Вы находитесь на странице: 1из 16

Atomic Bonding and Crystal Structures in Metals

The smallest particles of any material are electron, proton and neutron. Among
these three particles, neutrons are electrically neutral and other two particles (electron
and proton) are electrically charged. Proper arrangements of these particles make an
electrically neutral atom.
According to Bohr atomic model, electrons are assumed to revolve around the
atomic nucleus in various discrete orbits and the position of any particular electron is
more or less well defined in terms of its orbital. This model of the atom is represented in
Figure 1.

Figure 1: Arrangement of electron, proton and neutron in an atom (left) and


movement electron around nucleus (mixture of proton and neutron) is shown at right.

Each atom consists of a very small nucleus composed of protons and neutrons,
which is encircled by moving electrons. Both electrons and protons are electrically
charged, the charge magnitude being 1.6x10-19C, which is negative in sign for electrons
and positive for protons. However, neutrons are electrically neutral. These neutral
particles are, possibly, responsible to keep the atom in its stable condition. Masses for
these subatomic particles are infinitesimally small; protons and neutrons have
approximately the same mass, 1.67x10-27 kg, which is significantly larger than that of an
electron, 1.67x10-31 kg. Each chemical element is characterized by the number of
protons in the nucleus or the atomic number (Z). For an electrically neutral or complete
atom, the atomic number also equals the number of electrons. This atomic number
ranges in integral units from 1 for hydrogen to 92 for uranium, the highest of the
naturally occurring elements. The atomic mass (A) of a specific atom may be expressed
as the sum of the masses of protons and neutrons within the nucleus. Although the
number of protons is the same for all atoms of a given element, the number of neutrons
(N) may be variable. Thus, atoms of some elements have two or more different atomic
masses, which are called isotopes. The atomic weight of an element corresponds to the
weighted average of the atomic masses of the atom’s naturally occurring isotopes. The
atomic mass unit may be used for computations of atomic weight.

Atomic Bonding in Solids


Bonding Forces and Energies
An understanding of many of the physical properties of materials is predicated from
the knowledge of the interatomic forces that bind the atoms together. Perhaps the
principles of atomic bonding are best illustrated by considering the interaction between
two isolated atoms as they are brought into close proximity from an infinite separation.
At large distances, the interactions are negligible, but as the atoms approach, each
exerts forces on the other. These forces are of two types, attractive and repulsive and
the magnitude of each is a function of the separation or interatomic distance. The origin
of an attractive force depends on the particular type of bonding that exists between the
two atoms. The magnitude of the attractive force varies with the distance, as
represented schematically in Figure 2. Ultimately, the outer electron shells of the two
atoms begin to overlap and a strong repulsive force comes into play. The net force
between the two atoms is just the sum of both attractive and repulsive components.

Figure 2: Variation in attractive and repulsive forces with interatomic distance.


The magnitude of this bonding energy varies from material to material and depends
on the type of atomic bonding that influences a number of material properties. For
example, materials having large bonding energies typically also have high melting
temperatures; at room temperature solid substances are formed for large bonding
energies, whereas for small energies the gaseous state is favoured. Similarly, liquids
prevail when the energies are of intermediate magnitude. Three different types of
primary or chemical bond are found in solids: ionic, covalent and metallic. For each
type, the bonding necessarily involves among the valence electrons; furthermore, the
nature of the bond depends on the electron structures of the constituent atoms. In
general, each of these three types of bonding arises from the tendency of the atoms to
be stable electron structures, like those of the inert gases by completely filling the
outermost electron shell. Secondary or physical forces and energies are also found in
many solid materials they are weaker than the primary ones. But, nonetheless, they
influence the physical properties of some materials. The following sections will explain
the several kinds of primary and secondary interatomic bonds.

Ionic Bonding
Ionic bonding is perhaps the easiest to describe and visualize the bonding that
exists in various materials. It is always found in compounds that are composed of both
metallic and nonmetallic elements, elements that are situated at the horizontal
extremities of the periodic table. Ionic bonding is the complete transfer of valence
electron(s) between atoms. It is a type of chemical bond that generates two oppositely
charged ions. In ionic bonds, the metal loses electrons to become a positively charged
cation, whereas the non-metal accepts those electrons to become a negatively charged
anion. Ionic bonds require an electron donor (i.e. metal) and an electron acceptor
(preferably nonmetal). Ionic bonding is observed between metals and nonmetals,
because metals have few electrons in its outer-most orbital. By losing those electrons,
these metals can achieve noble-gas configuration and satisfy the octet rule more easily.
Similarly, nonmetals that have close to 8 electrons in its valence shell tend to readily
accept electrons to achieve its noble gas configuration, Figure 3 (ionic bonding in
sodium and chlorine). At the most outer shell, sodium atom has only one electron, which
is easy to lose for becoming sodium ion. On the other hand, chlorine has seven
electrons at the outermost shell, which is difficult to give up. In this situation, the most
comfortable option is to receive one electron from sodium atom to be negatively
charged ion. Thus, the formed positively charged sodium ion and negatively charged
chlorine ion attracts each other and they form sodium chloride by means of ionic bond.
In ionic bonding, more than 1 electron can be donated or received to satisfy the octet
rule. The charge on the anion and cation corresponds to the number of electrons
donated or received. In ionic bonds, the net charge of the compound must be zero.
Figure 3: Outer shell electrons in sodium and chlorine.

Sodium chloride (NaCl) is the classic ionic material. A sodium atom can assume the
electron structure of neon (and a net single positive charge) by a transfer of its one
valence electron to a chlorine atom (Figure 3). After such a transfer, the chlorine ion has
a net negative charge and an electron configuration identical to that of argon (Figure 3).
In sodium chloride, all the sodium and chlorine exist as ions. This type of bonding in
sodium chloride salt is illustrated schematically in Figure 4.

Figure 4: Molecular configuration of sodium chloride.

Ionic bonding is termed nondirectional; that is, the magnitude of the bond is equal in
all directions around an ion. It follows that for ionic materials to be stable, all positive
ions must have as nearest neighbours negatively charged ions in a three dimensional
scheme and vice versa. The predominant bonding in ceramic materials is ionic. Ionic
materials are characteristically hard and brittle and, furthermore, electrically and
thermally insulative.
Covalent Bonding
Covalent bonding is the sharing of electrons between atoms. This type of bonding
occurs between two of the same element or elements close to each other in the periodic
table. This bonding occurs primarily between nonmetals; however, it can also be
observed between nonmetals and metals as well. When molecules have similar
electronegativity, same affinity for electrons, covalent bonds are most likely to occur.
Since both atoms have the same affinity for electrons and neither is willing to donate
them, they share electrons in order to achieve octet configuration and become more
stable. In addition, the ionization energy of the atom is too large and the electron affinity
of the atom is too small for ionic bonding to occur. In the case of di-atomic gases,
molecules are formed by covalent bonding between two atoms, e.g. chlorine gas
(Figure 5).

Figure 5: Covalent bonding between chlorine atoms to form chlorine molecule.

Carbon does not form ionic bonds since it has 4 valence electrons, half of an octet.
To form ionic bonds, carbon molecules must either gain or lose 4 electrons. This is
highly unfavorable; therefore, carbon molecules share their 4 valence electrons through
single, double and triple bonds so that each atom can achieve noble gas configurations.
Covalent bonds lead to formation of single, double, triple and quadruple bonds.
Schematic diagram of covalent bond between carbon and hydrogen atoms are shown in
Figure 6.
Figure 6: Covalent bonding between hydrogen and carbon to form a methane molecule.

Many nonmetallic elemental molecules (H2, O2, F2, etc) as well as molecules
containing dissimilar atoms, such as H2O, HNO3 and HF are covalently bonded.
Furthermore, this type of bonding is found in elemental solids such as diamond
(carbon), silicon and germanium and other solid compounds composed of elements that
are located on the right-hand side of the periodic table, such as gallium arsenide
(GaAs), indium antimonide (InSb) and silicon carbide (SiC). In general, covalent bonds
are very strong and rigid with poor level of ductility.

Metallic Bonding
Most metals have very few electrons in their outermost energy shells and some
have vacant outer electron orbitals. What this means for the metal is that its valence
electrons are decentralized and free to move around. Remember that in ionic bonds,
the electrons transfer from one atom to another atom. In covalent bonds, the electrons
are shared between atoms. In metallic bonds, the electrons wander around and aren't
transferred or shared. It is more of a communal thing where they belong to all the
metal atoms around them.
Metals form compact and orderly crystalline structures. When metals are next to
each other, the valence electrons don't just stay on their own atom; they roam around
the whole metal complex. They float free as though floating through a sea of electrons
much like an individual water molecule floats free in the sea, Figure 7. This is why it is
called the electron sea model.
Figure 7: Floating positive iron cores in moving electron in sea of electrons in
metallic bond.

Each metal atom allows its electrons to roam freely. So, these atoms become
positively charged cations. In other way, these cations can be thought as positively
charged islands and are surrounded by a sea of negatively charged electrons. In most
cases, the outermost electron shell of each of the metal atoms overlaps with a large
number of neighbouring atoms. As a consequence, the valence electrons continually
move from one atom to another and are not associated with any specific pair of atoms.
In short, the valence electrons in metals, unlike those in covalently bonded substances,
are nonlocalized, capable of wandering relatively freely throughout the entire crystal.
The atoms that the electrons leave behind become positive ions and the interaction
between such ions and valence electrons gives rise to the cohesive or binding force that
holds the metallic crystal together.
Many of the characteristic properties of metals are attributable to the non-localized
or free-electron character of the valence electrons. This condition, for example, is
responsible for the high electrical conductivity of metals. The valence electrons are
always free to move when an electrical field is applied, Figure 8. The presence of the
mobile valence electrons, as well as the nondirectionality of the binding force between
metal ions, account for the malleability and ductility of most metals. When a metal is
shaped or drawn, it does not fracture, because the ions in its crystal structure are quite
easily displaced with respect to one another. Moreover, the nonlocalized valence
electrons act as a buffer between the ions of like charge and thereby prevent them
from coming together and generating strong repulsive forces that can cause the crystal
to fracture.
+
Aplied Electric
Field

Figure 8: Random (without electric field) and directional (in electric field) movement of
electron in Al.
van der Waals Bonding
Water molecules in liquid water are attracted to each other by electrostatic forces
and these forces have been described as van der Waals forces or van der Waals
bonds. Even though the water molecule as a whole is electrically neutral, the distribution
of charge in the molecule is not symmetrical and leads to a dipole moment; a
microscopic separation of the positive and negative charge centers, Figure 9. This leads
to a net attraction between such polar molecules which finds expression in the cohesion
of water molecules and contributes to viscosity and surface tension. Perhaps it is fair to
say that van der Waals forces are what hold water in the liquid state until thermal
agitation becomes violent enough to break those van der Waal bonds at 100°C.

Figure 9: van der Waal’s bonding in water molecules.

Crystal Structures in Solids

Introduction
The properties of some materials are directly related to their crystal structures. For
example, pure and undeformed magnesium and beryllium, having one crystal structure,
are much more brittle (i.e., fracture at lower degrees of deformation) than are pure and
undeformed metals such as gold and silver that have yet another crystal structure.
Furthermore, significant property differences exist between crystalline and
noncrystalline materials having the same composition. For example, noncrystalline
ceramics and polymers normally are optically transparent; the same materials in
crystalline (or semicrystalline) form tend to be opaque or, at best, translucent, Figure 10.
Figure 10: Change in transparency in aluminum oxide depending on processing such
as single crystal (left), polycrystal (middle) and polycrystal with many defects (right).

Solid materials may be classified according to the regularity with which atoms or
ions are arranged with respect to one another. A crystalline material is one in which the
atoms are situated in a repeating or periodic array over large atomic distances; that is,
long-range order exists. In crystalline materials, upon solidification, the atoms will
position themselves in a repetitive three-dimensional pattern. All metals, many ceramic
materials and certain polymers form crystalline structures under normal solidification
conditions. For those that do not crystallize, this long-range atomic order is absent.
Some of the properties of crystalline solids depend on the crystal structure of the
material, the manner in which atoms, ions, or molecules are spatially arranged. There is
an extremely large number of different crystal structures all having long range atomic
order; these vary from relatively simple structures for metals to exceedingly complex
ones. The present discussion deals with several common metallic crystal structures.

Metallic Crystal Structures


In the earlier section, it has been mentioned that electron, proton and neutron are
the smallest particles that make a neutral atom. In each atom, these smallest particles
occupy the position in periodic fashion. The atomic order in crystalline solids indicates
that small groups of atoms form a repetitive pattern that make a unit cell for the metallic
crystals. A unit cell is chosen to represent the symmetry of the crystal structure, wherein
all the atom positions in the crystal may be generated by translations of the unit cell
integral distances along each of its edges. Thus, the unit cell is the basic structural unit
or building block of the crystal structure. When these unit cells occupy positions in well
defined periodic fashion, then crystals are formed. We know that the atomic bonding in
metallic elements is nondirectional in nature. Consequently, there are restrictions as to
the number and position of nearest-neighbor atoms; this leads to relatively large
numbers of nearest neighbors and dense atomic packing for most metallic crystal
structures. There are various crystal structures, however, three relatively simple crystal
structures are found for most of the common metals, which are body-centered cubic,
face centered cubic and hexagonal close-packed.

Body-Centered Cubic (BCC) Crystal Structure


The crystal structure found for many metals has a unit cell of cubic geometry. BCC
is a common metallic crystal structure with cubic unit cell. Here atoms located at all
eight corners and a single atom at the centre of a cube. This is called a body-centered
cubic (BCC) crystal structure. Various features of a BCC crystal structure is shown in
Figure 11.

Figure 11: Various features of a BCC cell; location of atoms (right) and sharing of
atoms of a unit cell with the neighbouring cells.

In BCC crystal, center and corner atoms touch one another along cube diagonals.
There are eight corners in BCC crystal, where each corner atom is shared among eight
unit cell neighbor corners. So, each cell gets one atom from corner (8x1/8) and that the
body atom is not shared with any other unit cell as it is within the body of the cell, which
is out of contact of the surrounding neighbouring unit cells. Altogether, each BCC unit
cell has two atoms. Suppose, length of side of unit cell is a and radius of one atom is R.
So, the face diagonal length Df can be calculated as follows:

Similarly, the body diagonal length of the BCC unit cell is possible to calculate as
follows:
Now, volume of two atoms and the unit cell will be:
and

So, the atomic packing factor (APF) of BCC unit cell will be:

From the above calculation, it is mathematically established that there is 32%


vacant space in any BCC unit cell. It has already been mentioned that there is one atom
each corner and one atom at the centre of the unit cell. This information will be very
useful to explain properties of metals having BCC crystal.

Face-Centered Cubic Crystal Structure


Another type of cubic crystal structure found for many metals has atoms located at
each of the corners like BBC unit cell and the centers of all the cube faces unlike to that
of the BCC cell, where one atom is located inside the body (at the centre). Unit cell
having atoms at the centre of each faces is termed as face-centered cubic (FCC) crystal
structure. Some of the familiar metals having this crystal structure are copper,
aluminum, silver and gold. Figure 12 shows various features of a FCC unit cell.

Figure 12: Various features of a FCC cell; location of atoms (right) and sharing of
atoms of a unit cell with the neighbouring cells.

In FCC crystal, corner and face atoms touch one another along cube diagonals.
Similar to BCC cell, there are eight corners in FCC unit cell, where each corner atom is
shared among eight corners. So, each cell gets one atom from corner (8x1/8) and that
the face atom is shared between two adjacent faces. There are six faces in the cubic
cell. So, each unit cell will have three atoms from six faces (6x1/2). Altogether, each
BCC unit cell has four atoms.
From above figure it is clear that the atoms touch one another across a face-
diagonal the length of which is 4R. Since the unit cell is a cube, its volume is a3, where
a is the cell edge length. From the right triangle on the face of FCC unit cell presented in
Figure (right one),

The volume of the FCC cell Vcell may be computed from the following equation:

We know that FCC crystal has four atoms per unit cell. Now, the total volume V atom
of the four atoms will be:

Now, the atomic packing factor of FCC crystal is:

Similar to BCC crystal, it is now clear that there is empty space in the FCC crystal,
which is about 26%. From the above calculation, it is mathematically established that
FCC crystals are more closely packed than BCC one.

Hexagonal Close-Packed Cell


Not all metals have unit cells with cubic symmetry. There are many metals that
show hexagonal crystal structure. Figure 13 shows various features of this structure,
which is commonly termed as hexagonal close-packed (HCP) cell.
b
o
30
a/2
a/2

Figure 13: Various features of a HCP cell; location of atoms (right) and sharing of
atoms of a unit cell with the neighbouring cells.

The top and bottom faces of the unit cell consist of six atoms that form regular
hexagons and surround a single atom at the centre of each hexagon. Another plane that
provides three additional atoms to the unit cell is situated between the top and bottom
planes. The atoms in this mid-plane have as nearest neighbors atoms in both of the
adjacent two planes. The equivalent of six atoms is contained in each unit cell; one-sixth
of each of the 12 top and bottom face corner atoms, one-half of each of the 2 center
face atoms and all 3 mid-plane interior atoms. If a and c represent, respectively, the
short and long unit cell dimensions of the c/a ratio should be 1.633. However, for some
HCP metals this ratio deviates from the ideal value. The atomic packing factor for the
HCP crystal structure is the same as for FCC (0.74). The basic calculations for finding
APF of HCP cell is shown below.
So, the atomic packing factor of HCP cell is possible to calculate from the following
general equation:

The HCP metals include cadmium, magnesium, titanium and zinc. Some common
metals belonging to BCC, FCC and HCP cells are shown in Table 1.

Metal Crystal Structure Metal Crystal Structure


Aluminum FCC Molybdenum BCC
Cadmium HCP Nickel FCC
Chromium BCC Platinum FCC
Cobalt HCP Silver FCC
Gold FCC Tantalum BCC
Copper FCC Titanium ( ) HCP
Iron ( ) BCC Tungsten BCC
Lead FCC Zinc HCP

In the earlier section, it been mentioned that crystal structures control many
properties of the metallic materials. Steel is the most widely used engineering material,
which is an alloy of iron and carbon. In the iron carbon alloy, iron is the solute and
carbon is the solvent. We know that iron is an allotropic material, which mean that it can
stay in different crystallographic states (BCC or FCC) depending on temperature or
presence of other alloying elements. As per APC, BCC cell has more vacant space than
that of the FCC cell, however, solubility of carbon in FCC cell is significantly higher than
that of the FCC cell. The reason behind this is the arrangement of atoms in these two
cells, which will be discussed in the phase diagram chapter in details.
Another important property is the ductility or deformability of metals, which is the
key property of many metallic materials for their applications. For example, consider
gold and copper, which are very ductile in nature. High level of ductility or deformability
without any form of cracking is one of the key properties that enables to make fine wire
or ornaments from them. On the other hand, being a most useful engineering metallic
metal, iron is not as deformable as gold or copper. This is also possible to explain in
terms of their crystal structures, i.e. the position of atoms in their unit cells. Gold and
copper have FCC crystals, on the other hand iron has BCC cell. Because of presence of
atoms at each face of FCC crystal, they have large numbers of close packed plane (five
atoms in each plane), which is limited for BCC crystal, Figure 14.

Figure 14: Figure showing close-packed planes in FCC (left) and BCC (right) cells.

Because of presence of planes of high atomic density in almost all directions of


FCC crystals, they are more deformable. However, for BCC crystals, planes of high
atomic density are limited. They are deformable in certain direction. But in the case of
bulk material deformation, deformability in all directions is essential. As a result, BCC
crystals are, in general, deformable to a limited extent.

Вам также может понравиться