Thanks to visit codestin.com
Credit goes to www.scribd.com

0% found this document useful (0 votes)
10 views5 pages

Assignment 2

The document discusses the structure of solids, distinguishing between crystalline and amorphous forms, and explains how X-ray crystallography is used to determine atomic arrangements in crystalline materials. It details the principles of diffraction, the classification of crystal lattices, and the methodology of X-ray crystallography, including instrumentation and applications in various fields such as chemistry, biology, and industry. Overall, it emphasizes the importance of understanding solid-state structures and the role of X-ray crystallography in advancing scientific knowledge.

Uploaded by

aribaashiq098
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
10 views5 pages

Assignment 2

The document discusses the structure of solids, distinguishing between crystalline and amorphous forms, and explains how X-ray crystallography is used to determine atomic arrangements in crystalline materials. It details the principles of diffraction, the classification of crystal lattices, and the methodology of X-ray crystallography, including instrumentation and applications in various fields such as chemistry, biology, and industry. Overall, it emphasizes the importance of understanding solid-state structures and the role of X-ray crystallography in advancing scientific knowledge.

Uploaded by

aribaashiq098
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 5

Structure of Solids and X-ray Crystallography

Introduction

Solids are materials in which atoms, ions, or molecules occupy fixed positions in an extended
three-dimensional arrangement. When these constituents are arranged in a regular, repeating
pattern (a crystal lattice), the solid is crystalline; if the arrangement lacks long-range order, the
solid is amorphous. Crystalline solids exhibit flat faces and sharp melting points because the local
environment of each unit is identical, whereas amorphous solids have irregular surfaces and melt
over a broad range of temperatures. X-ray crystallography is a powerful analytical method to
determine the atomic arrangement of such crystalline materials. By directing monochromatic X-
rays (wavelength on the order of 1 Å) at a crystal, the X-rays are diffracted by the periodic lattice
planes. The resulting diffraction pattern – essentially a map of constructive interference – encodes
the positions of the atoms. Intensity and angle measurements of diffracted beams, combined with
Bragg’s law (2d sin θ = nλ), allow the determination of the crystal’s unit-cell dimensions and
atomic coordinates. Historically, the recognition that crystals act as three‐dimensional diffraction
gratings for X-rays (first demonstrated by von Laue in 1912) cemented X-ray crystallography as a
cornerstone of solid-state science.

Classification and Structure of Solids: Solids are broadly classified by their internal order and
chemical bonding. Crystalline solids have atoms arranged in highly ordered, repeating lattices.
This order yields characteristic physical properties: sharp melting points, distinctive crystal faces,
and clear X-ray diffraction patterns. For example, ionic crystals like sodium chloride cleave along
flat planes and diffract X-rays into sharp spot patterns, reflecting their uniform lattice
environments. In contrast, amorphous solids (often called glasses) lack long-range periodicity.
Their constituents are more randomly arranged, so cleavage produces irregular, curved surfaces
and X-ray exposure yields diffuse halos rather than sharp spots. Amorphous materials do not have
a well-defined melting point; they soften over a range of temperatures because the local bonding
environments vary throughout the solid. In practice, many materials can form either type: e.g. SiO₂
forms crystalline quartz when cooled slowly (ordered network of SiO₄ tetrahedra) but becomes
amorphous quartz glass when cooled rapidly. In summary, the key difference is that crystalline
solids have a periodic lattice with long-range order, whereas amorphous solids lack such order.

Types of Crystal Lattices: Crystalline order can take on a limited set of symmetric patterns called
crystal lattices. In three dimensions, there are fourteen distinct Bravais lattices formed by
combining seven crystal systems with lattice centering types. The seven lattice systems are defined
by constraints on the unit-cell parameters (edge lengths a, b, c and interaxial angles α, β, γ). These
systems and their characteristic parameter constraints are:

 Triclinic  a ≠ b ≠ c; α ≠ β ≠ γ (no angle is 90°).


 Monoclinic  a ≠ b ≠ c; α = γ = 90°, β ≠ 90°.
 Orthorhombic  a ≠ b ≠ c; α = β = γ = 90°.
 Tetragonal  a = b ≠ c; α = β = γ = 90°.
 Trigonal (Rhombohedral)  a = b = c; α = β = γ ≠ 90° (all angles equal, but not 90°).
 Hexagonal  a = b ≠ c; α = β = 90°, γ = 120°.
 Cubic  a = b = c; α = β = γ = 90°.

Each Bravais lattice is a specific arrangement of lattice points within these systems. For example,
in the cubic system there are three Bravais lattices: simple cubic (lattice points at cube corners
only), body-centered cubic (one additional point at the cube center), and face-centered cubic
(points at all face centers as well as corners). Many common materials adopt these lattices; e.g.
NaCl and Cu have face-centered cubic structures, and body-centered cubic is found in metals like
iron. In total 14 unique Bravais lattices exist in 3D, each giving rise to a characteristic symmetry
and repeat motif.

Unit Cell and Bravais Lattices: The unit cell is the fundamental repeat unit of a crystal lattice. It
is defined as the smallest volume of space that, when translated along the lattice vectors,
reproduces the entire crystal. In crystallography, a unit cell comprises the space between adjacent
lattice points plus any atoms (basis) located in that region. The simplest unit cell is the primitive
cell, which contains exactly one lattice point (allowing fractional counting if atoms are shared
between cells). A primitive cell is the smallest-volume cell that can generate the lattice by
translational repetition. By contrast, a conventional cell may contain multiple lattice points but is
chosen to reflect the symmetry of the crystal for easier visualization. For instance, the face-
centered cubic (FCC) lattice can be described by a conventional cubic cell containing four lattice
points (one at each corner shared among eight cells, plus one at each face center shared by two
cells).

In three dimensions, each unit cell is specified by six lattice parameters: three edge lengths (a, b,
c) and three interaxial angles (α between b and c; β between a and c; γ between a and b). From
these, the volume of the cell is given by the triple scalar product V = a⋅(b×c). The choice of unit
cell and lattice vectors defines the Bravais lattice. For example, carbon in diamond adopts the
diamond cubic structure, which can be viewed as two interpenetrating face-centered cubic lattices
offset by (¼,¼,¼) of the cell length. The accompanying figure illustrates the diamond cubic lattice:
each unit cell is a cube (a=b=c, α=β=γ=90°) with a two-atom basis at fractional coordinates.

Principles of X-ray Crystallography: The core principle of X-ray crystallography is diffraction:


waves scattered by the periodic array of atoms in a crystal interfere constructively at certain angles.
According to Bragg’s law (2d sin θ = nλ), constructive interference (diffraction peaks) occurs when
the path-length difference between waves reflected by successive lattice planes equals an integer
multiple of the X-ray wavelength. Equivalently, one can derive diffraction as the condition that
the reciprocal lattice vector 2π*(ha*+kb*+l**c*) must match the scattering vector of the X-ray
beam (the Laue condition). In practice, a monochromatic X-ray beam impinges on the crystal; as
the crystal is rotated, each set of lattice planes (indexed by Miller indices hkl) will satisfy Bragg’s
law at a specific angle θ. The detector records a series of diffracted beams (reflections), each
labeled by (hkl) and having intensity proportional to |F(hkl)|², where F(hkl) is the crystal’s structure
factor.

Mathematically, the structure factor F(hkl) is a sum over atomic scattering factors and phase
factors for the wave scattered by each atom in the unit cell. The collection of measured |F(hkl)|
values encodes the electron density in the crystal via a Fourier synthesis:

Where ρ(x,y,z) is the electron density at position (x,y,z) in the cell. From this relation, one obtains
a 3D electron density map of the crystal, and by locating the density peaks one identifies the atomic
positions. In summary, X-ray crystallography solves a crystal structure by measuring diffraction
angles/intensities, indexing them to Miller indices, and performing an inverse Fourier transform
to obtain the atomic arrangement. The key requirement is that the crystal be periodic: amorphous
materials yield only broad halos rather than discrete reflections, so they cannot be solved by these
methods.

Instrumentation and Methodology

A modern X-ray diffractometer consists essentially of an X-ray source (generator), wavelength


selector (monochromator or filter), goniometer (sample holder), and detector. Commonly, the X-
rays are produced in a sealed tube where a heated tungsten filament (cathode) emits electrons that
are accelerated at high voltage onto a metal anode (target), often copper or molybdenum. The
impact produces characteristic X-ray radiation (e.g. Cu Kα at 1.54 Å, or Mo Kα at 0.7107 Å).
Filters or crystal monochromators are used to isolate one wavelength (for example, a nickel filter
blocks Cu Kβ, yielding predominantly Kα radiation). The X-ray beam is then directed toward the
mounted crystal, typically centered on a goniometer that can precisely rotate the crystal about one
or more axes. As the crystal is rotated, different lattice planes satisfy Bragg’s law and diffract X-
rays into a detector. Modern detectors are highly sensitive photon counters or area detectors (e.g.
CCD or CMOS-based), which record the position and intensity of each diffracted beam.

Key procedural steps include sample preparation and data collection. Obtaining a suitable crystal
is often the first challenge: it must be sufficiently large (typically >0.1 mm) and free of defects
such as cracks or twinning. Small-molecule crystallography usually uses single crystals with well-
ordered structures, whereas powder X-ray diffraction (PXRD) involves many microcrystals
oriented randomly (useful for phase identification and materials analysis, though PXRD yields
averaged structural information). In a typical single-crystal experiment, the crystal is mounted
(often on a thin glass fiber) and cooled to reduce thermal motion. Monochromatic X-rays are then
directed at the crystal while it is rotated; a sequence of diffraction images is collected at different
orientations. Each image contains spots (reflections) whose angles (2θ) and intensities are
measured. These data are indexed to (hkl) indices and used in a computer algorithm (often
involving Fourier methods and least-squares refinement) to solve and refine the atomic structure.
In essence, the methodology is: (1) mount a pure crystal; (2) irradiate with monochromatic X-rays
while recording diffraction peaks; (3) use Bragg’s law to index the pattern; and (4) compute the
electron density to determine atomic positions.

Instrumentation details include the specific source and optics. For example, a typical sealed-tube
instrument uses Cu Kα radiation with a graphite or Ni monochromator, whereas high-throughput
systems may employ a rotating anode source (higher intensity) or even a synchrotron beamline for
intense, tunable X-rays. The goniometer design (three- or four-circle) allows precise alignment of
crystal axes relative to the beam. Modern detectors (e.g. CCD) count individual X-ray photons and
interface with computers for rapid data acquisition. In summary, an X-ray crystallographic
instrument integrates an X-ray generator, focusing optics, a sample stage, and a sensitive detector
to systematically probe the crystal’s diffraction.

Applications of X-ray Crystallography in Science and Industry

X-ray crystallography has profound applications across chemistry, biology, physics, geology, and
industry. In chemistry and materials science, it is the definitive method for determining
molecular and crystal structures. For instance, minerals and inorganic compounds were among the
first structures solved – menzer elucidated garnet in 1924, and systematic studies of silicate
frameworks quickly followed. Complex inorganic and organometallic molecules (e.g. fullerenes,
porphyrins, zeolites, superconductors) are routinely characterized by X-ray analysis. In the
Cambridge Structural Database (CSD) – a repository of small-molecule crystal structures – over
one million entries are catalogued, almost all determined by X-ray diffraction. In pharmaceutical
research, crystallography is essential for drug design and development. The three-dimensional
structures of drug targets (proteins, nucleic acids) determined by X-ray crystallography serve as
the “gold standard” for structure-based drug design. Over 90% of the >100,000 structures in the
Protein Data Bank come from X-ray studies, including many that directly guided drug discovery
(e.g. HIV protease and neuraminidase inhibitor complexes). X-ray data allow chemists to visualize
the exact binding mode of a drug candidate and rationally modify its interactions. For example,
Dorothy Hodgkin’s determination of penicillin and vitamin B₁₂ structures enabled better
understanding of these molecules. Today, crystallography is routinely used to refine drug leads: it
“reveals how a pharmaceutical drug interacts with its protein target and what changes might
improve it”.

In industry and engineering, XRD is used for quality control and characterization. Powder X-ray
diffraction identifies and quantifies crystalline phases in mixtures – crucial in cement
manufacturing, metallurgy (phase composition in steel), and battery cathode development. It is
employed to verify the polymorph of an active pharmaceutical ingredient (polymorphic form can
alter drug efficacy). In materials science, new compounds (e.g. novel semiconductors, polymers,
nanomaterials) are often characterized first by XRD to confirm their crystal structure. For example,
the structures of silicon, germanium, and other Group-14 semiconductors (which are diamond-
cubic lattices) were determined by X-ray methods and underlie modern electronics.

In geology and planetary science, X-ray crystallography has identified countless minerals and
elucidated their structure-composition relationships. Notably, the Mars Curiosity rover’s CheMin
instrument performed the first extraterrestrial XRD on Martian soil, identifying feldspars,
pyroxenes and olivine in situ. Terrestrially, XRD is fundamental in mineralogy: for instance, Linus
Pauling’s work on metal alloys (e.g. Mg₂Sn) used X-ray structures to develop theories of ionic
crystal stability.

Overall, X-ray crystallography’s impact is ubiquitous: it provides the definitive atomic picture
for research ranging from new drug compounds to advanced materials and planetary exploration.
According to Anton Paar, it is “one of the most important tools in materials science, chemistry,
biology, and physics” for studying atomic arrangements. Nearly every discovery of a new crystal
structure – whether an inorganic solid or a biological macromolecule – relies on X-ray diffraction
techniques. In industry, it ensures the identity and purity of products; in science, it continuously
uncovers the detailed architecture of matter.

Conclusion

The structure of solids – crystalline versus amorphous – governs many of their physical and
chemical properties, and understanding this structure is a central goal of solid-state science. X-ray
crystallography provides a direct window into this structure by exploiting the wave nature of X-
rays and the periodicity of crystals. Through precise measurement of diffraction patterns and
Fourier analysis, crystallography yields atomic positions with sub-Ångstrom accuracy. This has
enabled breakthroughs across disciplines: from confirming the tetrahedral bonding in diamond to
mapping complex protein active sites. In sum, the interplay of solid-state structure and X-ray
diffraction creates a virtuous cycle: knowledge of lattices and unit cells informs experimental
design, while crystallographic data refine our understanding of materials. As instrumentation and
computation continue to advance, X-ray crystallography will remain an indispensable, high-
resolution probe of the solid state for years to come.

You might also like