The Chemical Educator, Vol. 9, No.4, Media Reviews, © 2004 The Chemical Educator

 

Media Review

Metal-Ligand Bonding. By Rob Janes and Elaine Moore. Open University/Royal Society of Chemistry, U.K. ii + 104 pp. 263 ´ 210 mm. £24.95. ISBN 0854049797.

The remit of this book, published jointly by the Open University and the RSC, is to provide an understanding of metal–ligand interactions in transition metal complexes. The book is short, with just over 80 pages worth of content, and is divided into fifteen sections, which range in length from half a page to 13 pages. Following a brief introduction, the content is split into two approximately equal parts. The first of these deals with crystal field theory and the second with molecular orbital theory. Both are treated in a nonmathematical way with a strong emphasis on the use of figures. The text is very readable and adopts a question and answer style throughout to emphasise the key points. I found this to be a little annoying as it disrupts the flow, but students may consider it helpful in making the learning process more active.

The authors adopt a relatively conventional approach to crystal field theory with the model developed for octahedral geometry then extended to square-planar and tetrahedral complexes. This is similar to the strategy adopted by most books, and it is the most sensible and appropriate way of introducing this topic. Crystal field theory, once developed, is used to explain trends in electronic and magnetic properties in addition to those based on crystal field stabilization energies. The discussion of electronic spectra for complexes with more than one d electron is somewhat vague because terms lie outside the scope of the book. I liked the use of the magnetic quantum number, ml, in the explanation for when an orbital contribution to the magnetic moment occurs, though I felt this would have been clearer to the student if this quantum number had been introduced in the first section.

The strength of the second half of the book lies in the use of boundary-surface-type diagrams to show the interactions between orbitals. These figures are clear throughout and provide a very visual way for students to identify when interactions can and cannot occur; however, I was disappointed to see the phosphorus d orbitals rather than the P–C s* orbitals invoked as the p-acceptor orbitals for phosphine ligands. This part of the book also makes considerable use of molecular orbital energy-level diagrams. In these, the nonbonding orbitals have generally been omitted. Although this means that the figures are less cluttered, it ensures an imbalance in the number of atomic/ligand orbitals and molecular orbitals, which may confuse students. It also makes filling the orbitals through the aufbau principle harder to follow.

Group theory is introduced for the discussion of distorted octahedral and square-planar complexes. This section is the weakest in the book and includes both unconventional labelling, such as use of I rather than E for the identity, and mistakes, such as the assertion than sd planes do not contain axes other than the principal axis. This error is compounded by the two dihedral C2 axes being omitted from the square-planar symmetry elements that are depicted. The group theory is not fully developed, which is reasonable given the length of the book, but the symmetries for ligand-group orbitals are not given. This makes it difficult to appreciate which interactions between the metal and ligand orbitals can occur.

Would I recommend this book? By virtue of the length and topics covered, the book is unlikely to form the major source for many transition metal chemistry courses, especially given the reservations above; however it would provide a useful complement to a more comprehensive text, and students may find that the pictorial style, frequent summaries, and excellent index make this an approachable book and a useful revision aid.

Andrew D. Burrows

University of Bath, Bath, UK, a.d.burrows@bath.ac.uk

S1430-4171(04)04809-0, 10.1333/s00897040809a

Physical Chemistry: A Guided Inquiry—Thermodynamics. By James N. Spencer, Richard S. Moog, and John J. Farrell. Houghton Mifflin Company: Boston, 2004. ix + 341 pp + appendices, paperbound. $27.96 ISBN 0-618-30853-9.

As the introduction says, this is not a textbook on physical chemistry. It is a workbook designed to be used in the classroom in conjunction with a textbook. It is not, however, a simple workbook of problems to be solved. This is a guided-inquiry text in which students are led through a series of models (often with data) and questions requiring critical thinking skills to reach conclusions about chemical principles.

The book has four large sections (Gases, Thermodynamics, Electrochemistry, and Kinetics) as well as a section entitled “Mathematics for Thermodynamics,” which is a review of the relevant calculus necessary for the activities. Each of these sections is divided into ChemActivities. There are a total of 45 activities in the book with the largest proportion in the Thermodynamics section. The activities are integrative; that is, they require the student to understand and apply material covered in prior activities in order to complete later activities.

Each activity begins with a focus question that can be answered after the activity has been completed. A model or data is then presented, followed by several critical-thinking questions. These questions are the heart of the activity, leading the student (using the model or data) to arrive at conclusions about the concept being examined. The ChemActivity ends with a set of exercises that reinforce the concepts presented and give the students practice in the techniques used in physical chemistry. These ChemActivities are designed to be worked on in groups in the classroom with the instructor assisting, guiding, or encouraging the students in the right direction.

Although this methodology may be new to many instructors, the authors have provided excellent support materials, all available online at the publisher’s Web site, although you need to obtain a password from the publisher to access it. The Instructor’s Guide contains suggestions for implementing the course. These suggestions include ideas and examples of how to organize the class, choose groups, assign roles, and how to interact with students while using this method. There is also a sample schedule, a chapter correlation with five popular textbooks, and a description of each activity. These descriptions include an approximation of the time required to complete each activity as well as the concepts introduced and a discussion of some of the challenges that students encounter in each activity. A further supplement is the instructor’s annotated version of the text, which has the answers to the questions embedded in the text. Both of these will be helpful to anyone implementing this book into their teaching.

I have worked through several of the activities in preparation for the writing of this review. In general, they are straightforward and the models set up the material quite well. Students derive many of the equations used, allowing them the experience of obtaining the useful forms from the fundamental models. The exercises are well designed to fit with the concepts covered in that ChemActivity and can be challenging at times.

My concern with this book is the time it will take students to grasp the fundamental issues at hand. The authors admit in the instructor’s guide that their suggested time is an average and can vary by a factor of two, depending on a variety of variables, including group size, student preparedness, and student strength; however, using this book does allow for some flexibility. As the instructor’s manual suggests, the instructor can use this book in many ways, and problems can be overcome. If you are looking to incorporate active learning in your physical chemistry course, this book is a very useful way to integrate this methodology into your classroom.

Christopher J. Dunlap

Saint Mary’s College, Notre Dame, IN, 46556,

cdunlap@saintmarys.edu

S1430-4171(04)04810-1, 10.1333/s00897040810a

Van Nostrand’s Scientific Encyclopedia, Ninth Edition. Glenn D. Considine, Editor; Peter H. Kulik, Associate Editor. Wiley-Interscience: New York, 2002. 2 volumes. Figures, tables. xxvi + 3898 pp, hardcover. 22.4 ´ 28.3 cm. $375.00. ISBN 0-471-33230-5.

The latest edition of this well-known, concise, comprehensive, and accessible general science reference work continues its tradition of excellence maintained through six and a half decades. The first edition was published in 1938—before the Atomic Age—and the work has been extensively updated at intervals ever since. The unwieldy, one-volume sixth edition, published in 1983, was replaced by the two-volume seventh edition in 1989, and the two-volume format has been continued.

The editor of the ninth edition, Glenn D. Considine, is the son of the late Douglas M. Considine, who was editor of VNSE for more than three decades and who edited the fifth through eighth editions. Several hundred scientists, engineers, and university-level educators from around the world provided detailed information, graphics, and editorial guidance for the entries, and an abridged list (four double-column pages) includes more than 300 individuals and groups, and in some cases the titles of their articles, who contributed to the project.

Previous positive features of the series have been continued. The encyclopedia has been designed to be approachable by students of all ages. The discussion of each topic begins with a simple definition expressed in plain terms, followed by a more detailed treatment, augmented by sometimes extensive “Additional Reading” suggestions.

In addition to containing a preface and list of contributors, each volume includes a one-page “Representative Topical Coverage” that lists the 11 categories of the material, each subdivided into subcategories, designated here as numbers in parentheses: Animal Life (16); Biosciences (16); Chemistry (16); Earth and Atmospheric Sciences (12); Energy Sources and Power Technology (16); Mathematics and Information Sciences (8); Materials and Engineering Sciences (12); Medicine, Anatomy, and Physiology (24); Physics (16); Plant Sciences (12); and Space and Planetary Sciences (8). The Chemistry subcategories are: Acids and Bases, Catalysts, Chemical Elements, Colloid Systems, Corrosion, Crystals, Electrochemistry, Free Radicals, Inorganic Chemistry, Ions, Macromolecular Science, Organic Chemistry, Oxidation-Reduction, Photochemistry, Physical Chemistry, and Solutions and Salts.

Owners of earlier editions will be interested in the new features of this latest edition. The manner in which the editors wrote, gathered, and assembled articles differs fundamentally from their previous modus operandi. Whereas earlier generations of editors worked with typing, cutting, pasting, and telephoning experts, the current editors worked electronically via keyboarding, word processing software, the Internet, email query letters, and only occasionally by phone. Because of the now overwhelming amount of material, one of the greatest challenges was “separating the wheat from the chaff.”

Because we have entered what the editors call “a new Age of Discovery,” as shown by scores of new entries on topics that were only in their nascent stages by the time of the eighth edition, this latest edition features entirely new or completely rewritten or entire families of articles on the full array of topical coverage. These include, among others: Genetics Engineering, Human Genome Project, and Cloning; Bioprocess Engineering (Biotechnology); Space Shuttle, Space Stations, Spacecraft Missions, Satellites (Communications and Navigation), Cosmology, X-Ray Astronomy (family of articles), Astrobiology, and The Universe; Artificial Intelligence (family of articles); Medicine, Diseases, Vaccines, Vision (family of articles), AIDS, and STDs; Climate, Global Warming, and Acid Rain; Gerontology and Biochemical Theories of Aging; Computer Sciences and The Internet; and Flat Panel Display Technology (family of articles).

The suggested readings at the end of articles now include thousands of both updated print and Internet references, which provide additional knowledge on scores of topics. The Internet references enable users to find the “first places” to access for further information without first wading through thousands of “hits.”

Detailed time lines and glossaries have been added to some of the long home articles (Bioprocess Engineering, Artificial Intelligence, Vision and the Eye, Optical Fiber Systems, The Internet, and many others) to offer at a glance both historical perspective as well as information.

Finally, the editors have finally responded to criticism from reviewers of previous editions by including brief biographies of scores of scientists whose work is referred to in the encyclopedia. The coverage of chemists is uneven, but, of course, any biographies are welcome. Cavendish, Dalton, Lavoisier, Mendeleev, and van’t Hoff are included, but many of the other chemical “greats” are lacking. Watson and Crick are present but not Wilkins. Similarly, Kirchhoff is included but not Bunsen. For a general encyclopedia the coverage of fullerenes (pp 618–620) is quite detailed.

VNSE 9 contains more than 8,000 alphabetically arranged entries, ranging in length from a sentence to more than a dozen pages, the major ones of which are signed. Volume 1 contains those from “A.A.” through “Kyphosis,” and Volume 2 contains those from “Labaria” through “Zymolytic Reaction.” The new edition includes more than 9,000 cross-references for easy retrieval of information; an expanded alphabetical index of more than 19,500 lines on 78 triple-column pages (in Volume 2 only); 4,378 diagrams, graphs, and photographs; and more than 550 tables. The “interior” references, in which one entry refers to another entry that offers augmented or related coverage, the visual aids, and the index have all been completely revised, resulting in a greater facility in locating material.

VNSE 9’s only comparable competitor is the one-volume fourth edition of the McGraw-Hill Concise Encyclopedia of Science and Technology, which is older (published in 1998) and shorter (2,450 slightly smaller pages) and contains slightly fewer entries (7,800), but which contains 1,200 biographies and is less expensive ($150.00).(For a review of the second edition see Kauffman, G. B. Today’s Chemist, 1989 (December), 2(6), 16–18). VNSE 9, however, contains longer and more detailed entries.

VNSE 9’s presentation ranges from the introductory to the highly technical. These different levels make it useful for students at all levels, from elementary school through graduate school. This latest edition of a popular, standard, and critically acclaimed desktop reference source, featuring a host of state-of-the-art entries, belongs in personal, institutional, academic, professional, and industrial libraries. The eighth edition, which appeared in 1994, was later made available on CD-ROM. I hope that the ninth edition will soon be made similarly available.

George B. Kauffman

California State University, Fresno, georgek@csufresno.edu

S1430-4171(04)04811-0, 10.1333/s00897040811a

Experiments, Models, Paper Tools: Cultures of Organic Chemistry in the Nineteenth Century. By Ursula Klein. Stanford University Press: Stanford, CA, 2003. Figures, tables. xi + 305 pp, hardbound. 16.0 ´ 23.9 cm. $65.00. ISBN 0-8047-4359-2.

In January 1999 the History and Philosophy of Laboratory Sciences Research Group of the Max Planck Institute for the History of Science in Berlin began a project on “Representation as Practice,” directed by Ursula Klein, Director of MPI’s Research Group on the History and Philosophy of Chemistry and Biochemistry. The project involved the study of the tools and modes of representing invisible scientific objects in nineteenth- and twentieth-century laboratory sciences with the goal of historically and philosophically reconstructing such representational practices. By the end of 2001 most parts of the three-year project were complete. Among the results, aside from many journal articles, oral presentations, and international conferences, were a volume of edited contributions [1] and two monographs [2], one of which, Klein’s first book written (rather than edited) in English, is the subject of this review.

This book, a volume in the “Writing Science” series, edited by Timothy Lenoir and Hans Ulrich Gumbrecht, analyzes the representational practices of nineteenth-century chemists—models and “paper tools”—and demonstrates convincingly how organic chemistry ceased being a “natural” science, focused on the investigation of organic matter extracted from plant or animal tissues, and instead evolved into an experimental culture focused on synthesizing “artificial” organic compounds and reactions that do not occur “in nature.” Klein carried out research on several chapters while she was a Visiting Fellow at the Department for the History of Science at Harvard University (1996–1998) and a Resident Senior Fellow at the Dibner Institute for the History of Science and Technology (1997–1998). Her Habilitationsschrift (a paper embodying the results of original and independent research that is a requirement for lecturing at a university) at the Universität Konstanz was based on the much longer German version of the book, and several chapters draw on some of her previously published material.

In her book Klein details how, in the early nineteenth century, in a development that revolutionized the practice of science, chemistry emerged in Europe as a truly exceptional discipline. She explains how this process came about and evolved, and she argues that this experimentalization of chemistry was driven by what would seem to be a relatively innocuous tool—Swedish chemist Jöns Jacob Berzelius’ system of chemical formulas. She traces the history of this “paper tool” and demonstrates how chemistry rapidly lost its natural historical orientation and soon became a major productive force in industrial society.

According to Klein, Berzelius’ formulas were not only a convenient shorthand device of minor use in research but instead were productive “paper tools” for creating order from the chaos of early nineteenth-century organic chemistry—what Friedrich Wöhler characterized as a “jungle.” Chemists were able to use these formulas to create an entire world on paper, which they could then use to correlate with experiments and products in test tubes and flasks. These formulas made it possible to extend the order achieved in inorganic chemistry to the comparatively much more confusing and complex area of organic chemistry.

Berzelius’ formulas became tools for the experimental study of organic chemical reactions and for the construction of models of chemical reactions and the invisible constitution of organic compounds. By permitting manipulation on paper and by displaying visually possible recombinations of signs, formulas could suggest new possibilities to chemists, which they then tried to compare with experimental results. Klein’s semiotic approach to the Berzelian formulas themselves permits her to demonstrate in detail how their semantic and representational qualities made them exceptionally useful for productive application on paper.

Although a meticulously documented and cogently argued scholarly work, the book need not be read from cover to cover. Its organization reflects its interdisciplinary approach. Chapters organized along a historical narrative offering detailed historical analyses and descriptions alternate with those summarizing the historical details and linking them to the ongoing discourse in the philosophy and semiotics of science. Readers interested primarily in the latter can concentrate on Chapters 1, 2, 8, and 9 and read only the summaries of the other chapters, while readers interested mainly in historical issues may benefit more from Chapters 3–7, which also provide most of the empirical evidence for Klein’s broader contentions.

In her “Introduction” (8 pp) Klein explains her purpose in writing the book and outlines its contents. Eight of the nine chapters are prefaced by one or more pertinent quotations by chemists such as Berzelius, Berthelot, Liebig, and Wöhler and even by nonscientists like Karl Marx. In Chapter 1, “The Semiotics of Berzelian Chemical Formulas” (32 pp), she argues that Berzelius’ formulas had several layers of reference and meaning and were “compact” or “dense” signs able to convey simultaneously a plurality of information. She questions the widely held sharp distinction of Berzelian formulas as linguistic or logical signs from structural and stereochemical formulas as graphic or iconic signs. She contends that all the different kinds of chemical formulas introduced during the nineteenth century had a logical and iconographic form and that there is only a difference of degree between them. She compares the syntax of Berzelian formulas with alternatives such as ordinary language and John Dalton’s cumbersome pictorial diagrams, and she argues that the “graphic suggestiveness” and “maneuverability” of Berzelian formulas was a significant precondition for their application as paper tools beginning with the late 1820s.

Chapter 2, “Two Cultures of Organic Chemistry in the Nineteenth Century: A Structural Comparison” (45 pp, the longest chapter), compares the structure of the “pluricentered culture” of plant and animal chemistry, which overlaps with natural history, pharmacy, and chemical arts, with the “experimental culture” of synthetic carbon chemistry. Klein’s comparison involves the entire field of scientific objects, the reference and meaning of “organic” matter as a distinctive scientific object, the classification of organic substances, and the types of experiments. Her comparison shows that both forms of organic chemistry differ so much that they can be considered different cultures.

Chapter 3, “Experiments on the Periphery of Plant Chemistry” (32 pp), discusses the experiments carried out by French, German, and other European chemists on alcohol, so-called ethers, and other alcohol derivatives during the period 1794–1820. In the 1820s they moved from the periphery to the center of research, and they evolved from the context of commercial pharmacy to academic chemical laboratories.

The next four chapters discuss the application of Berzelian formulas as paper tools for studying and constructing interpretative models of organic reactions and for modeling the invisible constitution and the classification of organic compounds from 1827 to the mid-1830s.

Chapter 4, “Paper Tools for the Construction of Interpretive Models of Chemical Reactions” (12 pp, the shortest chapter), reconstructs how in 1827 French chemists Jean-Baptiste-André Dumas and Polydore Boullay used Berzelian formulas to construct a model of a chemical reaction underlying the production of ether and its by-products and to balance the masses of the reactants and products.

In Chapter 5, “Paper Tools for the Classification of Organic Substances” (19 pp), Klein describes Dumas’ and Boullay’s 1828 method of classifying organic compounds that became the model for all subsequent classifications. Based on experimental studies of the composition and constitution of organic compounds and on the later construction of formula models of constitution rather than on the observable properties of organic substances and their natural origins—the most important criteria of classification in the “pluricentered culture,” this new classification mode assimilated the classification of organic substances to that of inorganic ones.

Chapter 6, “Paper Tools for Modeling the Constitution of Organic Compounds” (39 pp), details the construction of constitutional formula models by Dumas, Justus Liebig, Berzelius, and other French and German chemists during the late 1820s and 1830s, when these models caused a vigorous controversy. From quantitative analytical results chemists constructed a Berzelian raw formula, which they manipulated to yield a binary constitutional formula model. These manipulations are related to the well-known controversy over radicals.

Chapter 7, “The Dialectic of Tools and Goals” (19 pp), examines the dialectics between both paper and laboratory tools and goals that resulted in the concept of substitution, which, between 1834 and 1840, accelerated the transformation of plant and animal chemistry into the new experimental culture of synthetic carbon chemistry. A characteristic aspect of this new culture was the creation of “robust fits” between Berzelian raw formulas and various kinds of formula models and the reflection of these fits in the new mode of justification.

Chapter 8, “The Historical Transformation Process” (24 pp), analyzes chemists’ actions during the transformation period described in Chapters 4–7 and shows how alcohol, the so-called ethers, and alcohol derivatives became “model objects” in the transformation of organic chemistry. The overall trajectories of the transformation process and the questions posed by its dynamics are also reconstructed.

Chapter 9, “Paper Tools” (17 pp), recapitulates the practical applications of Berzelian formulas, especially as paper tools for representing and constructing simple, interpretive models of chemical reactions. It also explores the advantages and limitations of Bruno Latour’s concept of “chains of inscriptions” and compares “paper tools” with laboratory tools or instruments.

The book is replete with 34 figures, diagrams, and tables featuring reproductions from pertinent classic papers, structural formulas, reaction schemes, laboratories, and apparatus such as Liebig’s “Kaliapparat” (incorporated into the logo of the American Chemical Society) and his device for saturating liquids with gases. The documentation includes 30 pages of “Notes” and 19 pages of “Literature Cited,” including 348 books and articles as late as 2001 (or 2003 if one considers Ramberg’s book, reference 2 below, in press at the time Klein was writing her book). An index of 7 double-column pages facilitates location of material.

This moderately priced monograph on the epistemology and history of chemistry will be of great interest to historians and philosophers of chemistry and of science as well as to practicing chemists, especially organic chemists, concerned with the history of their science.

References and Notes

1.       Klein, U., Ed. Tools and Modes of Representation in the Laboratory Sciences; Kluwer: Dordrecht, The Netherlands, 2001; a revised edition of contributions to an international conference held at the MPI in December 1999.

2.       Ramberg, P. J. Chemical Structure, Spatial Arrangement: The Early History of Stereochemistry, 1874–1914; Ashgate Publishing Co.: Burlington, VT, 2003. For a review see Kauffman, G. B. Chem. Educator 2004, 9, 139–141; DOI 10.1333/s00897040778a.

George B. Kauffman

California State University, Fresno, georgek@csufresno.edu

S1430-4171(04)04812-X, 10.1333/s00897040812a

Molecules, 2nd Edition. By Peter Atkins. Cambridge University Press: Cambridge, Great Britain, September 2003; 235 pp, ISBN: 0-521-5356-0 (softcover, $30.00) ISBN: 0-5218-2397-8 (hardcover, $85.00).

Atoms are to a chemist what cells are to a biologist. Molecules, therefore, are a chemist’s plants and animals. Taking this analogy one step further, in Molecules Peter Atkins, renowned textbook writer and master advocate of scientific thinking, takes us on a trip to the zoo, and a hilarious trip it is!

In the chemical zoo we meet familiar “animals” (water and ammonia) and exotic ones (methylenedioxymeth-amphetamine). We meet the chemical “skunk” (putrescine and hydrogen sulfide) as well as lovely species that please the nose (geraniol and cinnemaldehyde). There are the molecular elephants (rubber and proteins) and the tiny ants (ozone). Visitors to the Atkins menagerie will encounter the good (salicylic acid), the bad (cocaine), and the ugly (tetrachlorodibenzo-p-dioxin). Free-roaming pets (ethanol) are on exhibition alongside dangerous beasts that must be kept behind bars (nitroglycerine).

Ornamental birds (vanillin and limonene) live next to beasts of burden (glycerol and fatty acids). The twisted and the distorted (oleic acid) have their place in the molecular zoo, as does the beautifully symmetric (benzene). Naturally bred varieties (starch) contrast with those invented by the human chemist (nylon). And many more, of course. For a complete tour, get hold of a copy of Molecules, second edition.

Prior exposure to chemistry is hardly necessary, though at least a vague understanding that the material world is made up of atoms and—yes, of course—molecules may indeed be helpful. Molecules is a “coffee-table book” at its best. The author’s deep enthusiasm for his discipline and an ebullient drive to spread the “gospel of science” is unmistakably felt throughout this beautifully produced volume.

More than a hundred molecular species are introduced and explained in simple language that avoids academic jargon as much as possible. All are represented in space-filling as well as stick models, giving the advanced reader a more detailed idea of what the “creatures” look like. Being able to decipher the graphic code of the chemist is not a prerequisite to understanding the text, though. The uninitiated may just behold and wonder, while the apprentice may draw additional insight from the structures, inferring chemical activities and behavior.

Many high quality color photographs have been included that bridge the gap between the rather abstract microscopic world of molecules and the mesoscopic world of familiar objects that can be seen and touched directly. Examples are mostly taken from the living world, emphasizing the intimate relationship between chemistry (often misconstrued by lay people as a discipline opposing biology) and the science of life.

The take home message of Professor Atkins’s book is simple: molecules are everywhere, molecules are us! Admittedly a trivial statement for the chemist, yet undoubtedly a revealing, far-reaching message for a silent majority. Peter Atkins is to be congratulated for (and encouraged to continue) his efforts to reduce the scientifically illiterate part of society to as small a fraction as possible.

Thomas Lazar

Max Planck Institute for biophysical Chemistry, Goettingen, Germany, TLazar@web.de

S1430-4171(04)04813-9, 10.1333/s00897040813a