You are on page 1of 18

Downloaded from http://rsta.royalsocietypublishing.

org/ on February 26, 2017 On


the occurrence of
metallic
character in the periodic table
1
rsta.royalsocietypublishing.org Fachbereich Chemie, Philipps-Universität
Marburg, Hans-Meerwein-Strasse, Marburg
2
35032, Germany Inorganic Chemistry
Laboratory, Department of Chemistry,
University of Oxford, South Parks Road,
Oxford OX1 3QR, UK
Review

Cite this article: Hensel F, Slocombe DR,


Edwards PP. 2015 On the occurrence of metallic
character in the periodic table of the chemical
elements.Phil. Trans. R. Soc. A 373: 20140477.
http://dx.doi.org/10.1098/rsta.2014.0477

One contribution of 18 to a discussion meeting


issue ‘The new chemistry of the elements’.

The classification of a chemical element as either ‘metal’


or ‘non-metal’ continues to form the basis of an instantly
Subject Areas:
recognizable, universal representation of the periodic
physical chemistry, inorganic chemistry
table (Mendeleeff D. 1905 The principles of chemistry, vol.
II, p. 23; Poliakoff M. & Tang S. 2015 Phil. Trans. R. Soc. A
Keywords:
373, 20140211). Here, we review major, pre-quantum-
metals, metal to non-metal transition, periodic table
mechanical innovations (Goldhammer DA. 1913
of chemical elements
Dispersion und Absorption des Lichtes; Herzfeld KF. 1927
Phys. Rev. 29, 701–705) that allow an understanding of
Author for correspondence: the metallic or non metallic status of the chemical
Peter P. Edwards elements under both ambient and extreme conditions. A
e-mail:peter.edwards@chem.ox.ac.uk special emphasis will be placed on recent experimental

of the chemical advances that investigate how the electronic properties


of chemical elements vary with temperature and density,
and how this invariably relates to a changing status of
elements Friedrich Hensel ,
1 the chemical elements. Thus, the prototypical non-
metals, hydrogen and helium, becomes metallic at high
2
Daniel R. Slocombe and Peter P. densities; and the acknowledged metals, mercury,
rubidium and caesium, transform into their non metallic
2
Edwards forms at low elemental densities. This reflects the
fundamental fact that, at temperatures above the
absolute zero of temperature, there is therefore no clear chemical elements into metals and non-metals
dividing line between metals and non-metals. Our (Mendeleeff D. 1905 The principles of chemistry, vol. II, p.
conventional demarcation of chemical elements as metals 23).
or non-metals within the periodic table is of course
governed by our experience of the nature of the elements
under ambient conditions. Examination of these other
situations helps us to examine the exact divisions of the 2015 The Author(s) Published by the Royal Society. All
rights reserved.
Downloaded from http://rsta.royalsocietypublishing.org/ on February 26, 2017

108
2
metals
106
MoTc Rh Ag Al
Na Be Sc Hf Ta ReRu Ni Zn Sn (w)

)1

104 102 1
K
Li
Rb
Cs
Ca Mg Sr
Ba
Y
La
W
Cr Nb
V
ZrTi
Ir
Co Os
Fe
Mn
Cu
Pt Au Pd
In
Ga
Ti Cd Hg (l)
Pb

C (g) Sn (g)
As
Sb
Bi
Po Te

semimetals
..
..
..
...
..
...
r

ya
ls

m – u

non-metals Ge
c

10–2 He semiconductor s ..
b

Se
... l
p

i
1
s

Ω( v n g

it
H Li metalloids and metals B Ar Br Kr l Xe Si
–4
g

10
y

Be Si
..
.... .
t h

C N O F Ne P S Cl
i

semiconductors i

10–6
Na K
Mg Ca
Sc
Ti
V
Cr
Mn
Fe
Co
Ni
Cu
Zn
Al Ga
Ge As Sb
Se Te
At Rn
B
..
...
Ph
n

oc

10–8 10–10
Rb Cs Fr
Sr Ba Ra
Y
La Ac
Zr Hf Rf
Nb Ta Db
Mo W
Sg
Tc Re Bh
Ru Os Hs
Rh Ir
Mt
Pd Pt Ds
Ag Au Rg
Cd Hg

In Sn Po Tl Pb Bi

I
...
..

..
i

l.

10–12 10–14
.

P (w) non-metals
.
c
So

.
. .
A

Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Th
. .

37
. . 3

Pa U Np Pu AmCm Bk Cf Es Fm Md No Lr . . :

. 2
.

IA IIA IIIB IVB VB VIB VIIB VIII IB IIIAIIB IVA VA VIA VIIA 0 VIII 0

.
4

.
0

.
4

Figure 1. Electrical conductivities at room temperature for chemical elements of the periodic table
across s, p and d blocks. .

7
.

The inset represents one conventional designation of elements as metals, metalloids and
semiconductors, and non-metals. (Adapted from [4]). (Online version in colour.)

1. Metals and non-metals of the periodic table


Perhaps the most traditional and common classification of the chemical elements of the periodic table
in textbooks and in chemistry classroom charts—and elsewhere [1]—is in terms of their metallic and
non-metallic properties. Over a century ago, Mendeleeff noted ‘Many elements, although not all of
them, have the peculiar lustre, opacity, malleability and the high thermal and electrical conductivities
peculiar to the metals .... Those which do not possess the physical properties of metals are called non-
metals’ [2]. Perhaps, the most outstanding physical property of a metal is its superb ability to transmit
electricity. Metals are thus magnificent conductors of electricity and ultimately this is what
distinguishes them from non-metals; the broadest, modern definition of metals and the metallic state
is therefore of a substance transmitting electricity by electron transfer [3].
Even armed with room-temperature electrical conductivity data (as shown in figure 1), one can
fairly easily identify elements of the periodic table for which the designation ‘metal’ or ‘non metal’ is
clearly appropriate. The difference in electrical conductivities between a metallic element and a non-
metallic element (under ambient conditions) is itself a striking fact; between the most conductive
elements (copper and silver) and the most resistive (sulfur, for example) it amounts to 23 orders of
+23
magnitude [5]. This range of over 10 in electrical conductivity (or rather its inverse, resistivity, the
quantity customarily used in the application of Ohm’s law) has been cited as one of the most widely
ranging of all physical properties [5,6]. Furthermore—and remarkably—this range in the elements’
electrical property is even larger at low temperatures; the temperature dependence of the electrical
resistivity of non-metals is in every way in direct contrast to that of metals [7]. Then, surely, only the
most rigid purist would seriously question the ‘easy’ designation of mercury as a metal or sulfur, say,
as a non-metal in the periodic table.
However, the very basis of such a conventional classification over time is undoubtedly influenced
by our experience of the nature of the chemical elements existing under ambient conditions (generally
room temperature, room pressure) on our planet. It has, however, been surmised
Downloaded from http://rsta.royalsocietypublishing.org/ on February 26, 2017

very early (without a general proof) that metallic or non-metallic behaviour is not an inherent and 3
unchanging property of any particular chemical element of the periodic table [3]. Once one begins to
consider the extremes of temperature and pressure achievable in the
r

laboratory, or the even more exotic conditions of geophysical and astrophysical relevance, many t

a
.

.
.

more varieties of physical behaviour become possible [7]. o

.
a

This view was beautifully illustrated by the notable discovery of a metallic form of phosphorus l


a century ago by Bridgman, who subjected white phosphorus at 200 C to pressures of .

e
.

t
.

12–15 kbar [8]. This is most certainly the first experimental observation of the transition of an p

b
.

elemental solid from a non-metal to a metal. l


.

s
.

i
.

2. The metal–non-metal transition in fluid chemical elements .

.
o

g
.

(a) Concepts, models and criteria .

.
l

.
T

a
.

n
.

This fundamental issue of the rigorous distinction between the metallic and non-metallic states .
s

R
.

of matter becomes even more important if we include consideration of the fluid state of the .
.

S
.

chemical elements. It has been more than 100 years since Strutt [9] and Bender [10] pointed out the c
.

interesting implications of an undoubted, density-induced metal to non-metal transition for the .

phase behaviour of certain fluid metals at high temperatures and pressures. Strutt noted ‘Mercury .

.
2

vapour is an insulator, while liquid mercury is a conductor. Since the liquid and saturated .

.
4

.
0
vapour are indistinguishable above the critical temperature, one or both of these must undergo .
4

7
.

a remarkable change of electrical properties as that temperature is approached’ [9]. Remarkably, .

Strutt found a reduction of close to eight orders of magnitude in electrical resistance for saturated
mercury vapour at high temperatures and pressures as compared with its resistance at room
pressure, even though experimentally he could not fully approach the critical temperature (1751 K)
and pressure (1673 bar) at that time.
Such a metal to non-metal transition at finite temperature, in fact, raises fundamental questions of
definition of both of these canonical states of matter [7]. As Landau & Zeldovich [11] and Mott (e.g.
[12]) pointed out, there is a fundamental difficulty—theoretically and indeed experimentally—in
drawing a distinction between a metal and a non-metal except at the absolute zero of temperature [7]. We
now recognize that, at the absolute zero of temperature, a metal sharply and fundamentally differs
from a non-metal with respect to its accessible spectrum of electronic energy levels.
The ground state of a metal comprises a continuous spectrum of electronic energy states. This
explains the striking fact that in a metal the application of even the weakest electrical field produces
an electrical current due to transitions to adjacent electronic levels in the continuum of electronic
energy states. By contrast, the energy spectrum of a dielectric, i.e. a non-metal, is characterized by the
existence of a finite electronic energy gap between levels. Thus, a rigorous criterion to distinguish
between a metal and a non-metal can exist only at T = 0 K. There ... ‘.... a metal conducts and a
nonmetal doesn’t’ [13]. At any non-zero temperature, therefore, thermal excitation of conduction
electrons, no matter how few, blurs the difference between metal and non-metal.
The most common modern—that is, quantum mechanical—description of metals and non metals
is that metals contain a partially filled conduction band, whereas non-metals have a completely filled
valence band separated by a large electronic energy gap from an empty conduction band. The
modern theory of electronic band structure provides a comprehensive picture of the behaviour of
electrons in a solid; it derives completely from the underlying periodic lattice structure of atoms in
crystalline metals or non-metals [14]. However, predicting, rationalizing and applying such concepts
and band structures for amorphous or liquid substances require different models (figure 2).
The real difficulty—in either crystalline or amorphous substances—arises in the transition range
where metallic properties evolve continuously into those characteristic of non-metals. Various
approximate criteria exist for defining a boundary between metal and non-metal [15,16].
Downloaded from http://rsta.royalsocietypublishing.org/ on February 26, 2017

4
conduction
band
.

EG ~ .

kT .

EF EF
.

. .

. .

. .
r

. s

t
.

EG >>
r

ya
ls

kT
o

yp

partly filled band conduction band


b

EF .
l

i
.

.
hi

filled band filled band filled band n

.
o

g
.

.
.

metal intrinsic semiconductor insulator .

.
l

.
T

a
.

Figure 2. Electron energy bands for metals with a partially filled conduction band, intrinsic
semiconductors, and insulators .

.
s

with a large electronic energy gap or band [7]. .

.
.

S
.

o
.

c
.

One of the earliest attempts—perhaps the earliest attempt—to explain the occurrence of .

metallic versus non-metallic behaviour of the chemical elements, and of course with that the .
2

first discussion of the metal–non-metal transition, can be traced back to the pioneering, pre .
4

.
0

.
4

quantum-mechanical models of Goldhammer [17] and Herzfeld [18]. Herzfeld’s theory is based .

7
.

on the classical (Lorentz) oscillator model of an atom. Herzfeld pointed out that electrons localized
around atomic nuclei constitute polarizable objects and that their internal dynamics in the dense
assembly of the element leads to local corrections to the polarizing tendency of any external field
impressed on the system.
He proposed that the corresponding element (viewed obviously as a collection of atoms) becomes
metallic when the frequency of the oscillator placed in this dense, dielectric medium approaches zero;
i.e. the valence electron is set free and the element—under those critical conditions—acquires metallic
status.
For an isotropic substance, the Goldhammer–Herzfeld criterion is based upon the Lorenz– Lorentz
or Clausius–Mosotti relation
2
n −1 2
n +2=ε−1
ε + 2 = 4πα R
3VNA = V ,
where n is the index of refraction, ε is the dielectric constant, α is the atomic electronic polarizability,
4
V is the molar volume, NA is Avogadro’s number and R = ( 3πα) NA is the molar refractivity. Thus
for R/V = 1, the refractive index n or the dielectric constant ε diverge. This can only occur if electrons
are no longer bound to individual atoms in the solid or liquid, and now become free, as in a metal.
A simple physical interpretation of the Goldhammer–Herzfeld criterion becomes apparent when
one adopts the model of an isolated atom as a perfectly conducting sphere. It can be shown from
3
electrostatics that the isolated atom’s electronic polarizability is α = r , where r is the atomic radius
4 3
and the molar refractivity is then R = ( 3πNAr ); in reality, nothing more than a measure of the
volume occupied by one mole of these perfectly conducting spheres. On the Goldhammer– Herzfeld
view, metallization therefore occurs when the molar volume (the volume accessible to the atom at the
specific conditions of temperature, pressure and elemental density) becomes equal to or less than R,
the molar refractivity; that is, the system of isolated conducting spheres becomes one large,
macroscopic conductor of electricity.
The accuracy of the empirical Goldhammer–Herzfeld criterion with respect to the metallic
character of the chemical elements of the periodic table at ambient conditions is nicely illustrated in
figure 3. It is seen at a glance from figure 3 that, as one moves from left to the right across the periodic
table, the elements transform from metal to non-metal around the IVA group just as the R/V values
move below the dashed line at the critical R/V = 1 threshold.
Downloaded from http://rsta.royalsocietypublishing.org/ on February 26, 2017

10
5
..

Zr Sc Ti Cr Tc Ru Pt
metals
.
rs

Li Hf Mo
ta

Y Nb W Ta V Re Rh Pd Au
Os Ir

1
Na Cs K
Rb
Ba Be Ca Sr
Mg
Mn
Fe
Co
Ni
Ag Cu
Hg
Cd

Zn
Al
Ga (s)
Sn (β) In
Ga (l)
Pb
Ti
Si
Sn (α)
B
Ge
Bi
Sb
Po
As2
Te2
I2 (s)
.

..
...
..
...
.

ya
ls

yp

V/ l Ph c

H2 He C (d) (w) Cl2 (l) Ne i

Se8 Se2 (am.)


i
l. .

.. s

P4 (w) P2 (w)
R T
... A

Kr
..

S2
h
.... r

..
... a 3

O2 (l)
i

Br2 (I)
...
.. n
. n 7

S8 (y) Ar (s) ..
..
g
s 3

..
. .
.. :

–1 –2
..
o

Xe
R

10 10
..
2

N2 (l)
..

non-metals F2 (l)
. r
u .

Rn
0
g
So
b

IA IIA IIIB IVB VB VIBVIIB VIII VIII IB IIB IIIA IVA VA VIA VIIA O 1

.
4

.
0

.
4

Figure 3. The metallization


of chemical elements of the
periodic table under
ambient conditions imposed
on the Earth. The .

figure shows the


ratio (R/V) for
elements of the s,
p and d blocks of
the periodic table.
Here,Ris the molar
refractivity and V is
the molar volume.
The shaded circles
represent elements
for which both
Rand V are known
experimentally. The open circles are for elements in which Ris calculated and V is known
experimentally. (Adapted from [19,20].

Therefore, if we start with an element of the periodic table at low density—in the non-metallic
state—and continually increase its density (by reducing the spacing between atoms in the solid or
liquid), this leads ultimately to a dielectric catastrophe (ε = ∞) for R ≥ 1 and the constituent atoms are
no longer able to hold on to their valence electrons and metallic conductance will occur. Herzfeld
showed that a large number of normally existing metals and non-metals across the periodic system
do satisfy this condition. He also discussed—in 1927—the possible metallization of Xe and predicted
3 −1
the metallization volume 10.2 cm mol in close agreement with the modern experimentally
3 −1
observed value of 10.7 cm mol [21].

(b) Critical behaviour of fluid metals


As mentioned above, the metal–non-metal transition in elemental fluids is among the conceptually
simplest examples of such an electronic phase transition from the metallic to the non-metallic state,
and vice versa. A liquid metal near its triple point is in natural equilibrium with its corresponding low-
density, non-metallic vapour. There is, therefore, a metal–non-metal transition that coincides exactly
with the liquid–vapour transition, i.e. both the density and the electronic and molecular structures
change when a metallic liquid vaporizes.
For example, liquid mercury and caesium approaching their triple points are considered as
normal liquid metals with electronic properties typical of the condensed state. The remarkably small
changes of basic properties of almost all metals, such as the electrical conductivity or magnetic
susceptibility on melting [22], show that the electronic structure of the liquid metal is indeed quite
similar to that of the crystalline solid, and both liquid and solid can be reasonably well described by
the nearly free-electron approximation. This behaviour is usually rationalized by the fact that the
short-range atomic correlations and the atomic density in the liquid near melting are closely similar
to those of the crystals. Consequently, the liquid metallic phase is usually treated as a monatomic
state, which reflects the solid structure being regarded as single screened metal ions each diffusively
uncoupled from every other.
Downloaded from http://rsta.royalsocietypublishing.org/ on February 26, 2017

In contrast to melting, however, there are substantial, qualitative changes in the very nature 6
of chemical bonding upon evaporation of a liquid metal. At sufficiently low densities, in the vapour
phase, the valence electrons occupy spatially localized atomic or molecular orbitals. We
r

take the example of caesium and mercury for illustration. In this vapour state, many caesium t

a
.

.
.

atoms form chemically bound dimers with a dissociation energy of 0.45 eV, whereas in the o

.
a

vapour phase of mercury, chemically bonded dimers do not form because the ground state of the l

mercury atom arises from a closed-shell electronic configuration, which by itself cannot form an i

e
.

t
.

2
appreciable chemical bond owing to the inert character of the 6s shell. The interaction potential
.

b
.

between two mercury atoms is thus generally regarded as acting between highly polarizable, l
.

s
.

closed-shell systems involving very little electronic density migration from the constituent i
.

partners, in a remarkably similar way to noble gas atoms. In this sense, mercury vapour has .

.
o

g
.

been denoted ‘pseudo-helium’ [23], while the vapour of the alkali metal caesium resembles .

hydrogen [24]. h

.
l

.
T

However, this description applies only to conditions near the triple point, and whether .

a
.

n
.

the liquid–vapour phase change is always and necessarily accompanied by a metal–non-metal .


s

R
.

transition, even up to the liquid–vapour critical point, remains one of the major points at issue, .
.

S
.

o
.

dating back to the work of Landau & Zeldovich [11]. c


.

This almost certainly is not the case in mercury [25] but may be so in the alkali elements .

caesium, rubidium and potassium, for which extensive data exist [26]. Another point at issue is .

.
2

whether localized chemical bonding plays a significant role in liquid alkali metals as the density .

.
4

.
0

is lowered towards the critical density. .


4

7
.

(c) The metallization of fluid hydrogen: the lightest alkali metal


Renewed interest in these questions is motivated by the exciting progress in the search for metallic
character of the two most abundant chemical elements in our Universe, hydrogen and helium. Major
advances have come from the shock wave experiments of Weir et al. [27] in the case of fluid hydrogen
and from first-principles molecular dynamics calculation determined by Stixrude & Jeanloz [ 28] in
the case of fluid helium.
Weir et al. [27] found a highly conducting state of fluid hydrogen (and also for deuterium) at high
temperatures and in a pressure range above 1.4 Mbar. Their measurements at these extreme
−1 −1
conditions showed electrical conductivity values of about 2000 Ω cm , comparable, interestingly
and importantly, with those for expanded supercritical fluid caesium and rubidium [29]. Weir et al.
[27] interpreted their findings in terms of the existence of H 2 molecules within the highly conducting
fluid phase. It has so far, however, not been possible to investigate the structural properties of fluid
hydrogen at the extraordinarily high temperatures and pressures prevailing in the shock wave
experiment in order to prove whether the pairing character persists in the highly conducting state.
However, given the close similarity in the atomic electronic structure of hydrogen and the alkali
metals, an alternative, chemically intuitive approach is to compare and contrast the electrical and
structural properties of hydrogen at high temperatures and high pressures with those of the
experimentally accessible fluid alkali metals, which may serve as models for the metallization of
shock-compressed hydrogen [30–32].
We therefore compare in figure 4 the measured electrical conductivity for compressed fluid hydrogen
and fluid caesium at comparable temperatures (2000–4000 K) over a range of pressures. Importantly,
one can easily see that all three fluid elements reveal a continuous transition from a low-conductivity
to a highly conducting state with increasing density. In the case of fluid hydrogen at these high
temperatures, pressures of more than 1 Mbar are required to effect the transition at an atom density
23 −3
exceeding 3 × 10 cm . By contrast, the alkali metal caesium— recognized to be metallic under
−1 −1
ambient conditions of temperature and pressure—reaches the conductivity of 2000 Ω cm already
21 −3
at the relatively moderate pressure of 100 bar and a concomitant atom density of ca 4 × 10 cm . The
large difference in the transition pressures indicates that the elemental density is the dominating factor
for the transition to the metallic state. This
Downloaded from http://rsta.royalsocietypublishing.org/ on February 26, 2017

caesium
7
3
10
r

a
.

.
.

.
o

)1 t P

1 Cs hydrogen H yp

u
h

Rb
b
m
l.

l
c

.
s

... h
..

10–1
...
.. i
...
.. n
1
....
..
– ... g
.

rubidium
ya
ls
.
o
Ω
o


c

2
r

10 10
i g

01234 12345 .
T

a
.

×1021 ×1023 n
.

.
s

atom density, n (cm–3) .

.
.

S
.

o
.

c
.

Figure 4. Themeasured electrical conductivity of fluid caesium,rubidium and


hydrogen as a function ofthemolar atom density A

nat a temperature of approximately 2000 K. The arrows indicate the predicted


metallization densities for each element, based 7

.
2

on the Goldhammer–Herzfeld model. Adapted from [30–32]. .

.
4

.
0

.
4

7
.

is obvious from figure 4, where it is clear that all three elements undergo a ‘density-induced’
−1 −1
continuous transition to the highly conducting state of 2000 Ω cm .
It is noteworthy—and important—that this conductivity value is comparable to that calculated on
the assumption that high-temperature fluids metallize only, and remain metallic, when the mean free
path of the itinerant conduction electrons becomes comparable to, or exceeds, the mean distance
between the particles supplying the electrons—in this case, the constituent atoms of the various
chemical elements [31,33].
It is therefore reasonable to speculate that the differences in elemental densities at the transition to
metallic behaviour are related to the individual, characteristic atomic properties of the three elements.
One such property, for example, is the static electronic polarizability α of the isolated atom. The
3
relatively small value of α for atomic hydrogen (0.67 Å ) then indicates from the Goldhammer–
Herzfeld view that very high elemental densities are required for the transition to the metallic state of
3 3
hydrogen. By contrast, the high polarizabilities of rubidium (47.3 Å ) and caesium (59.7 Å ) ensure
that these elements are metallic at ambient conditions in the laboratory—which of course is common
experience. The metallization densities predicted by the ‘Goldhammer–Herzfeld criterion’ indicated
by vertical arrows in figure 4 are in remarkably good agreement with the observed onset of high
electrical conductivity for all three chemical elements.
Another property common to compressed fluid hydrogen and the expanded fluid alkali metals
points to their chemical similarity in the periodic table of the chemical elements. The characteristic
atomic property, for example, is the radius of the principal maximum in the electron charge density

of the ns valence orbital, aH , which enters into the venerable Mott criterion for the onset of the
metal–non-metal transition, viz.

1/3 ∗
n c aH = 0.26,

and nc is the critical carrier density for the transition to the metallic state at T = 0K[34,35]. It is
instructive to plot the measured electrical conductivity as a function of the product—the Mott scaling
1/3 ∗
parameter, n aH —where n is the valence electron density.
We include here also the data for expanded Hg as well as data for expanded rubidium, caesium
and compressed hydrogen [30–32].
Downloaded from http://rsta.royalsocietypublishing.org/ on February 26, 2017

non-metal metal
8
3
10
r

a
.

.
.

.
o

2
10 .

Rb .
a

)
.
l

1
s

o

m
i

e
.

t
.

1
10 y

– u .

–1 Cs
10
o

Ω r

Hg
l g
(

H
i

h Ph
.. i
...
.. l.
i
....

1
.. n
...

.
p
g

0.1
.

0.2 0.3 0.4 n1/3 aH*


.
T

r
n R
. .

. .
s .

. S
.
.

Figure 5. The transition to the metallic state for high-temperature (T > 1000 K) fluid caesium,
rubidium, mercury and o

c
.

hydrogen: the dependence of the electrical conductivity on thescaling parametern1/3aH∗. The dotted
line drawn atn1/3aH∗ = .

0.38 indicates the common metallization condition for these chemical elements. To the left of the
metallization condition, .

.
2

we have the non-metallic form of the four elements (non-metallic fluids); to the right, we have the
corresponding metallic .

.
4

state (the metallic fluids). Above the metallization threshold, we anticipate conduction based on the
theory put forward .
0

.
4

by Ziman [36]. .
7

One sees here (figure 5) a clear scaling behaviour for all of hydrogen, the alkali metals, Rb and Cs,
and Hg, and all show the density-induced transition to the conducting metallic state (recall the high
temperatures of these studies). The change in the slope at the conductivity value of approximately
−1 −1
2000 Ω cm (i.e. when the mean free path of the conduction electrons becomes comparable to the
mean distance between the scattering centres) leads one to conclude that all four of these high-
temperature fluids acquire metallic properties at a constant value (approx. 0.38) of the scaling
1/3 ∗
parameter n aH . Under the appropriate experimental conditions, one is therefore led naturally to
identify all three fluid elements, hydrogen, rubidium and caesium, as metallic, with hydrogen now
assuming the position of the lightest alkali metal of group IA of the periodic table.

(d) Molecular dimers: their role in the metallization of hydrogen


and the heavier alkali elements
The nature of molecular hydrogen at high pressures has been extensively investigated. Labet et al.
[37] examine computationally the transition from paired to monatomic hydrogen. They suggest the
possibility of a microscopically non-crystalline phase of hydrogen at elevated pressures, one in which
there is a substantial range of roughly equi-enthalpic geometries available to the system.
Weir et al. [27] assumed for fluid hydrogen that the tendency of the structure to dimerize persists
up to the densities and temperatures prevailing in the shock experiment. Then the question emerges
whether dimers and other clusters can be detected by structural measurements in expanded fluid
alkali metals. Naturally, to pose this question is to ask further ‘What is the relation of dimer
formation to the metal–non-metal transition?’
The underlying view [38,39] is that the transition in hydrogen and the alkali metals may begin
with metallization of the diatomic system by an overlap of the molecular valence and conduction
bands (figure 2) and this is followed at higher density by a gradual transition into the monatomic,
metallic state. Thus, the dimer molecules present in the vapour may well survive
Downloaded from http://rsta.royalsocietypublishing.org/ on February 26, 2017

104
9
3
10 r

a
.

.
.

102 r

.
o

.
)
a

.
l
1

.

10 .

c
e
.

t
.

.

1 .

u
Ω
.

b
.

l
.

σ
i

s
.

10–1 i
.

He g

–2
10 Hg
o

g
.

0.15 0.20 0.25 0.30 0.35 0.40 i

.
l

.
T

1/3
n aH* .

a
.

n
.

.
s

R
.

Figure6. The calculated conductivity values ofHe(at10 000 K) andthe experimentally determined
conductivity values ofHg(at .
.

S
.

o
.

1800 K) as a function of thescaling parameter,n1/3aH∗. Theshaded region indicates the transition


region to metallic behaviour. c
.

This transition range is indicated as the onset of band-structure effects [28,42]. .

.
2

.
4

.
0

the passage to a low-density liquid. Interesting insights regarding up to what density remnants .
4

7
.

of the diatomic unit, present in the dilute vapour phase, can survive have been obtained from .

measurements of the dynamic structure factor of fluid rubidium expanded over a relatively wide
1/3 ∗
density range [39]. At a scaled density n aH = 0.41, the observed dynamic structure factor is in
accord with molecular dynamics calculations [40] employing effective pair potentials derived by
pseudopotential theory, typical for monatomic metals. By contrast, at a lower scaled density of
1/3 ∗
n aH = 0.37, drastic changes in the dynamic structure factor are observed and a well-defined
excitation peak appears at about hω = 3.2 meV with the maximum intensity at a momentum transfer
−1
near 1 Å . This observation was interpreted as an optic-type mode [41] in which two atoms tend to
move in opposite directions.
The measured excitation energy of 3.2 meV is surprisingly close to that calculated for Rb 2
molecules on a simple cubic lattice [39].
Thus, the results of the calculation and the experiment support the view that rubidium, which is
stable as a diatomic molecule at low elemental density, may undergo a density-induced molecule-to-
metal transition at a scaled density lower than 0.38, which precedes the occurrence of monatomic Rb,
which is analogous to the situation that is suggested to occur in shock-compressed hydrogen in the
transition to the conducting metallic phase of hydrogen.
It has been discussed very early (specifically for mercury [42,43]) that there exists a major
difference in the way the metallic state develops in divalent systems like mercury and helium and
monovalent systems like alkali metals and hydrogen. As is well known, an expanded divalent metal
such as mercury is predicted by the Bloch–Wilson band model to undergo a metal– semiconductor
transition when the s valence and the p conduction bands no longer overlap. In a crystal, a real gap
appears and widens as the density decreases. Mott proposed, specifically for mercury, that the
disordered fluid retains features of the crystal’s electronic structure, i.e. a minimum in the density of
states at the chemical potential of the electrons develops with decreasing density. As the results of the
recent work of Stixrude & Jeanloz [28] for helium are completely consistent with Mott’s scenario
developed for mercury, it is certainly tempting to compare experimentally determined properties of
1/3 ∗ ∗
mercury with the calculated data for helium at scaled densities n aH with a now the s-orbital
radius of the isolated helium or mercury atoms, respectively.
Figure 6 shows the transition to the metallic state for the high-temperature conditions of the
chemical elements helium and mercury, plotting the dependence of the electrical conductivity on the
1/3 ∗
Mott scaling parameter, n aH [44]. The shaded region indicates the transition region to
Downloaded from http://rsta.royalsocietypublishing.org/ on February 26, 2017

metallic behaviour. It is not, however, trivial for the experimentalist to determine the metallization 10
condition in this instance, since the common (and often used) dependence of the sign of the
conductivity change with changing temperature is not obeyed by divalent metals. The sign of
r

the conductivity dependence is always positive at constant volume [36]. This is observed here for t

a
.

.
.

He and Hg, in contrast to the alkali metals at higher densities. o

.
a

Stixrude and Jeanloz [28] argue that the band gap is closed, as s–p hybridization causes the l

valence band to broaden with increasing temperature. Indeed, it is suggested that the valence i

e
.

t
.

band broadens by a factor of 2 from 0 to 50 000 K. The calculated scaled gap closing density p

b
.

of mercury is 0.329 at 10 000 K, which agrees remarkably with the measured scaled closing gap l
.

s
.

density 0.333 [44]. i


.

In Figure 6, the shaded region indicates the transition region to metallic behaviour. This .

.
o

g
.

transition range is indicated as the onset of band-structure effects [28,42] (pseudo-gaps or minima .

.
in the density of states, i.e. Mott‘s famous g-factor). For the lower limit of conductivity (about 200– h

.
l

.
T

250), both Mott and Stixrude assume that metallization occurs close to densities and temperatures .

a
.

n
.

of gap closing. The upper limit (about 2000) is the ubiquitous Ioffe–Regel limit (i.e. for a .
s

R
.

scaled density 0.358 in the case of Hg). Above this density, Hg is a nearly free-electron metal .
.

S
.

o
.

[25,36,42]. In the shaded range of figure 6, band-structure effects modify the Ziman formula c
.

via Mott’s g-value. Stixrude et al. claim that in the case of He the value of g is everywhere less .

than unity. .

.
2

Figure 6 is also illuminating, as it reflects the transition to—and from—the metallic state in .

.
4

.
0

helium, conventionally a non-metal, and mercury (‘pseudo-helium’), conventionally a metal in .


4

7
.

the periodic table of the chemical elements (figure 1). .

3. Concluding remarks
The development of the periodic table of the chemical elements is one of the greatest achievements in
science. The underlying periodic law is now universally recognized as the consequence of the
periodic variation in electronic configuration; its most fundamental ramification leads to either metals
or non-metals in the periodic table. It is remarkable to reflect that in the pre quantum-mechanical
descriptions of Goldhammer [17] and Herzfeld [18], one can understand, and rationalize, the metallic
or non-metallic status of the chemical elements of the periodic table under both ambient and extreme
conditions. The very title of Herzfeld’s [18] paper—‘On atomic properties which make an element a
metal’—beautifully emphasizes the pivotal role of atomic properties in dictating the nature—metal or
non-metal—of a chemical element in the condensed state (either amorphous or crystalline). Thus, the
instantly recognizable ‘diagonal relationship’ separating metals from non-metals in the periodic table
is now rationalized and forms the cornerstone of our conventional, universally recognized
classification of the chemical elements of the periodic table [1].

References
1. Poliakoff M, Tang S. 2015 The periodic table: icon and inspiration. Phil. Trans. R. Soc. A 373,
20140211. (doi:10.1098/rsta.2014.0211)
2. Mendeleeff D. 1905 The principles of chemistry, vol. II, p. 23, 3rd English edn. London, UK:
Longmans, Green.
3. Bernal JD. 1929 Part III. Metals. The problem of the metallic state. Trans. Faraday Soc. 1, 367–379.
(doi:10.1039/tf9292500367)
4. Logan DE, Edwards PP. 1985 in The metallic and non-metallic states of matter. p65. London, UK:
Taylor and Francis.
5. Ehrenreich H. 1967 The electrical properties of materials. Sci. Am. 217, 194–204. (doi:10.1038/
scientificamerican0967-194)
6. Bardeen J. 1940 Electrical conductivity of metals. J. Appl. Phys. 11, 88. (doi:10.1063/ 1.1712751)
Downloaded from http://rsta.royalsocietypublishing.org/ on February 26, 2017
7. Edwards PP, Lodge MTJ, Hensel F, Redmer R. 2010 A metal conducts and a non-metal doesn’t. 11
Phil. Trans. R. Soc. A 368, 941–965. (doi:10.1098/rsta.2009.0282)
8. Bridgman PW. 1914 Change of phase under pressure. I. The phase diagram of eleven r

substances with especial reference to the melting curve. Phys. Rev. 3, 153–203. (doi:10.1103/ s

a
.

PhysRev.3.153) .
.

.
o

9. Strutt RJ. 1902 LXVI. The electrical conductivity of metals and their vapours. Phil. Mag. (Ser. 6) .
a

.
l

4, 596–605. (doi:10.1080/14786440209462883) o

e
.

10. Bender J. 1918 Über die kritische Temperatur des Quecksilbers. Phys. Z. 19, 410. t
.

11. Landau L, Zeldovich G. 1943 On the relation between the liquid and the gaseous states of .

b
.

metals. Acta Physicochim. USSR 18, 194. l


.

s
.

12. Mott NF, Davis EA. 1979 Electronic processes in non-crystalline materials, 2nd edn. Oxford, UK: i
.

Clarendon Press. .

.
o

13. Mott NF. 1996 Communication to P. P. Edwards, cited in [7]. g


.

14. Ashcroft NW, Mermin ND. 1976 Solid State Physics. Belmont, CA: Brooks/Cole. P

15. Edwards PP. 1995 The metal–non-metal transition: the continued/unreasonable effectiveness .
l

.
T

of simple criteria at the transition. In Perspectives in solid state chemistry (ed. KJ Rao), pp. 250– a
.

n
.

.
s

271. New Delhi, India: Narosa. .

R
.

.
.

16. Edwards PP, Johnson RL, Rao CNR, Turnstall DP, Hensel F. 1998 The metal– S
.

o
.

insulator transition: a perspective. Phil. Trans. R. Soc. Lond. A 356, 5–22. (doi:10.1098/ c
.

rsta.1998.0146) .

17. Goldhammer DA. 1913 Dispersion und Absorption des Lichtes. Leipzig, Germany: B. G. 3

.
2

Teubner. 0

.
4

18. Herzfeld KF. 1927 On atomic properties which make an element a metal. Phys. Rev. 29, 701– .
0

.
4

705. (doi:10.1103/PhysRev.29.701) 7
.

19. Edwards PP, Sienko MJ. 1982 The transition to the metallic state. Acc. Chem. Res. 15, 87–93.
(doi:10.1021/ar00075a004)
20. Edwards PP, Sienko MJ. 1983 On the occurrence of metallic character in the periodic table of the
elements. J. Chem. Educ. 60, 691. (doi:10.1021/ed060p691)
21. Goettel KA, Eggert JH, Silvera IF, Moss WC. 1989 Optical evidence for the metallization of xenon
at 132(5) GPa. Phys. Rev. Lett. 62, 665–668. (doi:10.1103/PhysRevLett.62.665) 22. Shimoji M. 1977 Liquid
metals. London, UK: Academic Press.
23. Pyykkö P. 1978 Relativistic quantum chemistry. Adv. Quantum Chem. 11, 353–409.
(doi:10.1016/S0065-3276(08)60241-5)
24. Ashcroft NW. 1990 Pairing instabilities in dense hydrogen. Phys. Rev. B 41, 10 963–10 971.
(doi:10.1103/PhysRevB.41.10963)
25. Hensel F. 1995 The liquid–vapour phase transition in fluid mercury. Adv. Phys. 44, 3–19.
(doi:10.1080/00018739500101476)
26. Freyland W, Hensel F. 1985 The metal–non-metal transition in expanded metals. In The metallic
and non-metallic states of matter (eds PP Edwards, CNR Rao), pp. 93–121. London, UK: Taylor and
Francis.
27. Weir ST, Mitchell AC, Nellis WJ. 1996 Metallization of fluid molecular hydrogen at 140 GPa (1.4
Mbar). Phys. Rev. Lett. 76, 1860–1863. (doi:10.1103/PhysRevLett.76.1860)
28. Stixrude L, Jeanloz R. 2008 Fluid helium at conditions of giant planetary interiors. Proc. Natl Acad.
Sci. USA 105, 11 071–11 075. (doi:10.1073/pnas.0804609105)
29. Hensel F, Stolz M, Hohl G, Winter R, Götzlaff W. 1991 Critical phenomena and the metal–
nonmetal transition in liquid metals. J. Phys. IV (Paris) 1 (C5), 191–205. (doi:10.1051/ jp4:1991523)
30. Hensel F, Edwards PP. 1996 Hydrogen: the first metallic element. Science 271 (5256), 1692.
(doi:10.1126/science.271.5256.1692)
31. Hensel F, Edwards PP. 1996 Hydrogen, the first alkali metal. Chemistry, Eur. J. 2 (10), 1201–1203.
(doi:10.1002/chem.19960021005)
32. Hensel F, Edwards PP. 1996 The changing phase of liquid metals. Phys. World 9 (April), pp. 43–46.
33. Ioffe AF, Regel AR. 1960 Non-crystalline, amorphous and liquid electronic semiconductors. Prog.
Semicond. 4, 239.
34. Mott NF. 1961 The transition to the metallic state. Phil. Mag. 6, 287–309. (doi:10.1080/14786
436108243318)
35. Edwards PP, Sienko MJ. 1978 Universality aspects of the metal–nonmetal transition in condensed
media. Phys. Rev. B 17, 2575–2581. (doi:10.1103/PhysRevB.17.2575)
Downloaded from http://rsta.royalsocietypublishing.org/ on February 26, 2017

36. Ziman JM. 1961 A theory of the electrical properties of liquid metals. I. The monovalent 12
metals. Phil. Mag. 6, 1013–1034. (doi:10.1080/14786436108243361)
37. Labet V, Hoffmann R, Ashcroft NW. 2012 A fresh look at dense hydrogen under pressure. r

IV. Two structural models on the road from paired to monatomic hydrogen, via a possible s

a
.

.
.

non-crystalline phase. J. Chem. Phys. 136, 074504. r

.
o

38. Nellis WJ, Louis AA, Ashcroft NW. 1998 Metallization of hydrogen. Phil. Trans. R. Soc. Lond. .
a

.
l

A 356, 119–138. (doi:10.1098/rsta.1998.0153) .

e
.

39. Pilgrim WC, Ross M, Yang LH, Hensel F. 1997 Monatomic–molecular transition in expanded t
.

rubidium. Phys. Rev. Lett. 78, 3685–3688. (doi:10.1103/PhysRevLett.78.3685) u

b
.

40. Hoshino K, Ugawa H, Watabe J. 1992 Temperature dependence of the dynamical structure of .

s
.

expanded rubidium. J. Phys. Soc. Jpn 61, 2182–2185. (doi:10.1143/JPSJ.61.2182) i


.

41. Pilgrim WC, Ross M, Yang LH, Hensel F. 1997 Monatomic–molecular transition in expanded .

.
o

g
.
rubidium. Phys. Rev. Lett. 78, 3685. (doi:10.1103/PhysRevLett.78.3685) .

42. Mott NF. 1972 Conduction in non-crystalline systems. IX. The minimum metallic conductivity. h

.
l

Phil. Mag. 26, 1015–1026. (doi:10.1080/14786437208226973) .


T

a
.

43. Mott NF. 1972 Evidence for a pseudogap in liquid mercury. Phil. Mag. 26, 505–522. n
.

.
s

(doi:10.1080/14786437208230101) R
.

.
.

S
.

44. Hensel F. Warren WW. 1999 Fluid Metals. Princeton, NJ: Princeton University Press. o

c
.

.
2

.
4

.
0

.
4

7
.

You might also like