Sie sind auf Seite 1von 11

The Polarized Light Microscope: Should We Teach the use of a 19th Century Instrument in the 21st Century?

Mickey E. Gunter ABSTRACT


The polarizing light microscope (PLM) has no doubt contributed more to our knowledge of minerals and rocks than any other single instrument. Then why is the use of the PLM, and the teaching of optical mineralogy in general, decreasing? Probably one of the main reasons is educators seek to present the newest, most technologically advanced techniques and methods to their students. Also, the geoscience curriculum has changed to include environmental geology, essentially hydrogeology. However, in presenting these newer materials, or new courses, we must exclude something, and it appears that one of the things excluded is instruction in use of the PLM. Another possibility is the professors teaching these courses do not have an adequate understanding of optical mineralogy. Excluding the subject of optical microscopy will be the biggest mistake we ever make in the geosciences curriculum. This statement is justified because of the fundamentally important concepts presented in optical mineralogy: 1) three-dimensional visualization, 2) inquiry-based learning, and 3) hands-on use of an analytical instrument. No other single course in our curriculum provides so many of these fundamental skills to our students. An example of inquiry-based learning is determining the best technique to identify a mineral. For example, powder X-ray diffraction provides a diffraction pattern searchable in a database, but the result might not be correct. In the past 20-30 years, the spindle stage has allowed for more detailed single crystal studies than ever before. Optical techniques are also used to study such diverse mineralogical problems as cation-diffusion in zeolites and asbestos identification, or they may be incorporated into other areas of research, such as synchrotron experiments on oriented single crystals. Also, any serious petrologic study must begin with a thorough examination of the samples by PLM, before other types of characterization can proceed. A diversity of employment opportunities exist for students who are trained in use of the PLM, such as in the fields of forensics, material science, manufacturing, the food industry, medical technology, and the emerging field of environmental mineralogy. Department of Geological Sciences, University of Idaho, Moscow, Idaho 83844-3022, mgunter@uidaho.edu computer-based instruments (e.g., X-ray diffractometers and electron microscopes). The overarching theme is that students, and researchers, must understand the limitations of the various instruments and be able to select the most appropriate one for the problem at hand. Also, equally important, we should teach how to integrate several different analytical methods to answer the question at hand. The PLM is one of the few instruments students will ever get to use independently. This itself warrants its teaching as we strive for more interactive methods of instruction. No other single instrument provides chemical, structural, and morphological information on a single sample and with so little sample preparation, albeit the chemical and structural data are found indirectly. It is this indirect method of inferring chemical and structural data that is the double-edged sword for the use of the PLM. Microscopists must infer this information based on their experiences. When data are misinterpreted, it is often (incorrectly) the microscopic techniques and not the microscopist who is blamed. Clearly, the PLM cannot directly determine a materials crystal structure or chemical composition; diffraction methods are commonly used for the former and spectroscopic methods for the latter (e.g., energy dispersive spectroscopy in an electron beam instrument). However, once the structure and composition are known, generally speaking, they can be related to some feature that is observable or measurable with a PLM. For instance, refractive index is related to compositional variation and optical class is related to crystal system. Students can quickly confirm the identity of a mineral if they are taught how to use oil immersion and grain mounts. For example, the refractive index can provide the species of a plagioclase feldspar much easier than using power X-ray diffraction, which is difficult interpret. Also, chemical information from an electron microscope cannot differentiate among kyanite, sillimanite, and andalusite, or between quartz and opal. Working with minerals in thin sections, there are even simpler optical properties such as color, relief, twinning, mineral associations, and alteration that give qualitative chemical information and guide the pertologist in further characterization of the rocks. The examples are numerous, and several more will be presented below. Most importantly, the PLM excels in determining sample morphology, for instance to aid in differentiating amphibole from amphibole-asbestos in grain mounts or the textural relationships of minerals in rocks. The significance of these relationships could be expanded almost endlessly. And the cost of maintaining a PLM versus that of an X-ray diffractometer or electron-beam instrument is several orders of magnitude less. This lower cost allows institutions with modest budgets to still teach the use of the PLM, while the maintenance on other instruments may be cost prohibitive. Mineralogists often only think of using the PLM for geological materials, but the uses outside of our discipline often have important societal implications. Again, the examples are numerous, and several will be

INTRODUCTION
Our understanding of minerals and rocks is attributed more to use of the polarized light microscope (PLM) than any other single instrument. With the PLM, we can quickly identify and characterize minerals and determine textural relationships in rocks, including microstructures used in petrofabric analysis. Why, then, are we abandoning teaching these skills to our students in favor of other subject matter? One of the main problems using a PLM, though, is the answers do not come from a computer but must be determined by the user. Thus, more education and experience are required to use the PLM correctly than what is required for
34

Journal of Geoscience Education, v. 52, n. 1, January, 2004, p. 34-44

presented. Despite this, other disciplines have also discontinued instruction of the PLM. We not only should continue teaching polarized light microscopy but better integrate its use into our research, both in the study of geology and other areas. Probably the main reason for the decline in the teaching of optical mineralogy, and in fact mineralogy in general, is that many geologists view this discipline as a surrogate to igneous and metamorphic petrology, or geochemistry. That is to say, that mineralogy is not a stand-alone discipline but merely leads to more important areas in the field of geology. Unfortunately by thinking like this they ignore the geological information gained by consideration of the crystal chemical aspects of a mineral. A close corollary is that because so many petrologists have been trained in the field of optical petrography (i.e., using thin sections to study rock texture), they teach optical mineralogy courses without a full understanding of the power of optical mineralogy outside of the field of petrology (i.e., they would always favor cutting a thin section to examine a rock over using the immersion method to aid in identification of a mineral). This is also unfortunate because geological information could be gained by single crystal optical studies of, for instance, the structure state of feldspars (Su et al.,.1984). This paper will provide a brief history of the PLM and its historical uses in mineralogy and geology, followed by a review of some basic optical methods (e.g., what can be measured with a PLM with the aid of a spindle stage). Next, a series of mineralogical examples will be given, some of which could only be solved with a PLM. Finally, several examples of on-going integration of optical techniques to study the orientational dependence of physical properties of minerals will be highlighted. It is not the intention of this paper to discuss all the uses of the PLM but rather to concentrate on the authors research in an attempt to show the evolution of his work and how he has incorporated classical optical mineralogy into newer analytical methods in mineralogy. This paper is based on a seminar the author gave as a Mineralogical Society of America distinguished lecturer.

It was not until 1856 that the first true PLM was built, and a theoretical understanding of how light interacts in a mineral was quickly developed. Advances were made both in the immersion and thin section methods until the early 1900s, when these techniques were basically perfected. Little has changed in their use since. The immersion methods allow for a fairly rapid identification of single crystals, while thin section methods permit not only mineral identification but also the study of textural relationships of minerals in rocks. Currently, thin section techniques are more commonly taught than the immersion methods. This is probably because thin section techniques can be applied to the study of rocks, whereas the immersion method is basically used for identification of minerals. Also, many scientists incorrectly believe the commercially available refractive index liquids in use today contain PCBs and are carcinogenic; this was true of formulations used decades ago but not the newer formulations. In the early 1900s, the PLM was the main instrument used by mineralogists and petrologists to study minerals and rocks. The discovery of X-rays in 1912 led to a major breakthrough by which the crystalline properties of minerals could be directly studied by X-ray diffraction, instead of indirectly studied by the interaction of light with a mineral. The development of X-ray diffraction led many mineralogists to abandon use of the PLM, because now they could use powder diffraction to aid in the identification of minerals and single crystal X-ray diffraction to determine the crystal structures. Electron beam instruments were also being developed in the middle 1900s, but it was not until the invention of the microprobe in 1959 that these instruments became a dominant force in the field of mineralogy and petrology. In the 1960s, the chemical composition of micron-sized minerals could be determined with an electron microprobe. This replaced the laborious technique of using larger size samples and performing wet chemical analysis. Mineralogists then had at their disposal instrumentation capable of directly studying the atomic arrangement of atoms in minerals and precisely determining the chemical composition of minerals.

BRIEF EVOLUTION OF THE USE OF THE PLM


The main impetus for development of the PLM was to study crystalline materials. In the 1800s, scientists studying minerals noted light interacted with minerals in unique ways. They determined that certain minerals would cause light to become polarized, and they also observed that light traveled at different speeds in different directions for some minerals, while for others the speed of light was the same regardless of propagation direction. They attributed these differences to varying crystal structures of minerals and sought ways to study this in more detail. The PLM would become their main tool for studying minerals for the next hundred years. For a complete discussion of the development of the PLM and many of the accessories associated with it (e.g., the U-stage for study of petrofabrics) see Kile (2003). Two separate methods were developed to study minerals. One was the immersion method, where crystals of minerals were crushed and placed in a liquid then observed with a microscope. This was first done in 1815 using water as the liquid. The other method was development of thin sections, first crudely done in 1831.
Gunter - The Polarized Light Microscope

MAJOR USES OF THE PLM


Identification and characterization of minerals Mineral samples are identified based on their structural and chemical characteristics. X-ray and electron diffraction techniques directly provide data about the atomic structure of a mineral. Diffraction occurs when X-ray radiation or electrons pass through and interacts with the atoms of a crystalline material. Analysis of diffraction patterns are used to determine structural properties such as space group, unit cell parameters, and positions of atoms. Energy dispersive spectroscopy (EDS) and wavelength dispersive spectroscopy (WDS) directly provide information about the chemistry of a sample. A sample is placed in the path of a high-energy electron beam. These electrons collide with the inner shell electrons of the atoms comprising the sample, eject an inner shell electron, and create a vacancy. Outer shell electrons move into the vacant inner shell and, in the process, release energy in the form of X-ray radiation. The specific energy, or wavelength, of X-ray radiation produced by an atom is characteristic of each element. EDS methods determine the energy of these X-rays, and WDS determines the wavelengths of the X-rays. By using
35

Figure 1. A photomicrograph of amphibole and amphibole-asbestos from Libby, Montana showing differing morphologies. (The photo was taken in plane-polarized light with 1.58 refractive index liquid; the scale bar is 1 mm.)

EDS or WDS, it is possible to determine the quantity of each element in a sample. EDS is more rapid and requires less sample preparation than WDS; however, WDS produces more precise and accurate quantitative chemical analysis. Often, the correct identification of a mineral requires that both structural and chemical information be obtained. Structural and chemical characteristics can also be indirectly observed using polarized light microscopy. Structural properties are obtained by the correlation of a minerals optic class to its crystal system. Optical classification of minerals is based on whether they are isotropic or anisotropic. If a sample is isotropic, it has only a single refractive index value and belongs to the isometric crystal system, or it is amorphous. If a sample is anisotropic, it is further classified based on its optic class. Anisotropic minerals can be either uniaxial, which belong to either the hexagonal or tetragonal crystal system, or biaxial, which belong to the orthorhombic, monoclinic, or triclinic crystal system. Both of these classes are further divided into positive and negative groups based on the relationship of refractive indices and their principal vibration directions. Chemical characteristics of minerals can be related to the optical properties of the minerals (e.g., refractive indices, 2V, sign of elongation, extinction, pleochroism, etc.). A centurys worth of careful optical and chemical measurements has provided many empirical relationships, whereby a quick check of refractive index or a determination of 2V can provide an approximation of chemical composition often required for mineral identification. Thus, chemical data of minerals are found indirectly by measurement of some basic optic property, usually the refractive index.

SOME IMPORTANT SPECIFIC USES OF THE PLM


Asbestos - Without a doubt the most important use of the PLM for the past 20 to 30 years is in the identification of asbestos. This is true for two reasons. First, the size range of concern for identification of asbestos fits the
36

resolution of a light microscope. Second, the distinguishing characteristic of asbestos is its morphology, which is fibrous. It could be argued that the emerging filed of environmental mineralogy has its roots in the use of the PLM for asbestos identification and characterization. (For an overview of the asbestos issue in the U.S.A. refer to Gunter, 1994.) A material could have the exact same chemical composition and X-ray diffraction pattern and in one case its morphology might be fibrous and the other case it might be non-fibrous. So the only method to distinguish between fibers and non-fibers is the morphology, which can be observed in the magnification range of that for a PLM. For example, Figure 1 shows amphibole samples from Libby, Montana, in an oil immersion mount. Clearly, the particle on the left, labeled fragment, is non-fibrous and the particle on the right is fibrous. This distinction can only be made based on morphology. Chemical analysis of these two materials would be identical, and if we were to perform a powder X-ray diffraction experiment, the diffraction patterns would also be identical. Asbestos regulations are currently being revised in the U.S.A. This is mainly the result of the amphibole and amphibole asbestos minerals occurring in the former vermiculite mine near Libby, Montana (Bandli et al., 2003; Gunter et al., 2003). The first change in the regulations will no doubt be to regulate amphiboles based on group name and not species name. This is necessary because the amphiboles at Libby are winchites and richterites which, as of this time, are non-regulated but are just as harmful as the regulated amphibole minerals. At first this seems like a fairly simple modification to the regulations. However, the issue will then become how to distinguish asbestos from non-asbestos amphiboles. Many of the regulatory agencies have used the aspect ratio to make this distinction. For instance, if an amphibole particle has a length-to-width ratio greater than 3, it would be considered a fibrous particle. Thus, what would clearly be a non-fibrous sample (Figure 1) would, based on this aspect ratio rule, become regulated as an asbestos material. Clearly, to regulate these materials correctly there must be close interaction between those individuals trained in polarized light microscopy and the regulatory agencies. Today there are many more jobs in asbestos identification laboratories and in regulatory agencies than there are qualified microscopists to fill them. This scarcity of trained microscopists has occurred because we have stopped teaching classical optical mineralogy in our geology curriculum. The void has been filled in the private sector with one-week short courses in polarized light microscopy, optical mineralogy, and asbestos identification. For instance, the McCrone Research Institute in Chicago has trained hundreds of microscopists in the past 20 to 30 years. If the private sector is teaching these courses because of a nationwide demand, it would be logical for us to include them as part of our curriculum. Any geology student successfully completing a semester long course in optical mineralogy could be practically guaranteed a job in one of these labs. The spindle stage - One of the major advances in optical characterization of minerals has been the refinement of spindle stage techniques by Bloss and coworkers (for example, Bloss, 1982; Bloss 1999; Gunter et al., 1988; Gunter and Twamley, 2001; Su et al., 1987). The spindle stage, which is basically a one-axis rotation device, is a

Journal of Geoscience Education, v. 52, n. 1, January, 2004, p. 34-44

Figure 3. Photographs of three different types of spindle stages: Supper, detent, and poster board. The figure shows the X-ray goniometer head (with an affixed crystal) mounted onto the Supper spindle stage and the oil cell supplied with this spindle stage. The oil cell for the poster board and detent spindle is also shown.

Figure 2. A) The necessary equipment to mount single crystals for optical examination with a spindle stage. Under a binocular microscope, a small single crystal (ca. 50-200 mm) is glued with fingernail polish to a (ca.) 100 mm glass fiber. Then the glass fiber is placed in a brass pin that fits into the X-ray goniometer head, also pictured. Acetone can be used to remove the crystal for further study if desired. B & C) Photomicrographs of an amphibole single crystal mounted on a glass fiber. The sample has been rotated 90 on the spindle axis between B & C, thus allowing three-dimensional viewing of the single crystal. (The photos were taken in plane-polarized light with a 1.58 refractive index liquid; the scale bar is 100 mm.)

microscope attachment that aids in obtaining large amounts of optical data quickly and efficiently. A brief introduction to the spindle stage and its uses follows; for
Gunter - The Polarized Light Microscope

a thorough discussion of spindle stage techniques, see Bloss (1981; 1999). The basic design for the spindle stage is a single crystal mounted on a glass fiber, or sewing needle, that is placed on the rotation device (i.e., the spindle stage). The crystal can be placed in an immersion liquid, and routine optical measurements and observations can be made. Figure 2 is a photograph of three different spindle stages. The Supper spindle stage is often used for research purposes, while the small detent spindle stage (Bloss, 1999) and poster board spindle stage (Gunter, 1997) can be used for teaching purposes. In all cases, however, the setup is the same: small (50 to 200 mm) single crystals are mounted on the end of the spindle stage and placed on a PLM. Figure 3 shows the materials required to mount single crystals. Commonly, a single crystal is mounted, with the aid of a binocular microscope, on the end of a glass fiber with fingernail polish. The glass fiber is, in turn, placed in a 3 mm diameter brass pin, which is inserted into an X-ray goniometer head. Next, the X-ray goniometer head is screwed onto the Supper spindle stage shown in Figure 3. Now the crystal can be immersed in refractive index liquid contained in the oil cell mounted on the spindle stage. Figure 3B and C show an amphibole crystal affixed to the end of a glass fiber and immersed in refractive index oil. The crystal has been rotated 90 between B and C. Figure 4A shows the spindle stage X-ray goniometer head combination mounted on the stage of a PLM. Also shown in this image is a temperature-controlled oil cell used to record and change the oils temperature. At the same time, the wavelength of light can be changed with a monochrometer. By changing both temperature and wavelength, the refractive indices of a material are measured to within plus or minus two to five in the fourth decimal place by the double-variation method (Bloss, 1981; Su et al., 1987). One of the major advances of Bloss and coworkers is the ability to orient biaxial minerals easily on the spindle stage with the aid of the computer program EXCALIBR (Bloss, 1981; Gunter et al., 1988). All that is required is
37

Figure 4. Photographs of three different experimental setups for single crystal study; in all cases, an X-ray goniometer head is used and thus the same crystal can be easily moved between experimental setups. A) The Supper spindle stage, with goniometer head, is mounted on the rotating stage of a PLM to observe crystal morphology in three dimensions and to measure optical properties (e.g., refractive index, pleochroism, 2V, orientation of the indicatrix, etc.). B) A goniometer head mounted on a single-crystal diffractometer used to obtain structural data (e.g., unit cell parameters, crystal structure, orientation of the crystallographic axes, etc.). C) A modified spindle stage placed in a synchrotron beam source for studies of the orientational dependence of absorption spectra. 38

Figure 5. Sketches showing the optical orientation for andalusite and Mn-andalusite. Note that for andalusite the small refractive index (X) is parallel to the long dimension of the crystal, while for Mn-andalusite the optical orientation has changed and Z (the largest refractive index) is parallel to c. The lower part of the figure is a polyhedral representation of the structure of andalusite. The main structural unit is an edge-sharing chain of octahedral parallel to the c crystallographic axis.

Journal of Geoscience Education, v. 52, n. 1, January, 2004, p. 34-44

Figure 6. Relationship between refractive indices and Mn + Fe content for 20 andalusite samples (modified from Gunter and Bloss, 1982). Because there is a change in the optical orientation as a function of Mn + Fe content (x+y on graph), the refractive index curves are labeled n||a, n||b, and n||c to correspond to the refractive index directions in the a, b, and c crystallographic directions, respectively. Note that when the curves cross, a sample of that composition would appear isotropic.

extinction measurements are determined for different spindle stage settings. A mathematical relationship exists between these extinction data and the optical indicatrix. In other words, given extinction measurements, a computer program will provide the microscopist the spindle stage and microscope stage settings to place the crystal so any of its three principal refractive indices can be directly determined without appreciable error due to misorientation. Another major advantage of this method is that all refractive indices are determined on the same crystal. This, in turn, lends itself to studies whereby the optically characterized single crystal can also be used for X-ray diffraction or chemical analysis. All the physical properties of a mineral can be determined on the same single crystal. Isotropic andalusite - An example of an integrated optical, X-ray, and chemical study was performed on andalusites with varying compositions (Gunter and Bloss, 1982). Andalusite, Al2SiO5, can have appreciable substitution of Mn and Fe for the six-coordinated Al in its structure. There is also a change in the optical orientation in andalusite as this substitution occurs (Figure 5). Figure 5 also shows a polyhedral representation of a portion of the structure of andalusite. Andalusite is an orthosilicate mineral and the isolated Si tetrahedron is shown in Figure 5. There are two Al sites: a five-coordinated site labeled M2 and a six-coordinated site labeled M1. Mn
Gunter - The Polarized Light Microscope

and Fe substitute for Al in the M1 site that form edge-sharing chains of octahedrons parallel to the c crystallographic axis. Notice how the structure of andalusite controls its morphology; these edge-sharing chains are the main structural unit, thus elongating andalusite parallel to its c crystallographic axis. Gunter and Bloss (1982) obtained a suite of approximately 50 andalusite samples and selected 20 of varying compositions. For each of these 20 samples, we precisely measured the refractive indices with the PLM and spindle stage, cell parameters by X-ray diffraction, and chemical composition by electron microprobe. These different measurements were made on the same single crystals; thus, the change in the optical orientation could be tracked as a function of chemical composition. We found that the refractive index direction parallel to the c crystallographic axis was the smallest of the refractive indices for pure andalusite; however, it increases at a greater rate than the other two refractive indices with substitution of Mn and Fe for Al. Upon close examination of Figure 6, notice a void in our data in the area where these three curves of refractive index values cross. Two reasons were postulated for this: that samples of this composition do not exist in nature (although this did not seem reasonable to us), and that andalusites with this composition had been overlooked in thin section because they would appear isotropic, or near isotropic. Shortly after our work was published, Grambling and Williams
39

Figure 7. Relationships of the refractive of the fibrous zeolites - natrolite, mesolite, and scolecite (modified from Gunter and Ribbe, 1993). This graph was produced from data synthesized from the literature no data were collected. Based on this graph, it is easy to distinguish between these three samples by their sign of elongation.

Figure 8. Photomicrographs of an approximately 500 mm long natrolite sample capped with mesolite taken in both plane-polarized and cross-polarized light. The crystal is mounted on a glass fiber and immersed in index-matching fluid for natrolite. In the planepolarized photo, the mesolite tip has higher relief than the natrolite portion of the crystal. In the cross-polarized photo, the mesolite tip appears near (1985) found andalusites in this compositional range that isotropic because of the lower birefringence of mesolite (Figure 7). indeed appeared isotropic.

While this discovery may seem like a mere optical curiosity (i.e., that orthorhombic minerals may appear isotropic), it has significance to the metamorphic petrology community. Grambling and Williams (1985) showed how changes in the triple point for the aluminosilicates could be related to the Mn and Fe contents of those andalusites, thus helping to account for some of the variability in location of the triple-point in the aluminosilicate phase diagram.

Retardation and optical orientation of natrolite group zeolites - Based upon my previous work on andalusite, I searched the literature for other possible effects of changes in optical orientation that could provide worthwhile information. Also, in general, a portion of my research is directed toward the understanding of how the optical properties of materials are related to their crystal structure and chemical composition. I first investigated zeolites because of their many industrial applications and wide-spread geologic occurrences. Also, zeolites have low birefringence because they are framework silicates; thus, a small change in some structural or chemical property may cause a large change in the easily observable retardation. Gunter and Ribbe (1993) collected optical data from the existing literature for the natrolite group zeolites (natrolite, mesolite, and scolecite). After some obvious corrections to the optical orientations incorrectly provided in some of the standard reference books, a plot of the refractive indices for these three minerals could be made as a function of their Ca and Na content. In a similar manner to andalusite, the rate of change of the refractive indices were different for the different crystallographic directions. In this mineral group, the refractive index direction parallel to the c crystallographic axis, which corresponds to chains of tetrahedrons parallel to the c crystallographic axis,
40

increases at a lower rate than the refractive indices parallel to the a and b crystallographic axes (Figure 7). Thus, natrolite has its largest refractive index direction parallel to its long dimension, mesolite has its intermediate refractive index direction parallel to its long direction, and scolecite has its smallest refractive index parallel to its elongation. It is simple to distinguish these three minerals in thin section or grain mount based upon their sign of elongation. Figure 7 also shows that the retardation for natrolite and scolecite, while being low, would still appear anisotropic in thin section or grain mount. Mesolite, on the other hand, exhibits very low retardation and would appear isotropic, or near so. During the preparation of Gunter and Ribbe (1993), we collected samples of fibrous zeolites from a nearby basalt flow (Gunter et al., 1993). Figure 8 shows one of these fibrous zeolites affixed to a glass fiber. In the plane-polarized light photograph, the refractive index fluid matches natrolite, so it exhibits very low relief. Mesolite has a slightly higher refractive index and thus stands out in higher relief at the tip of the crystal. In cross-polarized light the photomicrograph reveals how the retardation clearly shows an anisotropic crystal, but because the retardation for mesolite is so low, it appears isotropic. The study again shows how subtle changes in chemistry can cause large effects on the retardation. Because retardation is so easily observable, it can be used for other purposes than simply to identify minerals in thin section or grain mount. Cation diffusion in heulandite group zeolites - In another study by Gunter et al. (1994) on Pb-exchange heulandite group zeolites, we used optical data to aid in determining the extent of Pb exchange. Pb substitution increases the birefringence, and thus the retardation. Figure 9A shows a non-exchanged sample exhibiting

Journal of Geoscience Education, v. 52, n. 1, January, 2004, p. 34-44

were able to determine the diffusion coefficients for Na, K, Rb, and Cs by making simple observations in changes of optical properties with a PLM. No other single method available could be used as easily and efficiently to qualitatively observe exchange in zeolites or to quantitatively determine their diffusion coefficients.

EXAMPLES OF CURRENT RESEARCH


A major theme in this paper is the orientational dependence of refractive indices and morphology of minerals. And, as discussed in detail above, the spindle stage is the perfect tool to study this dependence. What follows are three examples of how we have recently integrated the spindle stage into other analytical methods with the underlying theme to characterize the directional dependence of various physical properties. Single-crystal X-ray diffraction and optical orientation - The optical orientation of minerals relates the crystallographic axes to the optical indicatrix. It is also useful to identify the mineral either in immersion mounts or in thin section and is an important physical property of the mineral. While X-ray diffraction is used to locate the crystallographic axes, it cannot be used to locate the optical directions. Thus if one wants to collect spectroscopic data along the optical directions of a mineral, the spindle stage is used to determine these orientations. The optical orientation is basically nonexistent for isotropic minerals and trivial for uniaxial minerals. However, for biaxial minerals the relationship is harder to determine, especially for monoclinic and triclinic minerals. Determining the optical orientation of these minerals requires an integrated study using both X-ray diffraction to locate the crystallographic axes and the PLM to determine the orientation of the biaxial indicatrix. This entire process can be simplified by using a spindle stage equipped with an X-ray goniometer head (Figure 4A). The X-ray goniometer head, with crystal affixed, can be transferred between the spindle stage and an X-ray diffractometer (Figure 4B). Also, as described above, the orientation of the biaxial indicatrix can be determined with the measurement of extinction positions and use of the computer program EXCALIBR. Thus, the orientation of the biaxial indicatrix is found and cast into the coordinate system of the microscope. Next, the goniometer head is transferred to the X-ray diffractometer and the orientation matrix is obtained. The orientation matrix contains information on the orientation of the crystallographic axes a, b, c cast into the coordinate system of the X-ray diffractometer. Next, to fully determine the optical orientation, these two coordinate systems are transformed into each other and the relationship of the biaxial indicatrix and the crystallographic axes mathematically determined (Gunter and Twamley, 2001). This entire process takes about one hour, as compared to older methods that might require several days. Synchrotron radiation and absorption spectroscopy - The advent of synchrotrons with their high-energy white X-ray beams have created a major new analytical technique with which to examine minerals. X-ray absorption spectroscopy is one of many techniques that requires synchrotron radiation. In this method, the energy of the X-ray beam is changed in such a way as to encompass the absorption-edge of an element of interest.

Figure 9. A series of images of natural and Pb-exchanged heulandite zeolites showing changes in retardation (in the PLM images) and back-scattered electron (BSE) intensity (in the BSE images) as a function of Pb exchange (modified from Gunter et al., 1994, which are in color). (All scale bars are 100 mm.) A) PLM image of a natural sample exhibiting low retardation. B) PLM image of a fully Pb-exchanged sample showing a drastic increase in retardation. C) BSE image of a partially Pb-exchanged sample; the Pb has only diffused into the edges of the sample. D) PLM image showing how increased retardation can be used to monitor Pb exchange around the edges of the crystal; the greatest exchange is in the lower right hand corner of the grain. E) BSE image of a partially Pb-exchanged sample, with Pb diffusing in along the edge and cracks in the grain. F) PLM image showing Pb diffusion along the edges and cracks in a sample as shown by increases in retardation.

first-order grey retardation, and Figure 9B shows a fully Pb-exchanged sample exhibiting second-order red retardation. (A color rendition of this figure can be found in Gunter et al., 1994.) Another method that can be used to determine whether exchange has occurred, or to measure diffusion of Pb into the structure, is backscattered electron imaging (BSE) with a scanning electron microscope. Figure 9C shows a BSE image where the Pb has diffused a few tens of microns on the rim of the grain. Figure 9D shows a similar grain except this time as observed with a PLM. Note the high order retardation on the lower right portion of the grain; this indicates Pb has diffused into this portion of the mineral. Pb not only diffuses into the rims of these minerals, but also through cracks. Figure 9E is a BSE image showing areas of Pb diffusion, and Figure 9F is a PLM image showing areas of Pb diffusion. Thus, again, in this study optics is very sensitive in determining chemical changes and much more efficient than performing a chemical analysis with an electron beam instrument to see spatial variation in chemistry. A natural extension of this project was to measure diffusion rates of different cations in heulandite group zeolites by observing changes in optical properties. In a series of exchange experiments by Yang et al. (1997), we
Gunter - The Polarized Light Microscope

41

Figure 10. X-ray absorption near-edge spectroscopy (XANES) spectra of a garnet and olivine with the polarized synchrotron beam parallel to each of the three crystallographic axes. For garnet, all spectra exhibit the same features, while for olivine they vary as a function of orientation. Thus, the X-ray absorption indicatrix is similar to the optical indicatrix.

Figure 11. Photographs showing integration of spindle stage methods with a scanning electron microscope (SEM). A) Stage of SEM with a 3 mm brass pin used for crystal mounting in a goniometer head (Figure 3A) inserted into a normal SEM sample holder. B) Photomicrograph of an acicular amphibole particle mounted on the end of a glass fiber. C) SEM image of the same sample in B in approximately the same orientation. D) SEM image of a bundle of amphibole fibers mounted on a glass fiber. E) Magnified and reoriented image of the white box area in D to better observe the morphology of the amphibole. (Images modified from Bandli and Gunter, 2001.)

For example, for Fe absorption the energy of the X-ray beam is scanned over the main absorption edge at 7111 eV. Information about the coordination, valence, and other bonding characteristics can be obtained from interpretation of these spectra. The majority of the work in this area has been on millimeter-scale powders of minerals, as well as on single crystals of similar size. Our current research (Dyar et al., 2002a, b) is exploring new ways to make these measurements on micron-sized single crystals. However, one of the concerns in this area of research is the orientational dependence of the absorption spectra. Some previous researchers have tried
42

to ascertain the orientational dependence of these absorption spectra by using oriented thin section minerals. There are technical difficulties with this method, along with a considerable amount of time being required for sample preparation. For instance, study of an orthorhombic mineral would require at least two oriented samples. To study the orientational dependence of the spectra and fully characterize them on a 100 micron-size crystal, we modified a spindle stage so it could be placed in a

Journal of Geoscience Education, v. 52, n. 1, January, 2004, p. 34-44

polarized synchrotron beam (Figure 4C). We could then place an X-ray goniometer head on the spindle stage/synchrotron combination and make the required measurements. To test the system we developed, we made measurements on samples that should show no preferred orientation (e.g., an isotropic mineral, such as garnet) and samples that should show orientational dependence (e.g., an orthorhombic mineral, such as olivine). The garnet was oriented using X-ray diffraction as described above. The olivine was oriented using the spindle stage as described above. Based on the spectra in Figure 10, it is clear there is no orientational dependence of the absorption spectra for garnet, as would be predicted, and there is an orientational dependence of the absorption spectra for the olivine, as also would be predicted based on the analogy with the behavior of the minerals and polarized light. Prior to the collection of these spindle stage assisted spectra, it was unclear what the orientational dependence of the spectra would be. Thus, integration of the methods of classical optical mineralogy aids in understanding these new analytical methods and provides for better data collection methods. Scanning electron microscope and three-dimensional imaging - When higher magnifications are required, the scanning electron microscope (SEM) can be used in lieu of, or in conjunction with, a PLM. It is also possible to obtain chemical information directly if the SEM is equipped with an energy dispersive detector. The SEM has a greater depth of field at similar magnifications when compared to a PLM. Standard sample preparation for the SEM is placement of mineral grains on an aluminum stub or glass slide. With this type of mount, two-dimensional imaging of the sample is possible. However, if a single crystal is affixed to the brass pin/glass fiber combination that fits into the X-ray goniometer head, as used on the spindle stage, this brass pin can be placed in the sample holder stage of the SEM (Figure 11A) to achieve three-dimensional imaging of the particle (Bandli and Gunter, 2001). Combined optical and scanning electron microscope studies can be performed on the same single crystal; this is a great aid in identification of the sample because chemical information can also be obtained. It is also a very useful method to handle 100-micron size particles without losing them. Once a crystal has been affixed to a glass fiber and observed in the PLM (Figure 11B), it can then be easily transferred to the scanning electron microscope (Figure 11C). The main advantage of this method is to be able to manipulate the crystal in the SEM to gain morphological information in different directions. Figure 11D shows a bundle of amphibole asbestos fibers affixed to the end of a glass fiber. In Figure 11E the sample has been manipulated to view approximately down the long axis of the fibers. In this orientation the crystal faces of the amphibole particle can be seen. If the sample were mounted on a flat surface, one would not be able to manipulate it to observe the morphology in different directions. In the case of amphibole minerals, different surfaces may have an impact on the minerals health effect. For instance, amphibole asbestos particles tend to expose the 100 surface, amphibole cleavage fragments expose the 110 surface, and small single crystals of amphibole expose the 100 surface. Clearly, from a mineralogical perspective these different surfaces would have different reactivities. Using the SEM with a spindle stage style sample mount helps to characterize the morphology of these particles.
Gunter - The Polarized Light Microscope

CONCLUSIONS
I hope I have convinced the reader that the answer to the question posed in the title is yes. From the teaching perspective, the main reason we should continue teaching polarized light microscopy is that it is one of the few hands-on, interactive methods of learning about analytical instrumentation students get to participate in. From an employment perspective, there is no other single course students can take that will secure them a job as rapidly as one that teaches the use of a PLM in environmental applications, especially in the field of asbestos identification. In mineralogical research, there are still many new discoveries and innovations waiting to be made using well-established, existing methods in optical mineralogy, and efforts should be made in the mineralogical community not to lose sight of these methods in favor of the latest instrumentation, which may not provide the best and most accurate data.

ACKNOWLEDGMENTS
I would like to thank Emeritus Professor F.D. Bloss, VPI & SU, for his major contributions to the field of optical mineralogy as a textbook author, teacher, and researcher, and for mentoring me in the field of optical mineralogy. I would also like to thank the MSA (Mineralogical Society of America) for inviting me to be a distinguished lecturer in 2002-03; this paper was based on a seminar presented in that series. I would like to thank Professor Osamu Tamada of Kyoto University for providing a visiting professorship, which enabled me to contemplate the significance of polarized light microscopy in the geoscience curriculum and work on my MSA lecture. Lastly, I would like to thank Professor Charles Guidotti of the University of Maine for a thorough review of this manuscript, providing suggestions for how to improve it, and confirming my views of the importance of teaching the use of the PLM. This work was partially supported by funds from the National Science Foundation (NSF-CCLI 0127191).

REFERENCES
Bloss, F.D., 1981, The spindle stage: principles and practice, Cambridge University Press, Cambridge, England, 340 p. Bloss, F.D., 1999, Optical crystallography, Mineralogical Society of America, Washington, D.C., 239 p. Bandli, B.R. and Gunter, M.E., 2001, Identification and characterization of mineral and asbestos particles using the spindle stage and the scanning electron microscope, The Libby, Montana, U.S.A. Amphibole-asbestos as an example, The Microscope, v. 49, p. 191-199. Bandli, B.R., Gunter, M.E., Twamley, B., Foit, F.F., Jr., and Cornelius, S.B., 2003, Optical, compositional, morphological, and X-ray data on eleven particles of amphibole from Libby, Montana, U.S.A., Canadian Mineralogist, v. 41, p. 1241-1253. Dyar, M.D., Gunter, M.E., Delaney J.S., Lanzarotti, A., and Sutton, S.R. 2002a, Use of the spindle stage for orientation of single crystals for microXAS: Isotropy and anisotropy in Fe-XANES spectra, American Mineralogist, v. 87, p. 1500-1504.

43

Dyar, M.D., Gunter, M.E., Delaney, J.S., Lanzarotti, A., and Sutton, S.R., 2002b, Systematics in the structure and XANES spectra of pyroxenes, amphiboles, and micas as derived from oriented single crystals, Canadian Mineralogist, v. 40, p. 1347-1365. Grambling, J.A. and Williams, M.L., 1985, The effects of Fe3+ and Mn3+ on aluminum silicate phase reactions in North-Central New Mexico, U.S.A., Journal of Petrology, v. 26, p. 324-354. Gunter, M.E., 1994, Asbestos as a metaphor for teaching risk perception, Journal of Geological Education, v. 42, p. 17-24. Gunter, M.E., 1997, Laboratory exercises and demonstrations with the spindle stage, In Teaching Mineralogy, Mineralogical Society of America, editors J.B. Brady, D.W. Mogk, and D. Perkins III, p. 309-318. Gunter, M.E. and Bloss, F.D., 1982, Andalusite-kanonaite series: Lattice and optical parameters, American Mineralogist, v. 67, p. 1218-1228. Gunter, M.E. and Ribbe, P.H., 1993, Natrolite group zeolites: correlations of optical properties and crystal chemistry, Zeolites, v. 13, p. 435-440. Gunter, M.E. and Twamley, B., 2001, A new method to determine the optical orientation of biaxial minerals: A mathematical approach, Canadian Mineralogist, v. 39, p. 1701-1711. Gunter, M.E., Bloss, F.D., and Su, S.C., 1988, EXCALIBR revisited, American Mineralogist, v. 73, p. 1481-1482. Gunter, M.E., Bloss, F.D., and Su, S.C., 1989, Computer programs for the spindle stage and double-variation method, The Microscope, v. 37, p. 167-171.

Gunter, M.E., Knowles, C.R. and Schalck, D.K., 1993, Composite natrolite-mesolite crystals from the Columbia River Basalt Group, Clarkston, Washington, Canadian Mineralogist, v. 31, p. 467-470. Gunter, M.E., Armbruster, T., Kohler, T., and Knowles, C.R., 1994, Crystal structure and optical properties of Na- and Pb-exchanged heulandite group zeolites, American Mineralogist, v. 79, p. 675-682. Gunter, M.E., Dyar, M.D., Twamley, B., Foit, F.F.Jr., and Cornelius, S.B., 2003, Composition, Fe3+/Fe, and crystal structure of non-asbestiform and asbestiform amphiboles from Libby, Montana, U.S.A., American Mineralogist, v. 98, p. 1870-1942. Kile, D.E., 2003, The petrographic microscope: evolution of a mineralogical research instrument, Mineralogical Record, special publication number 1. Su, S.C., Bloss, F.D., Ribbe, P.H., and Stewart, D.B., 1984, Optic axial angle, a precise measure of Al,Si ordering in T1 tetrahedral sites of K-rich alkali feldspars, American Mineralogist, v. 69, p. 440-448. Su, S.C., Bloss, F.D., and Gunter, M.E., 1987, Procedures and computer programs to refine the double variation method, American Mineralogist, v. 72, p. 1011-1013. Yang, P., Armbruster, T., Stoltz, J., and Gunter, M.E., 1997, Na, K, Rb, and Cs exchange in heulandite single-crystals: Diffusion kinetics, American Mineralogist, v. 82, p. 517-525.

44

Journal of Geoscience Education, v. 52, n. 1, January, 2004, p. 34-44

Das könnte Ihnen auch gefallen