nitrogen group element

nitrogen group element

Introduction
 any of the chemical elements that constitute Group Va of the periodic table (see Figure—>). The group consists of nitrogen (N), phosphorus (P), arsenic (As), antimony (Sb), and bismuth (Bi). The elements share certain general similarities in chemical behaviour, though they are clearly differentiated from one another chemically, and these similarities reflect common features of the electronic structures of their atoms.

      Probably no other group of the elements is more familiar to the layman than this group. Although the five elements together make up less than 0.2 percent by weight of the Earth's crust, they assume an importance far out of proportion to their abundance. This is especially true of the elements nitrogen and phosphorus, which comprise 2.4 and 0.9 percent, respectively, of the total weight of the human body.

      The nitrogen elements have, perhaps, the widest range in physical state of any group in the periodic table. Nitrogen, for example, is a gas that liquefies at about −200° C and freezes around −210° C, whereas bismuth is a solid melting at 271° C and boiling at about 1,560° C. Chemically, too, the range in properties is wide, nitrogen and phosphorus being typical nonmetals; arsenic and antimony, metalloids; and bismuth, a metal. Even in appearance these elements exhibit great variety. Nitrogen is colourless both as a gas and as a liquid. Phosphorus exists in a variety of physical modifications, or allotropic forms, including the familiar white, highly reactive form that must be stored under water to prevent it from igniting in the air; a much less reactive red or violet form; and a black modification that, although least known, appears to be the most stable of all. Arsenic exists mainly as a dull gray metallic solid, but a more reactive yellow, solid form is also known, and there are indications that other forms exist under certain conditions. Antimony is a silver, metallic appearing, but somewhat brittle solid; and bismuth is a silver-white metal with a trace of pink in its lustre.

      Together with carbon, hydrogen, oxygen, and sulfur, the first two members of this group, nitrogen and phosphorus, are the principal chemical elements incorporated into living systems. Nitrogen and phosphorus are readily removed from the soil by plant growth, and therefore they are immensely important components of plant foods. Such designations as “5–10–5” on commercial fertilizers (fertilizer) represent the respective weight percentage composition of the material in terms of nitrogen, phosphoric oxide, and potassium oxide (potassium being the third principal element needed for healthy plant growth). Nitrogen in fertilizers may be in the form of sodium or potassium nitrates, ammonia, ammonium salts, or various organic combinations. Phosphorus is supplied chiefly as inorganic phosphate.

      These same elements, nitrogen and phosphorus, can also be used in ways less helpful to man. The explosives in conventional warfare are heavily dependent on their content of nitrogen compounds, and the deadly nerve gases are composed of organic compounds of phosphorus.

      On the other hand, arsenic, which is notorious for its toxicity, is most useful in agriculture, where its compounds are an aid in controlling harmful insect pests. Antimony and bismuth are used chiefly in metal alloys, because they impart unique and desirable properties to these alloys.

History

Nitrogen
      About four-fifths of the Earth's atmosphere is nitrogen, which was isolated and recognized as a specific substance during early investigations of the air. Carl Wilhelm Scheele (Scheele, Carl Wilhelm), a Swedish druggist, showed in 1772 that air is a mixture of two gases, one of which he called “fire air,” because it supported combustion, and the other “foul air,” because it was left after the “fire air” had been used up. The “fire air” was, of course, oxygen, and the “foul air” nitrogen. At about the same time nitrogen also was recognized by a Scottish botanist, Daniel Rutherford, and by the controversial British clergyman, Joseph Priestley (Priestley, Joseph), who, with Scheele, is given credit for the discovery of oxygen. Later work showed the new gas to be a constituent of nitre, a common name for potassium nitrate (KNO3); and, accordingly, it was named nitrogen by the French chemist Jean-Antoine-Claude Chaptal in 1790. Nitrogen also was considered a chemical element by Antoine-Laurent Lavoisier, whose explanation of the role of oxygen in combustion eventually overthrew the phlogiston theory, an erroneous view of combustion that became popular in the early 18th century. The inability of nitrogen to support life led Lavoisier to name it azote, still the French equivalent of nitrogen.

Phosphorus
      Arabian alchemists of the 12th century may have isolated elemental phosphorus by accident, but the records are unclear. Phosphorus appears to have been discovered in 1669 by Henning Brand (Brand, Hennig), a German merchant whose hobby was alchemy. Brand allowed 50 buckets of urine to stand until they putrified and “bred worms.” He then boiled the urine down to a paste and heated it with sand, thereby distilling elemental phosphorus from the mixture. Brand reported his discovery in a letter to Gottfried Wilhelm Leibniz, and, thereafter, demonstrations of this element and its ability to glow in the dark, or “phosphoresce,” excited public interest. Phosphorus, however, remained a chemical curiosity until about a century later when it proved to be a component of bones. Digestion of bones with nitric or sulfuric acid formed phosphoric acid, from which phosphorus could be distilled by heating with charcoal. In the late 1800's, James Burgess Readman of Edinburgh developed an electric furnace method for producing the element from phosphate rock, which is essentially the method employed today.

Arsenic
      Arsenic was known in the form of certain of its compounds long before it was clearly recognized as a chemical element. In the 4th century BC Aristotle wrote of a substance called sandarachē, now believed to have been the mineral realgar, a sulfide of arsenic. Then, in the 1st century AD, the writers Pliny the Elder and Pedanius Dioscorides both described auripigmentum, a substance now thought to have been the dyestuff orpiment, As2S3. By the 11th century AD three species of “arsenic” were recognized: white (As4O6), yellow (As2S3), and red (As4S4). The element itself possibly was first observed in the 13th century by Albertus Magnus (Albertus Magnus, Saint), who noted the appearance of a metallike substance when arsenicum, another name for As2S3, was heated with soap. It is not certain, however, that this natural scientist and scholar actually observed the free element. The first clearly authentic report of the free substance was made in 1649 by Johann Schroeder, a German pharmacist, who prepared arsenic by heating its oxide with charcoal. Later, Nicolas Lémery, a French physician and chemist, observed the formation of arsenic when heating a mixture of the oxide, soap, and potash. By the 18th century, arsenic was well known as a unique semimetal.

Antimony
      The ancients were familiar with antimony both as a metal and in its sulfide form. Fragments of a Chaldean vase made of antimony have been estimated to date from about 4000 BC. The Old Testament tells of Queen Jezebel using the naturally occurring sulfide of antimony to beautify her eyes. Pliny, during the 1st century AD, wrote of seven different medicinal remedies using stibium or antimony sulfide. Early writings of Dioscorides, dating from about the same time, mention metallic antimony. Records of the 15th century show the use of the substance in alloys for type, bells, and mirrors. In 1615 Andreas Libavius (Libavius, Andreas), a German physician, described the preparation of metallic antimony by the direct reduction of the sulfide with iron; and a later chemistry textbook by Lémery, published in 1675, also describes methods of preparation of the element. In the same century, a book summarizing available knowledge of antimony and its compounds was purportedly written by a Basil Valentine, allegedly a Benedictine monk of the 15th century, whose name appears on chemical writings over a span of two centuries. The name antimony appears to be derived from the Latin antimonium, in a translation of a work by the alchemist Geber, but its real origin is uncertain.

Bismuth
      Bismuth evidently was known in very early times, since it occurs in the native state as well as in compounds. For a long period, however, it was not clearly recognized as a separate metal, being confused with such metals as lead, antimony, and tin. Miners during the Middle Ages apparently believed bismuth to be a stage in the development of silver from baser metals and were dismayed when they uncovered a vein of the metal thinking they had interrupted the process. In the 15th-century writings of Basil Valentine this element is referred to as wismut. The name may have been derived from the German words wis mat, meaning white mass. In any case it was latinized to bisemutum by the mineralogist Georgius Agricola, who recognized its distinctive qualities and described how to obtain it from its ores. Bismuth was accepted as a specific metal by the middle of the 18th century, and works on its chemistry were published in 1739 by the German chemist Johann Heinrich Pott and in 1753 by the Frenchman Claude-François Geoffroy.

Comparative chemistry

Electronic configurations
Similarities in orbital arrangement
      In the periodic (periodic law) table, each of the nitrogen group elements occupies the fifth position among the main group elements of its period, a position designated Va. In terms of the electronic configuration of its atoms, each nitrogen group element possesses an outermost shell of five electrons. In each case, these occupy an outer s orbital completely (with two electrons) and contribute one electron to each of the three outer p orbitals (orbital) (the orbitals being electron regions within the atom and the letter designations, s, p, d, and f, being used to designate different classes of orbital). The arrangement of outer electrons in the atoms of the nitrogen elements thus provides three half-filled outer orbitals that, by interaction with half-filled orbitals of the atoms of other elements, can form three covalent bonds. The other atoms may attract the shared electrons either more or less strongly than do the nitrogen group atoms; therefore the latter may acquire either positive or negative charges and exist in oxidation states (oxidation number) of +3 or −3 in their compounds. In this respect, the nitrogen elements are alike.

      Another similarity among the nitrogen elements is the existence of an unshared, or lone, pair of electrons, which remains after the three covalent bonds (covalent bond), or their equivalent, have been formed. This lone pair permits the molecule to act as an electron pair donor in the formation of molecular addition compounds and complexes. The availability of the lone pair depends upon various factors, such as the relative size of the atom, its partial charge in the molecule, the spatial characteristics of other groups in the molecule, and the as-yet poorly understood phenomenon called the “inert pair effect.” This effect consists of a tendency for the paired s electrons in the outermost shell of the heavier atoms of a major group to remain chemically unreactive. Because of it, the electron pair-donating ability of the nitrogen group elements is not uniform throughout the group; it is probably greatest with nitrogen, less with the intermediate elements, and nonexistent with bismuth.

Variations in bonding capacity
      Significant differences in electronic configurations also occur among the elements of the nitrogen group with respect both to the underlying shell and to the outer d orbitals. Since the latter first appear with the third period of the table, they are present in all elements of the group but nitrogen. The possibility of utilizing these outer d orbitals for bonding thus exists for phosphorus, arsenic, antimony, and bismuth, but not for nitrogen.

      There are three principal ways in which the outer d orbitals can be used to increase the number of bonds or expand the valence octet. One is by providing a space to which one of the s electrons can be promoted. This creates two additional half-filled orbitals (one d and one s orbital), and it therefore generates the capacity to form two additional covalent bonds. This is exemplified by the production of phosphorus pentafluoride, PF5, by further fluorination of the trifluoride, PF3. Such promotion appears to be greatly assisted by the increase in outer d-orbital stability that results from the withdrawal of part of the screening electron and the attendant increase of the effective nuclear charge of the central atom. In PF5, for example, the fluorine atoms, being much more electronegative than the phosphorus atom, draw away a portion of the phosphorus electrons, leaving the outer d orbitals more exposed to the phosphorus nucleus and therefore more stable.

      A second way in which the outer d orbitals can become involved in the bonding is by their becoming sufficiently stable to attract a lone pair of electrons from a donor. For example, PF5 can serve as an electron pair acceptor through an outer d orbital to coordinate a fluoride ion donor and form the complex ion PF6.

      A third way of involving d orbitals in bonding is for them to become partially occupied in accommodating lone-pair electrons from another atom, which is already attached by a single bond, thereby strengthening the bond. The phosphorus oxyhalides, of general formula POX3, appear to be examples of this; their phosphorus–oxygen bonds are observed to be shorter and stronger than expected for ordinary single bonds.

The +5 oxidation state
      It thus is possible for an atom of phosphorus, arsenic, antimony, or bismuth to expand its valence octet to form five covalent bonds and one additional coordinate covalent bond. This is not possible for nitrogen, which exhibits a maximum coordination number of four: three single covalent bonds and a coordinate covalent bond with nitrogen acting as donor (through its lone pair). Nevertheless, the +5 oxidation state is formally applicable to nitrogen, so that all five elements can be found in this state. When compounds in the +5 oxidation state are studied, however, it is observed that their properties do not exhibit a uniform trend within the group. Rather, a certain degree of alternation is observed, the +5 states of nitrogen, arsenic, and bismuth appearing less stable and more strongly oxidizing than the corresponding states of phosphorus and antimony. In part this alternation may find explanation in the electronic differences among the atoms with respect to their underlying shells. The number of electrons in the shell just below the outermost level, is two for nitrogen, eight for phosphorus, and 18 for arsenic, antimony, and bismuth.

      Increasing the nuclear charge by 18 from phosphorus to arsenic may be accompanied by incomplete shielding of this extra charge by the ten 3d electrons also added. This would imply smaller size and a greater electronegativity for arsenic than for phosphorus and thus a greater similarity between the phosphorus and antimony atoms. This subject, however, is still controversial, and the widely used scale of electronegativities devised by Linus Pauling fails to make this distinction.

      An interesting anomaly is presented by the fact that nitrogen as a free element is in the form of gaseous diatomic molecules, while the elements immediately preceding it in its period of the table are solids, as are the other elements in its group. In surveying the elements of the second period, the most obvious difference in atomic structure found on reaching nitrogen is the appearance for the first time in compounds of the element of a lone pair of electrons not used in bonding with other atoms. Calculations suggest that the presence of this lone pair of electrons is associated with a considerable weakening of nitrogen to nitrogen single bonds in compounds where these bonds occur. In the diatomic nitrogen molecule, however, the bonding is of a different variety—triple bonds being found between the atoms. It is thought that the triple bond is unaffected (unweakened) by the lone pairs of electrons on the nitrogen atoms, and this is assumed to be the reason why nitrogen “prefers” to exist as triply bonded gaseous diatomic molecules rather than as a condensed singly bonded solid polymer.

      The same effect might be expected to be operable with the other elements of the nitrogen group, all of which also contain lone electron pairs in their outermost shells. Further calculations disclose, however, that the bond-weakening effect of the lone pair is far less pronounced with these elements than it is with nitrogen. As a result, with these elements, single bonds are favoured over multiple bonds, and the diatomic state of the molecules is not the preferred form.

Relative electronegativities
      It might also be expected that the weakening effect of the lone pair would be observed in compounds of the nitrogen group elements. The picture is more complicated here because the bonds under discussion are formed between different types of atoms. Since different elements differ in electronegativity, bonds between the atoms of different elements are inevitably polar. For purposes of discussion it can be assumed that polar bonds consist of blends of nonpolar covalent bonds and completely polar, ionic bonds. It can then be shown that a relatively small amount of ionic character will contribute a disproportionate share to the overall bond strength. Since the weakening effect of the lone pair is felt only on the covalent portion of the polar bond, rather than on the ionic portion, the less polar bonds will exhibit the greater lone-pair weakening effects.

Comparison of nitrogen group elements
      These considerations become important in comparing the chemical behaviour of the nitrogen group elements. The electronegativity of nitrogen itself, although lower than that of oxygen, is substantially higher than that of any of the other elements of this group. Bonds between nitrogen and oxygen, therefore, will be considerably less polar than those between oxygen and phosphorus, or oxygen and arsenic, antimony, or bismuth. Consequently, for this reason alone, the covalent contribution to the nitrogen–oxygen bond energy will be relatively more important than is the case with the bonds between oxygen and the heavier elements of the group. Thus, single-bond weakening by the lone pair—and a corresponding tendency toward bond multiplicity—is likely to be much greater with oxides of nitrogen than with oxides of the heavier nitrogen group elements.

       Some properties of the nitrogen group elementsFor a list of some of the chief properties of the nitrogen group elements, see Table (Some properties of the nitrogen group elements).

Individual nitrogen group elements

Nitrogen
Occurrence and distribution
      The atmosphere of the Earth consists of 78.03 weight percent of nitrogen; this is the principal source of nitrogen for commerce and industry. The atmosphere also contains varying small amounts of ammonia and ammonium salts, as well as nitrogen oxides and nitric acid (the latter substances being formed in electrical storms and in the internal combustion engine). Nitrogen occurs also in mineral deposits of nitre or saltpetre (potassium nitrate, KNO3) and Chile saltpetre (sodium nitrate, NaNO3), but these deposits exist in quantities that are wholly inadequate for man's needs. Another material rich in nitrogen is guano, found in bat caves and in dry places frequented by birds. Nitrogen constitutes on the average about 16 percent by weight of the complex organic compounds known as proteins, present in all living organisms. The natural abundance of nitrogen in the earth's crust is 0.3 parts per 1,000. The cosmic abundance—the estimated total abundance in the universe—is between three and seven atoms per atom of silicon, which is taken as the standard.

Commercial production and uses
      Commercial production of nitrogen is largely by fractional distillation of liquefied air. The boiling temperature of nitrogen is −195.8° C, about 13° below that of oxygen, which is therefore left behind. Nitrogen can also be produced on a large scale by burning carbon or hydrocarbons in air and separating (separation and purification) the resulting carbon dioxide and water from the residual nitrogen. Various laboratory reactions that yield nitrogen include heating ammonium nitrite (NH4NO2) solutions, oxidation of ammonia by bromine water, and oxidation of ammonia by hot cupric oxide.

      Elemental nitrogen can be used as an inert atmosphere for reactions requiring the exclusion of oxygen. In the liquid state, nitrogen has valuable cryogenic applications; except for the gases hydrogen, methane, carbon monoxide, fluorine, and oxygen, practically all chemical substances have negligible vapour pressures at the boiling point of nitrogen and exist, therefore, as crystalline solids at that temperature.

      In the chemical industry, nitrogen is used as a preventive of oxidation or other deterioration of a product, as an inert diluent of a reactive gas, as a carrier to remove heat or chemicals and as an inhibitor of fire or explosions. In the food industry nitrogen gas is employed to prevent spoilage through oxidation, mold, or insects, and liquid nitrogen is used for freeze drying and for refrigeration systems. In the electrical industry nitrogen is used to prevent oxidation and other chemical reactions, to pressurize cable jackets, and to shield motors. Nitrogen finds application in the metals industry in welding, soldering, and brazing, where it helps prevent oxidation, carburization, and decarburization. As a nonreactive gas, nitrogen is employed to make foamed—or expanded—rubber, plastics, and elastomers, to serve as a propellant gas for aerosol cans, and to pressurize liquid propellants for reaction jets. In medicine rapid freezing with liquid nitrogen (nitrogen fixation) may be used to preserve blood, bone marrow, tissue, bacteria, and semen.

      Although the other applications are important, by far the greatest bulk of elemental nitrogen is consumed in the manufacture of nitrogen compounds. The triple bond between atoms in the nitrogen molecules is so strong (226 kilocalories per mole; more than twice that of molecular hydrogen) that it is difficult to cause molecular nitrogen to enter into other combinations. Most living organisms cannot utilize nitrogen directly and must have access to its compounds. Therefore the fixation of nitrogen (the incorporation of elemental nitrogen into compounds) is vitally important. In nature, two principal processes of nitrogen fixation are known. One is the action of electrical (lightning) energy on the atmosphere, which dissociates nitrogen and oxygen molecules, allowing the free atoms to form nitric oxide, NO, and nitrogen dioxide, NO2. Nitrogen dioxide then reacts with water as follows:

      The nitric acid, HNO3, dissolves and comes to Earth with rain as a very dilute solution. In time it becomes part of the combined nitrogen of the soil. The other principal process of natural nitrogen fixation is that of certain plants and vegetables called legumes. Through a cooperative action with bacteria, legumes are able to convert atmospheric nitrogen directly into nitrogen compounds. Certain bacteria alone, such as Azotobacter chroococcum and Clostridium pasteurianum, are also capable of fixing nitrogen.

      The chief commercial method of fixing nitrogen is the Haber (Haber-Bosch process) process for synthesizing ammonia. This process was developed during World War I to lessen the dependence of Germany on Chilean nitrate. It involves the direct synthesis of ammonia from its elements.

Properties and reaction
      Nitrogen (see Table) is a colourless, odourless gas, which condenses at −195.8° C to a colourless, mobile liquid. The element exists as N2 molecules, represented as :N:::N:, for which the bond energy of 226 kilocalories per mole is exceeded only by that of carbon monoxide, 256 kilocalories per mole. Because of this high bond energy the activation energy for reaction of molecular nitrogen is usually very high, causing nitrogen to be relatively inert to most reagents under ordinary conditions. Furthermore, the high stability of the nitrogen molecule contributes significantly to the thermodynamic instability of many nitrogen compounds, in which the bonds, although reasonably strong, are far less so than those in molecular nitrogen. For these reasons, elemental nitrogen appears to conceal quite effectively the truly reactive nature of its individual atoms.

      A relatively recent and unexpected discovery is that nitrogen molecules are able to serve as ligands (ligand) in complex coordination compounds. The observation that certain solutions of ruthenium complexes can absorb atmospheric nitrogen has led to hope that one day a simpler and better method of nitrogen fixation may be found.

      An active form of nitrogen, presumably containing free nitrogen atoms, can be created by passage of nitrogen gas at low pressure through a high-tension electrical discharge. The product glows with a yellow light and is much more reactive than ordinary molecular nitrogen, combining with atomic hydrogen and with sulfur, phosphorus, and various metals, and capable of decomposing nitric oxide, NO, to N2 and O2.

      A nitrogen atom has the electronic structure represented by 1s22s22p3. The five outer shell electrons screen the nuclear charge quite poorly, with the result that the effective nuclear charge felt at the covalent radius distance is relatively high. Thus nitrogen atoms are relatively small in size and high in electronegativity, being intermediate between carbon and oxygen in both of these properties. The electronic configuration (molecule) includes three half-filled outer orbitals, which give the atom the capacity to form three covalent bonds. The nitrogen atom should therefore be a very reactive species, combining with most other elements to form stable binary compounds, especially when the other element is sufficiently different in electronegativity to impart substantial polarity to the bonds. When the other element is lower in electronegativity than nitrogen, the polarity gives partial negative charge to the nitrogen atom, making its lone-pair electrons available for coordination. When the other element is more electronegative, however, the resulting partial positive charge on nitrogen greatly limits the donor properties of the molecule. When the bond polarity is low (owing to the electronegativity of the other element being similar to that of nitrogen), multiple bonding is greatly favoured over single bonding. If disparity of atomic size prevents such multiple bonding, then the single bond that forms is likely to be relatively weak, and the compound is likely to be unstable with respect to the free elements. All of these bonding characteristics of nitrogen are observable in its general chemistry.

      Often the percentage of nitrogen in gas mixtures can be determined by measuring the volume after all other components have been absorbed by chemical reagents. Decomposition of nitrates by sulfuric acid in the presence of mercury liberates nitric oxide, which can be measured as a gas. Nitrogen is released from organic compounds when they are burned over copper oxide, and the free nitrogen can be measured as a gas after other combustion products have been absorbed. The well-known Kjeldahl method for determining the nitrogen content of organic compounds involves digestion of the compound with concentrated sulfuric acid (optionally containing mercury, or its oxide, and various salts, depending on the nature of the nitrogen compound). In this way, the nitrogen present is converted to ammonium sulfate. Addition of an excess of sodium hydroxide releases free ammonia, which is collected in standard acid; the amount of residual acid, which has not reacted with ammonia, is then determined by titration.

Biological and physiological significance
      As might be expected in view of the importance of the presence of nitrogen (nitrogen narcosis) in living matter, most—if not all—organic nitrogen compounds are physiologically active.

      Nitrogen itself, being inert, is innocuous (poison) except when breathed under pressure, in which case it dissolves in the blood and other body fluids in higher than normal concentration. This in itself produces a narcotic effect, but if the pressure is reduced too rapidly, the excess nitrogen evolves as bubbles of gas in various locations in the body. These can cause muscle and joint pain, fainting, partial paralysis, and even death. These symptoms are referred to as “the bends (decompression sickness).” Divers, and others forced to breathe air under pressure, must therefore be extremely careful that the pressure is reduced to normal very slowly following exposure. This enables the excess nitrogen to be released harmlessly through the lungs without forming bubbles. A better alternative is to substitute mixtures of oxygen and helium for air. Helium is much less soluble in body fluids, and the dangers are thus diminished.

Isotopes (isotope) of nitrogen
      Nitrogen exists as two stable isotopes, 14N (abundance 99.63 percent) and 15N (abundance 0.37 percent). These can be separated by chemical exchange or by thermal diffusion. Artificial radioactive isotopes have masses of 12, 13, 16, and 17. The most stable has a half-life of only about 10 minutes.

Occurrence and distribution
      Phosphorus is a very widely distributed element—12th most abundant in the Earth's crust, to which it contributes about 0.10 weight percent. Its cosmic abundance is estimated to be about one atom per 100 atoms of silicon, the standard. Its high chemical reactivity assures that it does not occur in the free state. The principal combined forms in nature are the phosphate salts. Nearly 190 different minerals have been found to contain phosphorus (phosphorite), but, of these, the principal source of phosphorus is the apatite series in which calcium ions exist along with phosphate ions and variable amounts of fluoride, chloride, or hydroxide ions, according to the formula [Ca10(PO4)6(F, Cl, or OH)2]. Commonly such metal atoms as magnesium, manganese, strontium, and lead substitute for calcium in the mineral; and silicate, sulfate, vanadate, and similar anions substitute for phosphate ions. Very large sedimentary deposits of fluoroapatite are found in many parts of the Earth. The phosphate of bone and tooth enamel is hydroxyapatite. (The principle of lessening tooth decay by fluoridation depends upon the conversion of hydroxyapatite to the harder, more decay-resistant, fluoroapatite.)

      Estimates of the total phosphate rock in the Earth's crust average about 50,000,000,000 tons, of which North Africa contains two-thirds, and Russia and the United States most of the remaining third. This estimate includes only ore that is sufficiently rich in phosphate for conversion to useful products by present methods. Vast quantities of material lower in phosphorus content also exist.

Commercial production and uses
      Two principal techniques for converting phosphate rock to usable materials are practiced. One involves acidulation of the crushed rock—with either sulfuric or phosphoric acids—to form crude calcium hydrogen phosphates that, being water-soluble, are valuable additions to fertilizer. The other method is the reduction of the phosphate with carbon in an electric furnace to give elemental phosphorus. The latter reaction is extremely complex, and its precise details depend upon the composition of the mineral phosphate. A charge of sand, coke, and phosphate rock is melted at about 1,500° C in an electric furnace. The calcium and impurities are left in the form of a complex fluorosilicate slag, and elemental phosphorus vapour, at about 300° C, distills out and is collected, condensed, and stored underwater as the white allotropic form of the element. More than half a million tons of phosphorus are made annually in the United States in this way. Most of the output is burned to phosphoric anhydride and subsequently treated with water to form phosphoric acid, H3PO4.

      Only about 5 percent of the 2,000,000 tons of phosphorus consumed per year in the United States is used in the elemental form. Pyrotechnic applications of the element include tracers, incendiaries, fireworks, and matches. Some is used as an alloying agent, some to kill rodents, and the rest is employed in chemical synthesis. A large amount is converted to sulfides used in matches and in the manufacture of insecticides and oil additives. Most of the remainder is converted to halides or oxides for subsequent use in synthesizing organic phosphorus compounds.

Properties and reactions
      The electronic configuration of the phosphorus atom can be represented by 1s22s22p63s23p3. The outer shell arrangement therefore resembles that of nitrogen, with three half-filled orbitals each capable of forming a single covalent bond and an additional lone-pair of electrons. Depending on the electronegativity of the elements with which it combines, phosphorus can therefore exhibit oxidation states of +3 or −3, just as does nitrogen. The principal differences between nitrogen and phosphorus are that the latter is of considerably lower electronegativity and has larger atoms, with outer d orbitals available. For these reasons, the similarities between nitrogen and phosphorus chemistry are largely formal ones, tending to conceal the actual, wide differences. The outer d orbitals in phosphorus permit an expansion of the octet, which leads to the +5 state, with five actual covalent bonds being formed in compounds, a condition impossible for nitrogen to achieve.

      The first striking difference in chemistry of the two elements is that elemental phosphorus exists under ordinary conditions in any of several modifications, or allotropic (allotropy) forms, all of which are solid. Phosphorus molecules of formula P2, structurally analogous to N2 molecules and evidently also triply bonded, exist only at very high temperatures. These P2 molecules do not persist at lower temperatures—below about 800°—because of the fact that three single bonds in phosphorus, in contrast to the situation with nitrogen, are energetically favoured over one triple bond. On cooling, the triply bonded P2 molecules condense to form tetrahedral P4 molecules, in which each atom is joined to three others by single bonds. These molecules further condense to form either hexagonal- or cubic-structured molecular solids, both called “white phosphorus.” Because of the relatively weak intermolecular attractions (van der Waals forces) between the separate P4 molecules, the solid melts easily at 44.1° C and boils at about 280° C. Formation of tetrahedra requires bond angles of 60° instead of the preferred 90°–109° angles, so that white phosphorus is a relatively unstable, or metastable, form. It changes spontaneously, but slowly, at temperatures around 200° or higher, to a polymeric form called “red phosphorus.” This substance is amorphous when formed at lower temperatures, but it can become crystalline, with a melting point of about 590° C. At higher temperatures and pressures, or with the aid of a catalyst, at ordinary pressures and a temperature of about 200° C, phosphorus is converted to a black crystalline form, which somewhat resembles graphite. This may prove to be the most stable form of phosphorus, despite the relative difficulty in its preparation. In both the red and the black forms, each phosphorus atom forms three single bonds, which are spread apart sufficiently to be relatively strain free.

      Consistent with the metastable condition of the white modification, and the crowding of its covalent bonds, this form is far more reactive chemically than the others. It is highly toxic, reacts vigorously with most reagents, and inflames in air at only 35° C, so that it must be stored under water or other inert liquid. White phosphorus dissolves readily in solvents such as carbon disulfide, in which it maintains the composition P4. In contrast, red phosphorus is insoluble and relatively inert, although large quantities of the usual commercial form can ignite spontaneously in air and react with water to form phosphine and phosphorus oxyacids. Black phosphorus is more inert and is capable of conducting electricity. Both these polymeric forms are insoluble and are very much less volatile than white phosphorus.

      Elemental phosphorus can be detected by its phosphorescence. It can also be converted to phosphine with boiling sodium hydroxide solution or with zinc and sulfuric acid; the phosphine is identified by means of test paper containing either silver nitrate or mercuric chloride, both of which are reduced to the free metal by phosphine, thereby darkening the paper. Phosphorus vapour also readily darkens silver nitrate test paper. Phosphorus is determined quantitatively by oxidation to phosphate, followed by any of several standard procedures. Phosphate, for example, may be precipitated as the magnesium ammonium salt, MgNH4PO4, converted by ignition to magnesium pyrophosphate, Mg2P2O7, and weighed. Alternatively, phosphate may be precipitated as ammonium phosphomolybdate; this can be weighed as such, converted to magnesium pyrophosphate and weighed, or titrated directly with sodium hydroxide solution.

Biological and physiological significance
      Phosphorus is an important constituent of bones and teeth, and it is essential to the growth of living organisms. In organisms the element usually appears as phosphate. In its other forms phosphorus is likely to prove very toxic. White phosphorus attacks the skin and, when ingested, causes a necrosis of the jawbone, called “phossy jaw.” Certain organic esters of phosphoric acid, used as lubricating-oil additives, have been found to cause permanent paralysis when accidentally ingested. Phosphine is extremely toxic, as are its organic derivatives. Some of the most toxic substances known to man, collectively termed nerve gas, are organic derivatives of phosphorus.

Isotopes of phosphorus
      The only naturally occurring isotope of phosphorus is that of mass 31. The other isotopes from mass 29 to mass 34 have been synthesized by appropriate nuclear reactions. All of these are radioactive with relatively short half-lives. The isotope of mass 32 has a half-life of about 14 days and has proven extremely useful in tracer studies involving the absorption and movement of phosphorus in living organisms.

Occurrence and distribution
      The abundance of arsenic in the Earth's crust is about five grams per ton; the cosmic abundance is estimated as about four atoms per million atoms of silicon. The element is widely distributed. A small amount exists in the native state, in 90–98 percent purity. Most, however, is combined in more than 150 different minerals, as sulfides, arsenides, sulfoarsenides, and arsenites. Mispickel, or arsenopyrite, FeAsS, is among the most common of arsenic-bearing minerals; others are realgar, As4S4; orpiment, As2S3; loellingite, FeAs2; and enargite, Cu3AsS4. Most commercial arsenic is recovered as a by-product of the smelting of copper, lead, cobalt, and gold ores.

Commercial production and uses
      Metallic arsenic forms when arsenopyrite is heated at 650°–700° C in the absence of air. The arsenic in arsenopyrite and the arsenic impurities in other metal ores unite readily with oxygen when heated in air, forming the easily sublimed oxide, As4O6, also known as “white arsenic.” The vapour of the oxide is collected and condensed in a series of brick chambers and later purified by resublimation. Most arsenic is prepared by carbon reduction of the arsenious oxide dust thus collected.

      World consumption of metallic arsenic is relatively small, only a few hundred tons per year. Most of what is consumed comes from Sweden. It is used in metallurgical (metallurgy) applications because of its metalloid properties. About one percent arsenic content is desirable in the manufacture of lead (lead processing) shot, for example, because it improves the roundness of the molten drops. Bearing alloys based on lead are improved in both thermal and mechanical properties when they contain about 3 percent arsenic. A small amount of arsenic in lead alloys hardens them for use in batteries and cable sheathing. Small concentrations of arsenic improve the corrosion resistance and thermal properties of copper and brass. Very highly purified arsenic finds applications in semiconductor technology, where it is used with silicon and germanium, as well as in the form of gallium arsenide, GaAs, for diodes, lasers, and transistors.

      In contrast to the small use of metallic arsenic, tens of thousands of tons of the element are consumed annually in the form of its compounds. These are used primarily in agriculture.

Properties and reactions
      In its most stable elemental state, arsenic is a steel-gray, brittle solid with low thermal and electrical conductivity. Other forms have been reported but are not well characterized, including especially a yellow, metastable form, which may consist of As4 molecules analogous to white phosphorus, P4. Arsenic sublimes at 613° C, and in the vapour it exists as As4 molecules, which do not begin to dissociate until about 800° C; dissociation to As2 molecules becomes complete at about 1,700° C.

      The electronic structure of the arsenic atom, 1s22s22p63s23p63d104s24p3, resembles those of nitrogen and phosphorus in that there are five electrons in the outermost shell, but it differs from them in having 18 electrons in the penultimate shell instead of two or eight. The addition of ten positive charges to the nucleus during the filling of the five 3d orbitals frequently causes a general contraction of the electronic cloud and a concomitant increase in electronegativity of the elements. In other groups of the periodic table this is clearly shown. Thus, it seems generally accepted that zinc is more electronegative than magnesium and, similarly, that gallium is more electronegative than aluminum. The difference diminishes, however, in the next groups, and many authorities do not agree that germanium is more electronegative than silicon, although an abundance of chemical evidence appears to indicate that this is so. The similar transition from penultimate 8-shell to 18-shell element in passing from phosphorus to arsenic might also produce an increase in the electronegativity of arsenic over phosphorus, but this remains controversial.

      The outer-shell similarity of the two elements suggests that arsenic, like phosphorus, can form three covalent bonds per atom, with an additional lone pair of electrons left unbonded. The oxidation state of arsenic should, therefore, be either +3 or −3 depending on the relative electronegativity values of arsenic and the elements with which it is combined. The possibility should also exist of utilizing the outer d orbitals to expand the octet, thereby allowing arsenic to form five bonds. This possibility is realized only in compounds with fluorine. The availability of the lone pair for complex formation (through electron donation) appears much less in the arsenic atom than in phosphorus and nitrogen, as evidenced by the chemistry of the element.

      Arsenic itself is stable in dry air, but in moist air it tends to become coated with a black oxide. Sublimed arsenic vapour readily burns in air to form arsenious oxide. The free element is essentially unaffected by water, bases, or nonoxidizing acids, but it can be oxidized by nitric acid to the +5 state. Halogens attack arsenic, as does sulfur, and the element will combine directly with many metals forming arsenides.

      Qualitatively, arsenic may be detected by precipitation as the yellow arsenious sulfide from hydrochloric acid of 25 percent or greater concentration. Trace amounts of arsenic are usually determined by conversion to arsine. The latter can be detected by the so-called Marsh test, in which arsine is thermally decomposed, forming a black arsenic mirror inside a narrow tube, or by the Gutzeit method, in which a test paper impregnated with mercuric chloride darkens when exposed to arsine because of the formation of free mercury.

Biological and physiological significance
      The toxicity of arsenic (arsenic poisoning) and its compounds varies widely, ranging from the exceedingly poisonous arsine—and its organic derivatives—to elemental arsenic itself, which is relatively inert. Arsenical compounds in general are skin irritants, which easily cause dermatitis. Protection against inhalation of arsenic-containing dusts is recommended, but most poisoning appears to come from ingestion. The maximum tolerable concentration of arsenic in dusts during an eight-hour day is 0.5 milligrams per cubic metre. For arsine, exposure of similar duration requires that the concentration be less than 0.05 parts per million in the air. In addition to the many uses of arsenic compounds as herbicides and pesticides, they have in several instances been employed as pharmacological agents. The first successful antisyphilitic agent (syphilis), for example, was an arsenic compound, “Salvarsan,” or “606,” or 3,3′-diamino-4,4′-dihydroxyarsenobenzene dihydrochloride.

Isotopes of arsenic
      Only one stable isotope of arsenic, that of mass 75, occurs in nature. Among the artificial radioactive isotopes is one of mass 76, which has a half-life of 26.4 hours.

Occurrence and distribution
      Antimony is about one-fifth as abundant as arsenic, contributing on the average about one gram to every ton of the Earth's crust. Its cosmic abundance is estimated as about one atom to every 5,000,000 atoms of silicon. Small deposits of native metal have been found, but most antimony occurs in the form of more than 100 different minerals. The most important of these is stibnite, Sb2S3. Small stibnite deposits are found in Algeria, Bolivia, China, Mexico, Peru, South Africa, and in parts of the Balkan Peninsula. Some economic value also attaches to kermesite (2Sb2S3 · Sb2O3), argentiferous tetrahedrite [(Cu,Fe)12Sb4S13], livingstonite (HgSb4S7), and jamesonite (Pb4FeSb6S14). Small amounts are also recoverable from the production of copper and lead. About half of all the antimony produced is reclaimed from scrap lead alloy from old batteries, to which antimony had been added to provide hardness.

Commercial production and uses
      High-grade or enriched stibnite reacts directly with scrap iron in the molten state, liberating antimony metal. The metal can also be obtained by conversion of stibnite to the oxide, followed by reduction with carbon. Sodium sulfide solutions are effective leaching agents for the concentration of stibnite from ores. Electrolysis of these solutions produces antimony. After further purification of the crude antimony, the metal, called regulus, is cast into cakes.

      About half of this antimony is used metallurgically, principally in alloys. It improves the hardness and corrosion resistance of lead. Most of the metal is used for lead alloys, largely for storage batteries, but also for chemical equipment such as tanks, pipes, and pumps. With tin, antimony forms such alloys as britannia metal and pewter, used for utensils, and Babbitt metal for bearings. Other applications of antimony alloys are in solder, type metal, and other special materials. Highly purified antimony is used in semiconductor technology to prepare the intermetallic compounds indium, aluminum, and gallium antimonide for diodes and infrared detectors.

Properties and reactions
      The most stable form of elemental antimony is a brittle, silvery solid of high metallic lustre. Electrolytic deposition of antimony under certain conditions produces an unstable, amorphous form called “explosive antimony,” because, when bent or scratched, it will change in a mildly explosive manner to the more stable, metallic form. There is also an amorphous black form of antimony that results from sudden quenching of the vapour, and a yellow form produced by low temperature oxidation of stibine, SbH3, with air or chlorine. Metallic antimony is not affected by air or moisture under ordinary conditions, but it can be oxidized easily by oxygen, sulfur, and the halogens, especially when heated.

      The electronic structure of antimony closely resembles that of arsenic. It is represented as 1s22s22p63s23p63d104s24p64d105s25p3, with three half-filled orbitals in the outermost shell. Thus it can form three covalent bonds and exhibit +3 and −3 oxidation states. The electronegativity of antimony, like that of arsenic, remains somewhat controversial. It is generally agreed to be lower than that of arsenic, but whether it is lower also than that of phosphorus is undecided. It can act as an oxidizing agent and reacts with many metals to form antimonides that, in general, resemble nitrides, phosphides, and arsenides but are somewhat more metallic. The promotion of one of the lone-pair electrons to an outer d orbital apparently occurs more easily with antimony than with arsenic, since antimony exhibits the +5 oxidation state in forming both the pentafluoride and the pentachloride.

      Antimony may be separated and weighed for analysis as the sulfide, Sb2S3. Alternatively, the sulfide may be converted to the oxide and, after careful ignition, weighed as Sb4O6. Numerous volumetric methods are also available, including several methods of oxidizing antimony in the +3 oxidation state with potassium permanganate, potassium bromate, or iodine. In the absence of arsenic, small amounts of antimony may be determined by a modified Gutzeit method.

Biological and physiological significance
      Antimony and a number of its compounds are highly toxic (poison). In fact, the use of antimony compounds for medicinal purposes was temporarily outlawed several centuries ago because of the number of fatalities they had caused. A hydrated potassium antimonyl tartrate called “tartar emetic” is currently used in medicine as an expectorant, diaphoretic, and emetic. The maximum tolerable concentration of antimony dust in air is about the same as for arsenic, 0.5 milligrams per cubic metre.

Isotopes of antimony
      Two stable isotopes, nearly equal in abundance, occur in nature. One has mass 121 and the other mass 123. Radioactive isotopes of masses 120, 122, 124, 125, 126, 127, 129, and 132 have been prepared.

Occurrence and distribution
      Bismuth is about as abundant as silver, contributing about 2 × 10−5 weight percent of the Earth's crust. Its cosmic abundance is estimated as about one atom to every 7,000,000 atoms of silicon. It occurs both native and in compounds. In the native state, it is found in veins associated with lead, zinc, tin, and silver ores in Bolivia, Canada, England, and Germany. Its naturally occurring compounds are chiefly the oxide (bismite or bismuth ochre), the sulfide (bismuthinite or bismuth glance), and two carbonates (bismutite and bismutosphaerite). Commercial bismuth, however, is produced largely as a by-product in the smelting and refining of lead, tin, copper, silver, and gold ores. Thus, it comes—for example—from tungsten ores in South Korea, lead ores in Mexico, copper ores in Bolivia, and both lead and copper ores in Japan.

Commercial production and uses
      Bismuth is volatile at high temperature, but it usually remains with the other metals after smelting operations. Electrolytic (electrolysis) refining of copper leaves bismuth behind as one component of the anode sludge. Separation of bismuth from lead by the Betterton–Kroll (Betterton-Kroll process) process involves the formation of high-melting calcium or magnesium bismuthide (Ca3Bi2 or Mg3Bi2), which separates and can be skimmed off as dross. The dross may be chlorinated to remove the magnesium or calcium, and finally the entrained lead. Treatment with sodium hydroxide then produces highly pure bismuth. An alternative separation, the Betts process, involves electrolytic refining of lead bullion (containing bismuth and other impurities) in a solution of lead fluosilicate and free fluosilicic acid, bismuth being recovered from the anode sludge. Separation of bismuth from its oxide or carbonate ores can be effected by leaching with concentrated hydrochloric acid. Dilution then precipitates the oxychloride, BiOCl. This, on heating with lime and charcoal, produces metallic bismuth.

      Metallic bismuth is used principally in alloys (alloy), to many of which it imparts its own special properties of low melting point and expansion on solidification. Bismuth is thus a useful component of type-metal alloys, which make neat, clean castings; and it is an important ingredient of low-melting alloys, called fusible alloys, which have a large variety of applications, especially in fire-detection equipment. A bismuth–manganese alloy has been found effective as a permanent magnet. Small concentrations of bismuth improve the machinability of aluminum, steel, stainless steels, and other alloys and suppress the separation of graphite from malleable cast iron. Thermoelectric devices for refrigeration make use of bismuth telluride, Bi2Te3, and bismuth selenide, Bi2Se3. Liquid bismuth has been used as a fuel carrier and coolant in the generation of nuclear energy.

      The principal chemical application of bismuth is in the form of bismuth phosphomolybdate, which is an effective catalyst for the air oxidation of propylene and ammonia to acrylonitrile. The latter is used to make acrylic fibres, paints, and plastics. Pharmaceutical uses of bismuth have been practiced for centuries. It is effective in indigestion remedies and antisyphilitic drugs. Slightly soluble or insoluble salts are utilized in the treatment of wounds and gastric disorders, and bismuth is sometimes injected in the form of finely divided metal, or as suspensions of its insoluble salts. Substantial quantities of the oxychloride, BiOCl, have been used to impart a pearlescent quality to lipstick, nail polish, and eye shadow.

Properties and reactions
      Bismuth is a rather brittle metal with a somewhat pinkish, silvery metallic lustre. It undergoes a 3.3 percent expansion when it solidifies from the molten state. Its electrical conductivity is very poor, but somewhat better in the liquid state than in the solid. With respect to thermal conductivity, it is the poorest of all metals except mercury. Bismuth is quite resistant to corrosion by air and moisture, but it is oxidized rapidly at its boiling point of 1,560° C. It is oxidized and dissolved by concentrated nitric acid.

      Bismuth atoms have the same electronic structure in their outermost shell as do the other elements of the nitrogen group. They can, therefore, form three single covalent bonds, exhibiting either a +3 or −3 oxidation state. The element has a somewhat lower electronegativity than the others, and its lone pair of electrons is evidently quite inert, causing the +5 state of bismuth to be rare and unstable.

Analytical and physiological chemistry
      Bismuth is usually determined gravimetrically, being precipitated and weighed as the phosphate or the oxychloride, BiOCl. To produce the latter, a suitable amount of hydrochloric acid is added to a nitric acid solution containing the bismuth, and the resulting solution is poured into a large volume of water, causing the oxychloride to precipitate. Volumetric and colorimetric methods of determination are also available.

      Bismuth is relatively nontoxic, the least so of the heavy metals. It is generally not an industrial hazard. Although bismuth and certain of its compounds find considerable therapeutic use, some authorities recommend that other remedies be substituted. Soluble inorganic bismuth compounds are toxic.

Isotopes of bismuth
      Bismuth forms only one stable isotope, that of mass 209. A large number of radioactive isotopes are known, as shown in the table, most of them being very unstable.

R. Thomas Sanderson Ed.

Additional Reading
A concise summary of nitrogen chemistry is contained in William L. Jolly, The Inorganic Chemistry of Nitrogen (1964). More detailed is the comprehensive compendium by C.A. Streuli and P.R. Averell (eds.), The Analytical Chemistry of Nitrogen and Its Compounds, 2 vol. (1970), which includes especially useful tables. A popularized, readable, introductory account is Isaac Asimov, The World of Nitrogen, rev. ed. (1962). Information about nitrogen and its compounds and about the other elements of the nitrogen group and their compounds is found in such standard works on inorganic chemistry as M. Cannon Sneed, J. Lewis Maynard, and Robert C. Brasted (eds.), Comprehensive Inorganic Chemistry, 8 vol. (1953–61); John C. Bailar, Jr., et al. (eds.), Comprehensive Inorganic Chemistry, 5 vol. (1973); and F. Albert Cotton and Geoffrey Wilkinson, Advanced Inorganic Chemistry, 5th ed. (1988); in R.T. Sanderson, Chemical Bonds and Bond Energy, 2nd ed. (1976), a brief book giving new insights into the nature of chemical bonds; and in Eugene G. Rochow, The Metalloids (1966), a concise, readable survey of the elements that border between metallic and nonmetallic. Detailed discussions on fixation are provided in W.J. Broughton (ed.), Nitrogen Fixation (1981– ). More detailed information on phosphorus may be found in D.E.C. Corbridge, The Structural Chemistry of Phosphorus (1974), and Phosphorus: An Outline of Its Chemistry, Biochemistry, and Technology, 4th ed. (1990); and Harold Goldwhite, Introduction to Phosphorus Chemistry (1981).R. Thomas Sanderson Ed.

* * *


Universalium. 2010.

Игры ⚽ Нужно сделать НИР?

Look at other dictionaries:

  • Nitrogen group — Group 15 Period 2 7 N 3 …   Wikipedia

  • carbon group element — ▪ chemical elements Introduction  any of the five chemical elements that make up Group 14 (IVa) of the periodic table namely, carbon (C), silicon (Si), germanium (Ge), tin (Sn), and lead ( …   Universalium

  • boron group element — ▪ chemical elements Introduction  any of the five chemical elements constituting Group 13 (IIIa) of the periodic (periodic law) table (see Figure >). The elements are boron (B), aluminum (Al), gallium (Ga), indium (In), an …   Universalium

  • Some properties of the nitrogen group elements — ▪ Table Some properties of the nitrogen group elements nitrogen phosphorus (white) atomic number 7 15 atomic weight 14.0067 30.9738 colour of element colourless white melting point (°C) −209.86 44.1 boiling point (°C) −195.8 280… …   Universalium

  • zinc group element — ▪ chemistry Introduction  any of the three metals that comprise Group 12 (IIb) of the periodic table of elements namely, zinc (Zn), cadmium (Cd), and mercury (Hg). (See Figure >.) They have properties in common, but they also differ in… …   Universalium

  • Main group element — The periodic table of the chemical elements. The columns represent the groups. Groups 1, 2 and 13 to 18 are the main group elements. In chemistry and atomic physics, main group elements are elements in groups (periodic columns) whose lightest… …   Wikipedia

  • Nitrogen — carbon ← nitrogen → oxygen ↑ N ↓ P …   Wikipedia

  • Element — El e*ment, n. [F. [ e]l[ e]ment, L. elementum.] 1. One of the simplest or essential parts or principles of which anything consists, or upon which the constitution or fundamental powers of anything are based. [1913 Webster] 2. One of the ultimate …   The Collaborative International Dictionary of English

  • nitrogen — Symbol: N Atomic number: 7 Atomic weight: 14.0067 Colourless, gaseous element which belongs to group 15 of the periodic table. Constitutes nitrogen78% of the atmosphere and is an essential part of the ecosystem. Nitrogen for industrial purposes… …   Elements of periodic system

  • nitrogen — /nuy treuh jeuhn/, n. Chem. a colorless, odorless, gaseous element that constitutes about four fifths of the volume of the atmosphere and is present in combined form in animal and vegetable tissues, esp. in proteins: used chiefly in the… …   Universalium

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”