confined in their research tools, and in both cases the leading techniques can be readily selected: NMR (43%) and XRD (56%), respectively. Complemented by 2–3 other techniques, they constitute 80–90% of the basic research instrumentation. There is a notice-

ably low impact of X-ray diffraction in organic research (~2%), which may change in the future owing to the progress in routine structure analysis and the accessibility of modern equipment and software. The situation is quite different in inorganic and especially physical chemistry, where the panoply of techniques is much broader. The NMR methods (both fluid and MAS variants) maintain their preponderant position (31%) in inorganic chemistry, whereas the leading tool of physical chemist appears to be IR (21%) with the impact of NMR at only 12%. This technique is then comparable to the use of UV–vis. As it may be expected, physical chemistry has clearly a multi-instrumental character with eight techniques exceeding the 5% level of counts and a remarkable preponderance (80%) of spectroscopy. With the Journal of the American Chemical Society, the dominant role of NMR (36%) is again observed. It is followed almost equally by IR, UV–vis, and XRD (10–12% for each one). Electron magnetic resonances (EPR/ESR, ENDOR) and mass spectroscopy emerge from the remaining methods (6–7%). The evaluation for Journal of Catalysis resembles that of Journal of Physical Chemistry, but instead of IR (14%), X-ray diffraction appears the most important (19%). One can appreciate also the high positions of GC (14%), MS (8%), microscopy (10%), and obviously surface analysis techniques (12%). Figure 5A gives the cumulative plot of the techniques (the counts of each journal have been added to give the total impact), which can be compared with the outcome of an analogous literature search in Journal of Chemical Education (Figure 5B). We surveyed the articles that have appeared using an advanced search option of the JCE Online. The general sequence of total counts for the techniques: NMR (28%) > IR (19%) > XRD, MS (11%) > UV–vis (8%) > EPR, RS, photoluminescence, GC, SEM/TEM (3–4%) qualitatively compares with that from Journal of Chemical Education with NMR (41%) > IR (18%) > UV–vis (11%), MS, GC (8–9%) >> EPR, RS, photoluminescence, XRD (2–3%). Despite the fact that the body of leading techniques is basically the same in both cases, the impact of some methods sub-

500

200

Journal of the American Chemical Society

400

Journal of Catalysis

150

300 100 200 50

100

0

GC

STM/AFM

SEM/TEM

SIMS

SAXS

LEED

EXAFS

MS, QMS

AES

XRD

Fluorescence

XPS, UPS, ESCA

UV–vis

Photoluminescence

IR

RS

Mössbauer

EPR/ENDOR

NMR, MAS-NMR

GC

STM/AFM

SEM/TEM

SIMS

SAXS

LEED

EXAFS

MS, QMS

AES

XRD

Fluorescence

XPS, UPS, ESCA

UV–vis

Photoluminescence

IR

RS

Mössbauer

EPR/ENDOR

NMR, MAS-NMR

0

Figure 4. Use of selected techniques in the articles published in general chemical journals in 2000.

938

Journal of Chemical Education  •  Vol. 85  No. 7  July 2008  •  www.JCE.DivCHED.org  •  © Division of Chemical Education 

In the Classroom

stantially differs. For instance XRD, EPR, photoluminescence spectroscopy tend to be distinctly under-represented in Journal of Chemical Education, whereas gas chromatography, perhaps because of its availability and conceptual simplicity, is overemphasized. There is also an apparent under-population of solid and surface characterization techniques (XPS, AES, TEM) that are rarely discussed in Journal of Chemical Education. Certainly, these are rather expensive methods and it is not likely that they will become part of student laboratory equipment. Nevertheless students should be aware of their use for analytical purposes and be guided how to select the most appropriate technique(s) to solve a particular problem. Curriculum Proposal and Suggestions of Some Practical Solutions The convenient classification tool of the instrumental techniques in the form of the Probst-type diagram, supported by the assessment of their importance, allowed us to propose an ensuing chemistry curriculum in advanced spectroscopy and instrumental analysis for graduating students. The backbone of such curriculum should undoubtedly cover spectroscopy, diffraction, imaging, and separation techniques as a standard. However, noting recent rapid developments in nanochemistry, biological chemistry, and chemistry of advanced materials, those methods need to be discussed not only at but also beyond the molecular scale, dealing with supramolecular assemblies, nanoparticles, extended solid or disordered, and soft materials. As it clearly emerges from the figures, the four principal spectroscopic methods such as magnetic resonance (NMR, possibly with an EPR detour), optical spectroscopy (UV–vis and photoluminescence), vibrational spectroscopy (IR and RS), and mass spectrometry along with X-ray diffraction, supplemented by chromatography (GC/LC), constitute a well-established basis, clearly reflected in their impact for current research. The courses on those techniques should provide students with the required theoretical and conceptual background and practical experience.

From a financial standpoint, one of the possible practical approaches in this issue, apart from direct access to instruments, is the application of spectroscopic workstations, computer networking to instruments, or use of computer spectroscopic simulators as discussed in more detail elsewhere (27). Relatively inexpensive software and spectral libraries are now easily available and can be used for development of modern laboratory exercises or for classroom innovations: a handy interactive EPR simulation program (28) with an intuitive windows menu can be obtained from the one of the authors (ZS). In turn, the spectroscopic workstation consists of optical, electronic, and mechanical components allowing for simple assembling of a variety of basic-grade instruments with quite modest expenditures (29). Relevant light sources, wavelength selectors and monochromators, detectors, readout and signal processing devices together with suitable sample-holder mounting systems are commercially available. This enables students to explore a wider range of techniques and measurement possibilities than otherwise would be possible. The refinement of the curriculum for advanced master studies may include more specialized courses on data processing and imaging, computational spectroscopy, and molecular modeling, whereas practical machinery with specific applications in a given field of chemistry can be developed within specialized modular courses. Despite the steadily increasing interdisciplinary character of most modern research, there is still an essential specific component for each of the particular disciplines (reflected in Figures 3 and 4) that also has an educational merit for curriculum construction. A working example may be provided, for example, by the European Erasmus Mundus Master Studies “Advanced Spectroscopy in Chemistry”, offered collectively by the consortium of seven European universities from Lille, Leipzig, Krakow, Bologna, Bergen, Helsinki, and Madrid (30). It is largely corroborated by the outcome of the literature search presented in this work. Student’s mobility scheme, supplied by the consortium, allows for expanding a direct access to more expensive and unique instrumental techniques and expertise that may be not available in partner home institutions.

500

2500

Journal of Chemical Education

Sum of the 7 journals 400

2000

300

1500

GC

STM/AFM

SEM/TEM

SIMS

SAXS

LEED

EXAFS

MS, QMS

AES

XRD

XPS, UPS, ESCA

UV–vis

Photoluminescence

IR

RS

Mössbauer

EPR/ENDOR

GC

STM/AFM

SEM/TEM

SIMS

SAXS

LEED

EXAFS

MS, QMS

AES

XRD

XPS, UPS, ESCA

UV–vis

Photoluminescence

IR

RS

0 Mössbauer

0 EPR/ENDOR

100

NMR, MAS-NMR

500

NMR, MAS-NMR

200

1000

Figure 5. Comparison of the cumulative (total of Figures 3 and 4) use of selected techniques with the number of articles devoted to each in Journal of Chemical Education.

© Division of Chemical Education  •  www.JCE.DivCHED.org  •  Vol. 85  No. 7  July 2008  •  Journal of Chemical Education

939

In the Classroom

Summary

10. Davis, D. S.; Moore, D. E. J. Chem. Educ. 1999, 76, 1617–1621.

Laboratory and practical courses where students become more familiar with the use of experimental techniques, learn to interpret data, and relate them to appropriate theory play an important role in chemical education. Undoubtedly, the range and level of sophistication of experimental techniques needed in chemistry is very broad. In a moving context of rapidly developing chemistry, with discovery of new techniques, the problem how to fill this crucial tutorial niche will certainly never find its ultimate solution. The Probst-type diagram may be used as a conceptually simple unifying tool to rationalize the techniques. Corroborated by the evaluation of their impact in current research, it may give a useful introduction to the subject. It is hoped that this article will provide a suitable inspiration for discussion of various experimental techniques and arguments for inclusion of some of them to curriculum. It may also stimulate more comprehensive surveys involving other branches of chemistry and additional methods as well, to yield more actual picture of contemporary chemistry techniques. A sketch of an exemplary curriculum in advanced spectroscopy was proposed and some possible practical solutions suggested.

11. Bose, R. N.; Al-Ajlouni, A. M.; Volckova, E. J. Chem. Educ. 2001, 78, 83–87.

Acknowledgment

21. Curtright, R. D.; Emry, R.; Markwell, J. J. Chem. Educ. 1999, 76, 249–252.

ZS is grateful to the Université Pierre et Marie Curie, Paris VI for an invited professorship. Literature Cited 1. Briois, V.; Belin, S.; Villain, F.; Bouamrane, F.; Lucas, H.; Lescouëzec, R.; Julve, M.; Verdaguer, M.; Tokumoto, M. S.; Santilli, C. V.; Pulcinelli, S. H.; Carrier, X.; Krafft, J. M.; Jubin, C.; Che, M. Physica Scripta 2005, T115, 38–44. 2. Novak, M.; Heinrich, J.; Martin, K. A. J. Chem. Educ. 1993, 70, A103–A110. 3. Reeves, P. C.; Pamplin, K. L. J. Chem. Educ. 2001, 78, 368– 370. 4. Atterholt, C.; Butcher, D. J.; Bacon, R. J.; Kwochka, W. R.; Woosley, R. J. Chem. Educ. 2000, 77, 1550–1551. 5. Bender, J. D.; Catino, A. J., III; Hess, K. R.; Lassman, M. E.; Leber, P. A.; Reinard, M. D.; Strotman, N. A.; Pike, C. S. J. Chem. Educ. 2000, 77, 1466–1468.

12. Hinde, R. J.; Kovac, J. J. Chem. Educ. 2001, 78, 93–98. 13. Mitchell, T. N.; Whewell, R. J. http://www.sfc.fr/Divens/Eurobachelor%20Brussels%20Versi.pdf (accessed Apr 2008). 14. Pinkerton, D. J. Chem. Educ. 2001, 78, 198–200. 15. Van Bramer, S. J. Chem. Educ. 2001, 78, 1167. 16. Park, R. L. Introduction to Surface Spectroscopies. In Experimental Methods in Catalytic Research; Anderson, R. B., Dawson, P. T., Eds.; Academic Press: New York, 1976; Vol. 3, pp 1–39. 17. Catalyst Characterization–Physical Techniques for Solid Materials; Imelik, B., Védrine, J. C., Eds.; Plenum: New York, 1994. 18. Thomas, J. M.; Thomas, W. J. Principles and Practice of Heterogeneous Catalysis; VCH: Weinheim, Germany, 1997; pp 145–255. 19. Hunger, M.; Weitkamp, J. Angew. Chem., Int. Ed. 2001, 40, 2954–2971. 20. van Eldik, R.; Hubbard, C. D. New J. Chem. 1997, 21, 825– 838.

22. Benson, G. A. J. Chem. Educ. 1982, 59, 344–346. 23. Kissinger, P. T.; Felice, L. J.; King, W. P.; Pachla, L. A.; Riggin, R. M.; Shoup, R. E. J. Chem. Educ. 1977, 54, 50–54. 24. Boyce, M.; Spickett, E. J. Chem. Educ. 2000, 77, 740–742. 25. Copper, C. L. J. Chem. Educ. 1998, 75, 343–347. 26. Williams, K. R. J. Chem. Educ. 1998, 75, 1079. 27. Winter 1999 CONFCHEM Teaching Spectroscopy. http://www. ched-ccce.org/confchem/1999/d/index.html (accessed Apr 2008). 28. Spalek, T.; Pietrzyk, P.; Sojka, Z. J. Chem. Inf. Model. 2005, 45, 18–29. 29. Long, G. R.; Ford, J. C. Winter 1999 CONFCHEM Teaching Spectroscopy.http://www.ched-ccce.org/confchem/1999/d/index. html (accessed Apr 2008). 30. Joint Master of Science. www.master-asc.org (accessed Apr 2008).

6. Templeton, D. M.; Ariese, F.; Cornelis, R.; Danielsson L.; Muntau, H.; van Leeuwen, H. P.; Lobinski, R. Pure Appl. Chem. 2000, 72, 1453–1470.

Supporting JCE Online Material

7. Kot, A.; Namiesnik, J. Trends Anal. Chem. 2000, 19, 69–79.

Abstract and keywords

8. Williams, K. R .; King , R . W. J. Chem. Educ. 1990, 67, A93–A99. 9. Doscotch, M. A.; Evans, J. F.; Munson, E. J. J. Chem. Educ. 1998, 75, 1008–1013.

940

http://www.jce.divched.org/Journal/Issues/2008/Jul/abs934.html

Full text (PDF)

Links to cited URLs and JCE articles



Color figures

Journal of Chemical Education  •  Vol. 85  No. 7  July 2008  •  www.JCE.DivCHED.org  •  © Division of Chemical Education