Category Archives: elements

Lead : It’s discovery

 

Lead discovery

Lead is very rarely encountered in a native state but is smelted fairly easily from ores. Lead become known to Egyptians simultaneously with iron and silver and was produced as early as the second millenium B.C. in India and china, In Europe production of lead began somewhat later although in the 6th century. B.C records we find mention of lead which was brought

Lead is a naturally occurring metal but its natural status doesn’t mean it’s healthy. In fact, lead is extremely toxic to humans and affects the liver, kidneys, reproductive system, and nervous system

to the Tyre trade fair. Lead was produced in great amounts during the reign of Hammurabi in Babylon. For a long time lead was confused with tin. Tin was named “plum bum album” and lead–“plumbum nigrum”. Only in the Middle Ages were they recognized as different metals.

Greeks and Phoenicians started many lead mines in Spain which later were taken over by Romans. In ancient Rome lead was widely used: for making crockery, styluses, and pipes for the famous Roman water-main. Lead was also used for manufacturing white lead. The island of Rhodes was the biggest exporter of white lead.

Lead Preparation :

The process of its preparation is still used as follows: lead pieces are immersed into vinegar and the salt thus obtained is boiled with water for a long time.

But red lead was first obtained unexpectedly. When a fire broke out in the Greek port of Piraeus barrels with lead were enveloped in flames. After the fire had been extinguished, red substance was found in the charred barrels–it was red lead.

Although in Russia lead has been known for a long time, up to the 18th century the process of lead production was very primitive. After the invention of firearms lead was used for making bullets and the military importance of lead is still great. But in addition to its “military” uses” lead has many peaceful ones; for instance, typographical types are made of its alloy with antimony. Lead is also used for protection against radiation in experiments.

Greeks named leadmolibdos; its chemical symbol Pb originates from Latin plumbum.

Silver

 

Silver Discovery

Silver is a more active metal than gold .

Although its abundance in the earth’s crust is about fifteen times that of gold, but it occurs much less frequently in a native state. It is not surprising that in antiquity silver was valued higher than gold. In ancient Egypt, for instance, the ratio between the costs of these metals was 2.5:1. Gold was used mainly for coins and jewelry; silver had other uses: for example, for making water vessels.

In the 4th century B.C the army of Alexander the Great conquered Persia and Phoenicia and invaded India. Here the Greek army was struck by an out break of a mysterious gastrointestinal disease and the men demanded to be sent home. Interestingly, the Greek military commanders fell victim to the disease far less frequently than their men, although they shared all the burdens of camp life with the soldiers. More than two thousand years had passed before scientists found an explanation of it. The soldiers drank from tin cups and their superiors from silver ones. It was proved that silver dissolves in water forming a colloid solution that kills pathogenic bacteria. And although the solubility of silver in in water in low, it is quite enough for disinfection.

Silver mines have known from time immemorial. The largest deposits of silver were in Greece, Spain, and Germany. After the discovery of America silver deposits were also found in Peru and Mexico. Lead minerals are often observed as constituents in silver ores.

Old process of Silver Extraction :

An old process of extracting silver from such ores is described as follows:-

Silver ore was ground, washed with water, and dried. Then it was fused together with flux and the alloy thus obtained was heated with charcoal. The resulting alloy of silver and lead was calcinated. On heating in air silver is practically unoxidized whereas lead transforms into oxide almost completely. The melting point of lead oxide is 896oC and that of silver, 960oC. Thus, practically pure silver was obtained.

At present more perfect processes of purifying silver are used.

Silver like gold was used in coins but the cost of silver compared to that of gold was gradually decreasing. In 1874 the cost of one pound of gold was equal to that of 15.5 pounds of silver but after the discovery of silver deposits in Australia this ratio fell to 1:46. In England bimetallism, i.e. the use of gold and silver jointly as a monetary standard, was discontinued in 1816. Later other countries followed this example.

In the last 40 years, silver has had two big bull runs, surrounded by even bigger bear markets. In the late 1970s, silver skyrocketed in price.

Russian words “rubl”’ (rouble) and “kopeika” (kopeck) owe their origin to silver. Rouble came into being in Kievan Russia in the 13th century –a silver bar weighing about 200 grams. It is believed that in the process of manufacturing roubles a long silver bar was cast and then hacked into parts (“rubit” is the Russian for “to hack”). The word “kopeika” appeared somewhat later (in 1534) when coins with an image of a horseman holding a speak (“kop’ e” in Russian) were first minted.

The name “silver” seems to stem from the Assyrian “serpu” or Gothic “silbur” the Latin argentum originates most likely from the Sanscrit arganta, which means “light, white”.

Hydrogen

Hydrogen

 

Discovery of Hydrogen : 

In 1671, Robert Boyle discovered and described the reaction between iron filings and dilute acids, which resulted in the production of hydrogen gas. In 1766-81, Henry Cavendish was the first to recognize that hydrogen gas was a discrete substance, and that it produced water when burned. He named it “flammable air”. In 1783, Antoine Lavoisier gave the element the name hydrogen (from the Greek υδρο- hydro meaning “water” and -γενης genes meaning “creator”) when he and Pierre-Simon Laplace reproduced Cavendish’s finding that water was produced when hydrogen was burned. 
Hydrogen was liquefied for the first time by James Dewar in 1898 by using regenerative cooling and his invention, the vacuum flask. He produced solid hydrogen the next year. Deuterium was discovered in December 1931 by Harold Urey, and tritium was prepared in 1934 by Ernest Rutherford, Mark Oliphant, and Paul Harteck. Heavy water, which consists of deuterium in the place of regular hydrogen, was discovered by Urey’s group in 1932.

The nickel hydrogen battery was used for the first time in 1977 aboard the U.S. Navy’s Navigation technology satellite-2 (NTS-2). It had two caesium atomic clocks on board and helped to show that satellite navigation based on precise timing was possible. In the dark part of its orbit, the Hubble Space Telescope is powered by nickel-hydrogen batteries, which were finally replaced in May 2009, more than 19 years after launch, and 13 years passed their design life.

from NASA (accessed 2 Feb 2015)

 

Isotopes of hydrogen

Hydrogen has three naturally occurring isotopes, denoted 1H, 2H and 3H. Other, highly unstable nuclei (4H to 7H) have been synthesized in the laboratory but are not observed in nature.

  • 1H is the most common hydrogen isotope with an abundance of more than 99.98%. Because the nucleus of this isotope consists of only a single proton, it is given the descriptive, but rarely used formal name of protium.
  • 2H, the other stable hydrogen isotope, is known as deuterium and contains one proton and one neutron in its nucleus. Essentially all deuterium in the universe is thought to have been produced at the time of the Big Bang, and has endured since that time. Deuterium is not radioactive, and does not represent a significant toxicity hazard. Water enriched in molecules that include deuterium instead of normal hydrogen is called heavy water. Deuterium and its compounds are used as a non-radioactive label in chemical experiments and in solvents for 1H-NMR spectroscopy. Heavy water is used as a neutron moderator and coolant for nuclear reactors. Deuterium is also a potential fuel for commercial nuclear fusion.
  • 3H is known as tritium and contains one proton and two neutrons in its nucleus. It is radioactive, decaying into helium-3 through beta decay with a half-life of 12.32 years. It is sufficiently radioactive that it can be used in luminous paint, making it useful in such things as watches where the glass moderates the amount of radiation getting out. Small amounts of tritium occur naturally because of the interaction of cosmic rays with atmospheric gases; tritium has also been released during nuclear weapons tests. It is used in nuclear fusion reactions, as a tracer in isotope geochemistry, and specialized in self-powered lighting devices. Tritium has been used in chemical and biological labeling experiments as a radiolabel.

Hydrogen is the only element that has different names for its isotopes in common use today. During the early study of radioactivity, various heavy radioactive isotopes were given their own names, but these names are no longer used, except for deuterium and tritium.

nuclide symbol Z(p) N(n) isotopic mass (u) half-life decay mode Daughter Isotope representative isotopic composition
1H 1 0 1.00782503207(10) Stable 0.999885(70)
2H – D 1 1 2.0141017778(4) Stable 0.000115(70)
3H – T 1 2 3.0160492777(25) 12.32(2) y β 3He <1 in 1017 atoms

Properties of hydrogen

The difference of mass between isotopes of most elements is only a small fraction of the total mass and so this has very little effect on their properties, this is not the case for hydrogen. Consider chlorine with Z=17, there are 2 stable isotopes 35Cl (75.77%) and 37Cl (24.23%). The increase is therefore 2 in 35 or less than 6%. Deuterium and tritium are about double and triple the mass of protium and show significant physical and chemical differences particularly with regard to those properties related to mass, e.g. rate of diffusion, density, etc.

Some physical properties of the hydrogen isotopes.
isotope MP /K BP /K ΔHdiss /kJmol-1 Interatomic Distance /pm
H2 13.99 20.27 435.99 74.14
D2 18.73 23.67 443.4 74.14
T2 20.62 25.04 446.9 74.14

Differences between H2O and D2O

 

Property H2O D2O
Melting point /K 273.15 276.97
Boiling point /K 373.15 374.5
Temperature of maximum density /K 277 284.2
Maximum density /g cm3 0.99995 1.1053
Relative permittivity (at 298 K) 78.39 78.06
Kw (at 298 K) 1 *1014 2 * 1015
Symmetric stretch, ν1 /cm-1
(gaseous molecule)
3657 2671

Given that the boiling point of D2O is 101.4 °C (compared to 100.0 °C for H2O), evaporation or fractional distillation can be used to increase the concentration of deuterium in a sample of water by the selective removal of the more volatile light water, H2O. Thus bodies of water that have no outlet, such as the Great Salt Lake in Utah, USA and the Dead Sea in the Jordan Rift Valley, which maintain their level solely by evaporation, have significantly higher concentrations of deuterated water than do lakes or seawater with at least one outlet.

 The isotopic ratio for H and D is not fixed and so a range is given
for the standard atomic weight in the IUPAC Periodic Table of isotopes.

Heavy water is 10.6% denser than ordinary water, a difference not immediately obvious since they are otherwise physically and chemically similar. The difference can be observed by freezing a sample and dropping it into normal water, where it sinks.

With respect to taste and smell, rats given a choice between distilled normal water and heavy water avoided the heavy water, based on smell, and it may be that they detected a different taste as well.

The difference in weight increases the strength of water’s hydrogen-oxygen bonds, and this in turn is sufficient to cause differences that are important to some biochemical reactions. The human body naturally contains deuterium equivalent to about five grams of heavy water, which is harmless. When a large fraction of water (> 50%) in higher organisms is replaced by heavy water, the result is cell dysfunction and death.

In normal water, about 1 molecule in 3,200 is HDO (one hydrogen in 6,400 is in the form of D), and heavy water molecules (D2O) only occur in a proportion of about 1 molecule in 41 million (i.e. one in 6,4002). Thus semiheavy water molecules are far more common than “pure” (homoisotopic) heavy water molecules.

Deuterium oxide was initially obtained by the electrolysis of ordinary water over a considerable period of time. This method of production requires a large cascade of stills or electrolysis chambers and consumes large amounts of power, so that chemical methods are generally now preferred. The most important chemical method is the Girdler sulfide process.

In this process, demineralised and deaerated water is trickled through a series of perforated (seive) plates in a tower, while hydrogen sulfide gas (BP -60 °C) flows upward through the perforations. Deuterium migration preferentially takes place from the gas to the liquid water. This “enriched” water from the cold tower (maintained at 32 °C) is fed to the hot tower (at 130 °C) where deuterium transfer takes place from the water to the hydrogen sulfide gas. An appropriate “cascade” setup accomplishes enrichment via the reversible reaction: 

H2O +HDS ⇄ HDO + H2

The equilibrium constant, K for, this reaction in terms of the concentrations, can be written as:

K = ([HDO][H2S]) / ([H2O][HDS]) or alternatively:

K = ([HDO]/[H2O]) / ([HDS]/[H2S])

If H and D were the same chemically, the equilibrium constant for the reaction would be equal to unity. However, what is found is that K is not equal to unity, and furthermore it is temperature dependent: 

at 25 °C,  K = 2.37
at 128 °C, K = 1.84

From the above information, at 32 °C, the equilibrium favours the concentration of deuterium in water. However, at around 130 °C, the equilibrium is now relatively more favorable to the concentration of deuterium in the hydrogen sulfide. In other words, the concentration of HDO in H2O is greater than the concentration of HDS in H2S but the relative concentration of HDS in H2S increases with increasing temperature, making it possible to separate D from H.

In the first stage, the gas is enriched from 0.015% deuterium to 0.07%. The second column enriches this to 0.35%, and the third column achieves an enrichment between 10% and 30% deuterium oxide, D2O. Further enrichment to “reactor-grade” heavy water (> 99% D2O) still requires distillation or electrolysis. The production of a single litre of heavy water requires ~340,000 litre of feed water.

In 1934, Norway built the first commercial heavy water plant with a capacity of 12 tonnes per year. From 1940 and throughout World War II, the plant was under German control and the Allies decided to destroy the plant and its heavy water to inhibit German development of nuclear weapons. In late 1942, a planned raid by British airborne troops failed, both gliders crashing. The raiders were killed in the crash or subsequently executed by the Germans. On the night of 27 February 1943 Operation Gunnerside succeeded. Norwegian commandos and local resistance managed to demolish small, but key parts of the electrolytic cells, dumping the accumulated heavy water down the factory drains. Had the German nuclear program followed similar lines of research as the United States Manhattan Project, the heavy water would not have been crucial to obtaining plutonium from a nuclear reactor, but the Germans did not discover the graphite reactor design used by the allies for this purpose.

ortho- and para-dihydrogen

In dihydrogen, the two electrons in the molecule will be spin paired but there is no similar requirement for the two nuclei; they may be parallel or opposed. There are therefore two nuclear spin isomers possible which are called ortho and para.


In the parahydrogen form the nuclear spins of the two protons are antiparallel and form a singlet (2I+1= 1) with a molecular spin quantum number, I, of 0 (½ – ½). In the orthohydrogen form, the spins are parallel and form a triplet state (2I+1= 3) with a molecular spin quantum number, I, of 1 (½ + ½). At standard temperature and pressure, hydrogen gas contains about 25% of the para form and 75% of the ortho form, also known as the “normal form”.

The para form has slightly lower energy:-

o-H2 ⇄ p-H2; ΔH = -1.5 kJmol-1

but due to the small difference this has little effect at room temperature.

The amount of ortho and para hydrogen varies with temperature:

  • At 20 K, hydrogen contains mainly para (singlet) hydrogen (99.8%) which is the more stable form.
  • At the temperature of liquefaction of air, ~80 K, the ratio of ortho and para hydrogen is 1 : 1.
  • At room temperatures, the ratio of ortho to para hydrogen is 3 : 1.
  • Even at very high temperatures, the ratio of ortho to para hydrogen never exceeds 3 : 1.

It is possible then to get pure para hydrogen by cooling ordinary hydrogen gas to very low temperatures (close to 20 K) but it is not possible to get a sample of hydrogen containing more than 75% of ortho (triplet) hydrogen. The first synthesis of pure parahydrogen was achieved in 1929.
This conversion of ortho- to para-hydrogen liberates some heat which can cause evaporation of hydrogen within storage vessels. Since orthohydrogen molecules make up 75% of “normal” hydrogen at room temperature, this can considerably complicate the performance of storing liquid hydrogen. Without an ortho-para conversion catalyst, (such as hydrous ferric oxide) extra refrigeration equipment is required to remove the heat generated by the natural conversion to para hydrogen.

Production of hydrogen

Laboratory preparations
In the laboratory, H2 can be prepared by the action of a dilute non-oxidizing acid on a reactive metal such as zinc, with a Kipp’s apparatus. 

Zn + 2Haq+ ⇄ Znaq2+ + H2

Aluminium can produce H2 upon treatment with bases:

2Al + 6 H2O + 2 OH ⇄ 2 Al(OH)4 + 3 H2

The electrolysis of water is another simple method of producing hydrogen. A low voltage current is passed through the water, and gaseous dioxygen forms at the anode while gaseous hydrogen forms at the cathode. Typically the cathode is made from platinum or other inert metal when producing hydrogen for storage. If, however, the gas is to be burnt on site, oxygen is desirable to assist the combustion, and so both electrodes would be made from inert metals. (Iron, for instance, would oxidize, and thus decrease the amount of oxygen given off.) The theoretical maximum efficiency (electricity used versus energetic value of hydrogen produced) is in the range 80-94%.

2 H2O(l) ⇄ 2 H2(g) + O2(g) 

In 2007, it was discovered that an alloy of aluminium and gallium in pellet form added to water could be used to generate hydrogen. The process creates alumina, but the expensive gallium, which prevents the formation of an oxide skin on the pellets, can be re-used. This has important potential implications for a hydrogen economy, as hydrogen could be produced on-site without the need of being transported.

Industrial preparation of hydrogen
Steam reforming is a method for producing hydrogen, carbon monoxide or other useful products from hydrocarbon fuels such as natural gas. This is achieved in a processing device called a reformer which reacts steam at high temperature with the fossil fuel.

At high temperatures (700 – 1100 °C) and in the presence of a metal-based catalyst (nickel), steam reacts with methane to yield carbon monoxide and hydrogen.

CH4 + H2O → CO + 3 H2

In order to produce more hydrogen from this mixture, more steam is added and the water gas shift reaction is carried out: 

CO + H2O → CO2 + H2

The mixture of CO and H2 is called “synthesis gas or syngas”. Syngas is used as an intermediate in producing synthetic petroleum for use as a fuel or lubricant via the Fischer-Tropsch process and previously the Mobil methanol to gasoline process.

Enzymatic route from xylose 
In 2013 a low-temperature, 50 °C, atmospheric-pressure, enzyme-driven process to convert xylose into hydrogen with nearly 100% of the theoretical yield was announced. The process employed 13 enzymes, including a novel polyphosphate xylulokinase (XK).

It was noted that: “Approximately 50 million metric tons of dihydrogen are produced annually from nonrenewable natural gas, petroleum, and coal. H2 production from water remains costly. Technologies for generating H2 from less costly biomass, such as microbial fermentation, enzymatic decomposition, gasification, steam reforming, and aqueous phase reforming, all suffer from low product yields.

Compounds of Hydrogen

The chemistry of hydrogen depends mainly on four processes:

  1. donation of the valency electron to form the hydrogen ion, H+
  2. accepting an electron to form the hydride ion H
  3. sharing the electron with a partner atom to form a pair bond (covalent bond) H-H
  4. sharing the electron with an ensemble of atoms to form a metallic bond H.

While H2 is not very reactive under standard conditions, it does form compounds with most elements. Hydrogen can form compounds with elements that are more electronegative, such as halogens (e.g., F, Cl, Br, I), or oxygen; in these compounds hydrogen takes on a partial positive charge. When bonded to fluorine, oxygen, or nitrogen, hydrogen can participate in a form of medium-strength noncovalent bonding called hydrogen bonding, which is critical to the stability of many biological molecules. Hydrogen also forms compounds with less electronegative elements, such as the metals and metalloids, in which it takes on a partial negative charge. These compounds are often known as hydrides.

The term “hydride” suggests that the H atom has acquired a negative or anionic character, denoted H-, and is used when hydrogen forms a compound with a more electropositive element. The existence of the hydride anion, suggested by Gilbert N. Lewis in 1916 for group I and II salt-like hydrides, was demonstrated by Moers in 1920 by the electrolysis of molten lithium hydride (LiH), producing a stoichiometry quantity of hydrogen at the anode.

Although hydrides can be formed with almost all main-group elements, the number and combination of possible compounds varies widely; for example, there are over 100 binary borane hydrides known, but only one binary aluminium hydride. A simple binary indium hydride has not yet been identified, although larger complexes exist.

The position of H in the Periodic Table

In some respects, H does not seem to have a perfect position in the Periodic Table and so many designers have it in more than one position, e.g. in Group 1 or Group 17 and even in Group 14.

Ionization energy of hydrogen

Hydrogen has a single outer electron, like the alkali metals, but they all form positive ions quite readily whereas hydrogen has little tendency to do so. Hydrogen often tends to share its electron with nonmetals rather than losing it to them.

The first ionization energies for H, Li, Na and K are 1312, 520.2, 495.8 and 418.8 kJmol-1. The high IE for H (even bigger than for Xe) can be attributed to the very small size of the atom and the strong attractive force between the proton and electron.

H(g) → H+(g) + e        ΔH = 1312 kJmol-1

Xe(g) → Xe+(g) + e      ΔH = 1170 kJmol-1

The free proton can only be obtained under extreme conditions such as by an electric arc or in a discharge tube and even then only exists for about half a second. H+ can be found in solvated form where the solvation energy provides the energy needed to overcome the very high ionization energy. Examples are in ammonia, alcohol or water with species like NH4+, ROH2+ and H3O4+ being formed.

Electron affinity of hydrogen

Hydrogen, like the halogens, exists as diatomic molecules and H atoms have electron configurations with one electron short of a filled outer shell hence the idea of placing H in Group 17. However unlike the halogens with large EA values, the EA for hydrogen is quite small. The formation of H is much less favourable than the formation of a chloride ion, as seen from the thermodynamic profiles below and it is rare whereas halide ions are common and stable. In addition H has a lower electronegativity value than any of the halogens.


Much more energy is required as well to break the H-H bond compared to the Cl-Cl bond where the steps for comparison are:

½H2 (g) → H. (g)                   ΔH = 218 kJmol-1
H. (g) + e → H (g)                ΔH = -72.8 kJmol-1
so overall for hydrogen
½H2 (g) + e → H (g)            ΔH = +145.2 kJmol-1
and
½Cl2 (g) → Cl. (g)                 ΔH = 121 kJmol-1
Cl. (g) + e → Cl (g)              ΔH = -348.6 kJmol-1
overall for chlorine
½Cl2 (g) + e → Cl (g)          ΔH = -227.6 kJmol-1

As a result, only the most active elements, whose Ionization Energies are low, can form ionic hydrides, e.g. NaH.

The covalent radius for H is 37 pm and the estimated radius for H is ~140 pm indicating a substantial increase. This comes about as a result of the interelectronic repulsion when a second electron is added to the 1s atomic orbital. All the alkali metal hydrides crystallize with the NaCl-type structure and are all considered ionic. They are sometimes called “saline” hydrides.

Saline hydrides

The instability of the hydride ion compared to the halide ions can be seen by comparison of the ΔHf for alkali metal hydrides and chlorides.

Cation ΔHf MH/ kJmol-1 ΔHf MCl/ kJmol-1
Li -90.5 -409
Na -56.3 -411
K -57.7 -436
Rb -52.3 -430
Cs -54.2 -433

Saline hydrides are formed by the group 1 and 2 metals when heated with dihydrogen (H2). They are white, high melting point solids that react immediately with protic solvents, for example:

NaH + H2O → NaOH + H2

(Their moisture sensitivity means that reaction conditions must be water-free.)

Evidence for the ionic nature of these hydrides is:
1) molten salts show ionic conductivity.
2) X-ray crystal data gives reasonable radius ratios expected for ionic compounds.
3) Observed and calculated Lattice Energies (from Born-Haber cycles etc.) are in good agreement (i.e. show little covalency).
NaH is capable of deprotonating a range of even weak Brønsted acids to give the corresponding sodium derivatives.

NaH + Ph2PH → Na[PPh2] + H2

Sodium hydride is sold by many chemical suppliers as a mixture of 60% sodium hydride (w/w) in mineral oil. Such a dispersion is safer to handle and weigh than pure NaH. The compound can be used in this form but the pure grey solid can be prepared by rinsing the oil with pentane or tetrahydrofuran, THF, care being taken because the washings will contain traces of NaH that can ignite in air. Reactions involving NaH require an inert atmosphere, such as nitrogen or argon gas. Typically NaH is used as a suspension in THF, a solvent that resists deprotonation but solvates many organosodium compounds.

Hydride reducing agents

LiH and Al2Cl6 gives lithium aluminium hydride (lithal LiAlH4), NaH reacts with B(OCH3)3 to give sodium borohydride (NaBH4). These find wide scope and utility in organic chemistry as reducing agents.

LiAlH4 is commonly used for the reduction of esters and carboxylic acids to primary alcohols; previously this was a difficult conversion that used sodium metal in boiling ethanol (the Bouveault-Blanc reduction). The solid is dangerously reactive toward water, releasing gaseous hydrogen (H2). Some related derivatives have been discussed for hydrogen storage.

NaBH4 is used in large amounts for the production of sodium dithionite from sulfur dioxide: Sodium dithionite is used as a bleaching agent for wood pulp and in the dyeing industry. NaBH4 consists of the tetrahedral BH4 anion in the crystalline form and is found to exist as three polymorphs: α, β and γ. The stable phase at room temperature and pressure is α-NaBH4, which is cubic and adopts an NaCl-type structure. Millions of kilograms are produced annually, far exceeding the production levels of any other hydride reducing agent.

NaBH4 will reduce many organic carbonyls, depending on the precise conditions. Most typically, it is used in the laboratory for converting ketones and aldehydes to alcohols. For example, reduction of acetone (propanone) to give propan-2-ol.

Molecular hydrides – covalent hydrides and organic compounds

Hydrogen forms a vast number of compounds with carbon, (the hydrocarbons), and an even larger array with heteroatoms that, because of their general association with living things, are called organic compounds. The study of their properties is covered in organic chemistry and their study in the context of living organisms is covered in biochemistry. By some definitions, “organic” compounds are only required to contain carbon. However, most of them also contain hydrogen, and because it is the carbon-hydrogen bond which gives this class of compounds most of its particular chemical characteristics, carbon-hydrogen bonds are required in some definitions of the word “organic” in chemistry. Millions of hydrocarbons are known, and they are usually formed by complicated synthetic pathways, which seldom involve direct reaction with elementary hydrogen.

Most molecular hydrides are volatile and many have simple structures that can be predicted by the VSEPR model. There are a large number of B hydrides known (boranes) and although the simplest BH3 has been found in the gas phase it readily dimerises to give B2H6

In inorganic chemistry, hydrides can serve as bridging ligands that link two metal centers in a coordination complex. This function is particularly common in group 13 elements, especially in boranes (boron hydrides) and aluminium complexes, as well as in clustered carboranes, (composed of boron, carbon and hydrogen atoms). The bonding of the bridging hydrogens in many of the boranes is explained in terms of 3 centre – 2 electron bonds.

Diborane is a colourless and highly unstable gas at room temperature with a repulsively sweet odour. Diborane mixes well with air, easily forming explosive mixtures. Diborane will ignite spontaneously in moist air at room temperature.

 

Metallic (interstitial) hydrides

Many transition metal elements form metallic (interstitial) hydrides, in which H2 molecules (and H atoms) can occupy the holes in the metal’s crystal structure. They are traditionally termed ‘compounds’, even though they do not strictly conform to the definition of a compound; more closely resembling common alloys such as steel. These systems are usually non-stoichiometric, with variable amounts of hydrogen atoms in the lattice.

Palladium is unique in its ability to reversibly absorb large amounts of H2 or D2 (up to 900 times its own volume of hydrogen, but no other gases, at room temperature) to form palladium hydride. Structural studies show that the absorbed H fits into octahedral holes in the cubic close packed Pd lattice with a non-stoichiometric formula approximating to PdH0.6 for the β-form. This material has been considered as a means to carry hydrogen for vehicular fuel cells. Interstitial hydrides show some promise as a way for safe hydrogen storage. During the last 25 years many interstitial hydrides have been developed that readily absorb and discharge hydrogen at room temperature and atmospheric pressure. At this stage their application is still limited, as they are capable of storing only about 2 weight percent of hydrogen, insufficient for automotive applications.

Hydrogen bonds

A hydrogen bond is the name given to the electrostatic attraction between polar molecules that occurs when a hydrogen (H) atom bound to a highly electronegative atom such as nitrogen (N), oxygen (O) or fluorine (F) experiences attraction to some other nearby highly electronegative atom. The name is something of a misnomer, as it represents a particularly strong dipole-dipole attraction, rather than a typical covalent bond.

The 2011 IUPAC definition specifies that “The hydrogen bond is an attractive interaction between a hydrogen atom from a molecule or a molecular fragment X-H in which X is more electronegative than H, and an atom or a group of atoms in the same or a different molecule, in which there is evidence of bond formation.

These hydrogen-bond attractions can occur between molecules (intermolecular) or within different parts of a single molecule (intramolecular). The hydrogen bond (5 to 30 kJ/mole) is stronger than a van der Waals interaction, but weaker than covalent or ionic bonds. This type of bond can occur in inorganic molecules such as water and in organic molecules like DNA and proteins.

Intermolecular hydrogen bonding is responsible for the high boiling point of water (100 °C) compared to the other group 16 hydrides that have no hydrogen bonds. Intramolecular hydrogen bonding is partly responsible for the secondary and tertiary structures of proteins and nucleic acids. It plays an important role in the structure of polymers, both synthetic and natural.

BP’s of MG hydrides with Noble gases for comparison /K

Hydrogen bonding in biological systems.

Base pairs, which form between specific nucleobases (also termed nitrogenous bases), are the building blocks of the DNA double helix and contribute to the folded structure of both DNA and RNA. Dictated by specific hydrogen bonding patterns, Watson-Crick base pairs (guanine-cytosine and adenine-thymine) allow the DNA helix to maintain a regular helical structure that is subtly dependent on its nucleotide sequence. The complementary nature of this based-paired structure provides a backup copy of all genetic information encoded within double-stranded DNA. The regular structure and data redundancy provided by the DNA double helix make DNA well suited to the storage of genetic information, while base-pairing between DNA and incoming nucleotides provides the mechanism through which DNA polymerase replicates DNA, and RNA polymerase transcribes DNA into RNA. Many DNA-binding proteins can recognize specific base pairing patterns that identify particular regulatory regions of genes.

Applications of hydrogen

Large quantities of H2 are used by the petroleum and chemical industries. The largest application of H2 is for the processing (“upgrading”) of fossil fuels, and in the production of ammonia. The key consumers of H2 in the petrochemical plant include hydrodealkylation, hydrodesulfurization, and hydrocracking. H2 has several other important uses. H2 is used as a hydrogenating agent, particularly in increasing the level of saturation of unsaturated fats and oils (found in items such as margarine), and in the production of methanol. It is similarly the source of hydrogen in the manufacture of hydrochloric acid. H2 is used as a reducing agent of metallic ores.

Nitrogen is a strong limiting nutrient in plant growth. Carbon and oxygen are also critical, but are more easily obtained by plants from soil and air. Even though air is 78% nitrogen, atmospheric nitrogen is nutritionally unavailable because nitrogen molecules are held together by strong triple bonds. Nitrogen must be ‘fixed’, i.e. converted into some bioavailable form, through natural or man-made processes. It was not until the early 20th century that Fritz Haber developed the first practical process to convert atmospheric nitrogen to ammonia, which is nutritionally available.

Fertilizer generated from ammonia produced by the Haber process is estimated to be responsible for sustaining one-third of the Earth’s population. It is estimated that half of the protein within human beings is made of nitrogen that was originally fixed by this process; the remainder was produced by nitrogen fixing bacteria and archaea.

Dozens of chemical plants worldwide produce ammonia, consuming more than 1% of all man-made power. Ammonia production is thus a significant component of the world energy budget. Modern ammonia-producing plants depend on industrial hydrogen production to react with atmospheric nitrogen using a magnetite catalyst or over a promoted Fe catalyst under high pressure (100 standard atmospheres (10,000 kPa)) and temperature (450 °C) to form anhydrous liquid ammonia. This step is known as the ammonia synthesis loop (also referred to as the Haber-Bosch process):

3 H2 + N2 ⇄ 2 NH3 (ΔH = -92.4 kJmol-1)

Nitrogen (N2) is very unreactive because the molecules are held together by strong triple bonds. The Haber process relies on catalysts that accelerate the cleavage of this triple bond.

At room temperature, the equilibrium is strongly in favor of ammonia, but the reaction doesn’t proceed at a detectable rate. Thus two opposing considerations are relevant to this synthesis. One possible solution is to raise the temperature, but because the reaction is exothermic, the equilibrium quickly becomes quite unfavourable at atmospheric pressure. Low temperatures are not an option since the catalyst requires a temperature of at least 400 °C to be efficient. By increasing the pressure to around 200 atm the equilibrium concentrations are altered to give a profitable yield.

The reaction scheme, involving the heterogeneous catalyst, is believed to involve the following steps:

1.   N2 (g) → N2 (adsorbed)

2.   N2 (adsorbed) → 2 N (adsorbed)

3.   H2(g) → H2 (adsorbed)

4.   H2 (adsorbed) → 2 H (adsorbed)

5.   N (adsorbed) + 3 H (adsorbed) → NH3 (adsorbed)

6.   NH3 (adsorbed) → NH3 (g)

Reaction 5 actually consists of three steps, forming NH, NH2, and then NH3. Experimental evidence suggests that reaction 2 is the slow, rate-determining step. This is not unexpected given that the bond broken, the nitrogen triple bond, is the strongest of the bonds that must be broken.

 

 

Iron

Iron discovery 

Iron is the second most abundant metal is nature after aluminium. But native iron is extremely rare. Probably, the first iron used by our forefathers was of a meteoritic origin.

Iron oxidizes readily in the presence of water and air and is found mainly in the form of oxides. Oxidation if iron is responsible for the fact that extent articles made of iron in antiquity are extremely rare. Man discovered iron about five thousand years ago. At first iron was very expensive and was valued much higher than gold; very often iron jewelry was set in gold.

People of all continents became aware of gold, silver, and copper approximately at the same time; but in the case of iron the situation is different. Thus, in Egypt and Meso-potamia the process of extracting iron from ores was discovered two thousand years B.C. ; in Trans-Caucasus, Asia Minor, and ancient Greece at the end of the second milleniumin; india in the middle of the second millenium; and in china much later, only in the middle of the first millenium B.C. in the countries of the New World Iron Age began only with the arrival of Europeans, i.e., in the second millennium A.D. ; some African tribes began to use iron skipping the Bronze Age period in development. This is due to the difference in natural conditions.

In countries where natural resources of copper and tin were small, a demand arose for replacing these metals. America had one of the largest deposits of native copper and, therefore, it was not necessary to search for new metals. Gradually, production of iron grew and iron began to pass from the category of precious metals into that of ordinary ones. By the beginning of the Christian era iron was already widely used.

Among all metals and alloys known by that time, iron was the hardest one. Therefore, as soon as iron grew relatively cheap, various tools and weapons were manufactured from it. At the beginning of the first millennium A.D. production of iron in Europe and Asia had made considerable progress; particularly great successes in smelting and processing iron had been achieved by India metallurgists.

It is interesting to have a look at the development of iron production methods. At first man used only meteoritic iron, which was very rare and therefore expensive. Then people learnt how to produce iron by intensively heating its ores with coal on windy sites.Iron thus obtained was spongy, of low grade, and with large inclusions of slag. An important step in iron production was made with the invention of a furnace open at the top and lined with a refractory material inside. Excavations of ancient towns in Syria indicate that iron of a rather good quality was produced in this way. Later, people noted that cast iron which had been considered to be a waste product could be transformed into iron, the process requiring much less coal and yielding highquality iron.

By the end of the 15th century first smelting furnaces appeared producing exclusively cast iron. Iron and steel smelting processes were rapidly improving. In 1885 there appeared the converter process of steel making which is still used. The Martin process developed in 1865 yield steel almost free of slags.

A chemical symbol Fe originates from the Latin ferrum, which means “iron”.

 

Copper

 

Copper

According to the French chemist M. Berthelot, mankind came to know copper more than five thousand years ago. Other scientists believe that this acquaintance is much older. Copper and its alloy with tin (bronze) had for a long time been the most widely used metals. These two materials marked a whole epoch in the history of mankind–the Bronze Age. Why did copper play such an important part? Copper is fairly abundant in nature and can readily be worked.

At first people used only native copper but later rising demand led to the processing of copper ores. It is comparatively simple to smelt the metal from ores with high copper content. As early as the third millennium B.C copper was widely used for manufacturing various tools. The Egyptian Pyramid of Cheops was built with gigantic stone blocks each of which was hewn with copper tools.

The raw material that is used to manufacture SIMCO copper products are LME grade ‘A’ cathodes of purity greater than 99.99% copper resulting in higher conductivity, crack free extrusion, excellent finish and longer life.

Among the copper mines of antiquity, the particularly famous ones were those on the island of Cyprus to which, as has been suggested, copper owes its name (cuprum in Latin). Only when man had learned to produce bronze, stone tools were completely replaced with bronze ones. Most likely bronze was first obtained by chance. This is evidenced by the archaeological finds on the island of create dating back to about 3500 B.C. which revealed not only copper but bronze articles as well. At first bronze was rather expensive and was used mainly for jewelry and luxury articles. In ancient Egypt mirrors were made from bronze. Bronze, like copper, proved to be an excellent material for relict makers and sculptors. As early as the 5th century B.C. man learned to cast bronze statues. Particular progress in bronze sculpture was made in ancient Greece beginning with the Mycenaean period. At our times copper and bronze still retain this role.

 

Besides bronze, another wonderful copper alloy, brass, has been known for a long time. It was prepared by fusing copper with zinc ore. Ancient Egyptians, Indians, Assyrians, Romans, and Greeks knew copper. Bronze, and brass. Both copper and bronze were used for making weapons. In excavations dated back to the 8th-6th centuries B.C. in Altai, Siberia, and Trans-Caucasus archaeologists found knives, arrow-heads, shields, and helmets made from bronze and copper. In ancient Greece and Rome copper and bronze were also used for making shields and helmets. Copper found other uses in firearms when they had been invented.

 

Gold

Gold

Karl Marx wrote: “Gold is in fact the first metal that man has discovered”.

This is really so. Gold articles were found in excavations together with stone tools dating from the Neolithic Age. But in those times people, evidently, used gold found by chance. Only after the emergence of classes in society first attempts were made to mine gold. The explanation is simple. Gold was particularly suited to play the function of money due to its properties of immutability, easy divisibility, and high cost.

As an ornamental material, gold began to be used from time immemorial. During excavations of pyramids of all dynasties in Egypt archaeologists found in great numbers not only gold jewelry but also household articles.

Gold bars on nugget grains 

When Gold was discovered is unknown. Gold was known not only in Egypt, as early as in the 10th century B.C. but also it was used in China, India states of Mesopotamia. In Greece gold coins circulated as far back as in the 8th –7th centuries B.C. In Armenia gold coins appeared in the 1st century B.C. Thus, gold was known to the peoples of ancient states in Europe and Asia. The oldest gold mines were found in India and Nubia (North-East Africa).

The processes of gold purification known in antiquity did not yield the pure metal but usually alloys consisting of gold silver which were named azem. A natural gold-silver alloy–electrum–was also known.

No other metal has played so sinister a role as gold in the history of mankind. Wars were waged, nations and states were annihilated, monstrous crimes were committed for the sake of gold. But possession of gold did not bring peace to man. On the contrary, sorrow and fear of losing this treasure filled his soul.

The alchemic period between the 4th and the 16th centuries was a gloomy one in the history of the search for gold. The efforts of alchemists were directed towards the search for the “philosophers’ stone” which, they held, possessed the property of transforming base metals into good. Alchemy did not start from scratch but had important precursors. Egypt’s fast rise was due to the fact that Egyptians possessed the secret of gold extraction. It was also known that iron articles that remained in copper mines for a long time became coated with copper. Iron was believed to transform into copper. If it was so, why could not other metals be transformed into gold? Native lead sulphide almost always contains an admixture of silver, which could sometimes be extracted. Could not silver be formed on lead? And, finally, progress in alchemy was facilitated by the idea about the unity of matter according to which all substances consist of the same components in different ratios.

All the attempts to find the “philosophers’ stone” turned out to be unsuccessful (as one should have expected), although many alchemists gave their lives for the idea. All reports about the discoveries of methods of preparing gold from other metals were pure charlatanism.

Alchemy was still flourishing in Europe when the first Spanish conquistadors set out for South and Central America. In the land of Incas they were amazed by the tremendous amounts of gold. For Incas gold was a sacred metal, the sun God’s metal, and colossal amount of gold had accumulated in the temples. When the Spaniards took Atahualpa, the Great Inca, prisoner, they promised him freedom for a fantastic ransom of almost 50 m3 of gold. But Francisco Pizarro thought it dangerous to free the Great Inca and, without waiting for the ransom, the Spaniards executed Atahualpa. When the Incas learned about the death of their leader, the caravan consisting of 1 100 Ilamas carrying gold had already been on its way. Incas hid the gold in the mountains of Azangaro (“the remotest place”). But they could not hide all their treasures. Spaniards captured and looted Cuzco, one of Peru’s richest cities. They, melted the priceless creations of ancient craftsmen into cold ingots and sent them to Spain. In Russia mining of gold began in 1600 but it was not until the 19th century that the large-scale extraction of this metal started.

The Latin name for gold, aurum, originates from the word Aurora (dawn).

 

Sulphur

Sulphur

Sulphur has been known to man for a very long time. Even in times of Homer ancient Greeks used the specific properties of Sulphur dioxide liberated in the burning of Sulphur for disinfection of homes. Deposits of native sulphur have also been known from ancient times. Thus, Pliny the Elder described the deposits of Sulphur in Italy and Sicily.

Sulphur was used for making dyes and treating fabrics. Like carbon, from the earliest times Sulphur was used in pyrotechnics. The composition known by the name of “Greek fire” and invented, apparently, in the 5th century A.D. in Byzantium was a mixture of finely ground Sulphur (one part), coal (two part), and saltpeter (six parts). It is interesting to note that this composition differs only slightly from that of black (smoky) gunpowder.

The fact that Sulphur is a good combustible material and combines readily with a great number of metals is responsible for its “privileged” position among other substances in the Middle Ages. Alchemists considered Sulphur as the element of combustibility and a constituent of all metals. Very unusual properties were often attributed to Sulphur, although some alchemists described its real properties rather accurately.

The elemental nature of Sulphur was established by A. Lavoisier. However, in spite of the fact that by the beginning of the 19th century Sulphur had already been recognized as an independent element, experiments had to be carried out to elucidate the exact composition of native Sulphur. In 1808 H. Davy suggested that Sulphur in its usual state is a combination of small amounts of oxygen and hydrogen with a great amount ofSulphur. This questioned the elemental nature of Sulphur but in 1809 Gay Lussac proved it beyond any doubt. In 1810 Davy pointed out that the presence of oxygen in Sulphur. The oxygen content in Sulphur varied depending on the deposit where the samples were taken. From the standpoint of modern chemistry one may say that oxygen found by Davy in Sulphur was not the oxygen of Sulphur oxides but that of oxysulphides of various metals, which are always present in Sulphur. The origin of the Latin word “Sulphur” is unclear.

 

Phosphorus

Phosphorus

Interestingly, among all elements of antiquity and Middle Ages only phosphorus has the exact (within a year) date of the discovery, namely 1669. There is no reliable information whether man had known phosphorus or its compounds before that or not. The unexpected discovery of phosphorus in the 17th century profoundly impressed the academic world a

match box are made using Phosphorus , Sulphur and Potassium Chlorate

nd was a real sensation owing to unusual property of the substance (it is too early to name it an “element”) it glowed in air at room temperature. Such compounds (for instance, Bologna stone–the product of calcination of baryta with coal and oil, i.e., barium sulphide BaS), were called “phosphorus: (from the Greek phos, light and phoro, to bear). Thus, the name appeared prior to the discovery of the element itself.

 

The history of its discovery was also unusual. There once lived in Hamburg a bankrupt merchant by the name of Hening Brand. At that time alchemy had already begun to lose ground but the belief in the “philosophers’ stone” was still alive. H. Brand was one of those who believed in it. With a view to mending his business, he began to search for primary matter in various compounds. Human urine was one of the materials he analysed. H. Brand evaporated urine up to a syrupy liquid, distilled it, and obtained a red liquid which he named urine oil. Having distilled this liquid once more, Brand saw a black precipitate at the bottom of his retort. After prolonged calcination the residue transformed into a white glowing substance precipitated on the walls of the vessel. Imagine the joy of the alchemist! He was sure that he had succeeded in isolating elementary fire. H. Brand tried to keep his discovery a secret and continued the work with phosphorus hoping to obtain gold from other metals. These efforts, as one might have expected, were in vain. But H. Brand could not keep his secret for a long time and he finally revealed it himself. Having failed to obtain gold from the other metals, Brand decided to put the new remarkable substance on sale keeping secret the method of its preparation. But in this attempt he also failed. As soon as phosphorus became known in Europe, it attracted attention of many scientists: the famous mathematician G. Leibniz, J. Kraft, J. Kunkel, R. Boyle, Ch. Huygens, and many other chemists and physicists. J. Kunkel, who was at that time the alchemist at the court of the Prince of Saxony, sent J. Kraft, his assistant, to Hamburg to get the secret phosphorus preparation from Brand. J. Kraft bought the secret for 200 thalers but it did not reach Kunkel. Kraft decided to keep the method of preparing the new substance to himself; he went on a trip of Europe to impress society with the marvelous substance’s glow. J. Kunkel tried to prepare phosphorus himself and after long work he succeeded in separating the new element.

The details of the method by which H. Brand prepared phosphorus did not reach is but the method of kunkel (1676) is known rather well. Fresh urine was evaporated forming a black precipitated which was heated at first carefully and then intensively with sand and coal. After removal of volatile and oily compounds, phosphorus precipitated on cold walls of the retort as a white deposit. The following chemical reaction were involved in the process.

(a)   NaNH4HPO4  → NAPO3  + NH3 +H2O

(b)   4NaPO3 + 2SiO2  → 2Na2SiO3 + P4O10

(c)    P4O10 + 10C →  P4 + 10CO

However, KunKel also decided not to make the method public. In 1680 R. Boyle became third scientist to obtain phosphorus by approximately the same method; he reported it in a private letter to the London Royal Society. A. Hanckewitz, Boyle’s assistant, organized production of phosphorus on a fairly large scale, deriving large profits since phosphorus was expensive.

It was believed for a long time that phosphorus existed only in one (white) allotropic modification but in 1847 A. Schroeter, heating white phosphorus up to 300oC without air, obtained red phosphorus which, in contrast to the white phosphorus, was neither toxic nor combustible in air. In 1934 P. Bridgeman obtained the third modification, namely, black phosphorus, having subjected phosphorus to heating under 

source of phosphorus in our food

high pressure.

Carbon

Carbon

The exact date of the discovery of carbon cannot be ascertained. However, it is not difficult to find out when carbon was identified as a simple substance. Let us direct our attention to “The Table of Simple Bodies” compiled by A. Lavoisier and published in 1789. Carbon appears as a simple substance in it. However, the time that carbon needed to occupy its place in the Table is measured not by years and even not by centuries but buy millennia. Man had met carbon even before he could make fire–in the form of woods burnt by lighting. After man had learnt how to start a fire, carbon became his constant “companion”.

Carbon played an important role in the progress of the phlogistic theory. According to this theory carbon was not a simple substance but pure phlogiston. By studying combustion of coal and other compounds, A. Lavoisier was the first to show that carbon is a simple substance. Here we are going to digress a little from the story about how carbon found its identity.

In nature carbon occurs in two allotropic modifications–diamond and graphite, both known to man for a long time. The fact that diamond burns without a residue at very high temperatures was also known long ago. Nevertheless, diamond and graphite

Pencil lead is Graphite

were believed to be two quite different substances. The discovery of carbon dioxide was an event which helped to establish that of diamond and graphite are modification of the same substance. After experimenting with the burning of diamond and charcoal, A. Lavoisier established that upon combustion both substances yield carbon dioxide. This prompted the conclusion that diamond and coal have the same origin. The name “carboneum” (carbon) appeared for the first time in the book “Methods of Chemical Nomenclature” (A. Lavoisier, L. Guyton de Morveau, C. Berthollet, and A. Fourcroy) in 1787.

A parallel can be drawn between the element itself, known from time immemorial, and its Latin name whose root originates from Sanscrit, one of the oldest known languages. In Sanscrit “cra” means “to boil”. The name “carbon” was suggested in 1824.

Diamond is also Carbon

In 1797 S. Tennant discovered that combustion of equal amounts of diamond and graphite liberates equal amounts of carbon dioxide; in 1799 L. Guyton de Morveau confirmed that carbon is the only constituent of diamond, graphite, and coke. Twenty years later he succeeded in transforming diamond into graphite and then into carbon dioxide by careful heating. But the reverse transformation of graphite into diamond was beyond the power of the science of the 18th and 19th centuries. It was only in 1955 that a group of English scientists obtained artificial diamonds for the first time in the world’s history. Synthesis was performed at 3000oC under a pressure exceeding 109 Pa.

Soon after the synthesis of diamond Soviet scientists prepared a new substance, carbine, which, as

has since been proved, is a new, third allotropic modification of carbon. The carbon atoms in it comprise long chains. This substance resembles soot.

The study of carbon and its compounds laid the foundation of a vast field of chemistry–organic chemistry.