Category Archives: Blog

Silver Protective metal
Silver : The Protective Metal : Ag Detection

 

SILVER, Ag (At : 107.868) Silver is a white, malleable, and ductile metal. It has a high density (10.5 g ml-1) and melts at 96.5oC. It is insoluble in hydrochloric, dilute sulphuric (M) or dilute nitric (2M) acid. In more concentrated nitric acid (8M) (a) or in hot, concentrated sulphuric acid (b) it dissolves:

    \[6Ag+8HN{{O}_{3}}\to 6A{{g}^{+}}+2NO\uparrow +6NO_{3}^{-}+4{{H}_{2}}\]

    \[2Ag+2{{H}_{2}}S{{O}_{4}}\to 2A{{g}^{+}}+SO_{4}^{2-}+S{{O}_{2}}\uparrow +2{{H}_{2}}O\]

Silver forms monovalent ion in solution, which is colourless. Silver (II) compounds are unstable, but play an important role in silver-catalysed oxidation reduction processes. Silver nitrate is readily soluble in water, silver acetate, nitrite and sulphate are less soluble, while all the other silver compounds are practically insoluble. Silver complexes are however soluble, Silver halides are sensitive to light; these characteristics are widely utilized in photography.

 

Reactions of silver (I) ions : A solution of silver nitrate (0.1M) can be used to study these reactions.

1. Dilute hydrochloric acid: (or soluble chlorides): white precipitate of silver chloride

    \[A{{g}^{+}}+C{{l}^{-}}\to AgCl\downarrow \]

With concentrated hydrochloric acid precipitation does not occur. Decanting the liquid from over the precipitate, it dissolves in concentrated hydrochloric acid, when a dichloroargentate complex is formed:

    \[AgCl\downarrow +\,C{{l}^{-}}\rightleftharpoons {{[AgC{{l}_{2}}]}^{-}}\]

By diluting with water, the equlibrium shifts back to the left and the precipitate reappears.

Dilute ammonia solution dissolves the precipitate, when diammineargentate complex ion is formed:

    \[AgCl\downarrow +2N{{H}_{3}}\to {{[Ag{{(N{{H}_{3}})}_{2}}]}^{+}}+C{{l}^{-}}\]

Dilute nitric acid or hydrochloric acid neutralizes the excess ammonia, and the precipitate reappears because the equilibrium is shifted back towards left.

Potassium cyanide (POISON) : dissolves the precipitate with formation of the dicyanoargentate complex:

    \[AgCl\downarrow +2C{{N}^{-}}\to {{[Ag{{(CN)}_{2}}]}^{-}}+C{{l}^{-}}\]

The safest way to study this reaction is as follows: decant the liquid from the Precipitate, and wash it 2-3 time with water by decantation. Then apply e reagent.

Sodium thiosulphate dissolves the precipitate with the formation of dithiosulphatoargentate complex:

    \[AgCl\downarrow +2{{S}_{2}}O_{3}^{2-}\to {{[Ag{{({{S}_{2}}{{O}_{3}})}_{2}}]}^{3-}}+C{{l}^{-}}\]

This reaction takes place when fixing photographic negatives or positive prints after development.

Sunlight or ultraviolet irradiation decomposes the silver chloride precipitate which turns to greyish or black owing to the formation of silver metal:

    \[2AgCl\downarrow \xrightarrow{(hv)}2Ag\downarrow +C{{l}_{2}}\uparrow \]

The reaction is slow and the actual reaction mechanism is very complicated. Other silver halides show similar behaviour. Photography is based on these reactions. In the camera these processes are only initiated the photographic material has to be Developed to complete the reaction.

Greyish or black silver particles appear on places irradiated by light; a negative image of the object is therefore obtained, The excess of silver halide has to be removed (to make the developed negative insensitive to light) by fixation.

 

2. Hydrogen sulphide: (gas or saturated aqueous solution) in neutral or acidic medium : back precipitate of silver sulphide

    \[2A{{g}^{+}}+{{H}_{2}}S\to A{{g}_{2}}S\downarrow +2{{H}^{+}}\]

Hot concentrate nitric acid decomposes the silver sulphide, and sulphur remains in  the form of white precipitate.

    \[3A{{g}_{2}}S\downarrow +8HN{{O}_{3}}\to S\downarrow +2NO\uparrow +6A{{g}^{+}}+6NO_{3}^{-}+4{{H}_{2}}O\]

The reaction can be understood better if written in two steps:

    \[3A{{g}_{2}}S\downarrow +2HN{{O}_{3}}\to S\downarrow +2NO\uparrow +3A{{g}_{2}}O+{{H}_{2}}O\]

    \[3A{{g}_{2}}S\downarrow +6HN{{O}_{3}}\to 6A{{g}^{+}}+6NO_{3}^{-}+3{{H}_{2}}O\]

If the mixture is heated with concentrated nitric acid for a considerable time sulphur is oxidized to sulphate and the precipitate disappears:

    \[S\downarrow +2HN{{O}_{3}}\to SO_{4}^{2-}+2NO\uparrow +2{{H}^{+}}\]

The precipitate is insoluble in ammonium sulphide, ammonium polysulphide, ammonia, potassium cyanide, or sodium thiosulphate. Silver sulphide can be precipitated from solutions containing diammine-, dicyanato- or dithiosulphato argentate complexes with hydrogen sulphide,

 

3. Ammonia solution: brown precipitate of silver oxide

    \[2A{{g}^{+}}+2N{{H}_{3}}+{{H}_{2}}O\to A{{g}_{2}}O\downarrow +2NH_{4}^{+}\]

The reaction reaches an equilibrium and therefore precipitation is incomplete at any stage. (If ammonium nitrate is present in the original solution or the solution is strongly acidic no precipitation occurs) The precipitate dissolves in excess of the reagent, and diammineargentate complex ions are formed:

    \[A{{g}_{2}}O\downarrow +4N{{H}_{3}}+{{H}_{2}}O\to 2{{[Ag{{(N{{H}_{3}})}_{2}}]}^{+}}+2O{{H}^{-}}\]

The solution should be diposed of quickly, because when set aside silver nitride Ag3N precipitate is formed, which explodes readily even in a wet form.

 

4. Sodium hydroxide: brown precipitate of silver oxide:

    \[2A{{g}^{+}}+2O{{H}^{-}}\to A{{g}_{2}}O\downarrow +{{H}_{2}}O\]

A well-washed suspension of the precipitate shows a slight alkaline reaction owing to the hydrolysis equilibrium:

    \[A{{g}_{2}}O\downarrow +{{H}_{2}}O\rightleftharpoons 2Ag{{(OH)}_{2}}\downarrow \,\rightleftharpoons 2A{{g}^{+}}+2O{{H}^{-}}\]

The precipitate is insoluble in excess reagent.

The precipitate dissolves in ammonia solution (a) and in nitric acid (b):

    \[A{{g}_{2}}O\downarrow +4N{{H}_{3}}+{{H}_{2}}O\to 2{{[Ag{{(N{{H}_{3}})}_{2}}]}^{+}}+2O{{H}^{-}}\]

    \[A{{g}_{2}}O\downarrow +2{{H}^{+}}\to 2A{{g}^{+}}+{{H}_{2}}O\]

 

5. Potassium iodide: yellow precipitate of silver iodide

    \[A{{g}^{+}}+{{I}^{-}}\to AgI\downarrow \]

The precipitate is insoluble in dilute or concentrated ammonia, but dissolve readily in potassium cyanide (POISON) (a) and in sodium thiosulphate(b):

    \[AgI+2C{{N}^{-}}\to {{[Ag{{(CN)}_{2}}]}^{-}}+{{I}^{-}}\]

    \[AgI+2{{S}_{2}}O_{3}^{2-}\to [Ag{{({{S}_{2}}{{O}_{3}}]}^{3-}}+{{I}^{-}}\]

 

6. Potassium chromate in neutral solution: red precipitate of silver chromate

    \[2A{{g}^{+}}+CrO_{4}^{2-}\to A{{g}_{2}}Cr{{O}_{4}}\downarrow \]

Spot test: place a drop of the test solution on a watch glass or on a spot plate, add a drop of ammonium carbonate solution and stir (this renders any mercury(I) or lead ions unreactive by precipitation as the highly insoluble carbonates). Remove one drop of the clear liquid and place it on drop-reaction Paper together with a drop of the potassium chromate reagent. A red ring of silver chromate is obtained.

Te reaction can be used for microscopic test, when a piece of potassium chromate crystal has to be dropped into the test solution. The formation of needle-like red crystals of silver chromate can be observed distinctly.

The precipitate is soluble in dilute nitric acid (a) and in ammonia solution (b):

    \[2A{{g}_{2}}Cr{{O}_{4}}\downarrow +2{{H}^{+}}\rightleftharpoons 4A{{g}^{+}}+C{{r}_{2}}O_{7}^{2-}+{{H}_{2}}O\]

    \[A{{g}_{2}}Cr{{O}_{4}}\downarrow +4N{{H}_{3}}\to 2{{[Ag{{(N{{H}_{3}})}_{2}}]}^{+}}+C{{r}_{2}}O_{4}^{2-}\]

The acidified solution turns to orange because of the formation of dichromate : ions in reaction (a).

 

7. Potassium cyanide (POISON) when added dropwise to a neutral solution of silver nitrate: white precipitate of silver cyanide:

    \[A{{g}^{+}}+C{{N}^{-}}\to AgCN\downarrow \]

When potassium cyanide is added in excess, the precipitate disappears owing to the formation of dicyanoargentate ions.

    \[AgCN\downarrow +C{{N}^{-}}\to {{[Ag{{(CN)}_{2}}]}^{-}}\]

 

8. Sodium carbonate: yellowish-white precipitate of silver carbonate:

    \[2A{{g}^{+}}+CO_{3}^{2-}\to A{{g}_{2}}C{{O}_{3}}\downarrow \]

When heating, the precipitate decomposes and brown silver oxide precipitate is formed:

    \[A{{g}_{2}}C{{O}_{3}}\downarrow \to A{{g}_{2}}O\downarrow +C{{O}_{2}}\uparrow \]

Nitric acid (a) and ammonia solution (b) dissolve the precipitate

    \[A{{g}_{2}}C{{O}_{3}}\downarrow +2{{H}^{+}}\to 2A{{g}^{+}}+C{{O}_{2}}\uparrow +{{H}_{2}}O\]

    \[A{{g}_{2}}C{{O}_{3}}\downarrow +4N{{H}_{3}}\to 2{{[Ag{{(N{{H}_{3}})}_{2}}]}^{2}}+CO_{3}^{2-}\]

Carbon dioxide gas is evolved in reaction (a).

 

9. Disodium hydrogen phosphate in neutral solution: yellow precipitate of silver Phosphate:

    \[3A{{g}^{+}}+HPO_{4}^{2-}\to A{{g}_{3}}P{{O}_{4}}\downarrow +{{H}^{+}}\]

Nitric acid (a) and ammonia solution (b) dissolve the precipitate:

    \[A{{g}_{3}}P{{O}_{4}}\downarrow +3{{H}^{+}}\to 3A{{g}^{+}}+{{H}_{3}}P{{O}_{4}}\]

    \[A{{g}_{3}}P{{O}_{4}}\downarrow +6N{{H}_{3}}\to 3[Ag{{(N{{H}_{3}})}_{2}}+PO_{4}^{3-}\]

Phosphoric acid, formed in reaction (a) is a medium-strong acid, which is only slightly dissociated if nitric acid is present in excess.

 

10. Hydrazine sulphate (saturated): when added to a solution of diammineargentate ions, forms finely divided silver metal, while gaseous nitrogen is evolving:

    \[4{{[Ag{{(N{{H}_{3}})}_{2}}]}^{+}}+{{H}_{2}}N-N{{H}_{2}}.{{H}_{2}}S{{O}_{4}}\to 4Ag\downarrow +{{N}_{2}}\uparrow +6NH_{4}^{+}+2N{{H}_{3}}+SO_{4}^{2-}\]

If the vessel in which the reaction is carried out is clean, silver adheres to the glass walls forming an attractive mirror.

MERCURY : LIQUID METAL HG DETECTION
Mercury : The Liquid Metal : Hg Detection

MERCURY, Hg (At: 200.59) – MERCURY(I) Mercury is a silver white, liquid metal at ordinary temperatures and has a density of 13.534gml-1 at 25oC. It is unaffected when treated with hydrochloric or dilute sulphuric : acid (2M), but reacts readily with nitric acid. Cold, medium concentrated (8m) nitric acid with an excess of mercury yields mercury(I) ions:

    \[6Hg+8HN{{O}_{3}}\to 3Hg_{2}^{2+}+2NO\uparrow +6NO_{3}^{-}+4{{H}_{2}}O\]


with an excess of hot concentrated nitric acid mercury(II) ions are formed:

    \[3Hg+8HN{{O}_{3}}\to 3H{{g}^{2+}}+2NO\uparrow +6NO_{3}^{-}+4{{H}_{2}}O\]


Hot, concentrated sulphuric acid dissolves mercury as well. The product is mercury(I) ion if mercury is in excess

    \[2Hg+2{{H}_{2}}S{{O}_{4}}\to Hg_{2}^{2+}+SO_{4}^{2-}+S{{O}_{2}}\uparrow +2{{H}_{2}}O\]


while if the acid is in excess, mercury(II) ions are formed:

    \[Hg+2{{H}_{2}}S{{O}_{4}}\to H{{g}^{2+}}+SO_{4}^{2-}+S{{O}_{2}}\uparrow +2{{H}_{2}}O\]


The two ions, mercury(I) and mercury(II) behave quite differently against reagents used in qualitative analysis, and hence belong to two different analytical groups.

Reactions of mercury (I) ions : A solution of mercury(I) nitrate (0.05m) can be used for the study of these reactions.

    \[Hg_{2}^{2+}+2C{{l}^{-}}\to H{{g}_{2}}C{{l}_{2}}\downarrow \]


1. Dilute hydrochloric acid or soluble chlorides: white precipitate of mercury(I) chloride (calomel)
The precipitate is insoluble in dilute acids.
Ammonia solution converts the precipitate into a mixture of mercury(II)amidochloride and mercury metal, both insoluble precipitates:

    \[HgC{{l}_{2}}+2N{{H}_{3}}\to Hg\downarrow +Hg(N{{H}_{2}})Cl\downarrow +NH_{4}^{+}+C{{l}^{-}}\]


the reaction involves disproportionation, mercury(I) is converted partly to mercury(II) and partly to mercury metal. This reaction can be used to differentiate mercury(I) ions from lead(II) and silver(I).
The mercury(II)amidochoride is a white precipitate, but the finely divided mercury makes it shiny black. The name calomel, coming from Greek (καονμελασ = nice black) refers to this characteristic of the originally white mercury(I) chloride precipitate.
Mercury(I) chloride dissolves in aqua regia, forming undissociated but soluble mercury(II) chloride:

    \[3H{{g}_{2}}C{{l}_{2}}\downarrow +2HN{{O}_{3}}+6HCI\to 3HgC{{l}_{2}}+2NO\uparrow \text{ }+4{{H}_{2}}O\]

2. Hydrogen sulphide in neutral or dilute acid medium: black precipitate, which is a mixture of mercury (II) sulphide and mercury metal

    \[Hg_{2}^{2+}+{{H}_{2}}S\to \text{ }Hg\downarrow +HgS\downarrow +2{{H}^{+}}\]


Owing to the extremely low solubility product of mercury (Il) sulphide the reaction is very sensitive.
Sodium sulphide (colourless). dissolves the mercury (II) sulphide (but leaves mercury metal) and a disulphomercurate (II) complex is formed:

    \[HgS+{{S}^{2-}}\to {{[Hg{{S}_{2}}]}^{2-}}\]


After removing the mercury metal by filtration, black mercury (II) sulphide can again be precipitated by acidification with dilute mineral acids:

    \[{{\left[ Hg{{S}_{2}} \right]}^{2-}}+2{{H}^{+}}\to HgS\downarrow +{{H}_{2}}S\uparrow \]


Sodium disulphide (yellow) dissolves both mercury (II) and mercury (II) sulphide:

    \[HgS\downarrow +Hg\downarrow +3S_{2}^{2-}\to 2{{[Hg{{S}_{2}}]}^{2-}}+S_{3}^{2-}\]


This rather complicated reaction can be understood more easily by breaking it down into the following steps:
First mercury is oxidized by the disulphide, yielding mercury (II)sulphide and (mono) sulphide ions:

    \[Hg\downarrow +S_{2}^{2-}\to HgS\downarrow +{{S}^{2-}}\]


Mercury (II) sulphide then dissolves in the (mono) sulphide formed in the previous reaction

    \[HgS\downarrow +{{S}^{2-}}\to {{\left[ Hg{{S}_{2}} \right]}^{2-}}\]


Mercury (II) sulphide, which was originally present in the precipitate, reacts with disulphide ions yielding disulphomercurate(II) and trisulphide ions:

    \[HgS+2S_{2}^{2-}\to HgS_{2}^{2-}+S_{3}^{2-}\]


Combining the reactions (a), (b) and (c) together we obtain the reaction described above.
Aqua regia dissolves the precipitate, yielding undissociatedmercury(II) chloride and sulphur:

    \[12HCl+4HN{{O}_{3}},+3Hg\downarrow +3HgS\downarrow =6HgC{{l}_{2}}+3S\downarrow \text{ }+4NO\uparrow +8{{H}_{2}}O\]


This reaction can be understood as the sum of the following steps:
When making up aqua regia chlorine atoms are formed:

    \[3HCI+HN{{O}_{3}}\to 3Cl+NO\uparrow +2{{H}_{2}}O\]


These react partly with mercury, forming mercury(Il) chloride:

    \[HgS\downarrow +2CI\to HgC{{l}_{2}}+S\downarrow \]


Another part of chlorine reacts with mercury(II) sulphide

    \[HgS\downarrow +2C1\to HgC{{l}_{2}}+S\downarrow \]


Combination of 4(a)+ 3(b)+ 3(c) yields the equation

    \[12HCl+4HN{{O}_{3}}+3Hg\downarrow +3HgS\downarrow =6HgC{{l}_{2}}+3S\downarrow +4NO\uparrow +8{{H}_{2}}O\]


When heated with aqua regia, sulphur is oxidized to sulphuric acid and the solution becomes clear:

    \[S\downarrow +6HCl+2HN{{O}_{3}}\to S_{4}^{2-}+6C{{l}^{-}}+8{{H}^{+}}+2NO\uparrow \]

3. Ammonia solution: back precipitate which is a mixture of mercury metal and basic mercury(II) amidonitrate, (which itself is a white precipitate)

     \[\begin{align} & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,N{{H}_{2}} \\ & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,/ \\ & 2Hg_{2}^{2+}+NO_{3}^{-}+4N{{H}_{3}}+{{H}_{2}}O\to HgO.Hg\,\,\,\,\,\,\,\downarrow +2Hg\downarrow +3NH_{4}^{+} \\ & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\backslash  \\ & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,NO \\ \end{align}\]

This reaction can be used to differentiate between mercury(I)and mercury(II) ions.

4. Sodium hydroxide: black precipitate of mercury(I) oxide

    \[Hg_{2}^{2+}+2O{{H}^{-}}\to H{{g}_{2}}O\downarrow +{{H}_{2}}O\]


The precipitate is insoluble in excess reagent, but dissolves readily in dilute nitric acid.
When boiling, the colour of the precipitate turns to grey, owing to disproportionation, when mercury(II) oxide and mercury metal are formed:

    \[H{{g}_{2}}O\downarrow \to HgO\downarrow +Hg\downarrow \]

5. Potassium chromate in hot solution : red crystalline precipitate of mercury(I) chromate

    \[Hg_{2}^{2+}+CrO_{4}^{2-}\to H{{g}_{2}}Cr{{O}_{4}}\downarrow \]

If the test is carried out in cold, a brown amorphous precipitate is formed with an undefined composition. When heating the precipitate turns to red, crystalline mercury(I) chromate.
Sodium hydroxide turns the precipitate into black mercury(I) oxide:

    \[H{{g}_{2}}Cr{{O}_{4}}\downarrow +2O{{H}^{-}}\to H{{g}_{2}}O\downarrow +CrO_{4}^{2-}+{{H}_{2}}O\]

6. Potassium iodide, added slowly in cold solution: green precipitate of mercury (I) iodide

    \[Hg_{2}^{2+}+2{{I}^{-}}\to H{{g}_{2}}{{I}_{2}}\downarrow \]


If excess reagent is added a disproportionation reaction takes place, soluble tetraiodomercurate(Il) ions and a black precipitate of finely divided mercury being formed:

    \[H{{g}_{2}}{{I}_{2}}\downarrow +2{{I}^{-}}\to {{[Hg{{I}_{4}}]}^{2-}}+Hg\downarrow \]


When boiling the mercury(I) iodide precipitate with water, disproportionation again takes place, and a mixture of red mercury(II) iodide precipitate and finely distributed black mercury is formed:

    \[H{{g}_{2}}{{I}_{2}}\downarrow \to Hg{{I}_{2}}\downarrow +Hg\downarrow \]

7. Sodium carbonate in cold solution: yellow precipitate of mercury (I) carbonate:

    \[Hg_{2}^{2+}+CO_{3}^{2-}\to H{{g}_{2}}C{{O}_{3}}\downarrow \]


The precipitate turns slowly to blackish grey, when mercury (II) oxide and mercury are formed:

    \[H{{g}_{2}}C{{O}_{3}}\downarrow \to HgO\downarrow +Hg\downarrow +C{{O}_{2}}\uparrow \]


The decomposition can be speeded up by heating the mixture.

8. Disodium hydrogen phosphate: white precipitate of mercury(I) hydrogen phosphate:

    \[Hg_{2}^{2+}+HPO_{4}^{2-}\to H{{g}_{2}}HP{{O}_{4}}\downarrow \]

9. Potassium cyanide (POISON): produces mercury(II) cyanide solution and mercury precipitate:

    \[Hg_{2}^{2+}+2C{{N}^{-}}\to Hg\downarrow +Hg{{(CN)}_{2}}\]


Mercury(II) cyanide, though soluble, is practically undissociated.

10. Tin(II) chloride: reduces mercury(I) ions to mercury metal, which appears in the form of a greyish-black precipitate:

    \[Hg_{2}^{2+}+S{{n}^{2+}}\to 2Hg\downarrow +S{{n}^{4+}}\]


Mercury(II) ions react in a similar way.

11. Potassium nitrite: reduces mercury metal from a solution of mercury(I) ions in cold, in the form of a greyish-black precipitate:

    \[Hg_{2}^{2+}+NO_{2}^{-}+{{H}_{2}}O\to 2Hg\downarrow +NO_{3}^{-}+2{{H}^{-}}\]


Under similar circumstances mercury(II) ions do not react. The spot test technique is as follows. Place a drop of the faintly acid test solution upon drop-reaction paper and add a drop of 50 per cent potassium nitrite solution, A black (or dark grey) spot is produced, The test is highly selective. Coloured ions yield a brown colouration which may be washed away, leaving the black spot.

12. Glossy copper sheet or copper coin: If a drop of mercury(I) nitrate is placed on a glossy copper surface, a deposit of mercury metal is formed:

    \[Cu+Hg_{2}^{2+}\to C{{u}^{2+}}+2Hg\downarrow \]


Rinsing drying. And rubbing the surface with a dry cloth, a glittery, silverish spot is obtained. Heating the spot in a Bunsen-flame, mercury evaporates and the red copper surface become visible again, Mercury(II) solutions react in a similar way.

13. Aluminium sheet: If a drop of mercury(I) nitrate is placed on a clean aluminium surface, aluminium amalgam is formed and aluminium ions pass into solution:

    \[3Hg_{2}^{2+}+2Al\to 2A{{l}^{3+}}+6Hg\downarrow \]


The aluminiun which is dissolved in the amalgam is oxidized rapidly by the oxygen of the air and a voluminous precipitate of aluminiun hydroxide is formed. The remaining mercury amalgamates a further batch of aluminium, which again is oxidized thus considerable amounts of aluminium get corroded.

14. Dry test : All compounds of mercury when heated with a large excess (7-8 times the bulk) of anhydrous sodium carbonate in a small dry test-tube yield a grey mirror, consisting of fine drops of mercury, in the upper part of the tube. The globules coalesce when they are rubbed with a glass rod.
Note : Mercury vapour is extremely poisonous, and not more than 0-1 gram of the substance should be used in the test.

detect poison
Lead : A Poison, How to Detect

LEAD, Pb (At : 207.19)

Lead is poisonous element. If you have any idea about rat poison then you can link how poisonous Lead is. It is a bluish-grey metal with a high density (11.48 g ml-1 at room temperature). It readily dissolves in medium concentrated nitric acid (8M), and nitrogen oxide is formed also:

    \[3Pb+\text{ }8HN{{O}_{3}}\text{ }\to \text{ }3P{{b}^{2+}}+\text{ }6NO_{3}^{-}\text{ }+2NO\uparrow +4{{H}_{2}}O\]

The colourles nitrogen oxide gas, when mixed with air, is oxidized to red- brown nitrogen dioxide:

2NO↑ (colourless) + O2 →+2NO2 ↑(red brown)

With concentrated nitric acid a protective film of lead nitrate is formed on the surface of the metal and prevents further dissolution. Dilute hydrochloric or sulphuric acid have little effect owing to the formation of insoluble lead chloride or sulphate on the surface.

Now Let’s see how to detect Lead by chemical reactions.

Reaction of lead (II) ions : A solution of lead nitrate (0.25M) or lead acetale (0.25) can be used for the study of these reactions.

1. Dilute hydrochloric acld (or soluble chlorides) : a white precipitate in cold and not too dilate solution:

    \[P{{b}^{2+}}+2C{{l}^{-}}\Leftrightarrow PbC{{l}_{2}}\downarrow \]

The precipitate is soluble in hot water (33.4 g l-1 at 100o C while only 9.9 g l‑1 at 20o C), but separates again in long, needle-like crystals when cooling. It s also soluble in concentrated hydrochoric acid or concentrated potassium chloride when the tetrachloroplumbate(II) ion is formed:

    \[PbC{{l}_{2}}\downarrow +2C{{l}^{-}}\to {{[PbC{{l}_{4}}]}^{2-}}\]

If the precipitate is washed by decantation and dilute ammonia is added, no visble change occurs [difference from mercury (I) or silver ions], though a precipitate-exchange reaction takes place and lead hydroxide is formed:

    \[PbC{{l}_{2}}\downarrow +2N{{H}_{3}}+2{{H}_{2}}O\to \text{ }Pb{{\left( OH \right)}_{2}}\downarrow +2NH_{4}^{+}\text{+}2C{{l}^{-}}\]

2. Hydrogen sulphide in neutral or dilute acid medium: black precipitate of lead sulphide:

    \[P{{b}^{2+}}+{{H}_{2}}S\to PbS\downarrow +2{{H}^{+}}\]

Precipitation is incomplete if strong mineral acids are present in more than 2M concentration. Because hydrogen ions are formed in the above reaction, it is advisable to buffer the mixture with sodium acetate.

Introducing hydrogen sulphide gas into a mixture which contains white lead chloride precipitate, the latter is converted into (black) lead sulphide in a precipitate-exchange reaction:

    \[PbC{{l}_{2}}\downarrow {{H}_{2}}S\to PbS\downarrow +2{{H}^{+}}+2C{{l}^{-}}\]

If the test is carried out in the presence of larger amounts of chloride [Potassium chloride (saturated), initially a red precipitate of lead sulpho- chloride is formed when introducing hydrogen sulphide gas:

    \[2P{{b}^{2+}}+{{H}_{2}}S+2C{{l}^{-}}\to P{{b}_{2}}SC{{l}_{2}}\downarrow +2{{H}^{+}}\]

This however decomposes on dilution (a) or on further addition of hydrogen sulphide (b) and black lead sulphide precipitate is formed:

    \[P{{b}_{2}}SC{{l}_{2}}\downarrow \to PbS\downarrow +PbC{{l}_{2}}\]

    \[P{{b}_{2}}SC{{l}_{2}}\downarrow +{{H}_{2}}S\to 2PbS\downarrow +2C{{l}^{-}}\]

Lead sulphide precipitate decomposes when concentrated nitric acid is added, and white, finely divided elementary sulphur is precipitated :

    \[3PbS\downarrow +8HN{{O}_{3}}\to 3P{{b}^{2+}}+6NO_{3}^{-}+3S\downarrow +2NO\uparrow +4{{H}_{2}}O\]

If the mixture is boiled, sulphur is oxidized by nitric acid to sulphate (a) which forms immediately white lead sulphate precipitate (b) with the lead ions in the solution.

    \[S\downarrow +2HN{{O}_{3}}+SO_{4}^{2-}+2{{H}^{+}}+2NO\uparrow \]

    \[P{{b}^{2+}}+SO_{4}^{2-}+2{{H}^{+}}+2NO\uparrow \]

Boiling lead sulphide with hydrogen peroxide (3%), the black precipitate turns white owing to the formation of lead sulphate:

    \[PbS\downarrow +4{{H}_{2}}{{O}_{2}}\to PbS{{O}_{4}}\downarrow +4{{H}_{2}}O\]

The great insolubility of lead sulphide in water (4.9 x 10-11 g l-1) explains why hydrogen sulphide is such a sensitive reagent for the detection of lead, and why it can be detected in the filtrate from the separation of the sparingly soluble lead chloride in dilute hydrochloric acid.

3. Ammonia solution : white precipitate of lead hydroxide

    \[P{{b}^{2+}}+2N{{H}_{3}}+2{{H}_{2}}O\to Pb{{(OH)}_{2}}\downarrow +2NH_{4}^{+}\]

The precipitate is insoluble in excess reagent.

4. Sodium hydroxide: white precipitate of lead hydroxide

    \[P{{b}^{2+}}+2O{{H}^{-}}\to Pb{{(OH)}_{2}}\downarrow \]

The precipitate dissolves in excess reagent, when tetrahydroxoplumbate(1l) ions are formed :

    \[Pb{{(OH)}_{2}}\downarrow +2O{{H}^{-}}\to {{[Pb{{(OH)}_{4}}]}^{2-}}\]

Thus, lead hydroxide has an amphoteric character

Hydrogen peroxide (a) or ammonium peroxodisulphate (b), when added to a solution of tetrahydroxoplumbate (II) forms a black precipitate of lead dioxide by oxidizing bivalent lead to the tetravelent state :

    \[{{[Pb{{(OH)}_{4}}]}^{2-}}+{{H}_{2}}{{O}_{2}}\to Pb{{O}_{2}}\downarrow +2{{H}_{2}}O+2O{{H}^{-}}\]

    \[{{[Pb{{(OH)}_{4}}]}^{2-}}+{{S}_{2}}O_{8}^{2-}\to Pb{{O}_{2}}\downarrow +2{{H}_{2}}O+2SO_{4}^{2-}\]

5. Dilute suphuric acid (or soluble sulphates): white precipitate of lead sulphate :

    \[P{{b}^{2+}}+SO_{4}^{2-}\to PbS{{O}_{4}}\downarrow \]

The precipitate is insoluble in excess reagent. Hot, concentrated sulphuric acid dissolves the precipitate owing to formation of lead hydrogen sulphate:

    \[PbS{{O}_{4}}\downarrow +{{H}_{2}}S{{O}_{4}}\to P{{b}^{2+}}+2HSO_{4}^{-}\]

Solubility is much lower in the presence of ethanol

Lead sulphate precipitate is soluble in more concentrated solutions of ammonium acetate (l0M) (a) or ammonium tartrate (6M) (b) in the presence of ammonia, when tetraacetatephumbate(I) and ditartratoplumbate(II) ions are formed:

    \[PbS{{O}_{4}}\downarrow +\,4C{{H}_{3}}CO{{O}^{-}}\to {{[Pb{{(C{{H}_{3}}COO)}_{4}}]}^{2-}}+\,\,SO_{4}^{2-}\]


    \[PbS{{O}_{4}}\downarrow +\,2{{C}_{4}}{{H}_{4}}O_{6}^{2-}\to [Pb{{({{C}_{4}}{{H}_{4}}{{O}_{6}}]}^{2-}}+\,\,SO_{4}^{2-}\]

The stabilities of these complexes are not very great; chromate ions, for example, can precipitate lead chromate from their solution.

When boiled with sodium carbonate the lead sulphate is transformed into lead carbonate in a precipitate-exchange reaction:

    \[PbS{{O}_{4}}\downarrow +CO_{3}^{2-}\to PbC{{O}_{3}}\downarrow +SO_{4}^{2-}\]

By washing the precipitate by decantation with hot water, sulphate ions can be removed and the precipitate will dissolve in dilute nitric acid

    \[PbC{{O}_{3}}\downarrow +2{{H}^{+}}\to P{{b}^{2+}}+{{H}_{2}}O+C{{O}_{2}}\uparrow \]

6. Potassium chromate in neutral acetic acid or ammonia solution : yellow precipitate of lead chromate

    \[P{{b}^{2+}}+CrCO_{4}^{2-}\to PbCr{{O}_{4}}\downarrow \]

Nitric acid (a) or sodium hydroxide (b) dissolve the precipitate:


    \[2PbCr{{O}_{4}}\downarrow +2{{H}^{+}}\rightleftarrows 2P{{b}^{2+}}+C{{r}_{2}}O_{7}^{2-}+2{{H}_{2}}O\]


    \[PbCr{{O}_{4}}\downarrow +4O{{H}^{-}}\rightleftarrows {{[Pb{{(OH)}_{4}}]}^{2+}}+CrO_{4}^{2-}\]

Both reactions are reversible: by buffering the solution with ammonia: acetic acid respectively, lead chromate precipitates again.

7. Potassium iodide : yellow precipitate of lead iodide

    \[P{{b}^{2+}}+2{{I}^{-}}\to Pb{{I}_{2}}\downarrow \]

The precipitate is moderately soluble in boiling water to yield a colourless solution from which it separates on cooling in golden yellow plates.

An excess of a more concentrated (6M) solution of the reagent dissolves the precipitate and tetraiodoplumbat(II) ions are formed:

    \[Pb{{l}_{2}}\downarrow +2{{I}^{-}}\rightleftarrows {{[Pb{{I}_{4}}]}^{2-}}\]

The reaction is reversible; on diluting with water the precipitate reappears.

8. Sodim sulphite in neutral solution: white precipitate of lead sulphite

    \[P{{b}^{2+}}+SO_{3}^{2-}\to PbS{{O}_{3}}\downarrow \]

The precipitate is less soluble than lead sulphate, though it can be dissolved by both dilute nitric acid (a) and sodium hydroxide (b).


    \[PbS{{O}_{3}}\downarrow +2{{H}^{+}}\to P{{b}^{2+}}+{{H}_{2}}O+S{{O}_{2}}\uparrow \]


    \[PbS{{O}_{3}}\downarrow +4O{{H}^{-}}\to {{[Pb{{(OH)}_{4}}]}^{2-}}+SO_{3}^{2-}\]

9. Sodium carbonate: white precipitate of a mixture of lead carbonate and lead hydroxide

    \[2P{{b}^{2+}}+2CO_{3}^{2-}+{{H}_{2}}O\to Pb{{(OH)}_{2}}\downarrow +PbC{{O}_{3}}+C{{O}_{2}}\uparrow \]

On boiling, no visible change takes place difference from mercury (I) and silver (I) ions. The precipitate dissolves in dilute nitric acid and even in acetic acid and CO2 gas is liberated:

    \[Pb{{(OH)}_{2}}\downarrow +PbC{{O}_{3}}\downarrow +4{{H}^{+}}\to 2P{{b}^{2+}}+3{{H}_{2}}O+C{{O}_{2}}\uparrow \]

10. Disodium hydrogen phosphate: white precipitate of lead phosphate

    \[3P{{b}^{2+}}+2HPO_{4}^{2-}\rightleftarrows \,\,P{{b}_{3}}{{(P{{O}_{4}})}_{2}}\downarrow +2{{H}^{+}}\]

The reaction is reversible; strong acids (nitric acid) dissolve the precipitate.

The precipitate is also soluble in sodium hydroxide.

11. Potassium cyanide (POISON): white precipitate of lead cyanide

    \[P{{b}^{2+}}+2C{{N}^{-}}\to Pb{{(CN)}_{2}}\downarrow \]

Which is insoluble in the excess of the reagent. This reaction can be used to distinguish lead (II) ions from mercury (I) and silver (I), which react in different ways.

Elements Atomic Size
Elements General Physical Properties : Atomic Size | Melting point | Boiling point

Elements General Physical Properties: Atomic Size | Melting point | Boiling point

 

 

 

 

1 H
53
He
31
2 Li
167
Be
112
B
87
C
67
N
56
O
48
F
42
Ne
38
3 Na
190
Mg
145
Al
118
Si
111
P
98
S
88
Cl
79
Ar
71
4 K
243
Ca
194
Sc
184
Ti
176
V
171
Cr
166
Mn
161
Fe
156
Co
152
Ni
149
Cu
145
Zn
142
Ga
136
Ge
125
As
114
Se
103
Br
94
Kr
88
5 Rb
265
Sr
219
Y
212
Zr
206
Nb
198
Mo
190
Tc
183
Ru
178
Rh
173
Pd
169
Ag
165
Cd
161
In
156
Sn
145
Sb
133
Te
123
I
115
Xe
108
6 Cs
298
Ba
253
*
 
Hf
208
Ta
200
W
193
Re
188
Os
185
Ir
180
Pt
177
Au
174
Hg
171
Tl
156
Pb
154
Bi
143
Po
135
At
127
Rn
120
7 Fr
 
Ra
 
**
 
Rf
 
Db
 
Sg
 
Bh
 
Hs
 
Mt
 
Ds
 
Rg
 
Cn
 
Nh
 
Fl
 
Mc
 
Lv
 
Ts
 
Og
 
Lanthanides *
 
La
226
Ce
210
Pr
247
Nd
206
Pm
205
Sm
238
Eu
231
Gd
233
Tb
225
Dy
228
Ho
226
Er
226
Tm
222
Yb
222
Lu
217
Actinides **
 
Ac
 
Th
 
Pa
 
U
 
Np
 
Pu
 
Am
 
Cm
 
Bk
 
Cf
 
Es
 
Fm
 
Md
 
No
 
Lr
 

 

Atomic

No.

  Atomic
Weight
Name Sym. M.P.
(°C)
 B.P.
(°C)
Density*
(g/cm3)
Earth
crust (%)*
Atomic  Radius in pm Electron
configuration
Ionization
energy (eV)
1 1.008 Hydrogen H -259 -253 0.09 0.14 1s1 13.60
2 4.003 Helium He -272 -269 0.18 1s2 24.59
3 6.941 Lithium Li 180 1,347 0.53 [He] 2s1 5.39
4 9.012 Beryllium Be 1,278 2,970 1.85 [He] 2s2 9.32
5 10.811 Boron B 2,300 2,550 2.34 [He] 2s2 2p1 8.30
6 12.011 Carbon C 3,500 4,827 2.26 0.09 [He] 2s2 2p2 11.26
7 14.007 Nitrogen N -210 -196 1.25 [He] 2s2 2p3 14.53
8 15.999 Oxygen O -218 -183 1.43 46.71 [He] 2s2 2p4 13.62
9 18.998 Fluorine F -220 -188 1.70 0.03 [He] 2s2 2p5 17.42
10 20.180 Neon Ne -249 -246 0.90 [He] 2s2 2p6 21.56
11 22.990 Sodium Na 98 883 0.97 2.75 [Ne] 3s1 5.14
12 24.305 Magnesium Mg 639 1,090 1.74 2.08 [Ne] 3s2 7.65
13 26.982 Aluminum Al 660 2,467 2.70 8.07 [Ne] 3s2 3p1 5.99
14 28.086 Silicon Si 1,410 2,355 2.33 27.69 [Ne] 3s2 3p2 8.15
15 30.974 Phosphorus P 44 280 1.82 0.13 [Ne] 3s2 3p3 10.49
16 32.065 Sulfur S 113 445 2.07 0.05 [Ne] 3s2 3p4 10.36
17 35.453 Chlorine Cl -101 -35 3.21 0.05 [Ne] 3s2 3p5 12.97
18 39.948 Argon Ar -189 -186 1.78 [Ne] 3s2 3p6 15.76
19 39.098 Potassium K 64 774 0.86 2.58 [Ar] 4s1 4.34
20 40.078 Calcium Ca 839 1,484 1.55 3.65 [Ar] 4s2 6.11
21 44.956 Scandium Sc 1,539 2,832 2.99 [Ar] 3d1 4s2 6.56
22 47.867 Titanium Ti 1,660 3,287 4.54 0.62 [Ar] 3d2 4s2 6.83
23 50.942 Vanadium V 1,890 3,380 6.11 [Ar] 3d3 4s2 6.75
24 51.996 Chromium Cr 1,857 2,672 7.19 0.04 [Ar] 3d5 4s1 6.77
25 54.938 Manganese Mn 1,245 1,962 7.43 0.09 [Ar] 3d5 4s2 7.43
26 55.845 Iron Fe 1,535 2,750 7.87 5.05 [Ar] 3d6 4s2 7.90
27 58.933 Cobalt Co 1,495 2,870 8.90 [Ar] 3d7 4s2 7.88
28 58.693 Nickel Ni 1,453 2,732 8.90 0.02 [Ar] 3d8 4s2 7.64
29 63.546 Copper Cu 1,083 2,567 8.96 [Ar] 3d10 4s1 7.73
30 65.390 Zinc Zn 420 907 7.13 [Ar] 3d10 4s2 9.39
31 69.723 Gallium Ga 30 2,403 5.91 [Ar] 3d10 4s2 4p1 6.00
32 72.640 Germanium Ge 937 2,830 5.32 [Ar] 3d10 4s2 4p2 7.90
33 74.922 Arsenic As 81 613 5.72 [Ar] 3d10 4s2 4p3 9.79
34 78.960 Selenium Se 217 685 4.79 [Ar] 3d10 4s2 4p4 9.75
35 79.904 Bromine Br -7 59 3.12 [Ar] 3d10 4s2 4p5 11.81
36 83.800 Krypton Kr -157 -153 3.75 [Ar] 3d10 4s2 4p6 14.00
37 85.468 Rubidium Rb 39 688 1.63 [Kr] 5s1 4.18
38 87.620 Strontium Sr 769 1,384 2.54 [Kr] 5s2 5.69
39 88.906 Yttrium Y 1,523 3,337 4.47 [Kr] 4d1 5s2 6.22
40 91.224 Zirconium Zr 1,852 4,377 6.51 0.03 [Kr] 4d2 5s2 6.63
41 92.906 Niobium Nb 2,468 4,927 8.57 [Kr] 4d4 5s1 6.76
42 95.940 Molybdenum Mo 2,617 4,612 10.22 [Kr] 4d5 5s1 7.09
43 * 98.000 Technetium Tc 2,200 4,877 11.50 [Kr] 4d5 5s2 7.28
44 101.070 Ruthenium Ru 2,250 3,900 12.37 [Kr] 4d7 5s1 7.36
45 102.906 Rhodium Rh 1,966 3,727 12.41 [Kr] 4d8 5s1 7.46
46 106.420 Palladium Pd 1,552 2,927 12.02 [Kr] 4d10 8.34
47 107.868 Silver Ag 962 2,212 10.50 [Kr] 4d10 5s1 7.58
48 112.411 Cadmium Cd 321 765 8.65 [Kr] 4d10 5s2 8.99
49 114.818 Indium In 157 2,000 7.31 [Kr] 4d10 5s2 5p1 5.79
50 118.710 Tin Sn 232 2,270 7.31 [Kr] 4d10 5s2 5p2 7.34
51 121.760 Antimony Sb 630 1,750 6.68 [Kr] 4d10 5s2 5p3 8.61
52 127.600 Tellurium Te 449 990 6.24 [Kr] 4d10 5s2 5p4 9.01
53 126.905 Iodine I 114 184 4.93 [Kr] 4d10 5s2 5p5 10.45
54 131.293 Xenon Xe -112 -108 5.90 [Kr] 4d10 5s2 5p6 12.13
55 132.906 Cesium Cs 29 678 1.87 [Xe] 6s1 3.89
56 137.327 Barium Ba 725 1,140 3.59 0.05 [Xe] 6s2 5.21
57 138.906 Lanthanum La 920 3,469 6.15 [Xe] 5d1 6s2 5.58
58 140.116 Cerium Ce 795 3,257 6.77 [Xe] 4f1 5d1 6s2 5.54
59 140.908 Praseodymium Pr 935 3,127 6.77 [Xe] 4f3 6s2 5.47
60 144.240 Neodymium Nd 1,010 3,127 7.01 [Xe] 4f4 6s2 5.53
61 * 145.000 Promethium Pm 1,100 3,000 7.30 [Xe] 4f5 6s2 5.58
62 150.360 Samarium Sm 1,072 1,900 7.52 [Xe] 4f6 6s2 5.64
63 151.964 Europium Eu 822 1,597 5.24 [Xe] 4f7 6s2 5.67
64 157.250 Gadolinium Gd 1,311 3,233 7.90 [Xe] 4f7 5d1 6s2 6.15
65 158.925 Terbium Tb 1,360 3,041 8.23 [Xe] 4f9 6s2 5.86
66 162.500 Dysprosium Dy 1,412 2,562 8.55 [Xe] 4f10 6s2 5.94
67 164.930 Holmium Ho 1,470 2,720 8.80 [Xe] 4f11 6s2 6.02
68 167.259 Erbium Er 1,522 2,510 9.07 [Xe] 4f12 6s2 6.11
69 168.934 Thulium Tm 1,545 1,727 9.32 [Xe] 4f13 6s2 6.18
70 173.040 Ytterbium Yb 824 1,466 6.90 [Xe] 4f14 6s2 6.25
71 174.967 Lutetium Lu 1,656 3,315 9.84 [Xe] 4f14 5d1 6s2 5.43
72 178.490 Hafnium Hf 2,150 5,400 13.31 [Xe] 4f14 5d2 6s2 6.83
73 180.948 Tantalum Ta 2,996 5,425 16.65 [Xe] 4f14 5d3 6s2 7.55
74 183.840 Tungsten W 3,410 5,660 19.35 [Xe] 4f14 5d4 6s2 7.86
75 186.207 Rhenium Re 3,180 5,627 21.04 [Xe] 4f14 5d5 6s2 7.83
76 190.230 Osmium Os 3,045 5,027 22.60 [Xe] 4f14 5d6 6s2 8.44
77 192.217 Iridium Ir 2,410 4,527 22.40 [Xe] 4f14 5d7 6s2 8.97
78 195.078 Platinum Pt 1,772 3,827 21.45 [Xe] 4f14 5d9 6s1 8.96
79 196.967 Gold Au 1,064 2,807 19.32 [Xe] 4f14 5d10 6s1 9.23
80 200.590 Mercury Hg -39 357 13.55 [Xe] 4f14 5d10 6s2 10.44
81 204.383 Thallium Tl 303 1,457 11.85 [Xe] 4f14 5d10 6s2 6p1 6.11
82 207.200 Lead Pb 327 1,740 11.35 [Xe] 4f14 5d10 6s2 6p2 7.42
83 208.980 Bismuth Bi 271 1,560 9.75 [Xe] 4f14 5d10 6s2 6p3 7.29
84 * 209.000 Polonium Po 254 962 9.30 [Xe] 4f14 5d10 6s2 6p4 8.42
85 * 210.000 Astatine At 302 337 0.00 [Xe] 4f14 5d10 6s2 6p5 9.30
86 * 222.000 Radon Rn -71 -62 9.73 [Xe] 4f14 5d10 6s2 6p6 10.75
87 * 223.000 Francium Fr 27 677 0.00 [Rn] 7s1 4.07
88 * 226.000 Radium Ra 700 1,737 5.50 [Rn] 7s2 5.28
89 * 227.000 Actinium Ac 1,050 3,200 10.07 [Rn] 6d1 7s2 5.17
90 232.038 Thorium Th 1,750 4,790 11.72 [Rn] 6d2 7s2 6.31
91 231.036 Protactinium Pa 1,568 0 15.40 [Rn] 5f2 6d1 7s2 5.89
92 238.029 Uranium U 1,132 3,818 18.95 [Rn] 5f3 6d1 7s2 6.19
93 * 237.000 Neptunium Np 640 3,902 20.20 [Rn] 5f4 6d1 7s2 6.27
94 * 244.000 Plutonium Pu 640 3,235 19.84 [Rn] 5f6 7s2 6.03
95 * 243.000 Americium Am 994 2,607 13.67 [Rn] 5f7 7s2 5.97
96 * 247.000 Curium Cm 1,340 0 13.50 5.99
97 * 247.000 Berkelium Bk 986 0 14.78 6.20
98 * 251.000 Californium Cf 900 0 15.10 6.28
99 * 252.000 Einsteinium Es 860 0 0.00 6.42
100 * 257.000 Fermium Fm 1,527 0 0.00 6.50
101 * 258.000 Mendelevium Md 0 0 0.00 6.58
102 * 259.000 Nobelium No 827 0 0.00 6.65
103 * 262.000 Lawrencium Lr 1,627 0 0.00 4.90
104 * 261.000 Rutherfordium Rf 0 0 0.00 0.00
105 * 262.000 Dubnium Db 0 0 0.00 0.00
106 * 266.000 Seaborgium Sg 0 0 0.00 0.00
107 * 264.000 Bohrium Bh 0 0 0.00 0.00
108 * 277.000 Hassium Hs 0 0 0.00 0.00
109 * 268.000 Meitnerium Mt 0 0 0.00 0.00
No.   Atomic
Weight
Name Sym. M.P.
(°C)
 B.P.
(°C)
Density*
(g/cm3)
Earth
crust (%)*
  Electron
configuration
Ionization
energy (e
PKA VALUES OF ACIDS
Ka and pKa Values of Acids , Phenols , Alcohols, Amines

Functional Group : Alcohols  

IUPAC Name Common Name Molecular Formula Ka pKa Melting Point (0C) Boiling Point (0C) Density

at

20 0C  gm/mL

1-alkanol Methanol Methyl alcohol 15.5 -98 65 0.791
1-alkanol Ethanol Ethyl alcohol 15.5 -114 78 0.789
1-alkanol 1-propanol Propyl alcohol 16.1 -124 97 0.804
1-alkanol 1-butanol Butyl alcohol 16.1 -89 118 0.810
2-alkanol 2-propanol Isopropyl alcohol, isopropanol 17.2 -88 82 0.785
2-alkanol 2-butanol sec-butyl alcohol 17.6 -88 99 0.806
2-alkanol 2-pentanol sec-amyl alcohol 17.8 -73 119 0.809
3-alkanol 3-pentanol Diethyl carbinol 18.2 -70 123 0.820
Functional Group :    Phenols 
Phenols Phenol Hydroxybenzene 9.98 41 181 1.132*
Phenols-alkanol 1,2-Benzenediol Catechol, pyrocatechol 9.45 102 245 1.344*
Phenols 1,3-Benzenediol Resorcinol 9.2 110 277 1.278
Phenols 1,4-Benzenediol Hydroquinone, 1,4-Dihydroxybenzene 10.9 171 285
Phenols 1,3,5-Benzenetriol Phloroglucinol, 1,3,5-trihydroxybenzene 8.45 216 s 1.460*
Phenols 2-Methylphenol o-cresol, 2-methylhydroxybenzene 10.29 31 191 1.135*
Phenols 3-Methylphenol m-cresol, 3-methylhydroxybenzene 10.09 12 202 1.03*
Phenols 4-Methylphenol p-cresol, 4-methylhydroxybenzene 10.26 35 202 1.154*
Phenols 2-Methoxyphenol Guaiacol 9.98 28 204 1.1287*
Phenols 3-Methoxyphenol Resorcinol, monomethyl ether 9.65 -18 244 1.145*
Phenols 4-Methoxyphenol 10.21 55 253
Phenols 4-Methyl-1,2-benzenediol 4-Methylcatechol 9.55 68 251 1.129*
Phenols 2-Ethylphenol 2-Ethylhydroxybenzene 10.2 -3 205 1.015*
Phenols 3-Ethylphenol 3- Ethylhydroxybenzene 9.9 -4 218 1.008*
Phenols 4-Ethylphenol 4-Ethylhydroxybenzene 10.00 45 218 1.05*
Phenols 2-Propylphenol o-propylphenol 10.47 7 223 1.015
Phenols 4-Propylphenol p-propylphenol 10.34 22 232 1.009*
Functional Group :  Carboxylic Acids
IUPAC Name Common Name Molecular Formula Ka pKa Melting Point (0C) Boiling Point (0C) Density

at

20 0C  gm/mL

Carboxylic acid Formic acid Methanoic acid 3.74 8 101 1.220
Carboxylic acid Acetic acid Ethanoic acid 4.76 17 118 1.0446*
Carboxylic acid Butanoic acid Butyric acid 4.82 -5 164 0.9528*
Carboxylic acid Pentanoic acid Valeric acid 4.86 -34 186 0.9339*
Carboxylic acid Propanoic acid Propionic acid 4.87 -21 142 0.9882*
Carboxylic acid Hexanoic acid Caproic acid 4.87 -4 202 0.9212*
Carboxylic acid Heptanoic acid Enanthic acid 4.89 -7 222 0.9124*
Carboxylic acid Octanoic acid Caprylic acid 4.89 17 240 0.907*
Carboxylic acid Nonanoic acid Pelargonic acid 4.96 12 256 0.905
Branched Carboxylic acid 2-Methylpropanoic acid Isobutyric acid 4.84 -46 155 0.943*
Branched Carboxylic acid 2,2-Dimethylpropanoic acid Trimethylacetic acid 4.78 35 164 0.9052*
Branched Carboxylic acid 2-Methylbutanoic acid 4.80 -70 176 0.934
Branched Carboxylic acid 3-Methylbutanoic acid Isovaleric acid 4.77 -30 176` 0.925
Branched Carboxylic acid 4-Methylpentanoic acid 4-Methylvaleric acid, Isocaproic acid 4.84 -33 200 0.923
Branched Carboxylic acid 2-Propylpentanoic acid Valproic acid 4.60 223 0.904*
Hydroxy acid Hydroxyethanoic acid Glycolic acid 3.88 80 d
Phenyl carboxylic acid Phenylethanoic acid α-Tolylic acid, Benzeneacetic 4.31 77 266 1.081*
Phenylcarboxylic acid 2-Phenylbutyric acid a-Ethyl-a-toluic acid 4.66
Benzoic acid Benzoic acid Benzenecarboxylic acid 4.20 122 249 1.2663*
Benzoic acid 2-Methyl-benzoic acid o-Toluic acid 3.91 107 258 1.0624*
Benzoic acid 3-Methyl-benzoic acid m-Toluic acid 4.25 111 1.0544*
Benzoic acid 4-Methyl-benzoic acid p-Toluic acid 4.37 182 275
Benzoic acid 2-Phenylbenzoic acid 3.46 112 344
Cinnamic acid trans-m-Methylcinnamic acid 4.44 115
Cinnamic acid trans-o-Methylcinnamic acid 4.50 175
Cinnamic acid trans-p-Methylcinnamic acid 4.56 199
Hydroxy acid 2-Hydroxy-benzoic acid Salicylic acid 2.97 159 1.443
Hydroxy acid 3-Hydroxy-benzoic acid 4.80 202
Hydroxy acid 4-hydroxy-benzoic acid 4.58 215
Dioic acid Ethanedioic acid Oxalic acid 1.25 d 190 s 157 1.93*
Dioic acid Propanedioic acid 1.59*Malonic acid 2.85 136 d 140 1.6193*
Dioic acid cis-Butenedioic acid Maleic 1.92 139 d 1.59*
Dioic acid trans-Butenedioic acid Fumaric 3.02 s 300 1.635
Dioic acid 2-Methylpropanedioic acid Methylmalonic acid 3.07 129 1.455
Dioic acid Butanedioic acid Succinic acid 4.21 185 d 235 1.572*
Dioic acid Pentanedioic acid Glutaric acid 3.22 98 273 1.4293*
Dioic acid 3-Methylpentanedioic acid 3-Methylglutaric acid 4.24 83
Dioic acid Hexanedioic acid Adipic acid 4.34 152 338
Dioic acid Heptanedioic acid Pimelic acid 4.71 104 357 d?
Dioic acid 1,2-Benzenedicarboxylic acid o-Phthalic 4.42 d 210
Dioic acid Octanedioic acid Suberic acid 4.52 142 230 d?
Dioic acid Nonanedioic acid Azelaic acid 4.53 110 336 d?
Dioic acid Decanedioic acid Sebacic acid 4.59 131 374 d? 1.2705
Dioic acid Undecanedioic acid 1,9-Nonanedicarboxylic acid 4.65 109
Dioic acid Dodecanedioic acid Decane-1,10-dicarboxylic acid 4.65 128 417 d?
Dioic acid Tridecanedioic acid Brassylic acid 4.65 113
Chlorocarboxylic acid Chloroacetic acid CH2ClCO2H 2.87 62 189 1.40435*
Chlorocarboxylic acid Dichloroacetic acid CHCl2CO2H 1.35 12 193 1.564
Chlorocarboxylic acid Trichloroacetic acid CCl3CO2H 0.66 59 198 1.612*
Chlorocarboxylic acid 2-Chloropropanoic acid C2H4ClCO2H 2.83 185 1.26
Chlorocarboxylic acid 3-Chloropropanoic acid C2H4ClCO2H 3.98 204 d
Chlorocarboxylic acid 2-Chlorobutanoic acid C3H6ClCO2H 2.86 1.18
Chlorocarboxylic acid 3-Chlorobutanoic acid C3H6ClCO2H 4.05 16 1.19
Chlorocarboxylic acid 4-Chlorobutanoic acid C3H6ClCO2H 4.52 16 1.22
Chlorocarboxylic acid 2-Chlorobenzoic acid C6H4ClCO2H 2.90 140 274 1.5445*
Chlorocarboxylic acid 3-Chlorobenzoic acid C6H4ClCO2H 3.84 154 283 1.4965*
Chlorocarboxylic acid 4-Chlorobenzoic acid C6H4ClCO2H 4.00 240
Bromocarboxylic acid Bromoacetic acid CH2BrCO2H 2.90 50 208 1.9335
Bromocarboxylic acid 3-Bromopropanoic acid C2H4BrCO2H 4.00 63 1.48
Bromocarboxylic acid 2-Bromobenzoic acid C6H4BrCO2H 2.85 149 295 1.929
Bromocarboxylic acid 3-Bromobenzoic acid C6H4BrCO2H 3.81 157 285 1.845
Bromocarboxylic acid 4-Bromobenzoic acid C6H4BrCO2H 3.96 254 1.894
Fluorocarboxylic acid Fluoroacetic acid CH2FCO2H 2.59 35 168 1.3682*
Fluorocarboxylic acid Trifluoroacetic acid CF3CO2H 0.52 -15 72 1.5351
Fluorocarboxylic acid 2-Fluorobenzoic acid C6H4FCO2H 3.27 124 1.46
Fluorocarboxylic acid 3-Fluorobenzoic acid C6H4FCO2H 3.86 124 0.47
Fluorocarboxylic acid 4-Fluorobenzoic acid C6H4FCO2H 4.15 184 1.48
Fluorocarboxylic acid Pentafluorobenzoic acid C6F5CO2H 1.75 103 220

 

Entropy Calculation
Entropy Calculation for Ideal Gas

Reversible Change: For reversible expansion or Compression-

    \[{{q}_{\text{rev}}}=-w=RT\ln \frac{{{V}_{2}}}{{{V}_{1}}}\]

[using ΔU = Q + w]

    \[\left( w=-RT\ln \frac{{{V}_{2}}}{{{V}_{1}}} \right)\]

    \[\Delta {{S}_{\text{system}}}=\frac{{{q}_{\text{rev}}}}{T}=R\ln \left( \frac{{{V}_{2}}}{{{V}_{1}}} \right)\]

qrev  is heat exchanged reversible between the system and the surrounding at temp T.

    \[\Delta {{S}_{\text{surrounding}}}=\frac{-{{q}_{\text{rev}}}}{T}\]

    \[\Delta {{S}_{\text{total}}}=\Delta {{S}_{\text{sys}}}+\Delta {{S}_{\text{surr}}}=0\]

Irreversible Change:

Case I : Free expansion: The gas expands into a vacuum for this process.

w = 0,   q = 0

Since entropy is a state function, the entropy change of a system in going from volume V1 to V2 by any path will same as that of a reversible change.

Therefore, 

    \[\Delta {{S}_{\text{sys}}}=R\ln \frac{{{V}_{2}}}{{{V}_{1}}}\]

It is because from surrounding no heat is supplied.

    \[\Delta {{S}_{\text{surr}\text{.}}}=0\]

    \[\Delta {{S}_{\text{total}}}=\Delta {{S}_{\text{sys}}}+\Delta {{S}_{\text{surr}}}=R\ln \frac{{{V}_{2}}}{{{V}_{1}}}+0=R\ln \frac{{{V}_{2}}}{{{V}_{1}}}\]

Intermediate Expansion:

    \[\Delta {{S}_{\text{sys}}}=R\ln \frac{{{V}_{2}}}{{{V}_{1}}}=\frac{{{q}_{\text{rev}}}}{T}\]

 

Note: The entropy change of the system ΔSsys  will be same in all three process as it is state function.

                       qrev  = Heat absorbed by the system if the process had been carried out reversibly.

            For irreversible expansion, work is done against constant pressure.

                  qrev = -w = p external × V2 – V1

                       

    \[\Delta {{S}_{\text{surr}\text{.}}}=\frac{-{{q}_{\text{irr}}}}{T}=-\frac{{{p}_{\text{ext}}}({{V}_{2}}-{{V}_{1}})}{T}\]

            Since magnitude of work done in the intermediate expansion is smaller than that involved in reversible expansion,

            Therefore

    \[{{q}_{\text{irr}}}<\text{ }{{q}_{\text{rev}}}\]

    \[\Delta {{S}_{\text{Total}}}=\Delta {{S}_{\text{sys}}}+\Delta {{S}_{\text{surr}\text{.}}}=\frac{{{q}_{\text{rev}}}}{T}-\frac{{{q}_{\text{irr}}}}{T}=+ve\]

Entropy Change in Adiabatic Expansion or Compression of an Ideal Gas

Entropy Change of System: Since in adiabatic processes q = 0, therefore

                   

    \[\Delta {{S}_{\text{surr}}}=0\]

   

Since in an adiabatic process, both temperature an volume (or pressure) change, the expression for the molar entropy change as given by 

                 

    \[\Delta {{S}_{\text{sys}}}={{C}_{V.\text{m}}}\ln \frac{{{T}_{2}}}{{{T}_{1}}}+R\ln \frac{{{V}_{2}}}{{{V}_{1}}}\]

                 

    \[\Delta {{S}_{\text{sys}}}={{C}_{p.\text{m}}}\ln \frac{{{T}_{2}}}{{{T}_{1}}}+R\ln \frac{{{p}_{1}}}{{{p}_{2}}}\]

Now, we proceed to evaluate the change in total entropy for the following categories.

Reversible change: In this case

 

    \[\Delta {{S}_{\text{sys}}}=0\]

                      

Since for the adiabatic reversible process,

                   

    \[{{C}_{V.\text{m}}}\ln \frac{{{T}_{2}}}{{{T}_{1}}}=-R\ln \frac{{{V}_{2}}}{{{V}_{1}}}\]

  

            And 

    \[{{C}_{V.\text{p}}}\ln \frac{{{T}_{2}}}{{{T}_{1}}}=-R\ln \frac{{{p}_{2}}}{{{p}_{1}}}\]

  

            Thus

    \[\Delta {{S}_{\text{Total}}}=\Delta {{S}_{\text{sys}}}+\Delta {{S}_{\text{surr}}}=0+0=0\]

      

In the present case of expansion (or compression), the increase (or decrease) in entropy due to the volume change just compensate the decrease (or increase) in entropy due to the fall (or rise) in temperature.

 Irreversible change: In this case,

                             

    \[\Delta {{S}_{\text{sys}}}=R\ln \frac{{{V}_{2}}}{{{V}_{1}}}+{{C}_{V.\text{m}}}\ln \frac{T_{2}^{'}}{{{T}_{1}}}\]

     

Where T2 is the temperature of the system in the final state. 

                         

    \[\Delta {{S}_{\text{sys}}}=-{{C}_{V.\text{m}}}\ln \frac{{{T}_{2}}}{{{T}_{1}}}+{{C}_{V.\text{m}}}\ln \frac{T_{2}^{'}}{{{T}_{1}}}\]

    

Where T2 is the temperature, if the process was reversible,

            Since we know that

                   

    \[{{w}_{\text{irr}}}>{{w}_{\text{rev}}}\]

           (including the sign of w)

            And moreover for adiabatic change

                 ΔU = w     

            It follows that

                 

    \[\Delta {{U}_{\text{irr}}}>\Delta {{U}_{\text{rev}}}\]

      

            Or       

    \[{{C}_{V.\text{m}}}\left( T_{2}^{'}-{{T}_{1}} \right)<{{C}_{V.\text{m}}}\left( {{T}_{2}}-{{T}_{1}} \right)\]

            Remembering that T2 < T1 in the expansion process and  T2 < T1  in the compression process, we have T2 < T1                                         

            That is, the decrease in temperature during the irreversible expansion will be lesser and the increase in temperature during the irreversible compression will be larger that the corresponding change in the reversible process. This, we have

                       

    \[{{C}_{V.\text{m}}}\ln \frac{T_{2}^{'}}{{{T}_{1}}}>{{C}_{V.\text{m}}}\ln \frac{{{T}_{2}}}{{{T}_{1}}}\]

            Substituting this relation ; we get

                       

    \[\Delta {{S}_{\text{sys}}}=+ve\]

            And this

    \[\Delta {{S}_{\text{total}}}=\Delta {{S}_{\text{sys}}}+\Delta {{S}_{\text{surr}}}=\text{+}\,\text{ve}\]

            In the present case of expansion (or compression), the increase (or decrease) in entropy due to the  volume change is larger (or smaller) that the decrease (or increases in entropy due to the temperature change and hence ΔSsys  is positive.

 

Discovery of Elements : Gallium

Gallium 

Eka – Aluminium Ea Gallium Ga
  • Atomic mass is about 68.
  • Element must have more melting point.
  • Density of the metal is close to 6.0
  • Atomic volume must be close to 11.5
  • Does not change in air.
  • Must decompose water on boiling
  • Forms alums but not so ready as aluminium
  • Ea2O3 must be readily reduced to matter
  • Ea is more volatile than aluminium it will be discovered by spectral analysis
 

  •  Atomic mass is 69.72
  • Melting point is 29. 75 degree centigrade.
  • Atomic volume is 11.8.
  • Oxidizes weakly upon  heating to redness.
  • Decompose water at high temperature
  • Gives alums of the formula NH4 Ga(SO4)2 12 H2O
  • Ga is readily reduced by calcination of  Ga2O3 in hydrogen inflow
  • Ga has been discovered by the spectroscopic method

Eka Aluminium properties were predicted by Mendeleev , while Ga properties are calculated using Modern Techniques .

Time of discovery of Gallium : 

The time of discovery of gallium is known to an hour. “One Friday of August 27, 1875, between 3 p.m. and 4 p.m. I discovered some signs that there can be a new simple body in the by-product of chemical analysis of zinc blende from the Pierfitt mine in the Argele valley (Pyrenees).” With these words P. E. Lecoq de Boisbaudran began his report to the Paris Academy of Sciences.

He described some of the new element’s properties and noted that its presence in the ores was ascertained by spectra; analysis just as predicted by Mendeleev five years before. Boisbaudran extracted an extremely small amount of the substance and, therefore, could not study its properties properly.

On August 29, 1875 , Boisbaudran suggested to name the element “gallium” after Gaul, the ancient name of France. The scientist continued the investigation of the new element and obtained additional information which he included into his report to the Paris Academy and then sent it to be academic journal. In the middle of November the journal with the article reached Petersburg where Mendeleev was impatiently waiting for it. There is every reason to believe that Mendeleev had already learnt about gallium though at second hand. Two weeks earlier the Russian Chemical Society had received a report from Paris signed by P. de Clermont. It recounted the discovery of gallium and contained a brief description of its properties. However, it was much more important for Mendeleev to read what the discoverer himself had written. Mendeleev’s reaction was prompt; on November 16, he delivered a report to the Russian Physical Society. According to the minutes of the session, Mendeleev declared that the discovered metal was, most probably, eka-aluminium. Next day he wrote an article in French entitled “Note on the discovery of Gallium”. And finally, on November 18, Mendeleev spoke about gallium at a session of the Russian Chemical Society. Such a spurt of activity is understandable: the great chemist saw an element predicted by him becoming a reality. Mendeleev believed that if further investigation confirmed the similarity of eka-aluminium properties of those of gallium, this would be an instructive demonstration of the periodic law’s usefulness.

Six days later (a surprisingly short time!) the “Note on the Discovery of Gallium” appeared in the journal of the Paris Academy of Sciences. Boisbaudran’s reaction to it is of particular interest. He continued his experiments and prepared the new results for publication. The next article by the French scientist was published on December 6. As before, he complained of the difficulties caused by the extreme scarcity of gallium, described the preparation of the metal by the electrochemical method and discussed some of its properties, and suggested that the formula of gallium oxide had to be Ga2O3.

Only at the end of the article were there a few words about Mendeleev’s note. Boisbaudran admitted that he had read it with great interest since classification of simple substances interested him for a long time. He had never known about Mendeleev’s prediction of eka-aluminium properties but it did not matter; Boisbaudran believed that his discovery of gallium was facilitated by his own laws of spectral lines of elements with similar chemical properties. In his opinion, spectral analysis played a decisive role. And not a word that Mendeleev in his prediction of eka-aluminium also underlined the prominent role of spectral analysis in the discovery of the new element. According to Boisbaudran, Mendeleev’s predictions had nothing to do with the discovery of gallium.

However, as Boisbaudran went on studying the properties of metallic gallium and its compounds, his results continued to coincide with Mendeleev’s predictions. For instance, in May 1876, the Franch scientist established that gallium was readily fusible (its melting point is 29.5oC), its appearance remained the same after storage in air, and it was slightly oxidized when heated to redness. The same properties of eka-aluminium were predicted by Mendeleev in 1870, who calculated the density of eka-aluminium to be 5.9-6.0 on the basis of the periodic system and the densities of eka-aluminium’s neighbours. Lecoq de Boisbaudranm, however, making use of his spectral laws, found that the density of eka-aluminium was 4.7 and confirmed the value experimentally. Such a difference (less than two units) might seem small to a layman but it was essential for the future of the periodic law. Up to that time only qualitative characteristics of the predicted properties had been confirmed and density was the first quantitative parameter. And it turned out to be erroneous.

There is a widely known story that Mendeleev, having received Boisbaudran’s article citing a low (4.7) density of gallium, wrote him that the gallium obtained by the French chemist was contaminated most likely by sodium used in the process of gallium preparation. Sodium has a very low density (0.98), which could substantially decrease the density of gallium. Hence, it was required to purify gallium thoroughly.

This letter has not been found either in French or in the Mendeleev’s archives. There is only indirect evidence from Mendeleev’s daughter and the eminent historian of chemistry B. Menshutkin that the letter did exist. However, that may be Mendeleev’s views became known to Boisbaudran who decided to repeat the measurements of gallium’s density. This time he took into account that Mendeleev’s calculations for the hypothetical element’s density this time he took into account that Mendeleev’s calculations for the hypothetical element’s density gave 5.9. And be obtained this value at the beginning of September, 1876. His report about this fact needs no comments.

The French scientist became firmly convinced of the extreme importance of the confirmation of Mendeleev’s predictions about the density of the new element. Sometime later Lecoq de Boisbaudran send his photo to the great Russian chemist with the inscription: “With profound respect and an ardent wish to count Mendeleev among my friends. L. de B.”

Mendeleev wrote under it: “Lecoq de Boisbaudran. Paris. Discovered eka-aluminium in 1875 and named it “gallium”, Ga=69.7.”

In autumn 1879, F. Engels became acquainted with a new detailed chemistry textbook by H. Roscoe and C. Shorlemmer. For the first time it contained the story about the prediction of eka-aluminium by Mendeleev and its discovery as gallium. In an article to be later included in his Dialectics of Nature Engels quoted the corresponding text from the book and concluded: “by means of the unconscious application of Hegel’s law of the transformation of quantity into quality, Mendeleev achieved a scientific feat which is not too bold to put on a par with that of Leverrier in calculating the orbit of the still unknown planet Neptune”.

Discovery of Americium and Curium

Americium and Curium

It is, perhaps, the only occasion in the history of transuranium elements that an element with a higher number (Z = 96) was identified earlier than its predecessor (Z = 95). In July 1944 the cyclotron of the University of California, which had already revealed to the world several synthesized elements, including plutonium, was geared to synthesize new transuranium elements. Seaborg and his coworkers bombarded a plutonium–239 target with accelerated alpha particles. One can readily reckon that as the alpha particle (the helium nucleus) has a charge of two the reaction product could be an isotope of element 96, provided that neutrons were emitted from the resulting nuclei. If the process mechanism was such that protons were emitted, rather than neutrons, then an isotope of element 95 could be synthesized. Indeed, various radioactive substances were produced in the plutonium target and it was difficult at first to identify “who was who”. Only skillful chemical analysis revealed that the mixture definitely contained the isotope 24296. To verify the discovery, the same isotope, plutonium–239, was bombarded with a high–intensity neutron beam so that the following chain of reactions took place :

    \[^{239}Pu+n{{\to }^{240}}Pu+n{{\to }^{241}}Pu{{\xrightarrow{{{\beta }^{-}}}}^{241}}95+n{{\to }^{242}}95{{\xrightarrow{{{\beta }^{-}}}}^{242}}96\]

After absorption of neutrons plutonium converted into element 95 via beta decay and this element absorbed a neutron and converted into element 96.

This final product was similar to that which the scientists had assumed to be the isotope of element 96 with a mass number of 242. The newly discovered element was named curium after the Curies. Another factor prompted this name. In the Mendeleev table element 96 was regarded as an analogue of gadolinium belonging to the rare–earth series the history of which had been started by J. Gadolin; in their turn, the Curies were the pioneers of the study of radioactivity whose development produced such amazing results. In January 1945 element 95 was extracted from plutonium bombarded with neutrons. The element was named americium in honour of America (and owing to its similarity to europium from the rare–earth series).

Though the researchers had accumulated considerable experience in syntheses the difficulties involved in producing americium and curium proved unusually great. It took a long time to distinguish definitely between americium–241 and curium–242. Both isotopes proved to be not the longest–lived ones. The longest–lived isotopes were americium–243 (a half–life of 7 950 years) and curium–247 (a half–life of 1.64 × 107 years), which were only synthesized in the fifties. The total of 11 americium isotopes and 13 curium isotopes are currently known. Here are a few more events in the history of these elements. Pure americium was extracted in 1945 and in 1951 it was prepared in a metallic form. The same year metallic curium was prepared.

The discovery of curium ends the first breakthrough period in the history of transuranium elements. The discoveries of neptunium, plutonium, americium, and curium were of great significance for science. It was for the first time that scientists artificially extended the boundaries of the periodic system. The properties of these elements proved to be quite different from those expected and chemists had to start seriously thinking how best to fit them into the periodic system.

Discovery of Astatine and Francium element

Astatine and Francium

In July 1925 the British scientist W. Friend went to Palestine but not as a pilgrim. Moreover, he was neither an archeologist nor a tourist visiting exotic lands. He was just a chemist and his luggage contained mostly ordinary empty bottles which he hoped to fill with samples of water from the Dead sea. Which has the highest concentration of dissolved salts on Earth. Fish cannot live in it and a man can swim in it without any danger of drowning–so high is the density of water in it.

The somber Biblical landscapes failed to dampen Friend’s hopes for success. His goal was to find in the water of the Dead Sea eka–iodine and eka–cesium which chemists had sought in vain. Sea water contains many dissolved salts of alkali metals and halogens and their concentration in the Dead Sea water must be exceptionally high. The higher the probability that they hide among them the unknown elements, namely the heaviest halogen and the heaviest alkali metal, even if in trace amounts.

Of course, Friend was not entirely original in choosing the direction of his search. As early as the end of the 19th century a chemist would not hesitate to answer the question where to look for eka–iodine and eka–cesium on Earth. The obvious answer was where natural compounds of alkali metals, in sea and ocean water, in various minerals, in deep well water, in some sea algae, and soon. In other words, the field of search was quite wide.

But all the attempts to find eka–iodine and eka–cesium failed and efforts of Friend were no exception. Now let us turn back to the last decades of the 19th century. When Mendeleev developed the periodic system of elements it contained many empty slots corresponding to unknown elements between bismuth and uranium. These empty slots were rapidly filled after the discovery of radioactivity. Polonium, radium, radon, actinium and finally protactinium took their places between uranium and thorium. Only eka–iodine and eka–cesium were late. This fact, however, did not particularly trouble scientists. These unknown elements had to be radioactive since there was not even a hint of doubt that radioactivity was the common feature of elements heavier than bismuth. Therefore, sooner or later radiometric methods would demonstrate the existence of elements 85 and 87.

The natural isotopes of uranium and thorium in long series of successive radioactive transformations give rise to secondary chemical elements. In the first decade of the 20th century scientists had in their disposal about forty radioactive isotopes of the elements at the end of the periodic system, that is, from bismuth to uranium. These radioelements comprised three radioactive families headed by thorium–232, uranium–235, and uranium–238. Each radioactive element sent its representatives to these families with the only exception of eka–iodine and eka–cesium. None of the three series had links that would correspond to the isotopes of element 85 or 87. This suggested an unexpected idea that eka–iodine and eka–cesium were not radioactive.But why? Nobody dared to answer this question. Under this assumption it was meaningless to look for these element in the ores of urnium and thorium which contained all the radioactive elements without exception.     

The assumption about stability of eka–iodine and eka–cesium was not confirmed. But all efforts to find isotopes of these elements in the radioactive families met with failure. But there remained one path of investigation which seemed promising. Does a given radioactive isotope have only one or two decay mechanisms? For instance, it emits both alpha and beta particles. If so the products of decay of this isotope are isotopes of two different elements and the series of radioactive transformations at the place of this isotope experiences branching. This problem was discussed for a long time and for some isotopes this effect seemed to take place.

In 1913 the British scientist A. Cranston worked with the radioelement MsTh–II (an isotope of actinium–228). This isotope emits beta particles and converts into thorium–228. But Cranston thought that he detected a very weak alpha decay, too. If that was true the product of the decay had to be the long–expected eka–cesium. Indeed, the process is described by

\[_{89}^{228}Ac{{\xrightarrow{\alpha }}^{224}}87\

But Cranston just reported his observation and did not follow the lead.

Just a year later three radiochemists from Vienna–S. Meyer, G. Hess, and F. Paneth–studied actinium–227, an isotope belonging to the family of uranium–235. They repeated their experiments and at last their sensitive instruments detected alpha particles of unknown origin. Alpha particles emitted by various isotopes have specific mean paths in air (of the order of a few centimetres). The mean path of the alpha particles in the experiments of the Austrian scientists was 3.5 cm. No known alpha–active isotope had such mean path of alpha particles. The scientists from the Vienna Radium Institute concluded that these particles were the product of alpha decay of the typically beta–active actinium–227. A product of this decay had to be an isotope of element 87.

The discovery had to be confirmed in new experiments. The Austrians were ready for this but soon the World War I started. They indeed observed alpha radiation of actinium–227 and this meant that atoms of element 87 were produced in their presence. But this fact had to be proved. It was easier to refute their conclusions. Sceptics said that the observed alpha activity was too weak and the results were probably erroneous. Others said that an isotope of the neighbouring element, protactinium, also emitted alpha particles with mean path close to 3.5 cm. Perhaps, an error was caused by an admixture of protactinium.

Elements 85 and 87 were discovered several times and given such names as dacinum and moldavium, alcalinium and helvetium, or leptinum and anglohelvetium. But all of them were mistakes. The fine–sounding names covered emptiness.

The mass numbers of all isotopes in the family of thorium–232 are divided by four. Therefore, the thorium family is sometimes referred to as the 4n family. After division by four of the mass numbers of the isotopes in the two uranium families we get a remainder of two or three. Re–spectively, the uranium–238 family is known as the (4n + 2) family and the uranium–235 family as the (4n + 3) family.

But where is the (4n + 1) family? Perhaps it is precisely in this unknown fourth series of radioactive transformations that the isotopes of eka–iodine and eka–cesium can be found. The idea was not unreasonable but not a single known radioactive isotope could fit into this hypothetical family by its mass number.

Sceptics declared, not without reason, that indeed there had been the fourth radioactive series at the early stages of Earth’s existence. But all the isotopes that comprised it including the originator of the series had too short half–lives and hence disappeared from the face of Earth long ago. The fourth radioactive tree had withered away long before mankind appeared.

In the twenties theorists attempted to reconstruct this family, to visualize its composition if it had existed. This imaginary structure had positions for the isotopes of elements 85 and 87 (but not for the radon isotopes). But this direction of search did not bring results, too. Perhaps the elusive elements did not exist at all?

But the goal was not that far. But before we start the tale about the realization of the scientists’ dreams let us turn back to the first synthesized element, namely, technetium.

Why was technetium the first? Primarily, because the choice of the target and the bombarding particles was obvious. The target was molybdenum, which could be made sufficiently pure at the time. The bombarding particles were neutrons and deutrons and accelerators were available for accelerating deutrons. This is why the discovery of technetium manifested the dawn of the age of synthesized elements. The work on promethium proved more complicated because in belonged to the rare–earth family and the main difficulties were met in determining its chemical nature.

But the task for elements 85 and 87 looked much more formidable. In their attempts to produce eka–iodine the scientists could only have one material for the target, namely, bismuth, element 83. The bombarding particles were a case of Hobson’s choice, too–only alpha particles could be used. Polonium, which precedes eka–iodine, could not be used as the material for the target. The elements with lower numbers than bismuth could not be used as targets because the scientists lacked appropriate bombarding particles to reach number 85.

Eka–cesium looked totally inaccessible for artificial synthesis. No suitable targets and bombarding particles existed in the thirties. But such is the irony of history that it was precisely element 87 that became the second after technetium reliably discovered element out of the four missing elements within the old boundaries of the periodic system. At this point in history the line of eka–iodine and eka–cesium, which had travelled parallel for such a long time, started to diverge and therefore we shall consider their discoveries separately.

Element 85 was synthesized by D. Corson, C. Mackenzie, and E. Segre who worked at Berkley (USA). The Italian physicist Segre by that time had settled in the USA and was the only one in the group who had an experience in artificial synthesis of a new element (technetium). On July 16, 1940, these scientists submitted to the prestigious physical journal Physical Review a large paper entitled “Artificial radioactive element 85”. They reported how they had bombarded a bismuth target with alpha particles accelerated in a cyclotron and obtained a radioactive product of the nuclear reaction . The product, most probably, was an isotope of eka–iodine with a half–life of 7.5 hours and a mass number of 211. Segre and his coworkers performed fine chemical experiments with the new element produced in negligible amounts and found that it was similar to iodine and exhibited weakly metallic properties.

The results seemed convincing enough. But the new element remained nameless for the time being. Further work on eka–iodine had to be delayed as the war started. It was resumed only in 1947 and the same group announced synthesis of another isotope with a mass number of 210. Its half–life was somewhat longer but still only 8.3 hours. Later it was found to be the longest–lived isotope of element 85. It was produced with a similar technique as the first isotope though the energy of the bombarding alpha particles was somewhat higher. As a result the intermediate composite nucleus (209Bi + α) emitted three rather than two neutrons and hence, the mass number of the isotope was lower by 1. Only now the new element was given the name astatine from the Greek for “unstable” (the symbol At).

But in the interval between the syntheses of the isotopes 211At and 210At a remarkable event occurred. The scientists from the Vienna Radium Institute B. Karlik and T. Bernert managed to find natural astatine. This was an extremely skillful study straining to the utmost the capacity of radiometry. The work was crowned with success and element 85 was born for the second time. As in the cases of technetium and promethium, we can name two dates in the history of astatine, namely, the year of its synthesis (1940) and the year of its discovery in nature (1943).

But when the Segre and his coworkers were preparing for irradiating a bismuth target with alpha particles the scientific community had known about the discovery of eka–cesium for more than a year. Transactions of the Paris Academy of Science published a paper headed “Element 87: AcK formed from actinium” and dated January 9, 1939. Its author was M. Perey, the assistant of the eminent radiochemist Debierne who had announced his discovery of actinium forty years earlier.

Marguerite Perey did not invent any fundamentally new methods and did not indulge in any vague and complicated speculations about possible sources of natural eka–cesium. In 1938 she came upon a paper published in 1914. The paper was signed by the Austrian chemists Meyer, Hess and Peneth. Perey attempted to prove their ideas. She obtained a carefully purified specimen of actinium–227. This isotope has a high beta–activity but sometimes it emits alpha particles, too. The mean path of such particles in air is 3.5 cm. This alpha radiation is by no means due to protactinium as the actinium specimen was sufficiently purified. Since alpha particles are emitted the eka–cesium isotope with a mass number of 223 must continuously be accumulated in the specimen. A series of experiments definitely demonstrated that, indeed, some substance with a half–life of 21 min is accumulated in the actinium specimen. Now it is the turn of chemical analysis to prove that this substance is a new element. Its properties proved to be similar to those of cesium. Perey named the new element francium in honour of her country. Only for a short period it was called actinium K (AcK) in accordance with the old nomenclature of radioelements.

The first description given by Perey to the newborn element was extremely brief: the element is formed with alpha decay of actinium –227 in the reaction

\[_{89}^{227}Ac{{\xrightarrow{\alpha }}^{223}}85\

and it is alpha–active with a half–life of 21 min. Then she spent several months studying its chemical properties and demonstrated convincingly that francium is similar to cesium in all its characteristics.

None of the natural radioactive elements had such a short half–life, even the artificially synthesized element 85 had a half–life measured in hours. There were hopes to find other natural isotopes of francium with longer half–lives. But in fact francium–223 proved to be the only francium isotope found on Earth.

The only remaining path to success was synthesis but it proved very difficult. More than ten years passed after the discovery of Perey when francium isotopes were artificially synthesized. The nuclear reaction giving rise to the francium isotope with a mass number of 212 can be written in short as

\[_{92}^{232}U(p,\,\,6p21n)\,\,_{87}^{212}Fr\

This reaction is the fission of uranium nucleus by protons accelerated to very high energies. When such a fast proton hits uranium nucleus it produces something like an explosion with ejection of a multitude of particles, namely, six protons and 21 neutrons. Of course, the reaction is not due to a blind chance but is based on careful theoretical predictions. Uranium may be replaced with thorium. The reaction product, francium–212, for some time was considered to be the longest–lived isotope (a half–life of 23 min) but later the half–life was found to be only 19 min.

Artificial synthesis of francium is much more difficult and less reliable method than extraction of francium as a product of decay of natural actinium. But natural actinium is rare. What to do? A current method is to irradiate the main isotope of radium with a mass number of 226 (its half–life is 1622 years) with fast neutrons. Radium–226 absorbs a neutron and converts into radium–227 with a half–life of about 40 min. Its decay gives rise to pure actinium–227 whose alpha decay in its turn produces francium–223. The symbols At and Fr were permanently installed in boxes 85 and 87 of the periodic table and their properties proved to be exactly the same as predicted from the table. But in comparison with their unstable mates born by nuclear physics, technetium and promethium, their position is clearly unfavourable.

According to estimates, the 20-km thickness of the Earth crust contains approximately 520 g of francium and 30 g of astatine (this is an overestimation in some respects). These amounts are of the same order as the terrestrial “resources” (quotation marks are more than suitable here) of technetium and promethium. We are probably making a mistake when we talk condescendingly about astatine and francium? Not at all. Technetium and promethium are produced in large amounts, kilograms and kilograms of them. The fact is that technetium and promethium have much longer half–lives and can therefore be accumulated in larger amounts. But accumulation of astatine and francium is just unfeasible. In fact, each time their properties have to be studied they have to be produced a new.

In the radioactive families the isotopes of astatine and francium are placed not on the principle pathways of radioactive transformations but at the side branches. Here is the branch on which natural francium is born:

         

The isotope Ac in 99 cases out of 100 emits beta particles and only in one case it undergoes alpha decay.

The situation is even less easy in the case of the branches responsible for the formation of astatine:

What may be said about these branches? The producers of natural astatine (the polonium isotopes) are by themselves extremely rare. For them alpha decay is not just predominant but practically the only radioactivity mechanism. Beta decays for them seem something like a mishap as can be clearly seen from the following data.

There is only one beta decay event per 5 000 alpha decays of polonium–218. Things are even sadder for polonium–216 (1 per 7 000) and polonium–215 (1 per 200 000). The situation speaks for itself. The amount of natural francium on Earth is larger. It is produced by the longest–lived actinium isotope 227Ac (a half–life of 21 years) and its content is, of course, much higher than that of the extremely rare polonium isotopes capable of producing astatine.

Discovery of element : Promethium

Promethium

The history of one rare–earth element is so unusual that it merits individual discussion. Promethium, as it is known now, is practically non–existent in nature (we write practically but not absolutely and the reason for that will be clear later). Event which can only be described as amazing preceded the discovery of element 61 by means of nuclear synthesis.

The work of Moseley made clear the existence of an unknown element between neodymium and samarium. But the situation proved to be not so clear and dramatic events rapidly followed in the history of element 61.

The New World was unlucky in discoveries of new elements. All the elements known by the twenties of this century (not counting the elements known from ancient times) had in fact been discovered by the European scientists. This is why the American scientific community was particularly happy to learn about the discovery of element 61 by the chemists from Chicago B. Hopkins, L. Intema, and J. Harris in 1926.

Starting from 1913 scientist from various countries had been searching intensely for the elusice rare–earth element and it seemed strange that they had not found it earlier. Indeed the elements of the first half of the rare–earth family known as the cerium elements (from lanthanum to gadolinium) had been shown by geochemists to be more abundant in nature than the yttrium elements of the second half of the family (from terbium to lutecium). But all the yttrium elements had been found while an empty box had remained in the cerium group between neodymium and samarium.

The straightforward explanation was that element 61 was not just rare but rarest element. Its abundance was assumed to be much lower than that of other rare–earth elements, and the available analytical techniques were not sensitive enough to identify its traces in the terrestrial minerals. New more sensitive methods were needed for the purpose.

The American chemists employed X–ray and optical spectral techniques to study the minerals where they hoped to find element 61. These well versed in the history of range earth elements could say that the path the Americans took was a troublesome one as spectral analysis not infrequently had acted as an evil genius of rare–earth studies despite all the benefits it had brought to them. But in the twenties the feet spectroscopy stood on were not so unsteady as a few decades earlier and the Moseley law could be used for predicting the X–ray spectra of any element.

The American chemists worked hard, analysed numerous specimens of various minerals and in april 1926 reported the discovery of element 61. But they did not extract even a grain of the new element and its existence was inferred from the X–ray and optical spectral data.

The discoverers University named the element illinium in honour of the Illinois University where they worked and the symbol Il took its place in box 61 of the periodic system but just a half–year later a new claimant of box 61 came into the limelight. It had been discovered by two Italian scientists L. Rolla and L. Fernandes who had named it florencium (Fl).  Allegedly, they had discovered element 61 two years earlier than the Americans but failed to report the discovery owing to some undisclosed reasons. They had sealed the report of their discovery into an envelope and left it for safe–keeping in the Florence Academy.

If different people obtain the same result with different means that would seem to prove that the result is genuine. Americans and Italians could be only too happy. As for the question of priority it was nothing new to science. But no one of the alleged discoverers of element 61 could imagine that their argument about periodic would soon become superfluous and both symbols, Il and Fl, would be shown to be illegal squatters in box 61 of the periodic table.

To trace the events now we have to go not further but some time back to the facts that were simply unknown at the time. The report of the discoverers of element 61 started with the words: “There had been absolutely no grounds for assuming the existence of an element between neodymium and samarium until it was demonstrated through the Mosely law”. Typical dry style of a scientific report, everything would seem to be correct. But….

The following remarkable conclusion in German (please, do not look it up in a dictionary yet) appeared in the margin of a hard–written manuscript of the element table found in the papers of certain scientist (we shall supply the name a little later): “NB. 61 ist das von mir 1902 vorhergesagte fehlende Elemente”.

The real history of element 61 should prominently feature the name we have already met on these pages. It is the Czech scientist Boguslav Brauner, Mendeleev’s friend and an eminent expert in the chemistry of rare–earth elements.

Illinium had been discovered, the discoverers accept congratulations and learn about the second, third, fourth confirmation of the discovery from the scientists of other countries. The pedigree of element 61 could be started thus: “Moseley had predicted and American chemists discovered”. But a discordant not unexpectedly sounded in November 1926 from the pages of Nature. It was none other than Brauner. He congratulated him American colleagues but voiced his disagreement with the above–cited beginning of their report. He argued that it was really not important who first discovered element 61 –American or Italians; in the twenties scientists became increasingly aware that the discovery by itself was a purely technical matter. The important issue is who predicted the new element. Was it Moseley? No, declared the Czech scientist. Who then? Of course, he himself, Boguslav Brauner…..

But nothing could be further from the truth if we thought that he was immodest. His claim was based on his vast experience of work with rare earths, on his profound understanding of the spirit of the periodic system, on his superb appreciation of slight changes of properties in the series of extremely similar rare–earth elements, and, finally, on his intuition of a dedicated researcher.

But these words of praise must be substantiated with facts. Let us turn back to 1882. The old didymium of K. Mosander is close to its death. P. Lecoq de Boisbaudran had already extracted a new element, samarium, from it. B. Brauner carefully analyses the residue and employing extremely complicated chemical procedures separates it into three fractions with different atomic masses. Owing to a number of reasons he has to discontinue his work and in 1885 K. Auer von Welsbach overtakes the Czech scientist. The old didymium is dead but praseodymium and neodymium have appeared, the first and the third fractions of Brauner. But what about the intermediate second fraction? No, its tine has not come. The chemistry of rare–earth elements is in a turmoil. The muddy stream of erroneous discoveries of new elements overflows with doubts the very periodic system. But life goes on. The chaos in rare earths gradually diminishes and the known rare–earth elements form an ordered series. Now Brauner notices that the difference between the atomic masses of neodymium and samarium is rather large; it is larger than the respective difference between any two neighbouring rare–earth elements. His brilliant knowledge of rare earths suggests to Brauner that there is a discontinuity in the variations of their properties in the part of the series between neodymium and samarium. At last, he recalls his work of 1882. The clues fit into a pattern leading to premonition and even certainty that an unknown element can be found between neodymium and samarium. But as his friend, Mendeleev, Brauner was never too hasty in his conclusions. It was only in 1901 that he placed an empty box between neodymium and samarium when he put forward his views on the place of the rare–earth elements in the periodic system.

Now we can give a translation of the note he wrote in margin of his hand–written table of elements. It reads: “61st element is the missing element predicted by me in 1902”.

His short letter to Nature was an attempt by Brauner to put the record straight. This would seem to simplify the task of science historians in writing the history of element 61. But a history is meaningful only if it treats a subject which really exists. As for illinium the element proved to be still–born.

While the hotheads kept trying to squeeze the symbol Il into box 61 of the periodic table meticulous critics tried to verify the discovery. The careful experiments by the first of them, Prandtl, could be doubted by nobody. But his results did not even hint at the existence of element 61.

In 1926 the Noddacks who had just announced their discovery of masurium and rhenium (Nos. 43 and 75) started their tests. They used all available techniques to analyse fifteen various minerals suspected of containing illinium. The processed 100 kilograms of rare–earth materials and could not detect a new element. The Noddacks claimed that if the American’s results had been correct they, the Noddacks, would undoubtedly extracted the new element. Even if the element were 10 million times rarer than niodymium or samarium they would still find it… There are two possible explanations: either element 61 is so rare that the existing experimental techniques are not fine enough to find it or wrong mineral specimens were taken.

Geochemists were against the first explanation. The abundances of rare–earth elements are more or less similar. There are no reasons to think that illinium is an exception. They suggested looking for illinium in minerals of calcium and strontium. All rare–earth elements are typically trivalent but some of them can exhibit a valence of two or four. For instance, europium rather easily gives rise to cations with a charge of two. Their size is closer to those of calcium and strontium cations and they can replace the letter in the respective alkaline–earth minerals. Perhaps, illinium has a similar more pronounced capacity and can be found in some rare natural compound of strontium. One hypothesis replaced another, one assumption stemmed from another, unsubstantiated one. Just in case, the Noddacks analysed several alkaline–earth minerals. Alas, they failed once more.

The search for illinium seemed to come to a dead end; though it still went on the reported results were little believed. Chemists failed in looking for element 61 in the terrestrial minerals it was theoretical physics whose fate it was to open up the “envelope” where nature had “sealed” element 61. But when the envelope was open the scientist (not for the first time!) were disappointed. The envelope was empty.

At this point the fate of element 61 directly involves the fate the element 43, that is, technetium. According to the law formulated by the German theoretical physicist Mattauch, technetium in principal cannot have stable isotopes. This law also forbids existence of stable isotopes of element 61. Illinium is dead but element 61 must survive.

But what if it really does not exist? I. Noddack put forward a daring idea that illinium (we shall use this name for the time being) had existed on Earth in early geological periods. But it had been a highly radioactive element with a short half–life and it had decayed fairly soon and disappeared from the face of Earth. If we agree with this idea we have to make two extremely unlikely assumptions. First, illinium which is at the centre of the periodic table has no stable isotopes. Second, the half–lives of its isotopes a e much shorter than the age of Earth.

Indeed, illinium neighbours in the periodic system (neodymium and samarium) have many (seven each) natural isotopes with a wide range of mass numbers–from 142 to 154. Any feasible isotopes of element 61 would have its mass number in this range. Thus, any illinium isotopes proves to be unstable in this range of mass numbers. The Mattauch law seem to bury for good the hopes to find element 61 on Earth. But then a gleam of hope appeared. All right, the illinium isotopes are all radioactive. But to what extent? Perhaps the half–lives of some of them are very long. At that time the theory had not learned how to predict half–lives of isotopes. The search for element 61 had to continue in the dark. Physicists believed that only nuclear synthesis could solve the riddle of element 61 the more so as the case of technetium was fresh in their minds.

As if trying to restore the honour of American science after its setback in 1926 two physicists from the University of Ohio conducted the first experiment of artificial synthesis of element 61 in 1938. They bombarded a neodymium target with fast deuterons (the nuclei of heavy hydrogen). They believed that the resulting nuclear reaction Nd + d → → 61 + n gave rise to an isotopes of element 61. Their results were inconclusive but nevertheless they thought that they obtained an isotope of the new element with the mass number of 144 and the half–life of 12.5 hours.

Again sceptics said that these results were erroneous and not without a reason since nobody could be sure that the neodymium target was ideally pure. The method of identification could hardly be considered reliable, too. Even uncomplicated optical and X–ray spectra evidenced the presence of element 61 as in the study of 1926; the conclusion was made from the radiometric data.

In fact, chemistry was not involved in this work and the chemical nature of the mysterious radioactive product was not determined. Therefore, one may ask whether 1938 can be regarded as the actual data of discovery of element 61. It can rather be said that only the consistent efforts to synthesize it started at the time.

As time passed the range of bombarding particles was extending, targets of other rare–earth elements were used, and the techniques of activity measurements were improved. Reports on other illinium isotopes started to appear in scientific journals. Element 61 was becoming a reality albeit an artificially created one. Its name was changed to cyclonium in commemoration of the fact that it was produced in a cyclotron but the symbol Cy did not remain for long in box 61 of the periodic table.

Researchers had detected the radioactive “signal” of cyclonium but nobody had seen even a grain of the new element and its spectra had not been recorded. Only indirect evidence of the existence of cyclonium had been obtained.

The history of science of the 20th century knows of many great discoveries and one of the greatest is the discovery of uranium fission under the effect of slow neutrons. The nuclei of uranium–235 isotopes are split into two fragments, each of which is an isotope of one of the elements at the centre of the periodic table. Isotopes of thirty odd elements from zinc to gadolinium can be produced in this way. The yield of the isotopes of element 61 has been calculated to be fairly high–approximately 3 per cent of the total amount of the fission products.

But the task of extracting the 3 per cent amount proved to be very difficult. The American chemists J. Marinsky, L. Glendenin, and Ch. Coryell applied a new chemical technique of ion–exchange chromatography for separation of the uranium fission fragments.

Special high–molecular compounds known as the ion–exchange resins are employed in this technique for separating elements. The resins act as a sieve sorting up elements in an order of the increasing strength of the bonds between the respective elements and the resin. At the bottom of the sieve the scientists found a real treasure–two isotopes of element 61 with the mass numbers 147 and 149.

At last, element 61 known as illinium, florencium, and cyclonium could be given its final name. According to recollections of the discoverers, the search for a new name was no less difficult than the search for the element itself. The wife of one of them, M. Coryell, resolved the difficulty when she suggested the name promethium for the element. In an ancient Greek myth Prometheus stole fire from heaven, gave it to man and was consequently put to extreme torture by Zeus. The name is not only a symbol of the dramatic way of obtaining the new element in noticeable amounts owing to the harnessing of nuclear fission by man but also a warning against the impeding danger that mankind will be tortured by the hawk of war, wrote the scientists.

Promethium was obtained in 1945 but the first report was published in 1947. On June 28, 1948, the participants at a symposium of the American Chemical Society in Syracuse had a lucky chance to see the first specimens of promethium compounds (yellow chloride and pink nitrate) each weighing 3 mg. These specimens were no less significant than the first pure radium salt prepared by Marie Curie. Promethium was born by the great creative power of science. The amounts of promethium prepared now weigh tens of grams and most of its properties have been studied.

The Mattauch law denied the existence of terrestrial promethium but this denial was not absolute. The search for promethium in terrestrial ores and minerals would be quite in order if promethium had long–lived isotopes with half–lives of the order of the age of Earth.

But in this respect nuclear physics proved to be a foe of natural promethium. With each newly synthesized promethium isotope a possible scope for search became increasingly narrow. The promethium isotopes were found to be short–lived. Among the fifteen promethium isotopes known today the longest–lived one had a half–life of only 30 years. In other words, when Earth had just formed as a planet not a trace of promethium could exist on it. But what we mean here is the primary promethium formed in the primordial process of origination of elements. What was discussed was the search for the secondary promethium which is still being formed on Earth in various natural nuclear reactions.

Technetium was finally found on Earth among the fragments of spontaneous fission of uranium. These fission products could contain promethium isotopes. According to estimates, the amount of promethium that can be produced owing to spontaneous fission of uranium in the Earth’s crust is about 780 g, that is, practically, nothing. To look for natural promethium would be tantamount to dissolving a barrel of salt in the lake Baikal and then trying to find individual salt molecules.

But this titanic task was fulfilled in 1968. A group of American scientists including the discoverer of natural technetium P. Kuroda managed to find the natural promethium isotope with a mass number of 147 in a specimen of uranium ore (pitchblende). This was the final step in the fascinating history of the discovery of element 61.

As in the case of technetium, we can name two dates of discovery of promethium. The first date is the date of its synthesis, that is, 1945. But under the circumstances synthesis was unconventional (it could be called fission synthesis). The first two promethium isotopes were extracted from the fragments of fission of uranium irradiated with slow neutrons rather than in a direct way as was the case with technetium, which was produced in a direct nuclear reaction. This makes promethium a unique case among all over synthesized elements.

The second date is the date of the discovery of natural promethium, that is, 1968. This achievement is of independent significance as it stretched to the utmost the capabilities of the physical and chemical methods of analysis. Of course, the achievement is of a purely theoretical significance since nobody can hope to extract natural promethium for practical uses.

The Boiling Point of Water

Water always boils at 100˚C, right? Wrong! Though it’s one of the basic facts you probably learnt pretty early on back in school science lessons, your elevation relative to sea level can affect the temperature at which water boils, due to differences in air pressure. Here, we take a look at the boiling points of water at a variety of locations, as well as the detailed reasons for the variances.

From the highest land point above sea level, Mount Everest, to the lowest, the Dead Sea, water’s boiling point can vary from just below 70 ˚C to over 101 ˚C. The reason for this variation comes down to the differences in atmospheric pressure at different elevations.

Atmospheric pressure the pressure exerted by the weight of the Earth’s atmosphere, which at sea level is simply defined as 1 atmosphere, or 101,325 pascals. Even at the same level, there are natural fluctuations in air pressure; regions of high and low pressure are commonly shown as parts of weather forecast, but these variances are slight compared to the changes as we go higher up into the atmosphere. As your elevation (height above sea level) increases, the weight of the atmosphere above you decreases (since you’re now above some of it), and so pressure also decreases.

In order to understand how this affects water’s boiling point, we first need to understand what’s going on when water boils. For that, we’ll need to talk about something called ‘vapour pressure’. This can be thought of as the tendency of molecules in a liquid to escape into the gas phase above the liquid. Vapour pressure increases with increasing temperature, as molecules move faster, and more of them have the energy to escape the liquid. When the vapour pressure reaches an equivalent value to the surrounding air pressure, the liquid will boil.

At sea level, vapour pressure is equal to the atmospheric pressure at 100 ˚C, and so this is the temperature at which water boils. As we move higher into the atmosphere and the atmospheric pressure drops, so too does the amount of vapour pressure required for a liquid to boil. Due to this, the temperature required to reach the necessary vapour becomes lower and lower as we get higher above sea level, and the liquid will therefore boil at a lower temperature.

This is, of course, a fact that’s true for all liquids, not just water. And it’s also not just atmospheric pressure that can affect water’s boiling point. Most of us are probably aware that adding salt to water during cooking increases water’s boiling point, and this is also related to vapour pressure. In fact, adding any solute to water will increase the boiling temperature, as it reduces the vapour pressure, meaning a slightly higher temperature is required in order for the vapour pressure to become equal to atmospheric pressure and boil the water.

Another factor that can affect the boiling temperature of water is the material that the vessel it’s being boiled in is made of. Experiments have shown that, at the same pressure, water will boil at different temperatures in metal and glass vessels. It’s theorised that this is because water boils at a higher temperature in vessels which its molecules adhere to more strongly – there’s much more detail on this phenomenon here.

So, water’s boiling point is anything but absolute, and it can be affected by a whole range of factors. Useful information if you ever find yourself wanting to make a cup of tea on Everest – the lower boiling point would mean the cup you end up with is rather weak and unpleasant

 

Discovery of element : Technetium

Technetium

The upper part of the periodic system down to the sixth period (where the rare–earth elements are located) always seemed relatively quiet, particularly after the discovery of the group of noble gases which harmoniously closed the right–hand side of the system. It was quiet in the sense that one could hardly expect any sensational discoveries there. The debates concerned only a possible existence of elements that were lighter than hydrogen and elements lying between hydrogen and helium. On the whole, we can say in the parlance of mathematicians that this part of the periodic system was an ordered set of chemical elements.

Therefore, the more awkward and confusing seemed to be the mysterious blank slot No. 43 in the fifth period and seventh group.

Mandeleev named this element eka–manganese and tried to predict its main properties. A few times the element seemed to have been discovered but soon it proved to be an error. This was the case with ilmenium allegedly discovered by the Russian chemist R. Hermann, back in 1846. For some time even Mendeleev tended to believe that ilmenium was eka–manganese. Some scientists suggested placing devium  between molybdenum and ruthenium. The German chemist A. Rang even put the symbol Dv into this box of periodic table. In 1896 there flashed and burned like a meteor lucium supposedly discovered by P. Barriere.

Mandeleev did not live to see the happy moment when eka–manganese was really found. A year after his death, in 1908, the Japanese scientist M.  Ogawa reported that he found the long–awaited element in the rare mineral, molybdenite and named it nipponium (in honour of the ancient name of Japan). Alas, Asia once more failed to contribute a new element to the periodic system. Ogawa, most probably, dealt with hafnium (which was also discovered later).

Chemists grew accustomed to a few chemical elements being discovered every year and they were at a loss in the case of eka–manganese. They began to think that Mendeleev could make a mistake and no manganese analogues existed. 

H. Moseley decisively refuted this skepticism in 1913. He clearly demonstrated that these analogues have their own place among the elements. In a paper dated September 5, 1925, W. Noddack, I. Tacke, O. Berg announced that they had discovered, together with element No. 75 (rhenium), its lighter analogue in the seventh group of the periodic system, namely, masurium whose number was 43. Two new symbols, Ma and Re, appeared in the periodic table, in chemical textbooks, and numerous scientific publications. The discoverers saw nothing odd in the fact that masurium and rhenium had not been discovered earlier. These elements were thought to be not too rare, however. The lateness of their discovery was attributed to another cause. A large group of trace elements in known to geochemistry. The trace elements are classified as those elements which have no or almost no own minerals but are spread in various amounts over minerals of other elements as if the nature has sprayed them with a giant atomizer. This is why the traces of masurium and rhenium were so hard to identify. Only the powerful eye of X–ray spectral analysis could distinguish them against the formidable background of other elements. There is an ancient saying that if two people do the same things this does not mean that the results will be identical. Two biographies started under the same conditions typically follow different paths. The same can be said about the fates of elements 43 and 75; one of them went a long way and found its proper place while the other’s way soon led it to a forest of errors, misunderstandings, and controversies. This was the path of masurium.

W. Prandtl got interested in the empty slots in the seventh group of the periodic table. He had his own outlook and put forward original ideas on the structure of the periodic system. He did not compile a new version of the table, though. He suggested placing the rare–earth elements each to a group though by that time most chemists had put down such an arrangement. But in Prandtl’s version of the table the seventh group happens to reveal as many as four empty slots below manganese corresponding to yet undiscovered elements (this was in 1924) whose numbers were 43, 61, 75, and 93. Prandtl believed this to do no chance occurrence but a result of a common cause that had prevented four elements from having been discovered. The German scientist, however, made his table structure too elaborate and artificial to be accepted. The final discovery of rhenium was the first indication of his errors, and his ideas on the first transuranium element (No. 93) were little thought of at the time. But he was intuitively right in thinking of a close common link between elements 43 and 61.

The belief in masurium’s existence gradually diminished. Only the original discoverers were firm. As late as the beginning of the thirties I. Noddack continued to say that in time element 43 would be commercially available as it happened with rhenium. But as the time passed and chemists again and again failed to find masurium in whatever minerals they analysed they came to believe that I. Noddack was right only by half, that is, only about rhenium. Rarest mineral specimens were tested for masurium. Some people even went as far as to claim that masurium minerals had yet to be found and would possess unheard of properties. Naturally, geochemists were quite sceptical. The imagination of some people went even further and masurium was suggested to be radioactive. That was too much, others said. But it was precisely this shot that did not go wild.

Let us talk about some concepts of nuclear physics. We have discussed isotopes. Now we meet another term–isobars–elements having the same atomic weight or mass numbers but different atomic numbers (from the Greek for “heavy”). Isobars, in other words, are isotopes of different chemical elements with different nuclear charges but identical mass numbers. Take, for instance, potassium–40 and argon–40 which have different nuclear charges (respectively, 19 and 20). Their mass numbers are identical because their nuclei contain different numbers of protons and neutrons but their total numbers are the same; potassium nucleus contains 19 protons 21 neutrons while argon nucleus has 20 protons and 20 neutrons.

Thus, the concept of isobars turned out to be the magic key that opened the door to the mystery of masurium.

When the majority of stable chemical elements were found to have isotopes–up to ten isotopes per element–the scientists started to study the laws of isotopism. The German theoretical physicist J. Mattauch formulated one of such laws at the beginning of the thirties (the basic premise of this law was noted back in 1924 by the Soviet chemist S. Shchukarev). The law states that if the difference between the nuclear charges of two isobars is unity one of them must be radioactive. For instance, in the 40K–40Ar isobar pair the first is naturally weakly radioactive and transforms into the second owing to the so–called process of K–capture. Then Mattauch compared with each other the mass numbers of the isotopes of the neighbours of masurium, that is, molybdenum (Z = 42) and ruthenium (Z = 44):

            Mo isotopes      94         95         96         97         98         –          100       –          –

            Ru isotopes         –           –          96         –          98         99         100       101       102

What did he deduce from this comparison? The fact that the wide range of mass numbers from 94 to 102 was forbidden for the isotopes of element 43 or, in other words, that no stable masurium isotopes could exist.

If that was really so that meant a peculiar anomaly linked to the number 43 in the periodic system. All the atom species with Z = 43 had to be radioactive as if this number was a small island of instability amidst a sea of stable elements. This, of course, would be unfeasible to predict within the framework of purely chemical theory. When Mendeleev predicted his eka–manganese he could never imagine that this member of the seventh group of the periodic system could not exist on Earth. Of course, in those times (the early thirties) Mattauch’s law was no more than a hypothesis, though one that looked like quite capable of becoming a law. And it became just that. The physicist’s idea opened the eyes of chemists who lost all hope of finding element 43 and they saw the source of their errors. However, the symbol Ma remained in box 43 of the periodic system for a few more years. And not without a reason. All right, all masurium isotopes are radioactive. But we know radioactive isotopes existing of Earth–uranium–238, thorium–232, potassium–40. They are still found on Earth because their half–lives are very long. Masurium isotopes are, perhaps, long–lived, too? If so, one should not be too hasty in dismissing the chances of successful search for element 43 in nature.

The old problem remained open. Who knows which way the biography of masurium would take if not for the dawn of a new age–that of artificial synthesis of elements.

Nuclear synthesis became feasible after invention of the cyclotron and the discoveries of neutrons and artificial radioactivity. In early thirties a few artificial radioisotopes of known elements were synthesized. Syntheses of heavier–than–uranium elements were even reported. But physicists just did not dare to take the challenge of the empty boxes at the very heart of the periodic system. It was explained by a variety of reasons but the major one was enormous technical complexity of nuclear synthesis. A chance helped. At the end of 1936 the young Italian physicist E. Segre went for a post–graduate work at Berkley (USA) where one of the first cyclotrons in the world was successfully put into operation. A small component was instrumental in cyclotron operation. It directed a beam of charged accelerated particles to a target. Absorption of a part of the beam led to intense heating of the component so that it had to be made from a refractory material, for instance, molybdenum.

The charged particles absorbed by molybdenum gave rise to nuclear reactions in it and molybdenum nuclei could be transformed into nuclei of other elements. Molybdenum is a neighbour of element 43 in the periodic system. A beam of accelerated deutrons could, in principle, produce masurium nuclei from molybdenum nuclei.

That was just Segre’s thought. He was a competent radiochemist and understood that if masurium really were produced its amount would be literally negligible and its separation from molybdenum would present an enormously intricate task. Therefore, he took an irradiated molybdenum specimen with him back to the University of Palermo where he was assisted in his work by the chemist C. Perrier.

They had had to work for nearly half a year before they could present their tentative conclusions in a short letter to the London journal nature. Briefly, the letter reported the first in history artificial synthesis of a new chemical element. This was element 43 the futile search for which on Earth wasted so much efforts of scientists from many countries. Professor E. Lawrence from the University of California at Berkley gave the authors a molybdenum plate irradiated with deutrons in the Berkley cyclotron. The plate exhibited a high radioactivity level which could hardly be due to any single substance. The half–life was such that the substances could not be radioactive isotopes of zirconium, niobium, molybdenum, and ruthenium. Most probably they were isotopes of element 43.

Though the chemical properties of this element were practically unknown Segre and Perrier attempted to analyse them radiochemically. The element proved to be closely similar to rhenium and exhibited the same analytical reactions as rhenium. However, it could be separated from rhenium with technique used for separating molybdenum and rhenium. The letter was written in Palermo and dated June 13, 1937. It was by no means a sensation. The scientific community regarded it as just the authors going on record. The reported data were too patchy while what was needed to be convincing was precisely the details and clear results of radiochemical analysis.

Only later Segre and Perrier were recognized as heroes; indeed, they extracted from the irradiated molybdenum just 10–10g of the new element–an amount formerly undetectable Never before radiochemists worked with such negligible amounts of material. The discoverers suggested naming the new element technetium from the Greek for “artificial”. Thus, the name of the first synthesized element reflected its origin. The name, though, became generally accepted only ten years later.

Perrier and Segre received new specimens of irradiated molybdenum and continued their studies. Their discovery was confirmed by other scientists. By 1939 it was understood that bombardment of molybdenum with deutrons or neutrons produces at least five technetium isotopes. Half–lives of some of them were sufficiently long to make possible substantial chemical studies of the new element. It no longer sounded fantastic to speak about “the chemistry of element 43”. But all attempts to measure accurately the half–lives of the technetium isotopes failed. The available estimates were disheartening since none of them exceeded 90 days and this put a stop to all hopes of finding the element on Earth.

So what was technetium in the late thirties and early forties? Nothing more than an expensive toy for curious scientists. Any prospects of accumulating it in a noticeable amount were, apparently, non–existent. The fate of technetium (and not only of it) was reversed when nuclear physics discovered an amazing phenomenon–fission of uranium by slow neutrons.

When a slow neutron hits a nucleus of uranium–235 it in effect breaks the nucleus down into two fragments. Each of the fragments is a nucleus of an element from the central part of the periodic table, including technetium isotopes. It is not without a reason that a fission reactor (a large–scale nuclear energy producer) is known as a factory of isotopes. Cyclotron made possible the first ever synthesis of technetium and fission reactor allowed the chemists to produce kilograms of technetium. But even before the first fission reactor started operating Segre in 1940 found the technetium isotope with a mass number of 99 in uranium fission products in his laboratory. Having found its new birthplace in a fission reactor technetium started to turn into an everyday (paradoxical as it may be) element. indeed, fission of 1 g of uranium–235 gives rise to 26 mg of technetium–99.

As soon as technetium ceased to be a rare bird scientists found the answers to many questions that had puzzled them, and first of all about its half–lives. In the early fifties it became clear that three of technetium isotopes are exceptionally long–lived in comparison with not only its other isotopes but also many other natural isotopes of radioactive elements. The half–life of technetium–99 is 212 000 years, that of technetium–98 is one and a half million years, while that of technetium–97 is even more, namely, 2 600 000 years. The half–lives are long but not long enough for primary technetium to be conserved on Earth since its origin. The primary technetium would survive on Earth if its half–life were not shorter than one hundred fifty million years. This makes obvious the hopelessness of all search for technetium of Earth.

But technetium can still be produced in the course of natural nuclear reactions, for instance, when molybdenum is bombarded by neutrons. How can free neutrons appear on Earth? They can be produced in spontaneous fission of uranium. The process occurs as described above, only spontaneously, and gives rise to a few neutrons, apart from two large fragments, i.e. nuclei of lighter elements.

The search for technetium in molybdenum ores failed and scientists turned their attention to another possibility. If technetium isotopes are produced in fission reactors why cannot they be born in natural processes of spontaneous uranium fission?

Using as a basis the Earth uranium resources (taking the figure for the mean abundance of uranium in the 20–km thickness of the Earth crust) and assuming the same proportion of produced technetium as in the case of reactor fission we can calculate that there are just 1.5 kg of technetium on Earth. Such a small amount (though it is larger than for other synthesized elements) could hardly be taken seriously. Nevertheless, scientists attempted to extract natural technetium from uranium minerals. This was done in 1961 by the American chemist B. Kenna and P. Kuroda. Thus, technetium acquired another birthday–the day when it was discovered in nature. If the methods of artificial synthesis of technetium had failed to materialize, even then it would, sooner or later, be brought to light from the bowels of the Earth.

But ten years earlier, in 1951, sensational news about element 43 was heard. The American Astronomer S. Moore found characteristic lines of technetium in the solar spectrum. The spectrum of technetium had been recorded immediately when it had become feasible, that is, when a sufficient amount of the element had been synthesized. The spectral data had been compared with those reported by the Noddacks and Berg for masurium. The spectra had proved to be quite different making ultimately clear the mistake of the discoverers of masurium. The spectrum of the solar technetium was identical to that of the terrestrial technetium. An analogy with helium was apparent–both elements sent messages from the Sun before to be found on Earth. True, astronomers questioned the data on the solar technetium.  But in 1952 the cosmic technetium once more sent a message when the British astrophysicist P. Merril found technetium lines in the spectra of two stars with the poetic names of R Andromedae and Mira Ceti. The intensities of these lines evidenced that the content of technetium in these stars was close to that of its neighbours in the periodic system, namely, niobium, zirconium, molybdenum, ruthenium, rhodium and palladium. But these elements are stable while technetium is radioactive. Though its half–life is relatively long it is still negligible on cosmic scale. Therefore, the existence of technetium on stars can mean only that it is still born there in various nuclear reactions. Chemical elements continue to be produced in stars on a gigantic scale. A witty astrophysicist named technetium the acid test of cosmogonic theories. Any theory of the origin of elements must elucidate the sequence of nuclear reactions in stars giving rise to technetium.

Reason for the formation of large number of organic compounds : Catenation

Reason for the formation of large number of organic compounds

 What makes the carbon so special ?

What is it that sets carbon apart from all other elements in the periodic table ?

Why are there so many organic compounds ?

The answer lies in carbon’s position in the periodic table. Carbon is in the centre of second row elements

Li > Be >  B > C > N > O > F

First think why molecules are formed from atoms ? It is because of the reason that atom combines with same or with other atoms to form molecule so as to complete its octet and attain lower energy stable and hence become stable. That is the reason why noble gases are considered as inert gases, they generally do not combine with itself or with other atoms because they have complete octet. But what about other atoms ? They have incomplete octet, so they must combines with same or other atoms to form molecule for better stability.

Elements on the left hand side of carbon have less than 4 electrons in the valence shell (Li-1, Be-2, B-3) so they have more tendencies to loose electron to attain noble gas configuration for stability. That’s why they generally forms compounds with Li+, Be2+, B3+ by losing 1, 2, 3 electrons respectively. Elements present downside in the same group too have similar tendency as that of Li, Be and B, hence form compounds in the following states; Na+, K+, Rb+, Cs+, Mg2+, Ca2+, Sr2+, Al3+, Ga3+, etc.

Elements on the right hand side of carbon have more than 4 electrons in the valence shell (N-5, O-6, F-7). To complete their octet, valance electron must be subtracted from 8 that’s why the valency of N is (8-5) i.e. 3, O is (8-6) i.e. and that of F is (8-7) i.e. 1. It is much easier to gain 3, 2, 1 electrons to complete their octet as compared to loosing 5, 6, 7 electrons to complete their octet. So these elements have more tendency to gain electrons and form compounds in the following states; N3, P3, As3, Sb3, Bi3, O2, S2, Se2, Te2, Po2, F, Cl, Br, I.

As elements present on the left hand side of carbon loose electrons to form compounds and elements of right hand gain electrons to form compounds so compounds formed are ionic in nature.

But think about carbon and the elements present down side, which are present in the middle of each period and have equal tendency to loose or gain electrons as they have 4 electrons in their octet. This led carbon and other elements of this group (Si, Ge, Sn & Pb) to share electrons with itself and other elements of periodic table to complete octet. As these compounds are formed by sharing of electrons so they are considered to be covalently bonded.

Carbon by sharing its electrons with other carbon atoms leads to formation of long chain carbon compounds which may be single, double or triple bonded, cyclic or acyclic, linear or branched. This self-linking property of carbon is called catenation. All the atoms of 14th group show the property of catenation but it decreases down the group because of weak overlapping due to large size and follows order :

C >> Si >> Ge > Sn > Pb

Carbon may also form multiple bonds with N, P, O, S etc. forming large number of functional group, which we will discuss later.

This is not the end of compound formation, carbon forms many abnormal compounds with elements of s, p & d blocks. So for sake of simplicity we are constructing an organic chemist’s periodic table with the most important elements emphasized.

Elements, which are in dark box, are generally involved in making organic compounds along with deuterium (D), which is an isotope of hydrogen (H).

As there are large number of atoms in periodic table which have valence electrons, atomic orbital of carbon may overlap with them and share its electron to form large number of compounds. But for that many other factors such as size, activation energy, electronegativity, electron affinity, catenation etc. are responsible which all come under one word “Position” i.e. position for carbon in the periodic table. This word “position” include everything related with molecule formation therefore the main reason behind large number of organic compounds is the position of carbon in the periodic table.

 

Baeyer’s strain theory

Baeyer’s strain theory : To compare stability of cycloalkanes 

 When we carefully look over the cyclic saturated compounds, we find that each atom is sp3  hybridized.  The ideal bond angle 109028’ but in cycloalkanes this angle is mathematically 180-(360/n) where n is the number of atoms making ring. 

for example Cyclopropane, angle is 600; in Cyclobutane it is 900 and so on.

Angle Strain : This difference in ideal bond angle and real bond angle, is called angle strain and it causes strain in bond which affects the stability of molecule. 

Greater is the deviation from the theoretical angle, greater is the Angle strain ; lesser the stability. 

To calculate the distortion or angle strain in cycloalkane we assume the atoms of ring in a plane, such as in cyclopropane, all the 3 carbon atoms occupy one corner of an equilateral triangle with bond angle 60o. As two corners bent themselves to form bond so strain too is divided equally. So strain in cyclopropane will be ½ (109o28’ – 600) = 24044’.

Deviation of bond angle in cyclopropane from normal tetrahedral angle

Distortion or strain = ½ (109028’ bond angle of ring). So angle strains in some cycloalkanes are listed in the table below.

Compound

No. of C in the ring

Angle between the C atoms

Distortion or strain

Cyclopropane

3

600

24o44’

Cyclobutane

4

900

9o44’

Cyclopentane

5

108o

0o44’

Cyclohexane

6

120o

-5o16’

Cycloheptane

7

128o34’

-9o33’

Cyclooctane

8

135o

-12o62’

 

From the table it is clear that cyclopropane has the maximum distortion, so it is highly strained molecule and consequently more reactive than any of one monocylic alkanes, which is clear from the reaction that ring can be opened very easily to relieve strain on reaction with Br2, HBr or H2/Ni at high temperature.

In contrast, cyclopentane & cyclohexane have least strain so they are found more readily and are very stable as compared to cyclopropane.

Baeyer strain theory satisfactorily explains the typical reactivity and stability of smaller rings (from C3 to C5) i.e. Stability order follows : Cyclopropane < Cyclobutane < Cyclopentane

But not valid for cyclohexane onwards because the strain again increases with the increase in number of carbon atom but actually large rings are more stable. So molecular orbital theory is also considered according to which covalent bond is formed by coaxial overlapping of atomic orbitals. The greater is the extent of overlap the stronger is the bond formed. In case of sp3 carbon, C – C bond will have maximum strength if the C-C-C bond have the angle 109o28’. If cyclopropane is an equilateral triangle then the bond angle of each C-C-C bond would be 60o. Therefore it was proposed by Couson that in cyclopropane the sp3 hybridized orbitals are not present exactly in one straight line due to mutual repulsion of orbital of these bonds resulting thereby loss of overlap. This loss of overlap weakens the bond and is responsible for its instability and strain in molecule. Similarly, in case of cyclobutane, there is also loss of overlap but the loss is less than in cyclopropane, so cyclobutane is more stable than cyclopropane. Overlapping of orbitals in large ring compound (5 more carbon atoms) is however much better which accounts for the greater stability of such compound.

It is natural that when a molecule has strain within it, it will affect the stability of molecule. The stability of molecules can be calculated easily by measuring heat of combustion which will give the measure of total strain and thermochemical stability which can be calculated mathematically.

Total strain = (No of C atom is the ring × observed heat of combustion/CH2) observed heat of combustion/CH2 for n alkane.

 

Experimental data of total strain for different cycloalkanes*

No. of C in the ring

Heat of combustion kJ per CH2 group

Total strain in kJ

3

697

115

4

686

109

5

664

27

6

659

0

7

662

27

8

663.8

42

664.6

54

* data from Organic chemistry solomons & Fryhle

From the data above it is clear that strain decreases from C3 to C6 i.e. stability increases, but stability again deteriorates from C7 to C9 ring system. 

 According to this theory, the carbon atoms in 5 membered and smaller rings can lie in one plane as explained by Baeyer but Sachse suggested that in six membered and higher rings the carbon atoms are non planar . In this way the ideal angle 109028’ is retained and the ring is free from angle strain. Thus Sachse proposed that cyclohexane exist in two puckered forms as boat and chair form. These forms are readily inter-convertible through half chair and twist boat forms simply by rotation about the single bonds.  

 

Discovery of Actinium element

Actinium

Was it just a chance that polonium and radium were the first to be discovered among radioactive elements? The answer is apparently no. Owing to its long half–life radium can be accumulated in uranium ores. Polonium has a short half–life (138 days) but it emits characteristic high–intensity alpha radiation. Though the discovery of polonium gave rise to a controversy it soon died off.

The third success of the young science of radioactivity was the discovery of actinium. Soon after they had discovered radium the Curies suggested that uranium ore could contain other, still unknown radioactive elements. They entrusted their collaborator A. Debierne with verification of this idea.

Debierne started his work with a few hundred kilograms of uranium ore extracting the “active principle” from it. After he had extracted uranium, radium, and polonium he was left with a small amount of a substance whose activity was much higher than the activity of uranium (approximately, by a factor of 100 000). At first, Debierne assumed that this highly radioactive substance was similar to titanium in its chemical properties. Then he corrected himself and suggested a similarity with thorium. Later, in spring of 1899 he announced the discovery of a new element and called it actinium (from the Greek for radiation).

Any textbook, reference book or encyclopedia gives 1899 as the date of the discovery of actinium. But in fact, to say that in 1899 Debierne discovered a new radioactive element–actinium–means to ignore very significant evidence to the contrary.

The real actinium has little in common with thorium but we did not mean this chemical difference as evidence against the discovery of actinium by Debierne. The main argument is as follows. Debierne believed that actinium was alpha–active and its activity was 100 000 times that of uranium. Now we know that actinium is a mild beta–emitter, that is, it emits beta rays of a fairly low energy which are hot that easy to detect. Of course, the primitive radiometric apparatus of Debierne was not capable of doing it.

Then what did Debierne discover? It was a complex mixture of radioactive substance including actinium. But the weak beta radioactive of actinium was quite indistinguishable against the background of the alpha rays emitted by the products of actinium decay. It took several years to extract the real actinium from this mixture of radioactive products.

In 1911 the outstanding British radiochemist F. Soddy published a book entitled chemistry of Radioactive Elements where he described actinium as an almost unknown element. He wrote that its atomic weight was unknown, the mean life time was also unknown, it did not emit rays (this shows how difficult it was to detect the beta radiation of actinium), and its parent substance was unknown. In a word, much about actinium was still vague.

The evidence presented by Debierne for his discovery of actinium did not seem convincing to his contemporaries. It is no wonder that soon another scientist–the German chemist F. Giesel–claimed a discovery of a new radioactive element. He also extracted a certain radioactive substance whose properties were similar to those of the rare–earth element. This fact is closer to the truth in the light of our current knowledge. Giesel named the new element emanium because it evolved a radioactive gas–emanation–which made a zinc sulphide screen to glow. Along with the radiotellurium vs. polonium controversy there appeared a similar controversy between the supporters of actinium and emanium. The first controversy ended by establishing identity between the elements in question. The second controversy proved to be more complicated and could not be speedily resolved since the behaviour of the third new radioactive element was too wayward. The name of Debierne went into the historical records as the name of the discoverer of actinium. However, the substance extracted by Giesel contained a significant proportion of pure actinium as was shown later. Giesel also succeeded in observing the spectrum of emanium. Many scientists believed that they proved identity of actinium and emanium. Gradually, the controversy lost its edge.

The British radiochemist A. Cameron was the first (1909) to place the symbol Ac into the third group of the periodic system (actually, he was the first to put forward the name radiochemistry for the relevant science). But only in 1913 was the position of actinium in the periodic system established reliably. As increasingly pure actinium preparations were obtained the scientist encountered an amazing situation–the radiation emitted by actinium proved to be so weak that some scientists even doubted if it emits at all. It has even been suggested that actinium undergoes an entirely new, radiation less, transformation. It was only in 1935 that beta rays emitted by actinium were reliably detected. The half–life of actinium was found to be 21.6 years.

For a long time extraction of metallic actinium was just out of question. Indeed, one ton of pitchblende contains only 0.15 mg of actinium while the content of radium is as high as 400 mg. A few milligrams of metallic actinium were obtained only in 1953 after reduction of AcCl3 with potassium vapour.

NTSE Scholarship : Path made simple and clear

Are you a student of Class X and looking for a platform to identify your hidden talent and evaluate your preparedness for further studies? The right answer and the platform catering to your needs is National Talent Search Examination (NTSE). NTSE identify and recognize students with high intellect and academic talent at the high-school level.

NTSE is a two-tier exam- Stage I (State Level)  is  conducted by  the States/ Union  Territories  and Stage-II(National Level) is conducted by NCERT. The objective of the two-tier  examination is to identify the talented students who have a special aptitude for sciences, maths, social sciences and questions based on analytical reasoning.

Stage I is for all interested students studying in class X to participate at state level While stage II can be written by stage I qualified students. The examination acts as unique platform for the students to check their capability and potential. It also points out the weaknesses and short comings in the domains covered by the examination. Every year, lakhs of students appear for this scholarship exam, out of which One Thousand scholarships are awarded. The scholarship is open only to the students of Indian nationality whether they are studying in India or a broad in class X or equivalent.

Scholastic Aptitude Test (SAT) SAT comprises of 100 multiple choice questions, where one alternative is correct.There are 40 questions from Science, 40 from Social Science and 20 from Mathematics. The idea is to assess the subject knowledge, reasoning ability and logical thinking of the candidates.

Mental Ability Test (MAT) : 

This also comprises of 100 multiple choice questions with only one out of four options correct. Each question carries 1 mark. There is no negative marking.  It has a variety of questions about analogies, classification, series, pattern perception, hidden figures, coding-decoding, block assembly, problem-solving, etc. Here, the goal is to gauge the power of reasoning, ability to judge, evaluate, visualize in space, spatial orientation, etc. of the candidates.

The second level test takes place every year in the month of May. As per the new pattern of Stage II announced by NCERT, each correctly answered question earns the candidate one mark with no negative marks. This level of exam tests the students’ potential concerning their mental ability and scholastic aptitude. income, government school, domicile, etc. However, no scholarships shall be available for studies abroad for any course. The rates of scholarship sat different stages of study are as under.

PREPARATION STRATEGY :It’s necessary to know your strengths and weaknesses. For example, if you are strong in math and science but average in MAT and weak in social scien­ces, then you can work on your weaknesses and turn them into strengths for an over all very good or excellent score.

Follow NCERT books : Follow NCERT books of classes IX and X for NTSE preparation. For some concepts in biology, refer NCERT Class XI Biology textbook (only portions that are relevant to Class IX and X curriculum). It is imperative to study the Social Studies NCERT book for NTSE thoroughly.

Strengthen your MAT section : MAT  and   SAT  carry  equal weight in NTSE. Therefore, it is critical to give due time to both the   papers  while   preparing. Some of us tend to overdo prepa­ration of 3 subjects but  in the process  tend to ignore MAT. However,   it is important to real­ize  that   MAT  carries  more weight than any of the subjects. The MAT section tests your men­tal ability.

Solving sample papers : Practice makes a man perfect is the mantra to success. So,if you are aspiring excellence, it is important that you solve lot of sample papers and mock tests at least thrice a week.  You should aim to solve these papers within the stipulated time so that you can improve your speed and know the sections that consume more time.

Analyse your performance : Make sure you minutely assess what you could do and what you had a hard time with. Was it the subject understanding   you lacked? Or did you miss out on the scores because of silly mistakes?

Proper guidance : Find a good mentor. There are times when in quest of achieving something, you tend to go off the track. To ensure you are on the right track, it is important  to have a mentor to guide you and show you the right path in your success journey.

Self-study : For clearing any exam,a decent amount of self-study is important. It is one of the most basic and important of all the tips to become an NTSE scholar. One should devote at least 3-4 hours to self-learning to crack this competitive exam.

Keep Practicing : Solve as many questions of men­tal ability as possible so that you are not shocked on the exam day. With practice, you’ll also get more confident about  your speed, accuracy and subject knowledge. Just remember that you don’t have to beak now-it-all to compete and excel. You can be just an ordinary candidate and follow these tips to become an NTSE scholar.





Discovery of Radon Elements

Radon

Radon Rn is the 86th element of the periodic system. It is the heaviest of the noble gases. It is highly radioactive and its natural abundance is so low that it could not be identified when W. Ramsay and M. Travers discovered other inert elements. Only application of the radiometric method made possible the discovery of radon.

What we know as radon at present is the combined name for the three natural isotopes of the element No. 86, which were discovered one by one and called emanations. Their appearance heralded a new stage in the studies of radioactivity as they were the first gaseous radioactive substances.

At the beginning of 1899 E. Rutherford (who lived at the time in Canada) and his collaborator R. Owens studied the activity of thorium compounds. Once Owens accidentally threw open the door to the laboratory where a routine experimenter was performed. There was a drought and the experimenters noticed that the intensity of radiation of the thorium preparations suddenly dropped. At first they ignored this event but later they observed that a slight movement of air seemed to remove a larger part of the activity of thorium. Rutherford and Owens decided that thorium continuously emitted a gaseous radioactive substance, which they called the emanation (from the Latin to flow) of thorium, or Theron.

By way of analogy, it was suggested that other radioactive elements could also evolve emanations. In 1900 the German physicist E.  Dorn discovered the emanation of radium and three years later Debierne observed the emanation of actinium. Thus, two new radioactive elements were found, namely, radon and action. An important observation was that all the three emanations differed only in their half–lives–51.5 s for thoron, 3.8 days for radon, and 3.02 s for action. The longest–lived element is radon and therefore it was used in all studies of the nature of emanations. All the other properties of emanations were identical. All of them lacked chemical manifestations, that is, they were inert gases (analogues of argon and other noble gases). Later they were found to have different atomic masses. But there was just a single slot for these three elements in the periodic system, immediately below xenon.

Such exclusive situation soon became a rule. Therefore, we shall have to discuss briefly some important events in the history of radioactivity studies. Now we must finish the story of radon. This name remained because radon is the longest–lived element among the radioactive inert gases. Ramsay suggested to name it niton (from the Latin for glowing) but this name did not take root.

Discovery of Radium Element

Radium

When the Curies and G. Bemont analysed pitchblende they noticed a higher radioactivity of one more fraction, apart from the bismuth fraction. After they had succeeded in extracting polonium they started to analyse the second fraction thinking that they could find yet another unknown radioactive element.

The new element was named radium from the Latin radius meaning ray. The birthday of radium was December 26, 1898. When the members of the Paris Academy of Sciences heard a report entitled “On a new highly radioactive substance contained in pitchblende”. The authors reported that they had managed to extract from the uranium ore tailings a substance containing a new element whose properties are very similar to those of barium. The amount of radium contained in barium chloride proved to be sufficient for recording its spectrum. This was done by the well–known French spectral analyst E. Demarcay who found a new line in the spectrum of the extracted substance. Thus, two methods–radiometry and spectroscopy–almost simultaneously substantiated the existence of a new radioactive element.

The position of radium among the natural radioactive elements (of course, excluding thorium uranium) almost immediately proved to be the most favourable one owing to many reasons. The half–life of radium was soon found to be fairly long, namely, 1 600 years. The content of radium in the uranium ores was much higher than that of polonium (4 300 times higher); this contributed to natural accumulation of radium. Furthermore, the intensity of alpha radiation of radium was sufficiently high to allow an easy monitoring of its behaviour in various chemical procedures. Finally, a distinguishing feature of radium was that it evolved a radioactive gas known as emanation (see p. 183). Radium was a convenient subject for studies owing to a favourable combination of its properties and therefore it became the first radioactive element (again, with the exception of uranium and thorium) to find its permanent place in the periodic system without long delay. Firstly, chemical and spectral studies of radium demonstrated that in all respects it belongs to the subgroup of alkaline earth metals; secondly, its relative atomic mass could be determined accurately enough. To do be obtained. The Curies worked ceaselessly for 45 months in their ill–equipped laboratory processing uranium ore tailings from Bohemian mines. They performed fractional crystallization about 10 000 times and finally obtained a priceless prize–0.1 g of radium chloride. The history of science knows no more noble examples of enthusiastic work. This amount was sufficient for measurements and on March 28, 1902, Marie Curie reported that the relative atomic mass of radium was 225.9 (which does not differ much from the current figure of 226.02). This value just suited the suggested position of radium in the periodic system.

The discovery of radium was the best substantiated one among the many alleged discoveries of radioactive elements, which soon followed. Every year more new discoveries were reported. Radium was also the first radioactive element obtained in the metallic form.

Marie Curie and her collaborator A. Debierne electrolyzed a solution containing 0.106 g of radium chloride. Metallic radium deposited on the mercury cathode forming amalgam. The amalgam was put into an iron vessel and heated under a hydrogen flow to remove mercury. Then grains of silvery whitish metal glistened at the bottom of the vessel.

The discovery of radium was one of the major triumphs of science. The studies of radium contributed to fundamental changes in our knowledge of the properties and structure of matter and gave rise to the concept of atomic energy. Finally, radium was also the first radioactive element to be practically used (for instance, in medicine).

Discovery of Polonium Element

Polonium

Polonium was the first natural radioactive element discovered with the radiometric technique. Back in 1870 the main properties of polonium were predicted by D. I. Mendeleev. He wrote: “Among heavy metals we can expect to find an element similar to tellurium whose atomic weigh is greater than that of bismuth. It should possess metallic properties, and give rise to an acid whose composition and properties should be similar to those of sulphuric acid and whose oxidizing power is higher than that of telluric acid…

The oxide RO2 cannot be expected to have acidic properties which tellurous acid still has. This element will form organometallic compounds but not hydrogen compounds…”

Nineteen years had passed and Mendeleev made a significant addition to his description of dvi–tellurium (as he called the unknown element). He predicted the following properties: relative atomic mass 212; forms oxide DtO3; in a free state the element is a crystalline low–melting non–volatile metal of grey colour with a density of 9.8; the metal is easily oxidized to DtO2; the oxide will have weak acidic and basic properties: a hydride of the element, if it exists at all, must be unstable; the element must form alloys with other metals.

Below readers will see for themselves how accurate were Mendeleev’s predictions of the properties of a heavy analogue of tellurium. But these predictions had only an indirect effect on the history of polonium, if any. The discovery of polonium (and then radium) proved to be a significant milestone in the science of radioactivity and gave an impetus to its development.

As one can see from the laboratory log–book of Marie and Pierre Curie they started to study the Becquerel rays, or uranium rays, on December 16, 1897. First the work was conducted by Marie alone and then Pierre joined her on February 5, 1898. He performed measurements and processed the results. They mainly measured the radiation intensities of various uranium minerals and salts as well as metallic uranium. The results of extensive experiments suggested that uranium compounds had the lowest radioactivity, the metallic uranium exhibited a higher radioactivity, and the uranium ore known as pitchblende had the highest radioactivity. These results indicated that pitchblende, probably, contained an element whose activity was much higher than that of uranium.

As early as April 12, 1898 the Curies reported this hypothesis in the proceedings of the Paris Academy of Science. On April 14 the Curies started their search for the unknown element with the assistance of the chemist G. Bemont. By the middle of July they finished the analysis of pitchblende. They carefully measured the activity of each product successively isolated from the ore. Their attention was focussed on the fraction containing bismuth salts. The intensity of the rays emitted by this fraction was 400 times that of metallic uranium. If the unknown element really did exist it had to be present in this fraction.

Finally, on July 18 Marie and Pierre Curie delivered a report to a session of the Paris Academy of Science entitled “On a new radioactive substance contained in pitchblende”. They reported that they had managed to extract from pitchblende a very active Sulphur compound of a metal that had previously been unknown. According to its analytical properties it was a neighbour of bismuth. The Curies suggested, if the discovery could be proved, to name the new element in honour of the country where Marie had been born and brought up, that is, polonium after Poland.

The scientists emphasized that the element had been discovered with a new research method (the term “radioactivity”, which later became conventional, was first introduced in this report).

The introduction of spectral analysis made it possible to reveal the existence in natural objects of elements that could not be seen, felt or weighed. Now the history repeated itself but the role of indicator was played by radioactive radiation, which could be measured with a radiometric technique. However, the results of the Curies were not faultless. They were wrong in suggesting a chemical similarity between polonium and bismuth. Even a brief look at the periodic system shows that the existence of a heavy analogue of bismuth is hardly possible. But one must not forget that the Curies did not extract pure metal, could not determine its relative atomic mass, and, finally, did not see differences in the spectra of polonium and bismuth. This is why they actually ignored a possible analogy between polonium and tellurium.

Thus, we may regard 18 July, 1898, as the date of just a preliminary discovery of polonium as substantiation of the discovery took quite a long time. The high intensity of radiation from polonium made difficult its study. The radiation was found to consist of only alpha rays with no beta or gamma rays. A strange finding was that the activity of polonium decreased with time and the decrease was rather noticeable; neither thorium nor uranium exhibited such behaviour. This is why some scientists doubted whether polonium existed at all. The sceptics said it was just normal bismuth with traces of radioactive substances.

But in 1902 the German chemist W. Marckwald extracted the bismuth fraction from two tons of uranium ore. He put a bismuth rod into a bismuth chloride solution and observed precipitation of a highly radioactive substance on it which he took for a new element and named radiotellurium. Later he recalled: “I named this substance radiotellurium just for the time being since all its chemical properties suggested placing it into the sixth group into the still unoccupied box for the element with a somewhat higher atomic weight than that of bismuth…. The element was more electronegative than bismuth but more electropositive than tellurium; its oxide should also have basic rather than acidic properties.

All this corresponded to radiotellurium…. The expected atomic weight for this substance was about 210”. Later he said that he had got his idea for extracting polonium when analysing the periodic system.

As for the polonium discovered earlier Marckwald promptly declared it a mixture of several radioactive elements. This led to a stormy discussion of the real nature of polonium and radiotellurium. Most scientists supported the Curies. A. later comparison of the two elements revealed their identity. The discovery was credited to the Curies and the name “polonium” was retained.

Though polonium was the first of the new natural radioactive elements its symbol Po did not appear in the appropriate box in the periodic system. The atomic mass of the element was very difficult to measure. The lines of the polonium spectrum were reliably identified in 1910. It was only in 1912 that the symbol Po occupied its place in the periodic table.

For almost half a century scientists had to be satisfied to work only with polonium compounds (usually in rather small amounts). The pure metal was prepared only in 1946. High density layers of metallic polonium prepared by vacuum sublimation have a silvery colour. Polonium is a pliable low–melting metal (melting point 254oC, boiling point 962oC), its density is about 9.3 g/cm3. When polonium is heated in the air it readily forms a stable oxide; its basic and acidic properties are weakly manifested. Polonium hydride is unstable. Polonium forms organometallic compounds and alloys with many metals (Pb, Hg, Ca, Zn, Na, Pt, Ag, Ni, Be). When we compare Mendeleev’s predictions with these properties we see how close they are to the truth.

Discovery of Rhenium element

Rhenium

As regards history, rhenium had an undoubted advantage over hafnium: nobody had ever questioned the fact that element No. 75 had to be an analogue of manganese, or tri-manganese in Mendeleev’s terminology. However, in all other respects there was no certainty.

Let us perform an experiment. If we select at random a few monographs and textbooks where rhenium is discussed we shall see that the authors agree on some things while sharply disagreeing on others. They all agree that rhenium was discovered in 1925 but when it comes to the source from which rhenium was extracted, they disagree. Among minerals mentioned as sources of rhenium are columbite and platinum ore, native platinum and tantalite, niobite and wolframite, alvite and gadolinite. Even an experienced geochemist will be at a difficulty finding his way among so varied a group of minerals.

After these introductory remarks, we may name the discoverers of rhenium: V. Noddack, I. Takke (who later married V. Noddack), and the spectroscopist O. Berg. Their authorship was never contested by anybody. This may be the only case when engineers became interested in the yet undiscovered element. They were aware of the uses of the periodic system. Since tungsten was widely used in electrical engineering, there was every reason to believe that element No. 75 would possess properties even more valuable for this industry. It is highly probable that the first attempts of the Noddacks to find this element were prompted by practical needs.

In 1922, after thorough preparations they set to work. First of all, they collected all reports on the discovery of manganese analogues. Since these discoveries remained unconfirmed, it was tempting to check them. The scientists drew up an extensive program of research: they were going to look for two elements at once since unknown manganese analogues included not only element No. 75 but also its lighter predecessor–element No. 43 with an unusual fate (see p. 200). The periodic table made it possible to predict many of their properties. We can now compare the Noddacks’ predictions on rhenium with the actual properties of the element.

            Prediction                                            Modern data

Atomic mass 187-188                                          186.2

Density 21                                                              20.5

Melting point 3300 K                                           3 323 K

The higher oxide formula X27                                  Re2O7

Melting point of the higher

Oxide 400-500oC                                                 220oC

The agreement is, indeed, excellent. Only the melting pint of the oxide proved to be much lower that

the expected one whereas on the whole Mendeleev’s classical method of prediction was fully confirmed. In other words the Noddacks had a perfectly good idea about what element No. 75 (and element No. 43) was going to be. Thus, the history of rhenium was closely related to the history of its light analogue.

But where to search for these element? Predicting the geochemical behavior of rhenium the Noddacks used to the full the capacity of theoretical geochemistry of that time; They even knew that it had to be a very rare element. They could not know, however, that it was a trace element and that, therefore, what seemed unquestionable to them was in effect open to doubt.

The scientists planned to investigate two groups of minerals: platinum ores and so called columbites (tantalites). Four years (from 1921–1925) were spent in searching for the wanted elements but in vain. Then a communication appeared about the discovery of hafnium whose existence in nature was proved by X-ray spectroscopy. Undoubtedly, this event gave the Noddacks the idea to use the same method in order to prove the existence of manganese analogues and they turned for help to O. Berg, a specialist in X-ray spectroscopy.

In June 1925, V. Noddacks, I. Takke, and O. Berg published an article about the discovery of two missing elements, Masurium (No. 43) and rhenium (No. 75). They were found in columbite and in the Uralian platinum and named after two German provinces. The elements X-ray spectra provided the main confirmation of their existence; but there was no question of extracting the elements and the reasoning of the German scientists was, in general, too involved. However, the article attracted attention and other scientists tried to reproduce the results.

However, no such reproduction followed. A year passed and the Soviet scientist O. E. Zvyagintsev and his colleagues proved irrefutably that the Uralian platinum ore contained no new elements. After that the German scientists continued to study columbites which varied considerably in composition but, according to the predictions, had to contain mysterious manganese analogues. They subjective the minerals to complex chemical treatment in order to concentrate the unknown elements and performed X-ray spectral analysis. The data obtained were reassuring but definite conclusions would have been premature: the scientists could not obtain any noticeable amounts of elements No. 43 and No. 75 and experimentally determine their properties.

Nobody could reproduce the results obtained by the Noddacks. Their compatriot W. Prandtl even sent his assistant. A Grimm to the Noddacks’ laboratory to watch them prepare manganese analogues Back home, A. Grimm reproduced the entire procedure, perfected it and…, we do not know the extent of his distress about the wasted time. The English scientists F. Loring and the Czechs Ya. Geirovskii and Y. Druce also doubted the Noddacks’ results. Later, Loring, Geirovskii, and druce claimed the priority of discovering element No. 75 by other methods and from other sources. History has retained their names but not as discoverers of rhenium.

The two German scientists believed to have also isolated element No. 43 (known later as technetium). Now we know that they by no means could detect the presence of technetium at the time but, nevertheless, the Noddacks were more sure of its discovery than of the discovery of rhenium (the fact which is hardly a feather in their cap). As time passed, the Noddacks became convinced that the range of the minerals for analysis had to be considerably enlarged. The previous geochemical prediction did not, apparently, come true. In the summer of 1926 and in 1927 the Noddacks went to Norway to collect minerals among which were: tantalite, gadolinite, alvite, fergusonite, and molybdenite. In the early 1928 the scientists, analysing the minerals, isolated about 120 mg of rhenium mainly from molybdenite (molybdenum sulphide). Earlier it had never been considered as a possible source of manganese analogues.

Thus, rhenium became, at last, a reality. An end was put to doubts and the symbol Re occupied forever box No. 75 in the periodic table; masurium, however, remained an enigma for a long time.

Hence, 1928 is the date of the reliable discovery of rhenium, the final step in the long process of search. As regards the widely accepted date, 1925, it is only a landmark in the prehistory of the element.

Having planned the directions of research, the Noddacks assembled all publications of supposed discoveries of eka-manganese. Their notes were lost during the second World War but, undoubtedly, the name of the Russian Scientists S. F Kern and the name of the element “devium” were mentioned in them. This may be the most reliable discovery of a new element of all unreliable discoveries. And it is equally possible that the history of element No.75 could have begun 50 years earlier.

The events were as follows. In 1877 reports appeared about the discovery of a new metal “devium” named after H. Davy. The reports aroused great interest and Mendeleev suggested inviting S. F.Kern to report to a session of the Russian Chemical Society. The scientists of Bunsen’s laboratory in Heidelberg decided to check Kern’s result carefully. Later his results were confirmed by two or three other scientists the most interesting fact was that some chemical reactions proved to be identical to those found later for rhenium. Does not it point to the identity of devium and rhenium?

For some reason or other S. F Kern lost interest in his discovery and never returned to the problem after 1878. He had extracted the element from platinum ores, which seems impossible from modern point of view (recall Zvyagintsev’s work in 1926). The fact is, however, that platinum ores have a complex and varied composition. The Uralian ore does not contain rhenium but its presence as traces in ores of other deposits has been proven.

  1. F. Kern studied a very rare sample of platinum ore from Borneo where by that time mines had already been abandoned. At the beginning of the 20th century the Russian chemist G. Chernik worked on the island. Analyzing platinum ores he found a constant mass loss in all samples and tried to explain it by the presence of an unknown element. This element- could well be Kern’s “devium”.

In 1950 Y. Druce devoted a large article to devium. He wrote that if rhenium would be discovered in platinum minerals, this would confirm Kern’s discovery. Samples of platinum ores from Borneo can be found now only in a few mineralogical museums of the world. It would be of interest to analyse them thoroughly. This is a case when the history of a chemical element could be partially changed.