Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

The Group 18 Elements: The Noble Gases

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 3

The Group 18 Elements:

The Noble Gases


The noble gases make up the least reactive group in the periodic table. In fact,
xenon is the only noble gas to form a wide range of compounds—and then
only with highly electronegative elements. It is doubtful that stable chemical
compounds will ever be made of helium and neon
Although as early as 1785 it had been noted that something else was in air besides oxygen and nitrogen, it was
not until 100 years later that Sir William Ramsay showed that this other gas produced a previously unknown
spectrum when an electric discharge was passed through it. Because every element has a unique spectrum, the
gas producing the new spectrum had to be a new element. He named it argon, from the Greek word for
“lazy,” because of its unreactive nature, and he suggested that it might be the first member of a new group in
the periodic table.
In fact, one other element in this group—helium—had already been discovered in 1868, but not on Earth.
Observations of the spectrum of the Sun had shown some lines that did not belong to any element known at
that time. The new element was named helium, the first part of the name indicating that it was first discovered
in the Sun (Greek, helios) and the ending indicating that it was expected to be a metal. The element was first
isolated on Earth in 1894 from uranium ores, and a few years later, it was realized that helium is produced
during the radioactive decay of uranium and its daughter elements. In 1926 it was suggested that the name of
the element be changed to helion to indicate that it was not a metal, but by that time the former name was too
well established. Every one of the noble gases was first identified by its unique emission spectrum. Hence, it
was really physical chemists rather than inorganic chemists who founded the study of this group of elements.
18.1 Group Trends
All the elements in Group 18 are colorless, odorless, monatomic gases at room temperature. They neither burn
nor support combustion; in fact, they make up the least reactive group in the periodic table. The very low
melting and boiling points of the noble gases indicate that the dispersion forces holding the atoms together in
the solid and liquid phases are very weak. The trend in the melting and boiling points, shown in Table 18.1,
corresponds to the increasing number of electrons and, hence, greater polarizability.
Table 18.1 Melting and boiling points of the noble gases
Noble gas Melting point (°C) Boiling point (°C) Number of electrons
He — -269 2
Ne -249 -245 10
Ar -189 -186 18
Kr -157 -152 36
Xe -112 -109 54
Rn -71 -62 86
Because the elements are all monatomic gases, there is a well-behaved trend in densities at the same pressure
and temperature. The trend is a simple reflection of the increase in molar mass (Table 18.2). Air has a density
of about 1.3 g*L-1; so, relative to air, helium has an extremely low density. Conversely, radon is among the
densest of gases at SATP.
TABLE 18.2 Densities of the noble gases (at SATP)
Noble gas Density (g?L21) Molar mass (g?mol21)
He 0.2 4
Ne 1.0 20
Ar 1.9 39
Kr 4.1 84
Xe 6.4 131
Rn 10.6 222
As we discussed in Chapter 10, Section 10.6, until 1962, the only known species involving the noble gases
were the clathrates, in which the noble gas atoms are trapped within ice cages. These clathrates have now
become of interest again to explain why the atmosphere of Saturn’s moon Titan lacks the krypton and xenon
that should be present from the forming of this large moon. Cosmochemists currently think that these gases
have been trapped within ice clathrates on the moon’s surface ever since the solar system formed and the
moon’s surface cooled. To date, chemical compounds have been isolated at room temperature for only the
three heaviest members of the group: krypton, xenon, and radon. Few compounds of krypton are known,
whereas xenon has an extensive chemistry. The study of radon chemistry is very difficult because all the
isotopes of radon are highly radioactive.
18.2 Unique Features of Helium
Helium is still a liquid at the lowest temperatures we can reach. In fact, even at a temperature of 1.0 K, a
pressure of about 2.5 MPa is required to cause it to solidify. However, liquid helium is an amazing substance.
At a pressure of 100 kPa, the gas condenses at 4.2 K to form an ordinary liquid (referred to as helium I), but
when cooled below 2.2 K, the properties of the liquid (now helium II) are dramatically different. For example,
helium II is an incredibly good thermal conductor, 106 times greater than helium I and much better than even
silver, the best metallic conductor at room temperature. Even more amazing, its viscosity drops to close to 0.
When helium II is placed in an open container, it literally “climbs the walls” and runs out over the edges.
These and many other bizarre phenomena exhibited by helium II are best interpreted in terms of quantum
behavior in the lowest-possible energy states of the element. A full discussion is in the realm of quantum
physics.
18.3 Uses of the Noble Gases
All the stable noble gases are found in the atmosphere, although only argon is present in a high proportion
(Table 18.3). Helium is found in high concentrations in some underground natural gas deposits, where it has
been accumulating from
TABLE 18.3 Abundance of the noble gases in the dry atmosphere
Noble gas Abundance (% by moles)
He 0.000 52
Ne 0.001 5
Ar 0.93
Kr 0.000 11
Xe 0.000 008 7
Rn Trace
the decay of radioactive elements in the Earth’s crust. As a result of the low molecular mass and hence high
average velocity of the helium molecules, atmospheric helium is rapidly lost into space. Gas reservoirs in the
southwestern United States are among the largest in the world, and the United States is the world’s largest
producer of helium. In fact, the discovery of the deposits in the 1920s caused the price of helium gas to drop
from $88 per liter (1915) to $0.05 per liter (1926).
Because it is the gas with the second-lowest density (dihydrogen having the lowest), helium is used to fill
balloons. Dihydrogen would provide more “lift,” but its flammability is a major disadvantage. Almost
everyone has heard of the Hindenburg disaster, the burning of a transatlantic airship in 1937. Yet few are
aware that the airship was designed to use helium. When the National Socialist Party came to power in
Germany in the 1930s, the U.S. government placed an embargo on helium shipments to Germany, fearing that
the gas would be used for military purposes. Thus, when the airship was completed, dihydrogen had to be
used. However, there is considerable evidence now that hydrogen was not the cause of the disaster and that it
was, in fact, the aluminum flakes in the varnish of the skin of the airship that caught fire.
Today, the public thinks of airships solely in their advertising role. However, they have also been used as
long-endurance flying radar posts by the U.S. Coast Guard to identify illegal drug-carrying flights. An airship
has also been used to study the upper canopy of the rain forest in the Amazon basin, a vital task that would be
very difficult to do by any other means. Using modern technology, new designs of airships are being
constructed for a variety of tasks, such as ecotourism and heavy lifting.
Helium is used in deep-sea-diving gas mixtures as a replacement for the more blood-soluble nitrogen gas in
air. The velocity of sound is much greater in low-density helium than in air. As many people know, this
property gives breathers of helium “Mickey Mouse” voices. It should be added that the combination of dry
helium gas and the higher frequency of vibrations in the larynx can cause voice damage to those who
frequently indulge in the gas for fun.
Of great scientific importance, liquid helium is commonly used to cool scientific apparatus close to 0 K.
Many pieces of equipment use superconducting magnets to obtain very high magnetic fields, but at present
the coils only become superconducting at extremely low temperatures.
All the other noble gases are obtained as by-products of the production of dioxygen and dinitrogen from air.
Some argon also is obtained from industrial ammonia synthesis, where it accumulates during the recycling of
the unused atmospheric gases. Argon production is quite large, approaching 10 6 tonnes per year. Its major use
is as an inert atmosphere for high-temperature metallurgical processes. Argon and helium are both used as an
inert atmosphere in welding; neon, argon, krypton, and xenon are used to provide different colors in “neon”
lights.
The denser noble gases, particularly argon, have been used to fill the air space between the glass layers of
thermal insulating windows. This use is based on the low thermal conductivities of these gases; for example,
that of argon at 0°C is 0.017J*S-1*m-1*K-1. Dry air at the same temperature has a thermal conductivity of
0.024 J*S-1*m-1*K-1.
The high abundance of argon in the atmosphere is a result of the radioactive decay of potassium-40, the
naturally occurring radioactive isotope of potassium. As mentioned in Chapter 11, Section 11.6, this isotope
has two decay pathways, one of which involves the capture of a core electron to form argon-40:
19 K+-1 e→18 Ar
40 0 40

18.4 A Brief History of Noble Gas Compounds


The story of the discovery of the noble gas compounds has become part of the folklore of inorganic
chemistry. Unfortunately, as with most folklore, the “true” story has been buried by myth. It was in 1924 that
the German chemist Andreas von Antropoff made the suggestion that is obvious to us today: because they
have eight electrons in their valence level, the noble gases could form compounds with up to eight covalent
bonds. Following that, in 1933 American chemist Linus Pauling predicted the formulas of some possible
noble gas compounds, such as oxides and fluorides. Two chemists at Caltech, Don Yost and Albert Kaye, set
out to make compounds of xenon and fluorine. At the time, they thought they had been unsuccessful, but there
is evidence that they did, in fact, make the first noble gas compound.
It was only after Yost and Kaye’s admitted failure that the myth of the inertness of the noble gases spread.
The “full octet” was claimed to be the reason, even though every inorganic chemist knew that many
compounds involving nonmetals beyond the second period violate this “rule.” So, things remained, with this
dogma being accepted by generation after generation of chemistry students, until the upsurge in interest in
inorganic chemistry in the 1960s. It was Neil Bartlett, working at the University of British Columbia, who
then approached the problem from a different direction in 1962.
Bartlett had been working with platinum(VI) fluoride, which he found was such a strong oxidizing agent that
it oxidized dioxygen gas to form the compound O24PtF6- While teaching a first-year chemistry class, he
noticed that the first ionization energy of xenon was almost identical to that of the dioxygen molecule. Despite
the skepticism of his colleagues and students, he managed to synthesize an orange-yellow compound that he
claimed was Xe4PtF6- This reaction was the first proven formation of a compound of a noble gas. However,
the compound did not have this simplistic formula, and it is now believed to have been a mixture of
compounds that contained the XeF1 ion.
Is It Possible to Make Compounds of the Early Noble Gases?
Why are there no chemical compounds of helium, neon, or argon? In fact, this statement is not quite correct. The gas-
phase ion HeH1 was first synthesized in 1925. However, it is true to say that no isolable stable compound of the early
noble gases has been synthesized so far. The known chemistry of krypton might be a guide to the potential chemistry of
argon. The only known binary compound of krypton is krypton difluoride, formed by the reaction of krypton and fluorine
in ultraviolet light at 2196°C. The compound decomposes at about –20°C. Krypton di-fluoride is an extremely strong
oxidizing agent. For example, it will oxidize/fluorinate metallic gold to give (KrF)1[AuF6]2. The fluorokrypton cation is
quite stable and will undergo reactions as a Lewis acid. Theoretical calculations have shown that the binding energy of the
analogous fluoroargon cation should be similar and that (ArF)1[AuF6]2 and similar compounds with large hexa-
fluoroanions should be capable of existence. The challenge is to find a synthetic route. At the time of writing, one argon
compound has been synthesized and conclusively identified, HArF. It was formed by irradiating a mixture of argon and
hydrogen fluoride at about 2255°C. Infrared spectroscopy was used to show that H —Ar and Ar—F covalent bonds indeed
were formed. Unfortunately, the compound decomposes above –245°C. Nevertheless, this first step shows that argon
chemistry may be possible. Will stable compounds of helium and neon ever be made? Theoretical calculations show that
HHeF might be more likely to form than HNeF, because the helium atom is small enough that a three-center bond might
be formed over the three atoms. However, the difficulty will be in finding a reaction pathway. There is a strong possibility
that room-temperature-stable helium and neon compounds will never be made. However, in chemistry, one should never
say never.

Unknown to Bartlett, Rudolf Hoppe, in Germany, had for some years been working with enthalpy cycles, and
he had come to the conclusion that, on thermodynamic grounds, xenon fluorides should exist. By passing an
electric discharge through a mixture of xenon and di-fluorine, he was able to prepare xenon difluoride.
Unfortunately for Hoppe, this discovery came a few weeks after Bartlett’s discovery. Since then, the field of
noble gas chemistry has blossomed. Xenon is the sole noble gas to form a rich diversity of compounds and
then only with electronegative elements, such as fluorine, oxygen, and nitrogen.

You might also like