Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

An Approach to Chemical Analysis: Its Development and Practice
An Approach to Chemical Analysis: Its Development and Practice
An Approach to Chemical Analysis: Its Development and Practice
Ebook559 pages7 hours

An Approach to Chemical Analysis: Its Development and Practice

Rating: 0 out of 5 stars

()

Read preview

About this ebook

An Approach to Chemical Analysis: Its Development and Practice provides an overview of the development of chemical analysis and its application in solving analytical problems in chemistry. The text is comprised of 19 chapters that are organized into two parts. In the first part, the text covers the historical aspects of chemical. The book then proceeds to tackling methods for analysis in which the final measurement is preceded by one or more chemical reactions. The first two chapters of the second part discuss distillation and chromatography, respectively. Next, the title details the physical methods that only occasionally and incidentally need to be preceded by chemical reactions. The text will be of great use for students, researchers, and practitioners of chemistry.
LanguageEnglish
Release dateOct 22, 2013
ISBN9781483185736
An Approach to Chemical Analysis: Its Development and Practice

Related to An Approach to Chemical Analysis

Related ebooks

Chemistry For You

View More

Related articles

Reviews for An Approach to Chemical Analysis

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    An Approach to Chemical Analysis - H. N. Wilson

    errors.

    PART I

    Outline

    CHAPTER 1: THE EARLY HISTORY OF CHEMICAL ANALYSIS

    CHAPTER 2: CLASSICAL ANALYSIS AND THE IMPACT OF PHYSICAL CHEMISTRY

    CHAPTER 3: THE FIRST ORGANIC REAGENTS

    CHAPTER 4: ORGANIC ANALYSIS

    CHAPTER 5: LARGE-SCALE INDUSTRY AND VOLUMETRIC ANALYSIS

    CHAPTER 6: THE MODERN APPROACH TO CHEMICAL ANALYSIS

    CHAPTER 7: ORGANIC MICRO-ANALYSIS

    CHAPTER 8: USE OF NON-AQUEOUS SOLVENTS IN ANALYSIS

    CHAPTER 9: POLAROGRAPHY

    CHAPTER 10: OTHER ELECTROCHEMICAL METHODS

    CHAPTER 1

    THE EARLY HISTORY OF CHEMICAL ANALYSIS

    Publisher Summary

    This chapter describes the early history of chemical analysis. The first demands for chemical analyses were made by the metallurgists, whose advances today are among the most insistent causes of progress in that science. As early as the fifth century B.C., there was some form of fire assay for gold and silver. The earliest fully documented analysis was performed by a purely physical method, and in the preliminary researches, the first sound physical law—that a body is buoyed up by the amount of fluid that it displaces—was discovered. In 1556, Agricola’s important treatise De Re Metallica was published; it included the assay of mercury ores by distillation and of gold ores by cupellation with silver-free lead, and it gave adequate descriptions of balances, furnaces, crucibles, and fluxes. Perhaps the real founder of analytical chemistry was Robert Boyle, 1627–1691. Besides his work on gases, he gave a satisfactory definition of an element. He used both a hydrostatical balance and a specific gravity bottle. He was familiar with several acid–base indicators and recognized copper both by the green color that it imparts to a flame and the blue color given to solutions containing copper by volatile alkali.

    Begin at the beginning, go on to the end, and then stop.

    C. L. DODGSON

    (Nineteenth-century Oxford logician and author)

    THE first demands for chemical analyses were made by the metallurgists, whose advances today are among the most insistent causes of progress in that science. As early as the fifth century B.C. there was some form of fire assay for gold and silver. I … will refine them as silver is refined, and try them as silver is tried (Zechariah, 13, v. 9). And he shall purify the sons of Levi, and purge them as gold and silver (Malachi, 3, v. 33). Both these texts indicate that the refining and trying (i.e., assaying) of silver and gold as such were well known. But determining one in presence of the other was—as far as is known—not solved until about 200 B.C. Hieron, king of Syracuse, suspected that his goldsmith had charged him the price of gold for a crown made of an alloy of gold and silver, and asked Archimedes to ascertain its true composition. Archimedes solved the problem by determining the density of the object, having ascertained that gold was about twice as heavy as silver. Thus the earliest fully documented analysis was performed by a purely physical method, and in the preliminary researches was discovered the first sound physical law—that a body is buoyed up by the amount of fluid that it displaces.

    In the vast body of alchemical literature from the early centuries of the Christian era until the fifteenth century there is nothing of importance to the history of analysis; it was speculative, fantastic and purely qualitative—when it meant anything chemical at all. But those hard-working and hard-headed men the goldsmiths and the miners needed to know the purity of gold and silver and the metal content of their ores. Assaying was a kind of trade secret, much must have been known but little was written down until the early sixteenth century, when a number of pamphlets on assaying appeared; the subject was also dealt with in two well-known treatises. Biringuccio (Pirotechnia, 1540) described the making of cupels from bone-ash, and the separation of the noble metals by cupellation. He also explains how to make nitric acid by dry distillation of a mixture of alum, sand, and saltpetre in an alembic, and how to free it from hydrochloric acid with silver. The purified acid was used to part gold and silver, the silver is dissolved and the gold remains as a black powder. It is a remarkable fact that this method of cupellation followed by parting is still in regular use for the assay of certain ores. In 1556 Agricola’s important treatise De Re Metallica was published; it included the assay of mercury ores by distillation, of gold ores by cupellation with silver-free lead, and gave adequate descriptions of balances, furnaces, crucibles, and fluxes. Among the metals recognized were bismuth, antimony, zinc, and probably cobalt. The thing to note about the early quantitative methods, however, is that they were scaled-down versions of the processes in use for winning the metals on a large scale, and as such postulated no theoretical background.

    But about a century later tests independent of industrial prior knowledge began to appear. Glauber (1604–1670) used filter paper, was familiar with the three common acids, and described a kind of borax bead test for several metals, using a soft glass as a flux. Perhaps the real founder of analytical chemistry was Robert Boyle (1627–1691). Besides his work on gases, he gave a satisfactory definition of an element. He used both a hydrostatical balance and a specific gravity bottle. He was familiar with several acid-base indicators (violet juice, litmus and cochineal) and recognized copper both by the green colour that it imparts to a flame, and the blue colour given to solutions containing copper by volatile alkali.

    The quite sensitive test for gold in acid solution with stannous chloride, which forms purple of Cassius dates from the same period. In skilled hands it can be made semi-quantitative, and was in regular use in gold mines using the cyanide process until quite recently, even if it is not still employed. There are few other analytical methods that have had a life of over 300 years.

    During the eighteenth century chemical knowledge grew apace. Magnesia was distinguished from lime by Hoffmann (1660–1742); he also recognized magnesium sulphate in a mineral water by the appearance under the microscope of the residue on evaporation—perhaps the first chemical identification by crystal form. He also noted the red colour of the face of victims of the fumes of burning charcoal. Joseph Black (1728–1799) must also be noted. He gave the first rigid demonstration of the identity of a gas—carbon dioxide— and demonstrated that the gas evolved on heating limestone or magnesia alba was identical with that obtained on treating carbonates with acids. This was also the first demonstration of a gas taking part in a chemical reaction.

    The blow-pipe was introduced by the Swedish school of chemists by the middle of the century. Its most famous exponent was Berzelius (1779–1848). In his time one might almost say that analytical and inorganic chemistry were the same. Chimie minérale was what its name implies, and when so little was really known about the composition of the earth, much of the chemist’s work was naturally the analysis of geological specimens. This led to the discovery of numerous elements; among those characterized before 1800 were manganese, molybdenum, tungsten (long known as an unidentifiable menace in tin smelting), chromium, zirconia, titania and strontia. Scheele (1748–1786) in his investigation of manganese discovered dephlogisticated marine air, i.e. chlorine.

    But the man who converted chemistry from an empirical art into a true science was Lavoisier (1743–1794). His famous Traité Elémentaire de Chimie with its insistence on the balance and on quantitative work not only solved the old puzzle of the nature of combustion, and sounded the death-knell of the phlogiston theory, it illuminated every aspect of chemistry as it was then known. But Lavoisier had been a fermier-général and this crime was not forgiven by a popular assembly which had no need of savants, so his reward was to be guillotined on 8th May, 1794. The increasing strain on the economy of the country led to a change of opinion, and before long the contributions of le Blanc, Chevreuil, Gay-Lussac, Berthollet, Dumas and others were recognized as useful, and France became the foremost chemical nation in the world.

    Chemistry and analysis were then almost inseparable; if a chemical investigation was in progress, almost always some analysis was entailed, and the investigator—then and for many years—had to do the analysis himself, for there was no-one else to do it; quite often he did not know how to do the analysis, so a divagation into analytical research was a part of many investigations. (It still happens today that a research throws up some knotty analytical problem and the researcher either has to stop following his original object, or—if in a large organization—to call in an expert. The second course is the better, if the two men respect each other’s limitations, and neither tries to run the whole job himself. The worst plan is for the original researcher to waste a lot of time messing about with the aid of an out-of-date textbook until he has used up all the material except a milligram or two which is rather the worse for wear, and then ask for a complete analysis. And even in these distressing circumstances it is surprising how often a modern microanalyst can rescue something from the wreck.)

    Gay-Lussac is a typical figure of the times. Besides showing (independently of Charles) that gases were expanded equally by the same degree of heat he also was responsible for the law of volumes, established the composition of nitrous and nitric acid, and (with Thénard) in 1811 isolated sodium by distilling caustic soda with iron. Also with Thénard he devised the first practicable method of determining carbon in organic compounds by oxidation with copper oxide in a combustion tube—the method subsequently improved by Liebig and known for a century as Liebig’s method (1831). Liebig passed a current of purified air or oxygen through the tube, the packing of copper oxide was heated and the sample slowly vaporized in the stream of air. The copper in fact acted as a catalyst, because when the oxide was reduced to metal by the sample, it was re-oxidized by the current of air. The water arising from the combustion of the hydrogen was caught and weighed in calcium chloride tubes, the carbon dioxide by strong caustic potash solution in a potash bulb. The process, like Dumas’ method for nitrogen (1833) must have been extraordinarily difficult to carry out. The chemist had to calibrate his own nitrometer, his furnace was heated by charcoal, his combustion tube was of hard Bohemian glass, and his rubber tubing, if any, was of very poor quality. (Rubber was first examined by la Condomine in the middle of the eighteenth century; Priestley first suggested its use as a pencil eraser. Rubber tubing of a kind was made in France in 1791, apparently for surgical use. The vulcanization process was invented in 1841, and even in 1871 Crookes said that when rubber tubing was used to make a joint it should be tightly tied as otherwise the joint would leak. Pig’s bladder and intestines were the usual materials in the pre-rubber days.) Even in 1920, using gas as heating agent, an elementary organic analysis was not easy, and the virtual supplanting of the old macromethods by the simple and rapid micro-methods is a good example of the way in which analysis has been made simpler and quicker, largely through the bringing into use of materials of construction and sources of power and heat unknown to the earlier chemists.

    At about the same period Chevreuil, better known for his work on fats and soaps and for the extraordinary span of his working life, introduced the use of the melting point as a criterion of purity.

    Volumetric analysis also began in France towards the end of the eighteenth century. As so often in the history of chemical analysis, the most important step was taken to meet the requirements of a new industry. In 1785 Berthollet had seen the possibility of using what we now know as chlorine as a bleaching agent; in 1789 he wrote Pour prévenir tous les accidens qui peuvent resulter de l’energie trop grande de la liqueur, il est important d’avoir un moyen pour en mesurer la force. These words were in a paper describing the first volumetric method. It was invented by Descroisilles, a pharmacist of Rouen. Descroisilles dissolved indigo in concentrated sulphuric acid on a water bath, diluted with water and used it as a titrant, adding it to his bleach liquor until the blue colour remained. He poured the titrant, finally in drops, from a graduated cylindrical measure, calibrated by means of a primitive pipette. Descroisilles called his graduated cylinder a burette, but the original meaning of the word is a cruet (e.g. for oil or vinegar). There is an interesting paper by Greenaway(1) on the early history of volumetric analysis. Vauquelan seems to have used litmus solution in 1801 in determining the strength of acids by volumetric titration with potassium carbonate as standard; Gay-Lussac proposed sodium carbonate as standard and in 1830 invented the titration of silver nitrate with sodium chloride.

    The story is interesting; it concerns the first intervention by a government department in an analytical matter, and the first of these standardized methods of analysis that have become so popular in our own times. It became known in the middle of the eighteenth century that the cupellation method was not suitable for the assay of silver alloys or fine silver.(2) Although there is a partial compensation of errors, retention of base metal by the silver compensating for losses of silver, there is a tendency for results to be low. Too high a temperature or too prolonged heating both cause losses, and when there was no method of measuring the temperature, and too low a temperature caused troubles of other kinds, this error could not be avoided, and results were from 5 to 15 parts (i.e. parts per 1000) low in silver 800–900 parts fine. Moreover, as always happens when a procedure is imperfectly understood, some laboratories tended towards high results and some towards low. This became known and smart dealers found it profitable to buy silver near a low assay office, pay the transport charges and sell it where they could get a high certificate. This became an embarrassment to the Administration, who in 1829 set up a commission to recommend a method of analysis for general and official use. The Commission requested the assistance of Gay-Lussac and in 1830—rapid work for a government department–reported against the continuance of the dry assay for fine silver and alloys, and made Gay-Lussac’s titration method official. The task given to Gay-Lussac has a very modern sound. He was asked to devise a procedure that should,

    (i) have an error of less than 4 or 5 parts per ten thousand.

    (ii) should be simple enough for use in control laboratories.

    (iii) should be rapid.

    A modern analyst cannot fail to be reminded of the old proverb Plus ça change, plus c’est la même chose.

    Gay-Lussac’s method is to take a sample that is expected to contain just over 1.000 gram of silver; the approximate assay is of course known, as is the case with the majority of samples in most laboratories. It is dissolved in nitric acid of specified strength and volume, and to it is added 100 ml of a solution that contains sodium chloride exactly equivalent to 1.000 g silver. After vigorous agitation the precipitate flocculates and falls to the bottom of the flask. The titration is continued with a chloride solution one-tenth as strong as the former, added in small volumes until no further turbidity is produced. As a final modern touch Gay-Lussac supplied tables giving the exact titre of the alloy in terms of the weight of the sample and the number of millilitres of dilute solution used. An experienced operator can complete the assay in 15 minutes, but it is usual to work in large batches.

    The only improvement made to this method in 133 years is the substitution of a Potentiometrie method of end-point detection for use with deeply-coloured solutions and in the presence of tin, which gives a turbidity due to metastannic acid in nitric acid solution.

    The determination of reducing sugars by means of an alkaline solution of a cupric salt containing tartrate was published in 1844 (Barreswil), permanganate as a titrant for ferrous salts was described by Marguaritte in 1846, and in 1850 the first collection of volumetric methods was published–Über die Massanalysen by H. Schwartz. Mohr’s much more influentual work Lehrbuch der Chemische-analytischen Titrir-methoden appeared in two volumes in 1855–1856. It is to Mohr that we owe the familiar run-out burette, but the replacement of a rubber tube with a pinch-cock or a glass bead by a ground glass stopcock is due to Geissler, who also invented the suction pump to attach to a water-tap. Dichromate as a volumetric reagent for iron was first employed by Penny in 1850; an interesting paper by Irving(3) reviews the early history of this important method.

    REFERENCES

    1. GREENAWAY, F. Endeavour. 1962; 21:91.

    2. FAUCONNIER, P. Chimie Anal. 1960; 42

    3. IRVING, H. Science Progress. 1951; 63.

    CHAPTER 2

    CLASSICAL ANALYSIS AND THE IMPACT OF PHYSICAL CHEMISTRY

    Publisher Summary

    This chapter describes the impact of physical chemistry on classical analysis during 20th century. The chapter also discusses the application of physical chemistry to analytical problems. Faraday’s law of electro-equivalence, 1833, is sometimes regarded as the beginning of physical chemistry. Faraday was also responsible for the words anion and cation, but electrochemical ideas of a qualitative go back to Berzelius, and soon led to the placing of the metals in order of increasing electropositiveness. Ostwald’s dilution law was used for calculating approximately the distribution of the various components involved in an equilibrium as a function of the hydrogen ion concentration. Of equal importance was the work of Arrhenius, who in 1877 first clearly expressed the idea that electrolytes in solution must be dissociated into free ions even when the solution was not subject to an externally applied electric potential.

    (The author uses the word classical with some hesitation; it does not seem to carry the same meaning when used about analytical methods as when applied to art or literature, a classical method isone I have often used but I can’t remember the reference It also seems to imply a preference for long gravimetric methods of a respectable antiquity, and it is in that sense that it is used here.)

    IF WE read the older textbooks, we find that up to about 1880 or even much later, the general aspect of classical analytical chemistry is rather like a vast patchwork quilt, some of the pieces being very old indeed and taken from the empirical practices of primitive and often secretive industries such as the alum industry, and others from recent discoveries of scientific workers. But—like the contemporary chemistry—all this information is ill-organized, and follows no discernible pattern, though the amount of information is already vast. Already in 1864—the year Fresenius’ Chemical Analysis was published—qualitative analysis is very similar to that taught 60 years later, though equations are hardly used, water is still HO and Professor Bunsen’s improved gas lamp is a novelty recommended for use with the spectral apparatus, the best for the detection of calcium, strontium, and barium. Part of the vast mass of facts contained in the Fresenius volumes or in Crookes’ Select Methods of Chemical Analysis was not quite accurate, but the analysts of the period certainly knew how to make the best of what they had, and they managed to advance both science and their bank accounts just as successfully as the later generations. Some of this earlier knowledge was subsequently forgotten; Sir J. J. Fox was asked in about 1940 whether he could recollect any method of separating cadmium and zinc, not involving the use of H2S. No, he said, I can’t. But you go and look it up in the earliest edition of Fresenius that you can find, and you are sure to find something that will give you a lead, and so it turned out. Most of the processes were lengthy, difficult and complex, and almost all the final determinations were gravimetric. Volumetric analysis in general was regarded as something rather commercial and low, and not very accurate, despite the evidence of Gay-Lussac’s silver method that it was potentially more accurate than any gravimetric method then known. Now, when we have a fuller understanding of dissociation constants, oxidation potentials, solubility products and so forth, it can be maintained a priori that it is potentially more accurate than gravimetry. The initial error arose because weighing was such an accurate operation, and it was felt that if the weight of the precipitate was accurately known, the result was accurate. Of course it is not, because only too often the precipitate is impure. This must have been accentuated in the early days by the imperfect separations then available, some derived from qualitative analysis, others from processes used in preparative chemistry. To these sources of error must be added the use of many imperfectly understood reactions. Crookes himself said that analyses made and published by the most eminent chemists vary between 99.1 and 100.7 per cent; many analyses yield results between 97 and 102 per cent, while the rest never see the light at all.

    The root of the trouble was that improvement was hindered by lack of sound generalizations, could only be empirical, and it was not possible to proceed far by deduction or even by analogy. Progress was therefore by small advances on very narrow fronts, and so continued until the value of the concepts of physical chemistry became recognized by the analysts, when the new methods of thought led to new techniques and the improvement of old processes.

    An early example is the electrodeposition of copper. Faraday had put forward his law of electrochemical equivalents in 1833, and copper was first deposited by electrolysis in 1837 (but not for analytical purposes). The first known analytical application was in 1867. The story as told by Crookes (Select Methods) is interesting. The copper ores mined at Mansfeld were very variable, and the directors of the mines, being dissatisfied with their analyses which were even more variable, offered a cash prize for a new method. A committee of three (two practical assayers and the well-known Dr. Böttger) decided that

    (i) Lengthy processes, and those that required the operator to have too much scientific training should be excluded.

    (ii) Processes in which the operator judged from the appearance of the ore how large a sample to take, should be excluded.

    (iii) Evaporation of large volumes of acid, the evolution of quantities of SO2 or H2S, and potentially violent reactions to be excluded.

    (iv) Expensive reagents to be excluded.

    (v) Dry assay to be excluded, on grounds of too great expense and inaccuracy.

    (vi) Precipitation of iron, alumina, etc., to be excluded as the precipitate always carries down copper.

    These conditions have a strangely modern look; this might almost be an instruction for a modern chief analyst in a rather old-fashioned industry. (But he would be unlikely to win a cash prize for his efforts.) Item (vi) is particularly interesting, as this is one of the facts that has been repeatedly forgotten and rediscovered in the last 100 years. Out of sixteen processes submitted, two were chosen for further study. Both were electrolytic in nature. In the first the copper was precipitated direct from an acid solution by a rod of zinc fastened to a piece of platinum foil. After removal of the zinc rod, the precipitated metal was washed, dissolved in water and nitric acid, the solution made ammoniacal and titrated with potassium cyanide solution to the disappearance of the blue colour. In the second it was demonstrated for the first time that copper could be deposited from a nitric acid solution in a coherent form by the galvanic current. Mercury, silver and bismuth accompanied the copper, but were not present in interfering amounts. The copper was plated on to a platinum foil cathode, washed, dried and weighed as in present practice. But as the current used was small, the assay required ten hours to complete, so the referees chose the other method as it was rapid, needing only four hours. One would think that the much simpler iodide method (described by de Haën in 1854) would have been much quicker; perhaps the reagent was too expensive. It certainly was much used in copper mines for very many years, and could be completed in far less than an hour. By 1955 X-ray fluorescence spectrometry could do the work in about five minutes; today it could provide a continuous record of the copper content of an ore or concentrate—wet or dry—on a conveyer band or in a stream of sludge almost instantaneously.

    THE ROLE OF PHYSICAL CHEMISTRY—FROM 1870 ONWARDS

    The above account, bringing us to about 1870, also brings us to the application of physical chemistry to analytical problems—a development impossible to overestimate. In systemizing and explaining the mass of facts huddled together under the old roofs of organic and inorganic chemistry a new architecture was created, in which every stone has a proper part to play in support of its neighbours. One may think that some old courtyard or stable is complete, and perhaps it is; but discovery or application of some powerful theory or hypothesis opens out a new area, and development begins again.

    This book is not a history of physical chemistry, but it is worth while recording some of the theories that had the greatest impact on chemical analysis. These are not always the most important; analysts most of the time get on very well without thinking about the phase rule for example, though one cannot imagine a phase rule investigation that does not need large numbers of

    Enjoying the preview?
    Page 1 of 1