You are on page 1of 5

Cerámicos

As with metals, the number of different ceramics is vast. But there is no need to
remember them all: the generic ceramics listed below (and which you should remember)
embody the important features; others can be understood in terms of these.
Although their properties differ widely they all have one feature in common: they
Although their properties differ widely, they all have one feature in common: they
are intrinsically brittle, and it is this that dictates the way in which they can be used.
They are, potentially or actually, cheap. Most ceramics are compounds of oxygen,
carbon or nitrogen with metals like aluminium or silicon; all five are among the most
plentiful and widespread elements in the Earth’ss crust. The processing costs may be
plentiful and widespread elements in the Earth crust. The processing costs may be
high, but the ingredients are almost as cheap as dirt: dirt, after all, is a ceramic.

Tipos de cerámicos

1. Vidrios, hechos la mayoría de SiO2 (interesante hacer notar el concepto de


Temperatura de Glass o de transición vítrea)
2. Cerámicas vítreas, o productos de arcillas (tazas, sanitarios, tejas, ladrillos, etc).
F
Fases cristalinas
i t li que se elaboran
l b en fforma pástica
á ti con agua, y luego
l se secan y
se hornean formado una película vitrosa basada en SiO2.
3. Cemento y concreto (Cemento: Mezcla de cal (CaO),silica (SiO2) y alumina
(Al2O3). Concreto: Cemento más arena y piedras).
4 Rocas
4. R y minerales
i l (i
(incluido
l id ell hi
hielo)
l )
5. Cerámicas de alta perfomance:
1. Glasses
Glasses are used in enormous quantities: the annual tonnage is not far below that of aluminium. As much as 80% of the
surface area of a modern office block can be glass; and glass is used in a load-bearing capacity in car windows, containers,
di i bells
diving b ll and
d vacuum equipment.
i t All important
i t t glasses
l are based
b d on silica
ili (SiO2).
(SiO2) T
Two are off primary
i i t
interest:
t common
window glass (soda glasses), and the temperature-resisting Borosilicate glasses.

2. Vitreous ceramics
Potters have been respected members of society since ancient times. Their products have survived the ravages of time better
than any other; the pottery of an era or civilisation often gives the clearest picture of its state of development and its customs.
customs
Modern pottery, porcelain, tiles, and structural and refractory bricks are made by processes which, though automated, differ
very little from those of 2000 years ago. All are made from clays (arcillas), which are formed in the wet, plastic state and then
dried and fired. After firing, they consist of crystalline phases (mostly silicates) held together by a glassy phase based, as
always, on silica (SiO2). The glassy phase forms and melts when the clay is fired, and spreads around the surface of the
inert, but strong, crystalline phases, bonding them together.

3. Cement and concrete


Cement and concrete are used in construction on an enormous scale, equalled only by structural steel, brick and wood.
Cement is a mixture of a combination of lime (CaO), silica (SiO2) and alumina (Al2O3), which sets when mixed with water.
Concrete is sand and stones (aggregate) held together by a cement.

4. Natural ceramics
Stone is the oldest of all construction materials and the most durable. The pyramids are 5000 years old; the Parthenon 2200.
Stone used in a load-bearing capacity behaves like any other ceramic; and the criteria used in design with stone are the
same. One natural ceramic, however, is unique. Ice forms on the Earth’s surface in enormous volumes: the Antarctic ice cap,
instance, is up to 3 km thick and almost 3000 km across; something like 1013 m3 of pure ceramic
for instance ceramic. The mechanical
properties are of primary importance in some major engineering problems, notably ice breaking, and the construction of
offshore oil rigs in the Arctic

5. High-performance engineering ceramics:


Diamond, of course, is the ultimate engineering ceramic; it has for many years been used for cutting tools, dies, rock drills,
and as an abrasive. But it is expensive. The strength of a ceramic is largely determined by two characteristics: its toughness
(tenacidad) (KIC), and the size distribution of microcracks it contains. A new class of fully dense, high-strength ceramics is
now emerging which combine a higher fracture toughness with a much narrower distribution of smaller microcracks, giving
properties which make them competitive with metals, cermets, even with diamond, for cutting tools, dies, implants and engine
parts. And (at least potentially) they are cheap.
The microstructure of ceramics:

Crystalline ceramics form polycrystalline


microstructures, very like those of metals. Each
grain is a more or less perfect crystal, meeting its
neighbours at grain boundaries. The structure of
ceramic grain boundaries is obviously more
complicated than those in metals: ions with the
same sign of charge must still avoid each other and,
as far as possible, valency requirements must be
met in the boundary, just as they are within the
grains.
i B t none off this
But thi is i visible
i ibl att theth
microstructural level, which for a pure, dense
ceramic, looks just like that of a metal. Many
ceramics are not fully dense. Porosities as high as
20% are a common feature of the microstructure.
The pores weaken the material, though if they are
well rounded, the stress concentration they induce
is small. More damaging are cracks; they are much
harder to see, but they are nonetheless present in
most ceramics,
ceramics left by processing,
processing or nucleated by
differences in thermal expansion or modulus
between grains or phases. Recent developments in
ceramic processing aim to reduce the size and
number of these cracks and pores, giving ceramic
bodies with tensile strengths as high as those of
high‐strength steel.
Most ceramics are intrinsically hard; ionic or covalent bonds present an 
enormous lattice resistance to the motion of a dislocation. Take the 
covalent bond first. The covalent bond is localised; the electrons which 
form the bond are concentrated in the region between the bonded
form the bond are concentrated in the region between the bonded 
atoms; they behave like little elastic struts joining the atoms. When a 
dislocation moves through the structure it must break and reform
these bonds as it moves: it is like traversing a forest by uprooting and 
then replanting every tree in your path.

Most ionic ceramics are hard, though for a slightly different reason. 
The ionic bond, like the metallic one, is electrostatic: the attractive 
force between a sodium ion (Na+) and a chlorine ion (Cl−) is simply 
proportional to q2/r where q is the charge on an electron and r the 
separation of the ions. If the crystal is sheared on the 45°
i f h i If h li h d h 45° plane shown 
l h
in Fig. (c) then like ions remain separated: Na+ ions do not ride over 
Na+ ions, for instance. This sort of shear is relatively easy – the lattice 
resistance opposing it is small. But look at the other shear – the 
horizontal one. This does carry Na+ ions over Na+ ions and the 
electrostatic repulsion between like ions opposes this strongly. The 
l l b lk h l h
lattice resistance is high. 
In a polycrystal, you will remember, many slip systems are necessary, 
and some of them are the hard ones. So the hardness of a 
polycrystalline ionic ceramic is usually high (though not as high as a 
covalent ceramic), even though a single crystal of the same material 
might – if loaded in the right way – have a low yield strength.
So ceramics, at room temperature, generally have a very large lattice resistance. The stress required to make dislocations 
move is a large fraction of Young’s modulus: typically, around E/30, compared with E/103 or less for the soft metals like 
copper or lead This gives to ceramics yield strengths which are of order 5 GPa – so high that the only way to measure them 
copper or lead. This gives to ceramics yield strengths which are of order 5 GPa so high that the only way to measure them
is to indent the ceramic with a diamond and measure the hardness.  
This enormous hardness is exploited in grinding wheels which are made from small particles of a high‐performance 
engineering ceramic (Table 15.3) bonded with an adhesive or a cement. In design with ceramics it is never necessary to 
consider plastic collapse of the component: fracture always intervenes first. 
Fracture
Most commonly the production method leaves small
Most commonly the production method leaves small 
holes: sintered products, for instance, generally contain 
angular pores on the scale of the powder (or grain) size. 
Thermal stresses caused by cooling or thermal cycling 
can generate small cracks. Even if there are no 
processing or thermal cracks, corrosion (often by water) 
or abrasion (by dust) is sufficient to create cracks in the 
surface of any ceramic. And if they do not form any other 
way, cracks appear during the loading of a brittle solid, 
nucleated by the elastic anisotropy of the grains or by
nucleated by the elastic anisotropy of the grains, or by 
easy slip on a single slip system.
Cracks in compression propagate stably, and twist out 
of their original orientation to propagate parallel to the 
compression axis. Fracture is not caused by the rapid
compression axis. Fracture is not caused by the rapid 
unstable propagation of one crack, but the slow 
extension of many cracks to form a crushed zone. 

The design strength of a ceramic, then, is determined by 
i l f
its low fracture toughness and by the lengths of the 
h db h l h f h
microcracks it contains. As we shall see, there are two 
ways of improving the strength of ceramics:
1) decreasing the length of the microcrack by careful 
q
quality control, and 
y ,
2) increasing the fracture toughness  by alloying, or by 
making the ceramic into a composite.

You might also like