Transcript
Page 1: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

Synthesis of

ORGANOPHOSPHORUS COMPOUNDS

For the chemist who would engage in synthesis

or study theory, the organic chemistry

of phosphorus is a fertile area

DR. MARTIN GRAYSON, American Cyanamid Co., Stamford, Conn.

Organophosphorus chemistry is a field which has experi­enced a sharp upsurge of interest during the past decade. Contributions to our knowledge of organophosphorus com­pounds have come from many sources, and both academic and industrial laboratories have been prominent in making those contributions. Much activity has been concerned with new methods for preparing the compounds—in some cases for the sake of their unique physical and chemical properties, and in others for theoretical studies related to a variety of new reactions.

Since the field is so broad, we must necessarily limit our discussion. Therefore, my intention is to offer a brief out­line of methods for forming the carbon-phosphorus bond, with emphasis on those reactions where the phosphorus atom functions as a nucleophilic reagent.

It is probably no coincidence that a renaissance in this field began some 12 years ago, at about the same time that the encyclopedic compendium, "Organophosphorus Com­pounds," by Prof. Gennedy M. Kosolapoff, appeared. Cer­tainly, no discussion of synthetic organophosphorus chem­istry should proceed before at least mentioning this and acknowledging our indebtedness to the book for its role in sparking current interest.

Much of this interest is due to some very special proper­

ties of phosphorus which contrast sharply with those of related organonitrogen compounds. Phosphorus forms stronger bonds than nitrogen does with oxygen, carbon, and halogens, but forms a weaker bond with hydrogen. The phosphorus-carbon bond is as strong as the carbon-carbon bond. The phosphorus-oxygen single bond is stronger than the carbon-oxygen single bond by some 10 kcal., and the phosphorus-oxygen "double bond" is some 55 kcal. stronger still, with a dissociation energy of about 135 to 140 kcal. This affinity for oxygen provides much of the driving force for many of the reactions of phosphorus compounds, such as the Wittig olefin synthesis in which a new P = 0 group is formed:

R3P±=CHR'+ R"CHO—>R3P = 0 + R'CH=CHR"

The bond dissociation energies in phosphine, ammonia, and methane increase in the order given (76, 84, and 91 kcal.), and this in part accounts for the ability of phosphines to participate in free radical and ionic additions to multiple bond systems. Furthermore, phosphorus is a larger atom than nitrogen, with more readily lost or polarized non-bonded electrons. This, coupled with nearly equal basicity in analogous compounds, produces greater nucleophilicity— greater reactivity toward centers of electron deficiency.

90 C&EN DEC. 3, 1962

Page 2: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

chemical science series

A point of major difference, compared to the first row elements, which phosphorus has in common with a number of other second-row metalloids, is its ability to accommo­date up to 10, and sometimes 12, rather than only eight, valence electrons. This outer shell expansion results from vacant 3d orbitals of low energy, in addition to the s and ρ orbitals to which nitrogen and other first-row elements are limited. This availability of vacant bonding orbitals, coupled with the nucleophilicity of phosphines, permits the formation of strong phosphorus-metal bonds. Trivalent or triply connected phosphorus compounds function as power­ful ligands in many inorganic complexes. When we write multiple bonds between phosphorus and metals, carbon, oxygen, or nitrogen, we are actually describing a valence shell expansion with interaction between the vacant d or­bitals of phosphorus and the p-electrons of the other atom (άπ-ρ-π bonding). This type of interaction is sometimes called back bonding.

Along these lines, phosphorus also forms stable, five-covalent compounds such as pentaphenylphosphorane— (C6H5)5P—in which d, s, and ρ orbitals have rehybridized to give the dsps system, which can be geometrically a bi-pyramid or a square pyramid. Recent x-ray work indicates the square pyramid structure to be correct for (C 6 H 5 ) 5 P, although the geometry for most reactive phosphorane in­termediates and transition states remains in doubt. Studies of the stereochemistry and reaction mechanisms involving pentacovalent intermediates are therefore of great theo­retical interest. This is the area in which much of the most exciting fundamental work in phosphorus chemistry is being carried out today.

Many of these characteristics of phosphorus confer unusual and desirable properties on its compounds. Ther­mal stability and fire resistance are often associated with organophosphorus compounds. They can be used in gaso­line to prevent preignition. In mining, they can be used for flotation and extraction of metals, including uranium. A host of other applications could be mentioned, including adhesives, lubricants, chelating agents, textile finishes, plasticizers, and so forth. But all in all, the industrial ap­plication of organophosphorus chemistry still seems to be in its earliest stages, with much yet to come.

It is both convenient and instructive to group methods of forming phosphorus-carbon bonds into those in which the phosphorus atom functions as a nucleophilic, or electron-donating, reagent, those in which it acts as an electrophilic, or electron-accepting, reagent, and those wherein the reac­tions involve free radicals. Unfortunately, space does not permit a complete discussion of the electrophilic and free radical reactions, and in these areas I shall merely indicate the newer reactions of major, preparative value.

Nucleophilic Reactions

This class includes most of the reactions having widest application and the greatest utility for forming a phos­phorus-carbon bond. Many of these reactions are cata­lyzed by bases, or they require formation of a negatively charged phosphide ion or metal salt of a phosphine, phosphine oxide, or dialkyl phosphonate. On the other hand, some of the most useful reactions involve the acid-catalyzed formation of an electron-deficient center, such

DEC. 3, 1962 C & E N 91

C&EN

as a carbonium ion, from an alkyl halide, olefin, or carbonyl compound. This is then attacked by the nonbonded pair of electrons of the nucleophilic phosphorus atom. Base-Catalyzed Nucleophilic Reactions. Compounds con­taining a phosphorus-hydrogen bond are common starting materials for nucleophilic reactions. *Much of our own early work at the Cyanamid Laboratories made use of phos-phine (PH3) generated from metal phosphides and water. The use of aluminum or magnesium-aluminum phosphides, rather than calcium phosphide, eliminates by-product generation of biphosphine (H 2P-PH 2) , which apparently has been largely responsible for the reported hazards of this gas. Of course, even pure phosphine is toxic and flam­mable, but handling in conventional equipment presents few difficulties.

The discovery of the base-catalyzed nucleophilic addi­tion of phosphine to acrylonitrile by I. Hechenbleikner and M. M. Rauhut at Cyanamid made the interesting series of mono-, bis-, and tris-2-cyanoethylphosphines available:

In this reaction, product distribution can be altered by appropriate changes in reactant ratios, to give high yields of any one of the tNree phosphines. These materials served as useful, easily prepared models for much of our early exploratory work.

Uncatalyzed addition of acrylonitrile to phenylphosphine, as previously reported in the literature, required tempera­tures of at least 130° C. Repetition of this reaction with the ION KOH catalyst at 2C° C. gave 82% bis(2-cyano-ethyl) phenylphosphine in a t tw hours. This reaction is probably a Michael addition process, involving the nucleo­philic phosphide ion:

The absence of acrylonitrile polymers or hydrolyzed prod­ucts clearly demonstrates the greater nucleophilicity of the phosphide intermediates, compared to hydroxide ion.

Similar base-catalyzed addition reactions of dialkyl phos-phonates, probably involving the ion [ ( R O ) 2 P = 0 ] - , were reported by a number of workers in the early 1950's. The enhanced acidity of the phosphonates was probably a major factor in the early recognition of their nucleophilic properties. This highly nucleophilic ion is also the basis for the well-known and useful Michaelis-Becker-Nylen re­action with alkyl halides to give dialkyl alkylphosphonates:

Secondary phosphine oxides also contain a relatively acidic phosphorus-hydrogen bond, and they readily form the nucleophilic anion with bases. At Cyanamid, for ex­ample, the author and Patricia T. Keough found that ethylene carbonate is attacked by secondary phosphine oxides in the presence of bases such as heptamethylbigua-nide (HMBG), producing reactive 2-hydroxyethyl phos­phine oxide intermediates. These are comparable to the

Page 3: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

BASE-CATALYZED NUCLEOPHILIC REACTIONS

Secondary phosphine oxides, having a relatively acidic phosphorus-hydrogen bond, readily form a nucleophilic anion with bases. They attack ethylene carbonate in the presence of heptamethylbiguanide,

Interestingly, ethylene carbonate can be used to oxidize secondary phosphines as a starting material in

centered mechanism, as in the Wittig olefin synthesis from aldehydes and methylene phosphoranes:

Ethylene carbonate and other cyclic carbonates will also oxidize tertiary phosphines by way of a four-

92 C & E N DEC. 3f 1962

the above reactions. Secondary phosphine oxide, then, is rapidly formed concurrently with the olefin:

for example, to give 2-hydroxyethyl phosphine oxide intermediates. These react, through dehydration and Michael addition of another secondary phosphine ox­ide, to give ethylenebis-disubstitutedphosphine oxides:

Page 4: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

2-hydroxyethyl phosphines obtained by others from metal phosphides and ethylene oxide.

These materials can be isolated in some cases, but they generally react further to give ethylenebis-disubstituted phosphine oxides. This takes place through a dehydration to vinylphosphine oxide, followed by Michael addition of a second mole of starting material. The dehydration and Michael addition processes are quite rapid, and this again emphasizes the phosphinyl anion's great nucleophilicity in adding to activated olefins, as well as in displacement on saturated carbon.

The ethylene carbonate reaction is an interesting one, since this reagent can also serve as an oxidizing agent and permits the use of secondary phosphines as starting mate­rials in the reaction. The phosphine is rapidly oxidized to the secondary phosphine oxide, with concurrent olefin formation. The mechanism of this oxidation involves a phosphorus-carbon bond cleavage in an intermediate in a way that is quite analogous to the cleavage involved in the Wittig olefin synthesis. With a tertiary phosphine, the oxi­dation reaction can be isolated. The oxidation reaction may have synthesis applications in converting 1,2-glycols to olefins via the cyclic carbonate.

Nucleophilic addition of phosphinyl anions, derived from secondary phosphine oxides, to active olefins, such as acry-lonitrile, can also be carried out. Such reactions have been reported by R. C. Miller, J. S. Bradley, and L. A. Hamil­ton of Socony-Mobil and Temple University, and also by M. M. Rauhut and H. A. Currier at Cyanamid. In this connection, I should point out that, contrary to earlier reports, secondary phosphine oxides can be prepared in high yield by air oxidation of the phosphines in neutral organic solvents; phosphinic acids are formed only when strong oxidizing agents or alkaline conditions are used. Therefore, secondary phosphine oxides are readily avail­able as starting materials, obtainable from olefins and phos­phine by nucleophilic or free radical addition reactions. Furthermore, work by S. A. Buckler and M. Epstein at Cyanamid has led to the discovery of hitherto unknown primary phosphine oxides, R P ( 0 ) H 2 , and these materials also act as nucleophilic reagents in base-catalyzed addition reactions.

With the added driving force of ring formation, base catalysis becomes unnecessary. This was shown by R. P. Welcher and Nancy E. Day, of Cyanamid, in their diaddi-tion of primary phosphines to conjugated dienones. The product 4-phosphorinones are six-membered rings, contain­ing carbonyl and tertiary phosphine functions.

In terms of commercial potential, nucleophilic addition of phosphine to active olefins, such as acrylonitrile, offers perhaps the cheapest route to trivalent organophosphorus compounds containing functional groups. A considerable body of knowledge concerning the reactions of these ma­terials has been developed, but discussion of that chem­istry is beyond the scope of this article. Acid-Catalyzed Nucleophilic Reactions. In general, nucleo­philic attack by phosphorus compounds on unsaturated centers, such as olefins and carbonyl groups, is promoted by acid-catalyzed formation of an electron-deficient center, such as a carbonium ion. Generation of a carbonium ion is essential in reactions involving the weakly nucleophilic phosphorus trihalides. For example, alkyl halides are con­verted to carbonium ions by aluminum chloride in a Friedel-Crafts type of reaction with phosphorus trichloride, giving complex salts containing a new phosphorus-carbon bond:

Organic phosphorus compounds have valuable flame-retardant properties. Helen C. Gillham, American Cyanamid, compares organophosphorus-treated plas­tic strip with a burning untreated strip (right)

These salts can be converted to alkyl phosphonic dichlorides [RP(0)C12] or the corresponding acids or esters, depend­ing on work-up procedures. Reducing agents such as metals or other phosphorus compounds, including elemental phosphorus, can also be used to convert the complexes to alkyl phosphonous dichlorides (RPG12).

In 1951, G. O. Doak and L. D. Freedman, at the Univer­sity of North Carolina, reported an equally general and useful method for preparing aromatic phosphonic or phos­phinic acids. Although not acid-catalyzed, this method is formally a nucleophilic substitution by phosphorus tri­chloride or RPC12 on an aryl cation, formed by copper-catalyzed loss of nitrogen from a diazonium salt. The re­action may well involve radical intermediates, however, as in the related conversion of tertiary phosphines to quater­nary phosphonium compounds by diazonium salts.

In 1952, H. C. Brown of Purdue, in a patent to Standard Oil ( Ind.) , disclosed a clear-cut case of acid-catalyzed nu­cleophilic addition of phosphine to olefins. This was am­plified by M. C. Hoff and P. Hill of the same company in 1959. Presumably, the olefin is protonated on the least substituted carbon to give a secondary or tertiary car­bonium ion, and carbon-phosphorus bond formation fol­lows. Primary phosphines are the chief products, as the result of salt formation with the acid. Furthermore, at the University of Mainz, L. Horner and H. Hoffmann have shown that stable quaternary phosphonium compounds can be formed by protonation of the zwitterion produced by adding tertiary phosphines to multiple bond systems. In the absence of protonation, anionic polymerization can be initiated with active olefins; and with highly electro­negative groups on the olefin, such as fluorine, acid catalysis may not be required. D. C. England and G. W. Parshall of Du Pont, for example, disclosed the noncatalyzed ad­dition of phosphine to tetrafluoroethylene in 1959. In addition, compounds that ionize to stable carbonium ions do not require catalysis either. Triphenylmethylcarbinol

DEC. 3, 196 2 C&EN 93

Page 5: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

falls into this class, and with phosphorus trichloride it forms triphenylmethylphosphonic dichloride.

One of the most interesting acid-catalyzed nucleophilic reactions involves adding phosphines to carbonyl com­pounds. Early work with phosphine or phosphonium io­dide and aldehydes in organic solvents was carried out in the past century. In the 1920's, acid was used to prepare tetr akis ( hydroxy methyl ) phosphonium chloride ( THPC ) from formaldehyde and phosphine. In the general reac­tion, R can be either an alkyl group or hydrogen:

Recently, the utility of metal salt catalysts in the absence of acid, as well as the uncatalyzed reaction of aqueous formalin and phosphine under slight pressure, was reported for this reaction.

At Cyanamid, this area has been reopened by S. A. Buckler and M. Epstein, with interesting results. They found, for example, that ketones and aromatic aldehydes differ from the simple aliphatic compounds. Apparently, transfer of oxygen from carbon to phosphorus, in addition to phosphorus-carbon bond formation, takes place in strong mineral acid solutions. The primary phosphine oxides formed with ketones can predominate, if large groups block the carbonyl group. Cyclotetraphosphines (RP)4 , are also formed in small amounts in a few cases. A carbonium ion intermediate following normal carbonyl addition has been proposed as the mechanism, since strong acid and the presence of groups that can stabilize the positive charge are required. The reaction is completed by water addition and suitable proton shifts. A hydride shift from phos­phorus to carbon was excluded as a possibility, since S. Trippett of Leeds University has shown that 1-hydroxyalkyl tertiary phosphines give a related reaction in strong acids.

In addition to demonstrating that the reaction of un-branched aldehydes to give tetr akis 1-hydroxy alky lphos-phonium salts is quite general, Buckler and Epstein, along with V. P. Wystrach, reported formation of a number of interesting cyclic, spirocyclic, and polycyclic products when branched aldehydes, dialdehydes, β-diketones, and pyruvic acid were used. For example, isobutyraldehyde reacted with phosphine in aqueous hydrochloric acid to give 2,4,6-triisopropyl-l,3-dioxa-5-phosphacyclohexane; glutaralde-hyde gave 1,5,7,1 l-tetrahydroxy-6-phosphoniaspiro [5,5] undecane chloride; 2,5-pentanedione gave the polycyclic acetal 1,3,5,7-tetramethyl-2,4,8-trioxa-6-phosphaadaman-tane; and pyruvic acid reacted to give 2,2',2"-phosphini-dyne trilactic acid trilactone. Reactions of this type are not limited to phosphine; primary, secondary, and tertiary phosphines can be reacted with simple aldehydes to give quaternary phosphonium compounds, and reactions lead­ing to more complex products, such as those described above, are possible with some primary and secondary phosphines.

Oxidized phosphorus compounds containing a phos­phorus-hydrogen bond can also be used. For example, hypophosphorous acid and pyruvic acid react to give bis ( 1-carboxy-1 -hydroxyethyl ) phosphinic acid ( CHEPA ) ,

which is also obtained by the nitric acid oxidation of the pyruvic acid—phosphine adduct. CHEPA is a potent sé­questrant for ferric ions at high basicity, a property which very few commercial chelating agents have.

In addition to these methods, aldehydes—especially form­aldehyde—and ketones react thermally with phosphorus-chlorine and phosphorus-hydrogen compounds to give products containing phosphorus-carbon bonds. Chloro-methylphosphonic dichloride is prepared commercially this way.

It is clear that carbonyl addition reactions of phosphine offer broad, general techniques for synthesizing a host of new phosphorus-carbon compounds. Compounds having both oil and water solubility can be made. The reactivity of the phosphorus-carbon-hydroxyl group in esterification, ether formation, dehydration, and so on, has been largely unexplored. This appears to be a potentially fruitful area for development activity and theoretical study. Simple al­dehydes and ketones are low-cost commercial materials, and their acid-catalyzed nucleophilic reactions with phos­phorus-hydrogen compounds may become major building blocks for a new family of industrial chemicals. Nucleophilic Quaternization Reactions. Some of the best known methods for forming carbon-phosphorus bonds in­volve alkylation of trivalent phosphorus compounds by a bimolecular, nucleophilic substitution (SN2) reaction to give quaternary phosphonium compounds. Tertiary phos­phines and alkyl halides provide perhaps the simplest ex­ample of this, and recent studies of the reaction provide new insight into the effect of structure on the nucleo-philicity of phosphorus.

Aside from the alkylation of phosphines, which goes back over 100 years to A. W. Hofmann's pioneering work, the most important example of the quaternization process for phosphorus-carbon bond formation is the classical Michaelis-Arbusov reaction. This is a two-step process. Quaternization of a tertiary phosphite is followed by de-alkylation by halide ion, to give a dialkyl alkylphosphonate ester. The quaternary intermediates can be isolated with triarylphosphites, and the dealkylation step is greatly aided by simultaneous formation of the stable phosphoryl group.

This type of driving force is the key to many important reactions of phosphonium compounds. For example, at Cyanamid we found that tertiary phosphine oxides, ethers, and hydrocarbons can be obtained by reacting benzylphos-phonium salts with alkoxides. The slow step in this reac­tion again is a dealkylation, this time by alkoxide ion, to give P = 0 and ROR.

In much the same way, the rate of the interesting hy­droxide cleavage of phosphonium salts is governed by the phosphine oxide formation step. This reaction has been shown by W. E. McEwen of the University of Massachu­setts, by H. Hoffmann at Mainz, and by recent work at Cyanamid with THPC, to be effectively second order in hydroxide ion. The basic mechanism involving pentaco-valent phosphorus was proposed in 1929 by G. W. Fenton and C. K. Ingold of University College, London.

Arylation of phosphines is, of course, much more difficult to achieve than alkylation. However, a number of useful free radical procedures, involving diazonium salts and Grignard reagents, have been developed in recent years. The use of aluminum or nickel salts to prepare tetraaryl-phosphonium salts from aryl halides has also been reported. But the simple nucleophilic displacement reaction provides the greatest insight into the effect of structure on the reac-

94 C & E N DEC. 3, 1962

Page 6: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

ACID-CATALYZED NUCLEOPHILIG REACTIONS

The acid-catalyzed reaction of phosphine with ketones aliphatics—besides formation of the Ρ—C bond, oxygen or aromatic aldehydes differs from that with the simple apparently transfers from carbon to phosphorus:

Ketones give predominantly primary phosphine oxides, tion of a carbonium ion intermediate, after normal car-if large groups block the carbonyl group. Forma- bonyl addition, has been proposed as the mechanism:

l,3,5,7-tetramethyl-2,4,8-trioxa-6-phosphaadamantane

2,2' ,2"-phosphinidyne trilactic acid trilactone

DEC. 3, 1962 C & E N 95

2,4,6-triisopropyl-l ,3-dioxa-5-phosphacyclohexane

1,5,7,11 -tetrahydroxy-6-phosphoniaspiro [5,5] undecane chloride

Page 7: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

NUCLEOPHILIC QUATERNIZATION The classic Michaelis-Arbusov reaction is the most quaternization of a tertiary phosphite, which is important example of phosphorus-carbon bond for- followed by dealkylation by halide ion to give a mation by quaternization. The process involves dialkyl alkylphosphonate ester:

The dealkylation step is helped by simultaneous type of driving force is the key to many of the reac-formation of a stable phosphoryl group, and this tions of the phosphonium compounds:

tivity of trivalent phosphorus, and this is the one that most merits our attention.

W. A. Henderson and S. A. Buckler, at the Cyanamid laboratories, found that the reaction rate of a variety of tertiary phosphines with ethyl iodide could be correlated with the inductive effects of the substituents on phosphorus. This means that resonance interactions between aromatic groups and phosphorus are weak. Here again, phosphorus and nitrogen differ considerably, and this is an example of the general notion that second-row elements form weak ρπ-ρπ double bonds. Deviations from the inductive effect

correlation were observed with methylphosphines, which reacted faster or were more nucleophilic than predicted. Likewise, such deviations were also found with sterically hindered phosphines, such as triisobutyl, which was less reactive than predicted on the basis of inductive effects alone.

The explanation for the enhanced nucleophilicity of methylphosphines is complex and tentative. Certainly, more work is needed to understand the factors governing this case. Geometry and bond hybridization are both im­portant in this respect. Understandably, electron-supply-

96 C & E N DEC. 3, 1962

The basic mechanism for this type of reaction, Ingold of London in 1929, involves pentacovalent which had been proposed by G. W. Fenton and C. K. phosphorus:

Page 8: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

ing groups generally enhance the nucleophilic reactivity of phosphines by increasing the negative charge or density of the free electron pair on phosphorus. Further, the reaction goes faster in more polar solvents, as might be expected from the increase in charge separation in the transition state. Greater solvation, of course, reduces the activation energy in reactions producing charged species, and this is primarily responsible for the rate increase.

In a related study, W. A. Henderson and C. A. Streuli found that the phosphine basicity is also correlated pri­marily by inductive effects with deviations, again princi­pally the result of steric factors. Since we are dealing in this case with protonation, rather than interaction with a more or less bulky alkylating agent, the steric effect is prob­ably related to solvation energy factors. This is apparent from the operation of such an effect in the secondary and primary phosphines, as well as in the tertiary series. It is also interesting that, in contrast to the nitrogen bases, arylphosphines fall on the same straight-line correlation of basicity with inductive effect as do the alkylphosphines. Finally, it is noteworthy that, although the amines are gen­erally stronger bases than the phosphines, the order with respect to the degree of substitution is reversed with equiv­alent substituents ( R N H 2 > R 2 H N > R 3 N > R 3 P > R 2 P H > -RPH2).

Quaternary phosphonium salts themselves may serve as useful synthetic intermediates for the formation of new carbon-phosphorus bonds. Our work has shown that 2-cyanoethylphosphonium salts undergo elimination with base (E2); and the resultant tertiary phosphine, less a 2-cyanoethyl group, can be requaternized with a new alkyl halide to replace these groups successively, in effect, with alkyl substituents. Similar work has been done by several groups with hydroxymethylphosphonium salts. In this case, bases reverse the aldehyde addition reaction, and stepwise quaternization and base treatment procedures to give new alkylphosphines can be carried out, starting with commercially available THPC. Stoichiometry is important with THPC and some alkyl tris (hydroxymethyl) phospho­nium salts, since excess aqueous caustic gives hydrogen gas and a tertiary phosphine oxide instead of a phosphine.

In terms of intrinsic properties, quaternary phosphonium compounds show interesting utility in a number of areas where the unique properties of the phosphorus atom, coupled with an ionic character, seem to play a key role. Incorporation of these properties in plastics and other poly­mers may have far-reaching commercial significance. As far as the chemistry of organophosphorus compounds is concerned, quaternary compound reactions offer some of the most fertile subjects of study concerning the stereo­chemistry and mechanisms of fundamental organic proc­esses in this field. Not the least of these is the "ylid" chemistry of the conjugated bases (methylene phospho-ranes) that take part in the Wittig and similar reactions.

Elemental Phosphorus

Because of its importance as the lowest-cost source of the element in a reactive form, the reactions of white phosphorus merit a special place in any survey of general methods for forming phosphorus-carbon bonds. Most of the interesting new reactions involve nucleophilic attack on the P4 tetrahedron by some anionic reagent, followed by reaction of the resulting phosphide nucleophile with an un­saturated or potentially electron-deficient center. The

Monsanto scientist Dr. Gail Birum {center) dis­cusses an organopolyphosphorus molecule with G. A. Richardson {left) and J. L. Dever. Organophosphorus compounds show promise as additives and plasticizers, as well as for many other uses

new phosphorus-carbon bond may be formed at either stage of the process. Fittingly, then, this class of reactions comes between the nucleophilic and electrophilic reactions of phosphorus.

An interesting old reaction between olefins, white phos­phorus, and oxygen—which may involve free radicals—was re-examined in 1958 by C. Walling of Columbia University. In this process, intermediate products (which may be poly­mers) containing carbon-phosphorus and carbon-oxygen-phosphorus bonds are formed in organic solvents, but hydrolysis or oxidation with nitric acid gives a 2-hydroxy-alkylphosphonous or -phosphonic acid. A number of even older reactions of limited value are in the literature. These include high temperature, sealed-tube reactions of alcohols and alkyl halides. In all of these reactions, mixtures of products containing from one to four carbon-phosphorus bonds are formed.

More recently, the reaction of phosphorus with trifluoro-methyliodide was reported to give a mixture consisting of (CF3)3P, (CF3)2PI, and CF3PI2 with the tertiary phos­phine predominating. In 1959, L. Maier of Monsanto found that, in general, lower alkyl halides react with red phosphorus and copper powder in a hot tube to give pre­dominantly the phosphonous dihalide, RPC12, derivative. Perhaps the only other useful method available in the past for preparing organophosphorus compounds from the ele­ment involves the use of sodium metal. This method has not been applied extensively, and the possibilities of form­ing carbon-phosphorus bonds this way remain to be fully explored. It appears likely that metal-phosphorus systems may be significantly useful in future synthetic methods.

As a part of our over-all phosphorus research effort at Cyanamid Central Research Laboratory, M. M. Rauhut and A. M. Semsel, in their study of the chemistry of elemental phosphorus, discovered several new reactions for preparing organophosphorus compounds. These reactions are re­markable for their mild conditions and simple products.

The first type of reaction involves strong aqueous base-promoted addition of phosphorus to an active olefin at temperatures of 30° to 35° C. An organic solvent, such as acetonitrile or ethanol, is used; and tertiary phosphine

DEC. 3, 196 2 C&EN 97

Page 9: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

SUGGESTED ADDITIONAL READING

Berlin, K. D., and Butler, G. B., "The Preparation and Properties of Tertiary and Secondary Phosphine Oxides," Chem. Rev., 60, 243 (1960).

Crofts, P. C , "Compounds Containing Carbon-Phos­phorus Bonds," Quart. Rev. (London), 12, 341 (1958).

Frank, A. W., "The Phosphorus Acids and Their Deriva­tives," Chem. Rev., 61, 389 (1961).

Freedman, L. D., and Doak, G. O., "Preparation and Properties of Phosphonic Acids," Chem. Rev., 57, 479 (1957).

Kosolapoff, G. M., "Organophosphorus Compounds," John Wiley & Sons, New York, N.Y., 1950.

Schollkopp, U., "The Wittig Reaction," Angew. Chem., 71,260 (1959).

Trippett, S., "The Wittig Reaction," Advan. Org. Chem., Vol. I, 83, Interscience (John Wiley & Sons), New York, N.Y., 1960.

VanWazer, J. R., "Phosphorus and Its Compounds," Interscience (John Wiley & Sons), New York, N.Y., 1958.

oxide, the sole organic product, separates as a solid from the reaction mixture. Acrylonitrile is converted to tris (2-cyanoethyl)phosphine oxide (TPO) in 72% yield:

P4 + CH2=CHCN + KOHaq ^ ? , N > (NCCH2CH2),P=0 +

KH2P02 + K2HP03

Hypophosphite and phosphite salts are by-products, since about 50% of the phosphorus is converted to the organic product. Acrylamide also reacts in this way in ethanol as solvent to give tris(2-carbamoylethyl)phosphine oxide (CARPO) in high yield. The reaction probably occurs by a series of cleavages of phosphorus-phosphorus bonds by hydroxide ion, followed by nucleophilic addition of the phosphide intermediates to the olefinic group. CARPO and TPO, or their derivatives, are of considerable interest as components in certain resin formulations, where fire resistance and heat stability are imparted.

The second type of reaction involves formation of the carbon-phosphorus bond in the initial stage of attack on the P4 tetrahedron. Organometallic reagents are the nu-cleophiles in this case, and the type of product is deter­mined by the nature of the electrophilic reagent used to quench the first organophosphides produced. For example, reaction with aryl Grignard or lithium reagents in ether or tetrahydrofuran, followed by hydrolysis with water gives primary aryl phosphines in 20 to 40% conversions of phos­phorus, with occasional trace amounts of secondary and tertiary phosphine as coproducts. If alkyl halides instead of water are added to the reaction mixture, the phosphorus is converted in 50 to 80% yields into about equal amounts of two unsymmetrical tertiary phosphines. These can be derived from 2 moles of organometallic and 1 mole of the alkyl halide, or vice versa:

98 C & E N DEC. 3, 1962

With alkyl organometallics, such as butyl lithium or sodium and butyl halides, secondary and tertiary phos­phines are formed. However, butylmagnesium bromide reacts with phosphorus and butyl bromide in tetrahydro­furan to give tetrabutylcyclotetraphosphine as the major product. Because of their high phosphorus content, cyclo-tetraphosphines may become a future source of valuable gasoline additives:

From studies made of the reaction and the forma­tion of amorphous organopolyphosphides, especially with hydrolytic quenching, it is evident that the reaction in­volves complex intermediates that contain phosphorus-phosphorus bonds with varying numbers of organic groups per phosphorus atom. These reactions provide access to organophosphorus compounds that would be difficult to prepare by conventional methods. Also, they illustrate the concept that nucleophilic cleavage of phosphorus-phosphorus bonds provides a general approach to the chemistry of elemental phosphorus.

Electrophilic Reactions of Phosphorus

Reactions in which nucleophilic reagents such as car-banions, olefins, and aromatic compounds attack phos­phorus, with the formation of a new phosphorus-carbon bond, provide many useful synthetic methods. Generally, the new bond results from halogen replacement by an organometallic reagent, since the electrophilic reactivity of phosphorus is enhanced by electronegative groups. Alkoxy groups can also be replaced. In addition, other general methods include the addition of phosphorus pentachloride to olefins and acetylenes, to give unsaturated phosphonic acids, as well as phosphonation of aromatic hydrocarbons and olefins by phosphorus trichloride and aluminum chlo­ride. Aromatic substitution may also be achieved with P4S10 and A1C13, or with P4S10 or P0O5 alone at higher temperatures. Aryl phosphonic acids are the ultimate products. More recently, phosphorus trichloride has been added to olefins in the presence of aluminum chloride to obtain products of yet undetermined structure containing a phosphorus-carbon bond. In all of these Friedel-Crafts-like reactions, a PCL^AIC^ complex containing P C l 2

e -A1C14^ is probably the active reagent.

Free Radical Phosphorus-Carbon Bond Formation

Free radical methods for making organophosphorus com­pounds have considerable preparative value, and they also serve to illustrate the chemistry that prevails in many oxidation and sulfur abstraction reactions of trivalent phosphorus.

One fundamental underlies the free radical chemistry of phosphorus. This is the ready homolysis of the phos­phorus-hydrogen and phosphorus-chlorine bonds. This is probably a dual function of their relatively low bond strengths in trivalent, as well as tetravalent, or quadruply connected compounds, and of the ability of phosphorus to sustain an odd electron. The latter is especially significant with trivalent compounds such as PC13, P (OR) 3 , and PR8, where the phosphoranyl radical, Ξ Ρ - Ζ , forms readily by addition of radical Z*, which may be alkyl, aryl, alkoxy, thiyl, halo, or some other group. Presumably, empty d

Other electrophilie quenching reagents can be used. For example, propylene oxide leads to 2-hydroxypropylphos-phines and phosphonium salts.

Page 10: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

CONVERSION OF ELEMENTAL PHOSPHORUS Most interesting new reactions for forming phos­phorus-carbon bonds, starting with elemental phos­phorus, involve a nucleophilic attack on the P4 tetra­hedron by an anionic reagent, followed by reaction of

orbitals of phosphorus play an important part in giving some stability to this reactive, nine-electron configuration.

Perhaps the most useful reactions in this class are those involving the peroxide, ultraviolet, or x-ray catalyzed ad­dition of phosphorus-hydrogen and phosphorus-chlorine compounds to olefins. The peroxide-induced addition of phosphorus trichloride to an olefin to give 2-chloroalkyl-phosphonous dichlorides was reported in 1945 by M. S.

the new phosphide nucleophile with an unsaturated or potentially electron-deficient center. Acrylonitrile is converted to tris(2-cyanoethyl) phosphine oxide, with hypophosphite and phosphite salts as by-products:

Kharasch, E. V. Jensen, and W. H. Urry at the University of Chicago. Homolysis of the phosphorus-chlorine bond was proposed as the initiation and chain-carrying steps. Addition of the dichlorophosphinyl radical to the olefin provides the carbon-phosphorus bond formation.

The use of phosphorus-hydrogen compounds was dis­closed 10 years ago by A. R. Stiles, F. F. Rust, and W. E. Vaughn of the Shell Development Co., who reported addi-

DEC. 3, 1962 C&EN 99

Page 11: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

DR. MARTIN GRAYSON is a group leader in the chemical research depart­ment of American Cyanamid's Central Research Division, where he has been doing organophosphorus research for the past six years. Previously, he had been with Allied Chemical's Nitrogen Division. He completed his under­graduate work with honors at New

York University and received his Ph.D. from Purdue Uni­versity in 1952, where he had been an AEC predoctoral fellow. As an undergraduate he was elected to Sigma Xi and Phi Beta Kappa and was awarded the AIC medal for chemistry. Dr. Grayson is coeditor of "Progress in Phos­phorus Chemistry,>y with Dr. E. J. Griffith of Monsanto Chemical, a series of critical reviews to be published by Interscience (John Wiley b- Sons).

tion to olefins in the presence of free radical initiators, such as di-f-butylperoxide. Primary, secondary, and tertiary phosphines were produced with phosphine. The avail­ability of low-cost olefins makes this one of the most attrac­tive routes to phosphines for commercial exploitation.

A study of the scope and utility of the phosphine-olefin reaction was carried out at Cyanamid by M. M. Rauhut, H. A. Currier, A. M. Semsel, and V. P. Wystrach. These workers found that a,a'-azobisisobutyronitrile (AIBN) initiated the reaction more effectively than peroxides. Ex­cellent yields of primary or tertiary phosphines could be obtained with appropriate PH3/olefin reaction ratios. Further, they noted a steric effect with shielded double bonds and, with cyclohexene, no tertiary phosphine was formed. Unsymmetrical products were obtained with primary or secondary phosphines as reactants, and re­active olefins such as acrylonitrile, styrene, and ethyl acry-late were used successfully. Reaction with acetylenes gave olefinic phosphines in moderate yields.

In 1960, E. K. Fields and R. J. Rolih of Standard Oil (Ind.) reported an interesting free radical phosphonation of aromatic compounds involving addition of a phosphonyl radical, generated by di-f-butyl peroxide and a dialkyl phosphonate, to the aromatic ring.

Radical reactions involving addition to phosphorus, to give the phosphoranyl radical as an intermediate, include the formation of phosphonium salts from triarylphosphines and phenyl or halomethyl radicals, and the unusual oxida­tive chlorophosphonation reaction. The latter was first reported by J. O. Clayton and W. L. Jensen of California Research Corp. in 1948, and it entails the formation of alkylphosphonyl chlorides, RPOCl2, from phosphorus tri­chloride, oxygen, and saturated hydrocarbons. Excess phosphorus trichloride is used, since much of this reagent is oxidized directly to POCl3. Yields are generally poor, and isomeric mixtures similar to those obtained in chlorina-tion are produced. The reaction works with many organic materials as well as substituted phosphorus halides.

Mechanistic studies have been reported by C. E. Boozer and R. L. Flurry of Emory University and by F. R. Mayo of Stanford Research Institute. The reaction is obviously complex; but formation of the carbon-phosphorus bond probably proceeds by direct addition of a hydrocarbon radical to phosphorus trichloride, which can compete effec­

tively with oxygen as a scavenger. The hydrocarbon radical is probably generated by chlorine atoms, which in effect serve as chain carriers. Oxidation is no doubt partly accomplished by peroxy radicals, although largely so by oxygen, with a phosphorus-peroxy structure, Ph3RP—O—O, as a possible intermediate.

The radical chemistry of trivalent phosphorus is a fascinating subject for the theoretical investigator, as well as for the practical polymer chemist. Polymers with phosphorus in the backbone have been made by radical reactions—the reactions with oxygen, disulfides, and many others not involving carbon-phosphorus bond formation are radical in nature. Much remains to be done in elucidating mechanisms and stereochemistry of these re­actions, and the potential in this area of organophosphorus chemistry appears to be considerable.

It is apparent that organophosphorus chemistry has come through a period of vigorous activity during the past few years, with many general principles being uncovered and many novel and convenient synthesis methods being found. Clearly, this broad base of fundamental knowledge places us in a position to advance farther in several technical and theoretical directions. Applying nucleophilic addition re­actions of phosphine to unsaturated systems in an industrial way should bring on a variety of new commercial materials.

This kind of development requires little more than a good commercial process for phosphine production. Such a process might be based on elemental phosphorus, or better still, on phosphate minerals. In commercial terms, direct synthesis of organophosphorus compounds from elemental phosphorus has many advantages. But funda­mental work still needs to be done to develop selective, low cost procedures for trivalent compounds. On the other hand, results obtained so faf clearly indicate that such reactions are attainable.

For the investigator in synthesis and theory, the or­ganic chemistry of phosphorus is represented by a map of predominantly unexplored territory. With respect to new phosphorus-carbon bond formation, for example, nu­cleophilic addition to heterocyclic systems seems wide open for study. In the entire area of radical addition and pentavalent stereochemistry much remains to be done. We know something about diphosphine and cyclotetraphos-phine chemistry, but little about other linear and cyclic organic poly phosphines. Then, of course, there are the carbon-like metalloid elements—silicon, germanium, and others—whose phosphorus chemistry is still in its infancy. In all of these studies, nuclear magnetic resonance will provide clear guideposts to the structural chemist who incorporates phosphorus-31 into new molecules.

REPRINTS . . . . . . of this article on organophosphorus chemistry are available at the following prices:

One to nine copies—75 cents each 10 to 49 copies—15% discount 50 to 99 copies—20% discount Prices for larger quantities on request

Address orders to Reprint Department, ACS Applied Pub­lications, 1155 16th St., N.W., Washington 6, D.C.

100 C&EN DEC. 3f 196 2

Page 12: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

CHEMICALS EXCHANGE

5-CYANO INDOLE WRITE FOR C A T A L O G #4 The most complete price list

of Research Chemicals.

ORGANOMETALLICS Dimethylmercury $16/25 gm Triphenylarsine Oxide $11/25 gm Tetraethylgermane $32/25 gm Tetraethylt in $10/25 gm For our complete list of organometallics write to

METALLOMER LABORATORIES Box 152, M a y n a r d . Mass . __

W O E L M A d s o r b e n t s for Column Chromatography:

Alumina basic, neutral, and acid New for TL Chromatography:

three special types of Alumina, Silica Gel, Magnesium Silicate,

and Polyamide

1,1,1,4,4,4,Hexafluorobutyne-2 CF3CsCCF3

Write for Catalogue of Organic Research Chemicals

TRADEMARK

NEW SCINTILLATORS DIMtTHYLPOPOP*

Better than POPOP: Fl. Max.—120A longer Three times more soluble Pulse height identical

LIQUID SCINTILLATOR HF Safe, ready-to-use solution of PPO and Dimethyl -POPOP*

in a high flash-point solvent.

Flash Point—140 to 150°F. Fl. Max.—4300A Relative Pulse Height—Equivalent to PPO: POPOP in Toluene

Write Dept. " D " for Data Sheets and further information.

'"Arapahoe's exclusive new Scintillator—Patent applied for.

ARAPAHOE CHEMICALS, INC. 2 8 5 5 WALNUT STREET · BOULDER, COLORADO PRODUCERS O F FINE ORGANIC CHEMICALS

DIRECTORY SECTION CHEMICALS EXCHANGE—Chemicals, Resins, Gums, Oils, Waxes, Pigments.

EQUIPMENT MART—New and Used Equipment, Instruments. TECHNICAL SERVICES—Consultants, Professional Services.

Rates: $50 per inch. Lower rates on contract basis. An "inch" measures Vs" deep on one column. Additional space in even lineal inch units. Maximum space—4". Closing date 21 days in advance of publication. Flush mount cuts.

CHEMICALS EXCHANGE

VINYL FORMATE B.P. 44-46 °C

LINDEN LABORATORIES, INC. P. O . Box 302, Stale Coll«s«, Pennsylvania

0-AMINOCROTONITRILE /?-AMIN0CR0T0NIC ACID ESTERS

Literature upon request

GALLARD-SCHLESINGER CHEMICAL MFG. CORP. 1 0 0 1 F R A N K L I N A V E . , G A R D E N C I T Y , L. Ι . , Ν. Υ .

TECHNICAL SERVICES

CLARK MICR0ANALYTICAL LABORATORY

C H , N, S, Halogen, Fluorine, Oxygen, Alkoxyl, Alkimide, Acetyl, Terminal Methyl, etc., by special­ists in organic microchemical analysis.

Howard S. Clark, Director P. O . Box 17, Phone: 217-367-8406, Urbana, III.

Ο M A D I S O N 1 . W I S C O N S I N P. O . Box 1 1 7 5

A L p i n e 6 - 5 5 8 1

• H O U S T O N 6 , TEXAS 2 4 0 5 N o r f o l k S t r e e t

J A c k s o n 6 - 3 6 4 0

• W A S H I N G T O N .6, D. C. 1 7 0 0 Κ S t r e e t , N . W .

S T e r l i n g 3 - 6 5 1 0

SCHWARZKOPF MICROANALYTICAL LABORATORY 56-19 37th Ave., Woodside 77, New York

Telephone: HAvemeyer 9-6248, 9-6223 Complete Analysis of Organic Compounds.

Results within one week. Elements, Functional Groups, Molecular Weight

Physical Constants, Spectra ANALYSIS OF ORGANO METALLICS,

BORO-FLUORO AND SILICON COMPOUNDS Trace Analysis Microanalytical Research

EQUIPMENT MART

How To Get Things Done

BOARDMASTER VISUAL CONTROL Y o u r operat ions are pictured at a g lance .

Y o u save t ime, money and prevent mixups by See ing W h a t is Happen ing at all t imes. Ideal for Product ion, Maintenance , x^wîièr^v Inventory, Schedul ing, Sa les , E tc .Γ $ Î Q M ) Easy to U s e . Y o u wr i te on cards, ̂ ϋ - ^ snap on meta l board . Over 7 5 0 , 0 0 0 in U s e .

24-Page BOOKLET No.50-K Mailed Without Obligation

G R A P H I C S Y S T E M S 925 Danville Road · Yanceyyille, N.C

Bench-Type ,

white chemical porcelain

Two piece construct ion Sturdy and

simple design Two sizes 2 and 7 gal

working capacity Both sizes less than

10" high

U.S. STONEWARE AKRON 9, OHIO 10 H

DEC. 3, 1962 C& EN 101

LABORATORIES, INC. 177 10 93rd AVENUE JAMAICA 33. Ν Υ

S t a n d a r d i z e d

A L U P H A R M C H E M I C A L S JA 5-3477 · P. Ο. Β. 30628

New Orleans 30. La.

Peninsular (_ hem Research, Inc. P.O. Box 3 5 9 7

Gainesville,Florida, U.S.A.

O R G A N O · D ISPERSIONS · A L K O X I D E S

OF ALKALI METALS CUSTOM REACTIONS

R E S E A R C H - D E V E L O P M E N T I^A.IirXJI^A.CTXJR.IN'G

Hiifwicw 3mokpohatm ^ « . ^ - • ^ • ^ « . . . . . . « ^ . ^ i g n r e n s .

ÎLABORATORY SERVICES Applied Research and Development, Test­ing and Consultation · Food, Feed, Drug nd Chemical Analyses, Animal Studies,

Pesticide Screening, Pesticide and Addi­tive Residue Analyses

For price schedule and specific work proposals, write !

W A R F P . O . B o x 2 2 1 7 · M a d i s o n 5, W i s c o n s i n

Fully Automatic OILATOMETER

• electronic recording · 200 to +2200 C • continuously variable magnification to 2000X and over • inductive receiver oscillates at 5000 cycles • automatic compensation for expansion cf specimen holder

FOR PARTICULARS OR DEMONSTRATION. CONTACT

W I L L I A M J . H A C K E R & C O . , I N C . B O X 6 4 6 / W C A L D W E L L . Ν J C A 6 - 8 4 5 0

VACUUM FILTER

Page 13: Synthesis of ORGANOPHOSPHORUS COMPOUNDS

DIRECTORY SECTION This section includes: CHEM­ICALS EXCHANGE—Chemicals, Resins, Gums, Oils, Waxes, Pig­ments, etc.; EQUIPMENT MART —New and Used Equipment, Instru­ments; Facilities for Plant and Lab­oratory; TECHNICAL SERVICES —Consultants, Professional Services.

EQUIPMENT MART

• Teflon® construction • Standard taper joints • Supplied for 6mm or 10mm

shafts and glass tubing • Vacuum tight to 10~6mm Hg. • Inner and Outer " 0 " ring seals • Operating temperatures as

high as 200° C. • No contamination to reactants • No breakage of parts or freezing • Will not wear loose

PRICE, Gland Only 10/30 6 mm. only S13.25ea. 24/40 6, 8, or 10 mm 13.25 ea. 29/42 6, 8, or 10 mm 1675ea. 34/45 6, 8, or 10 mm 18.35 ea. 45/50 6, 8, or 10 mm 21.75ea.

Direct order shipped immediately airmail prepaid

ARTHUR F. S M I T H COMPANY 311 ALEXANDER ST.» ROCHESTER 4, N.Y.

STAINLESS STEEL TANKS A l l Brand N e w in original fac­tory cartons. Oxygen cylinders.

-—v.. Type 304, 400 P.S.I. H" pipe r ^a i i thread openings at each end ex­

cept A - 4 . J - 1 , 24 X 48* , 80 gal., 10

gauge, 18,000 cu. in. W t . 150 lbs., ship w t . 250 lbs., $QQ 50 F.O.B. Baton Rouge, La., $300 value - - · G - 1 , 12 X 2 4 " , 2100 cu. in. 16 gauge, w t . 19 lbs., $1Q 50 F.O.B. Chicago ' . " 'A, Wi X 1 8 " , 16 gauge, 1,000 cu. i n . , wt . 10 %\k 50 bs., F.O.B. Baton Rouge, L a . , , ΑΖ*~Λ 3-2, 6 X 24* , carbon steel, $111 50 M . 6 lbs., F.O.B. Chicago . m'"' A - 4 , 5 H X 8" , carbon steel, Η opening one end, w t . , $ Q 95 \Vi lbs., F.O.B. Chicago M · A i r Force surplus. Ideal for chemicals, gases, pneumatic use, hydraulic, liquids, etc. Terms 2%, 10 days, net 30. Prompt shipment.

ILLINOIS M F G . * SUPPLY C O . t 1829 S. State St., Chicago 16, I I I . , Dept. C E N

Stainless Steel and Monel Utensils

ORDER DIRECT We ship anywhere

Highest quality, durable, corrosion-resistant uten­sils for lab and plant. Also beakers, batch cans, stock pots, shov­els. Over 3 0 years serving process indus­tries. Writ· for catalog-price list

METALSMITHS 567 Willi St., Orange. N.J.

EQUIPMENT MART

M E L - T E M P ® MEASURES CAPILLARY MELTING POINTS

• room temperature to 400°C • 500°C with special thermometer • rapid heating and cooling • one to three samples MEL-TEMP is an integrated capillary melting point apparatus whose heater is controlled by a variable transformer. Excellent viewing is provided by a built in light and a 6-power lens. The attrac­tive gray hammertone base occupies A" x 5".

only $ 97 5 0 'SSSSJU* 0-400 °C thermometer. f.o.b. Cambridge

Pat'd. in U.S.A. and Canada

SPECIAL THERMOMETER 100-500°C in 1 C°, fits MEL-TEMP, borosilicate

ass, Helium filled, 76 im imm. $7.50 each

Write for Bulletin SON

L A B O R A T O R Y D E V I C E S P.O. BOX 6 8 , C A M B R I D G E 3 9 , M A S S .

1 8 0 0 ° C I N A I R TUBE FURNACE AND CONTROLLER

For high temperature research: • Room température to 1800°C in less

than 2 hours. • Hot zone 10 inches by V/z inches. • Control better than ±0.3°C. • Linear programmed heating and cooling. • Compact, fully portable. Operates from

standard 110V, 20 amp. house lines. for technical data and quotation write:

.MOHAWK SCIENTIFIC CORPORATION

1 85 Southern Hill Circle. Henrietta. N.Y.

NATIONAL pH METER

CHECKS p H IN MINUTES TO ± 1 % ACCURACY

N O B U F F E R S O L U T I O N S

N O G L A S S E L E C T R O D E S

N O C A L I B R A T I N G

Just insert metal electrodes into sample; wait 3 minutes; turn switch to "on" position; READ pH ON CALI­BRATED SCALE.

Write for complete details to:

NATIONAL INSTRUMENT CO., INC. 4123 Fordleigh Road, Baltimore 15, Md.

PREVENTS """SSET™ IN PNEUMATIC CONVEYING SYSTEMS

SAVES MONEY TOO! • Lowest Original Cost of any coupling • One Man, One Wrench, One Minute Installation • Fast takedown and reassembly

SAFE—-Tests prove LOW 0.004 ohms resistance GET ALL THE FACTS . . . write or phone

* MORRIS C O U P L I N G A N D CLAMP C O M P A N Y

PEOPLE Continued from page 89

D E A T H S

Dr. Raymond C. Archibald, research supervisor, Shell Development Co., Emeryville, Calif., Oct. 31.

Arthur R. Cade, retired, Rutherford, N.J., Oct. 16. Joined ACS in 1916; emeritus member.

Dr. Kenneth A. Clendenning of the research staff, Institute of Marine Re­sources and the Scripps Institution of Oceanography at University of Cali­fornia, La Jolla, Oct. 12.

Arthur V. Davis, 95, retired since 1957 as chairman of the board of Aluminum Co. of America, Nov. 17.

Frank J. Dugan, 73, retired since 1956 as chief compounder of Good­year Tire & Rubber Co. plants, Nov. 1 in Hollywood, Fla.

Dr. B. G. Engel, 46, general man­ager, Swiss Federal Institute of Tech­nology, Zurich, Switzerland, Sept. 17.

Ford W. Harris, senior partner, Harris, Kiech, Foster & Kern, Los Angeles, Oct. 27.

Charles E. Ives, 60, research asso­ciate, Eastman Kodak, Rochester, Oct. 16. Joined ACS in 1926.

Dr. Peter D. Johnson, 49, v.p. and technical director, Beryllium Tech­nology Corp., Newbury Park, Calif., Oct. 30.

102 C & E N D E C . 3, 1 9 6 2

A S C O

G L A N D

For Rotary Seals

Thermometer Ports

Entry Tubes _

^ S T A T I C CONDUCTINC

COUPLING,

2 3 4 5 War fe l A v e n u e · Erie, Pennsy l van ia ι

Couplings and Repair Clamps to meet YOUR Size and Service Needs


Top Related