Skip to Main Content
Skip Nav Destination

Physical chemistry is the part of chemistry that seeks to account for the properties and transformations of matter in terms of concepts, principles, and laws drawn from physics. This glossary is a compilation of definitions, descriptions, formulae, and illustrations of concepts that are encountered throughout the subject. This section describes the concepts that begin with the letter S; where appropriate, the entries also describe subsidiary but related concepts. Refer to the Directory for a full list of all the concepts treated.

An s orbital is an atomic orbital with l = 0. A consequence of it having no angular momentum and therefore no angular nodes is that it has nonzero probability amplitude at the nucleus. An s orbital has a spherical boundary surface centred on the nucleus (Figure S.1). If it belongs to a shell with quantum number n, it has n – 1 radial nodes.

Figure S.1

Three s orbitals.

Figure S.1

Three s orbitals.

Close modal

A σ orbital is a molecular orbital that has cylindrical symmetry and therefore zero orbital angular momentum around the internuclear axis of the atoms it joins (Figure S.2). Two electrons in a σ orbital with paired spins constitute a σ bond. In valence bond theory, a σ bond arises from the pairing of electrons in atomic orbitals with cylindrical symmetry about the internuclear axis.

Figure S.2

(a) The formation of a σ orbital from s orbitals, (b) the orbital itself, and (c) its formation from p orbitals.

Figure S.2

(a) The formation of a σ orbital from s orbitals, (b) the orbital itself, and (c) its formation from p orbitals.

Close modal
The Sackur–Tetrode equation is an expression for the molar entropy of a perfect monatomic gas:
For the standard value, set p equal to p . In terms of the molar volume:

The salting-out effect is the reduction of the solubility of a gas or other nonelectrolyte in water by the addition of a salt. The salting-in effect is the opposite.

Scanning tunnelling microscopy (STM) is the study of surfaces with atomic resolution by making use of the ability of electrons to tunnel across a narrow gap. In one variety of the technique, a platinum, rhodium, or tungsten needle with a fine point is attached to a piezoelectric ceramic rod, which expands and contracts in response to an applied potential difference. Electrons tunnel across the gap between the needle tip and the surface and the resulting current is monitored. In the adaptation atomic force microscopy (AFM), the force between needle point and atoms on the surface is monitored and interpreted in terms of the variation in height of the surface. An advantage of AFM over STM is that the surface need not be electrically conducting.

In collision theory, the scattering cross-section (or collision cross-section, σ) is the area presented by a target molecule to an incoming projectile molecule such that a ‘collision’ is counted if the centre of the latter molecule passes through the area. If the diameters of the two molecules treated as hard spheres are dA and dB (Figure S.3), then
Figure S.3

The collision cross-section.

Figure S.3

The collision cross-section.

Close modal
In molecular beam studies, the differential scattering cross-section, σ, is the constant of proportionality between the intensity of molecules, dI, that are scattered and the intensity of the incident beam, I, the number density of target molecules, N , and the path length, dx, through the medium:
The scattering factor, f, of an atom is a measure of its ability to scatter X-rays of wavelength λ through an angle θ:
where ρ(r) is the electron probability density. For scattering in the forward direction (θ = 0), for a neutral atom f = Z, the atomic number of the atom. For an atom treated as a uniform sphere of radius R,

This function is plotted in Figure S.4 and S.5. If the scattering factor is large, then the atom contributes strongly to the scattering of X-rays that gives rise to a diffraction pattern.

Figure S.4

The dependence of the scattering factor on k.

Figure S.4

The dependence of the scattering factor on k.

Close modal
Figure S.5

The dependence of the scattering factor on θ.

Figure S.5

The dependence of the scattering factor on θ.

Close modal

The Schoenflies system of notation for point groups is as follows:

  • Cn. Groups that include only one n-fold axis.

  • Cnv. Groups that include one n-fold axis and n vertical mirror planes, σv.

  • Cnh. Groups that include one n-fold axis and a horizontal mirror plane, σh.

  • Sn. Groups that include an n-fold axis of improper rotation. If n is odd, the groups are the same as Cnh.

  • Dn. Groups that include n 2-fold axes perpendicular to an n-fold principal axis.

  • Dnd. Groups that in addition to the elements characteristic of the group Dn also include n dihedral mirror planes, σd, bisecting the 2-fold axes.

  • Dnh. Groups that include the elements of Dn together with a horizontal mirror plane, σh.

  • T, O, I. These are the cubic groups. T is the group of the tetrahedron, O the group of the octahedron, and I the group of the icosahedron.

  • Th. The group of a tetrahedron in which h denotes the presence of a centre of inversion arising from the presence of three horizontal mirror planes.

  • Td. The group of a tetrahedron in which d denotes the presence of three S4 axes of improper rotation arising from the presence of three diagonal mirror planes.

  • Oh. An octahedral group in which the h denotes the presence of a centre of inversion arising from the presence of horizontal and vertical mirror planes.

  • Ih. An icosahedral group in which the h denotes the presence of a centre of inversion arising from the presence of horizontal and vertical mirror planes.

See also the Hermann–Mauguin system. For the classification of molecules, use the charts in Figure S.6 and S.7.

Figure S.6

A flow chart for the classification of molecules.

Figure S.6

A flow chart for the classification of molecules.

Close modal
Figure S.7

Shapes and their symmetry classification.

Figure S.7

Shapes and their symmetry classification.

Close modal
The Schrödinger equation is a (partial) differential equation for calculating the wavefunctions of quantum mechanical systems. The time-dependent Schrödinger equation is
where H ˆ is the Hamiltonian operator for the system (and r stands for the spatial coordinates of all the particles in the system). It is used to calculate the variation with time of the wavefunction. If the Hamiltonian is independent of time, the wavefunction can be calculated from the time-independent Schrödinger equation:
In each case the wavefunction must satisfy the appropriate initial (in time) and boundary (in space) conditions. The time-independent equation can be regarded as an eigenvalue equation in which the wavefunction is the eigenfunction and E is the corresponding eigenvalue of the Hamiltonian operator.
The time-independent equation is normally written in the position representation, and for a particle of mass m in a one-dimensional region where its potential energy is V(x) is
In three dimensions
In certain cases (among them when V = 0 or is centrosymmetric), the partial differential equation can be separated into ordinary differential equations. See spherical polar coordinates for the form of the laplacian operator (∇2) in cartesian and spherical polar coordinates. For examples of the solutions of the Schrödinger equation; see harmonic oscillator, hydrogenic atoms, and particle in a box. Many-electron systems must be solved numerically: see self-consistent field.

The empirical Schultze–Hardy rule states that hydrophobic colloids are flocculated most efficiently by ions of opposite charge type and high charge number.

The Second law of thermodynamics has a variety of formulations. Two observation-based statements (Figure S.8) are

Figure S.8

The processes denied by the Second law.

Figure S.8

The processes denied by the Second law.

Close modal
  1. Kelvin statement: no process is possible in which the sole result is the absorption of heat from a reservoir and its complete conversion into work.

  2. Clausius statement: no process is possible in which the sole result is the transfer of energy from a cooler to a hotter body.

These formulations inspired the introduction of the entropy as a state function and the equivalent statement of the Second law in the form:

  • 3

    A spontaneous process in an isolated system is accompanied by an increase in entropy.

A spontaneous process is one that can occur without work needing to be done in the surroundings to bring it about. This statement leaves open the possibility that in one region of the isolated system there might be a decrease in entropy, provided that elsewhere in the system the entropy increases in compensation.

The molecular interpretation of the thermodynamic statement is based on the identification of entropy with molecular disorder: an increase in disorder can stem from the chaotic dispersal of matter and the chaotic dispersal of energy. Thus, the Second law is an expression of the view that chaotic dispersal is natural and that the emergence of order is unnatural.

A second order reaction is one with a rate, v , proportional to the square of the concentrations of the reactants, A:
A reaction between the reactants A and B is overall second-order if its rate law is
The integrated form of the first of these rate laws is either of the following expressions:
See Figure S.9.
Figure S.9

The time dependence of the concentration of a reactant in a second-order reaction.

Figure S.9

The time dependence of the concentration of a reactant in a second-order reaction.

Close modal
The half-life of A is
If the stoichiometry of the reaction is A → P, then the concentration of the product increases as
Provided the reaction stoichiometry of the overall second-order rate law is A + B → P, the concentrations of A and B are given by the solution of
and the concentration of product increases as
Note that the condition [B]0 > [A]0 implies that A is the limiting reagent. Both these expressions apply if [B]0 = [A]0, provided in that case they are modified by taking the limit [B]0 → [A]0. For instance, the last equation becomes
When a wavefunction is written as a linear combination of N basis functions, as in
then application of the variation theorem results in a set of N simultaneous equations, the secular equations, of the form
This set of N simultaneous equations for the coefficients has a solution only if the secular determinant, the determinant formed from the factors multiplying the coefficients, vanishes:
Expansion of the determinant gives an Nth-order polynomial in E, from which the N roots may be determined and identified with the energies of the N linear combinations.
A simple case is N = 2 with Sii = 1 and other elements zero. The two secular equations are
The secular determinant is
which expands to a quadratic equation in E. The two roots may then be substituted into the secular equations and solved for the two coefficients after each substitution (that is, for each energy). A form of solution when H12 = H21, which is often useful, is
with
See also Hückel method.
Sedimentation is the settling of particles in a gravitational or centrifugal field. The drift speed, s, is the speed particles reach when the accelerating force is balanced by the frictional force. In an ultracentrifuge operating at an angular velocity ω it is
where m is the mass of the particles, b the buoyancy factor, r the distance of the point from the axis of rotation, f the frictional coefficient (see diffusion), ρ the mass density of the solvent, and vs the specific volume of the particles (their volume divided by their mass). For spherical molecules of radius a that obey Stokes’ law in a solvent of viscosity η,
The sedimentation constant, S, is defined as
Therefore, for such molecules
where n is the number-average molar mass of the particles. Alternatively,
where D is the diffusion coefficient.
When the sedimenting particles have reached dynamic equilibrium in the presence of thermal motion, which tends to redistribute them, they adopt a characteristic distribution in the gravitational or centrifugal field. The mass-average (or weight-average) molar mass, w, can then be determined from
where c2 and c1 are the concentrations of the particles at the radii r1 and r2 from the axis of rotation.

A selection rule is a specification of the allowed spectroscopic transitions of an atom or molecule. A gross selection rule is a statement about the general properties that a molecule must possess in order to undergo a certain class of transitions. Thus, the gross selection rule for rotational transitions is that a molecule must be polar, and for rotational Raman transitions a molecule must have an anisotropic polarizability. A specific selection rule is a statement about the changes that may occur in terms of quantum numbers or other labels of the states of an atom or molecule. The following specific selection rules are often encountered for transitions that occur by an electric dipole mechanism:

  1. Atomic transitions: Δl = ±1; ΔL = 0, ±1; ΔS = 0; ΔJ = 0, ±1

  2. Vibrational transitions (infrared and Raman): Δ v = ± 1

  3. Rotational transitions (microwave, branch structure of vibrational transition): ΔJ = 0, ±1

  4. Rotational transitions (Raman): Δl = 0, ±2 for linear rotors; Δl = 0, ±1, ±2; ΔK = 0 for symmetric rotors

  5. Electronic transitions: g → u and u → g (the Laporte selection rule).

In most cases these rules can be traced to the spin (S = 1) of the photon and the conservation of angular momentum. They are established by considering the relevant transition dipole moment and noting the criteria for it to be nonzero either by using the properties of the relevant mathematical functions or in terms of group theory. In the latter case, the transition dipole moment is zero unless the integrand spans the fully symmetric irreducible representation of the molecular point group.

A self-consistent field method (an SCF method) is an iterative computational procedure that continues until certain calculated properties remain unchanged in a cycle (to within a predetermined criterion). It is commonly applied to the numerical solution of the Schrödinger equation for many-electron atoms and molecules, when it is commonly referred to as a Hartree–Fock method (HF method, or HF-SCF method). The method typically begins by proposing the wavefunctions to be used in the calculation (such as various atomic orbitals with variable parameters or linear combinations with variable coefficients). The potential energy of all the electrons is calculated using the electron distribution characteristic of that basis and a numerical solution of the Schrödinger equation is found for one of the electrons. The procedure is continued for all the electrons in the molecule and a refined set of orbitals is obtained. The refined set of orbitals will in general differ from the original orbitals, and the procedure is continued until no significant change occurs in the course of one iteration. At that stage, the orbitals are self-consistent and are accepted as the best solution for the system.

In a semi-empirical method, certain integrals that occur in a numerical solution of the Schrödinger equation for a many-electron system are either ignored (according to various criteria) or approximated by empirical data. For instance, various integrals may be adjusted to optimize the calculation of enthalpies of formation.

In the separation of variables procedure, a partial differential equation in several variables is separated into a series of (possibly linked) ordinary differential equations in each variable. For example, a two-dimensional particle-in-a-box system for which the wavefunctions depend are functions of x and y, can be expressed as ψ(x,y) = ψ(x)ψ(y) and the Schrödinger equations for ψ(x) and ψ(y) solved separately. Similarly, the system of a proton and an electron can be separated into equations for the translational motion of the centre of mass of the joint system and another for the motion of a particle with a reduced mass around that centre of mass. Moreover, the latter motion, which is described by the wavefunction ψ(r,θ,ϕ) can be separated further into radial and angular factors: ψ(r,θ,ϕ) = ψ(r)ψ(θ)ψ(ϕ), each of which satisfies an ordinary differential equation; in this case the three solutions are linked and consequently the quantum numbers that arise have ranges of values that depend on each other (for instance, l = 0, 1,…, n − 1). The viability of the separation of variables procedure depends on the form of the potential energy of the system.

The shell of an atom consists of all the n 2 orbitals with the same principal quantum number, n. A closed shell is a shell in which all the orbitals are doubly occupied by electrons. This definition is sometimes relaxed when d orbitals are available and incompletely occupied.

Shielding is the modification of a field. An important example is the role of inner-shell electrons in shielding the electric field due to the nucleus of an atom. Thus, an electron (the ‘test electron’) in a many-electron atom is repelled by all the other electrons, and its net attraction to the nucleus is thereby reduced. Note that the other electrons do not actually ‘shield’ the field of the nucleus: they simply provide an opposing contribution to the total experienced by the test electron. If the shielding electrons are distributed spherically, and the test electron is at a radius r, then it experiences a net repulsion only from the electrons within a sphere of radius r (Figure S.10). The net attraction to the nucleus is expressed by replacing the actual nuclear charge, Ze, by the effective nuclear charge, Zeffe. See penetration.

Figure S.10

Shielding by other electrons.

Figure S.10

Shielding by other electrons.

Close modal

A singlet state is an atomic or molecular state with total electron spin quantum numbers S = 0, MS = 0; a triplet state is one with S = 1, MS = 0, ±1. The spin wavefunctions for two electrons in these states are as follows:

MS = −1 MS = 0 MS = +1
S = 0    1 2 1 / 2 { α ( 1 ) β ( 2 ) β ( 1 ) α ( 2 ) }    
S = 1  β ( 1 ) β ( 2 )   1 2 1 / 2 { α ( 1 ) β ( 2 ) + β ( 1 ) α ( 2 ) }   α ( 1 ) α ( 2 )  
MS = −1 MS = 0 MS = +1
S = 0    1 2 1 / 2 { α ( 1 ) β ( 2 ) β ( 1 ) α ( 2 ) }    
S = 1  β ( 1 ) β ( 2 )   1 2 1 / 2 { α ( 1 ) β ( 2 ) + β ( 1 ) α ( 2 ) }   α ( 1 ) α ( 2 )  

Figure S.11 shows the vector model representations of these combinations.

Figure S.11

Vector models of singlet and triplet spin states.

Figure S.11

Vector models of singlet and triplet spin states.

Close modal

Singlet states of two electrons (and therefore with ‘paired’ spins) are commonly represented as ↓↑ and triplet states (with two ‘unpaired’ electrons) as ↑↑. An atom with a singlet term has one level (J = L) and one with a triplet term, provided L > 0, has three levels (J = L + 1, L, L − 1).

A triplet term of a given configuration is typically lower in energy than the singlet term of the same configuration. See Hund’s rules. Electric dipole transitions between singlets and triplets are forbidden. Transitions between them are enabled by spin–orbit coupling. See intersystem crossing and phosphorescence.

A smectic phase is a liquid-crystalline mesophase in which molecules are ordered in two dimensions but not in a third (Figure S.12).

Figure S.12

The smectic mesophase.

Figure S.12

The smectic mesophase.

Close modal
The solubility of a substance in a specified solvent is the concentration of solute in a saturated solution. A saturated solution is one in which the dissolved and undissolved solute are in dynamic equilibrium. For a solute and solution that form an ideal solution, the solubility of the solute in terms of its mole fraction is given by
where the standard enthalpy of fusion and the fusion (melting) temperature are those of the solute. This expression shows that the solubility increases as the temperature approaches the melting point of the solute and is greater for solids with low enthalpies of fusion. In practice, the identity of the solvent is also important. The solubility curves in Figure S.13 are labelled with the value of Δ fus H /RTf.
Figure S.13

The temperature dependence of ideal solubility.

Figure S.13

The temperature dependence of ideal solubility.

Close modal
The solubility constant (the solubility product, the solubility product constant), KM, for a solid MpXq that dissolves to give p cations and q anions is defined as
where the aJ are the (dimensionless) activities in a saturated solution. In elementary applications, the activities are replaced by [J]/c :
This approximation is particularly hazardous in solutions of electrolytes as the interactions between ions are so strong.

Space quantization is the restriction of the component of an angular momentum to discrete values mjħ, with mj = j, j − 1,…, −j, and therefore in the vector model with the vector representing the angular momentum at discrete angles to the z-axis.

The sphalerite structure (or zinc-blende structure) takes its name from the structure of ZnS and is one of the reasonably common crystal types, especially when covalent bonding is important (Figure S.14).

Figure S.14

Three representations of the sphalerite structure.

Figure S.14

Three representations of the sphalerite structure.

Close modal
The spherical harmonics (or surface harmonics), Y l , m l ( θ , ϕ ) , are solutions of the partial differential equation
subject to cyclic boundary conditions, which imply that l = 0, 1, 2,… and ml = 0, ±1,…, ±l; Λ 2 is the legendrian operator:

The spherical harmonics occur widely in dealing with systems with spherical symmetry. Some are listed in Table S.1.

Table S.1

The spherical harmonics, up to l = 3.

l ml Y l , m l ( θ , ϕ )
( 1 4 π ) 1 / 2  
( 3 4 π ) 1 / 2 cos θ  
  ±1  ( 3 8 π ) 1 / 2 sin θ e ± i ϕ  
( 5 16 π ) 1 / 2 ( 3 cos 2 θ 1 )  
  ±1  ( 15 8 π ) 1 / 2 cos θ sin θ e ± i ϕ  
  ±2  ( 15 32 π ) 1 / 2 sin 2 θ e ± 2 i ϕ  
( 7 16 π ) 1 / 2 ( 5 cos 3 θ 3 cos θ )  
  ±1  ( 21 64 π ) 1 / 2 ( 5 cos 3 θ 1 ) sin θ e ± i ϕ  
  ±2  ( 105 32 π ) 1 / 2 sin 2 θ cos θ e ± 2i ϕ  
  ±3  ( 35 64 π ) 1 / 2 sin 3 θ e ± 3i ϕ  
l ml Y l , m l ( θ , ϕ )
( 1 4 π ) 1 / 2  
( 3 4 π ) 1 / 2 cos θ  
  ±1  ( 3 8 π ) 1 / 2 sin θ e ± i ϕ  
( 5 16 π ) 1 / 2 ( 3 cos 2 θ 1 )  
  ±1  ( 15 8 π ) 1 / 2 cos θ sin θ e ± i ϕ  
  ±2  ( 15 32 π ) 1 / 2 sin 2 θ e ± 2 i ϕ  
( 7 16 π ) 1 / 2 ( 5 cos 3 θ 3 cos θ )  
  ±1  ( 21 64 π ) 1 / 2 ( 5 cos 3 θ 1 ) sin θ e ± i ϕ  
  ±2  ( 105 32 π ) 1 / 2 sin 2 θ cos θ e ± 2i ϕ  
  ±3  ( 35 64 π ) 1 / 2 sin 3 θ e ± 3i ϕ  
The functions are orthogonal and normalized in the sense that
Spherical polar coordinates, {r,θ,ϕ}, are the natural coordinates to use in the description of spherical systems: r is the radius, θ is the colatitude, and ϕ is the azimuth (Figure S.15). They are related to the Cartesian coordinates, {x,y,z}, as follows:
Figure S.15

Spherical polar coordinates.

Figure S.15

Spherical polar coordinates.

Close modal
The volume element in Cartesian coordinates, dτ = dxdydz, in spherical polar coordinates becomes
The ranges of the variables to cover all space are
The laplacian is
with the legendrian

Spin is the intrinsic angular momentum of a particle. Each fundamental particle has a characteristic, unchangeable spin, with spin quantum number s. The magnitude of the spin is {s(s + 1)}1/2ħ and its z-components are msħ, with ms = s, s – 1,…, −s. The spin of nuclei is specified similarly, with quantum numbers I and mI. Fundamental particles and nuclei with integer spin quantum number are bosons; those with half-integer spin quantum number are fermions. Particles with nonzero spin and nonzero mass have a magnetic moment; particles with spin quantum number greater than 1 2 may also have an electric quadrupole moment.

Electrons and protons have s =  1 2 and therefore ms = ± 1 2 . Those with ms = + 1 2 are commonly denoted either ↑ or α; those with ms = − 1 2 are likewise denoted wither ↓ or β. See photon, for its spin properties. The total spin angular momentum of a many-electron atom is denoted S, MS, with S obtained by using a Clebsch–Gordan series. See singlet and triplet states.

Spin correlation is the quantum mechanical tendency for electrons with parallel spins to stay apart. If two electrons are in spatial orbitals ψa and ψb with parallel spins (αα, for instance), then according to the Pauli principle, their overall antisymmetrical wavefunction is
This function approaches zero as r 2 →  r 1, so there is zero probability of finding the electrons in the same region of space. See Hund’s rules.

A spin echo is a pulse of magnetization that arises when individual nuclear magnetic moments come back into phase after having dispersed after an earlier pulse. To produce a spin echo, a pulse of magnetic field is applied in the x-direction and the magnetization vector is rotated through 90° into the xy-plane (Figure S.16). The individual spins fan out on account of their slightly different Larmor frequencies. After an interval, a 180° pulse is applied in the y-direction. Now the spins begin to regroup and reassemble into a magnetization vector that give a pulse, the echo, in a detector. The amplitude of the echo decays exponentially with a time constant equal to the transverse relaxation time, T2.

Figure S.16

The generation of a spin echo.

Figure S.16

The generation of a spin echo.

Close modal
Spin–orbit coupling is the magnetic interaction between the spin and orbital angular momenta of an electron in an atom or molecule. The contribution to the Hamiltonian that represents the interaction has the form
where A ˜ is the spin–orbit coupling constant (here, a wavenumber). The interaction removes the degeneracy of levels with different total angular momenta. In a hydrogenic atom when the spin, orbital, and total angular momentum quantum numbers are s, l, and j, the energies of the levels of the term 2{l}j, the eigenvalues of the Hamiltonian, are
and the constant for a hydrogenic atom of atomic number Z is
where α is the fine-structure constant and is the Rydberg constant. For a many-electron atom the energies of the levels of the term 2S+1{L}J are similarly of the form
with now a characteristic of the atom. A general rule is that the lowest level of an atom of a given term is the level with the lowest value of J (Figure S.17). The opposite is true for an atom with a shell that is more than half full, when the lowest level is that with the highest value of J.
Figure S.17

High and low energy levels and their correlation with the total angular momentum.

Figure S.17

High and low energy levels and their correlation with the total angular momentum.

Close modal

The spin–orbit coupling constant increases strongly with atomic number, so it is most important in heavy atoms. Spin–orbit coupling is also responsible for the process of intersystem crossing and the breakdown of the selection rule ΔS = 0 in electronic transitions.

Spin–spin coupling is a coupling between nuclear spins; it gives rise to the fine structure of NMR spectra. The interaction between nuclei A and B is commonly written
with J here a frequency. When only the z-components of the nuclear spin angular momenta need be considered, their energies (the eigenvalues of the Hamiltonian with only the I ˆ zA I ˆ zB term remaining) are
If J > 0, the lower energy is obtained if the spins are antiparallel (that is, the mI have opposite signs).

The coupling constant for nuclei A and B that are separated by N bonds is denoted N JAB. Its magnitude depends on a variety of factors as well as the value of N. These factors include the angles that the intervening bonds make with each other (see Karplus equation) and the hybridization of the intervening atoms. The mechanism of coupling is a combination of the non-Coulombic interaction between electrons and the bond and spin correlation: see polarization mechanism.

A spontaneous process is a change that can occur without an external agency having to do work to bring it about. Spontaneity in thermodynamics is a tendency which might not be realized in practice due to kinetic considerations. A process is spontaneous if the total entropy change that accompanies it is positive: ΔStotal > 0. At constant temperature and pressure, a spontaneous process is accompanied by a decrease in Gibbs energy (of the system): ΔG < 0. See also Einstein coefficients.

Standard ambient temperature and pressure (SATP) is 298.15 K and 1 bar. Distinguish it from standard temperature and pressure (STP) which is 0 °C and 1 atm.

The standard potential, E (Ox, Red), is the standard, zero-current potential of the cell
where SHE is the standard hydrogen electrode. It refers to the half-reaction
Standard potentials are also widely called standard reduction potentials to distinguish them from an earlier definition as standard oxidation potentials (which have the opposite sign). The standard potential of the cell Left||Right is
A standard potential is an indication of the relative reducing power of the redox couple in that Red1 has a thermodynamic tendency to reduce Ox2 if E ( Ox 1 , Red 1 ) is more negative than E ( Ox 2 , Red 2 ) . Standard potentials are used to predict the equilibrium constant of the corresponding cell reaction:
by using
The standard potentials of couples that are related only by changes in oxidation number (and therefore have different values of ν) are related as follows:
The standard state of a substance (denoted ) is the pure substance at 1 bar. For a gas, the standard state is a hypothetical state in which the fugacity of the gas is 1 bar and is behaving perfectly. For a solute, the standard state is at unit activity and behaving perfectly. Properties of a species in its standard state and changes in which all the participants are in their standard states are denoted X and ΔX , respectively. The advantage of these somewhat obscure definitions is that all the deviations from ideality due to molecular and ionic interactions are carried by the fugacity coefficient or the activity coefficient, respectively. The biochemical standard state (denoted ⦵′) includes the requirement that pH = 7. For a reaction of the form A + νp H+(aq) → B the two standard values of the reaction Gibbs energy are related by

The Stark effect is the removal of the degeneracy of the rotational states of polar molecules by the application of an electric field. It can be used to determine the electric dipole moment of a molecule. Stark modulation is the modulation of the rotational energy levels of polar molecules by the application of an oscillating electric field and is used in the detection of rotational transitions.

The Stark–Einstein law states that one photon is absorbed by each molecule responsible for the primary photochemical process. The law fails when the intensity of radiation is very high.

A state function, X, is a thermodynamic property that depends only on the current state of the system and is independent of its previous history. State functions include the internal energy and the entropy and properties derived from these properties, such as the enthalpy and Gibbs energy. The differentials, dX, of state functions are exact in the sense that the integral between two specified states is independent of the path of integration and therefore
where denotes integration around a closed cycle.

Statistical thermodynamics is a theory of thermodynamic properties in terms of the average behaviour of large assemblies of molecules. The basic concept used to evaluate averages is the ensemble, a collection of hypothetical replications of the system, and the evaluation of the partition function, a function that contains all the thermodynamic information about the system. In certain idealized cases, the partition function can be calculated from computed or spectroscopic properties of the molecules that constitute the system. See partition function and ensemble.

The steady-state approximation of chemical kinetics is that the concentrations of all intermediates in a proposed reaction mechanism are low and constant throughout the main part of the reaction. By setting d[I]/dt = 0, where I is an intermediate, the differential equation for its net production is turned into an algebraic equation which can be used to formulate the overall rate law corresponding to the proposed mechanism.

The sticking probability, s, is the proportion of collisions with a surface that successfully lead to adsorption:

In simple cases the sticking probability decreases with the extent of surface remaining uncovered and s ∝ 1−θ, where θ is the fractional coverage of the surface (Figure S.18).

Figure S.18

Typical variations of the sticking probability with surface coverage.

Figure S.18

Typical variations of the sticking probability with surface coverage.

Close modal
Stirling’s approximation for the factorial n! = n(n − 1)…1 when n ≫ 1 is
It is normally adequate to use the simpler form
A stoichiometric coefficient, ν, is the (positive) dimensionless coefficient in a balanced chemical equation. A stoichiometric number, νJ, is the signed, dimensionless coefficient for the participant J, with positive signs for products and negative signs for reactants. More formally, the νJ are the numbers in the chemical equation written as
where sJ is the chemical formula of J.

The stoichiometric point (formerly the closely related equivalence point) of a titration is the point at which the stoichiometric quantity of titrant has been added to the analyte solution. It should be distinguished from the end point, the point at which an indicator changes colour.

Stokes’ law states that the frictional retarding force on the sphere of radius a travelling at a speed s in a medium of viscosity η is

See also diffusion.

In the stopped-flow technique, reagents are impelled into a mixing chamber as a piston in the chamber is withdrawn suddenly. The composition of the chamber is then monitored. The technique is used to study fast reactions.

A structure factor, Fhkl, is the sum of the scattering factors, fi, of all the atoms in a unit cell weighted by a phase factor that depends on their location:

The intensity of the {hkl} reflection is proportional to |Fhkl|2. For the interpretation of diffraction intensities in terms of electron density in the unit cell, see Fourier synthesis and the phase problem.

Sublimation is the direct conversion of a solid into a vapour. The reverse of sublimation is vapour deposition. The enthalpy of sublimation is the sum of the enthalpies of fusion and vaporization at the same temperature:

The subshell of an atomic shell of principal quantum number n consists of all 2l + 1 orbitals of the same orbital quantum number l in that shell. The following notation is used:

l   … 
  … 
l   … 
  … 

A supercritical fluid is a fluid at a temperature higher than its critical temperature but with a density comparable to that of a normal liquid. A supercritical fluid has properties that are intermediate between those of a liquid and a gas.

A superfluid is a liquid phase of matter that flows without viscosity. The only known examples are helium-4 and helium-3. The transition to the superfluid phase occurs below the λ-transition temperature (Figure S.19).

Figure S.19

The phase diagram of helium-4 at low temperatures.

Figure S.19

The phase diagram of helium-4 at low temperatures.

Close modal
In quantum mechanics, a superposition is the addition of the probability amplitudes of an event. If the possible outcomes of observations of the event are described by the wavefunctions ψi, then the superposition is the linear combination
and the probability that the event described by ψi will be observed is proportional to |ci|2. The description of multiple possible outcomes in terms of a superposition is a principal difference of quantum mechanics from classical mechanics, for in the latter it is the probabilities that are added, not their amplitudes. The superposition of amplitudes allows for the possibility of interference between them.

The most common application of superposition in chemistry is in the construction of molecular orbitals as linear combinations of atomic orbitals (LCAO) and, in valence bond theory, of resonance hybrids.

Surface reconstruction is the modification of the structure of a solid at and close to its surface, perhaps due to the attachment of an adsorbate.

The surface tension, γ, of a liquid is the constant of proportionality between the work, dw, required to create a surface and the area created, dσ:
The dimensions of γ are those of energy/area and it is normally reported in either J m−2 or (equivalently) N m−1. At constant temperature, the work of surface formation is equal to the change in Helmholtz energy, dA, of the system:
For applications, see capillary action and the Kelvin equation.

The svedberg (Sv) is a non-SI unit encountered in sedimentation studies of macromolecules: 1 Sv = 10−13 s. It is commonly used to report the value of the sedimentation constant, S.

A symmetry operation is an action that leaves an object apparently unchanged. Examples are rotations, reflections, and inversions. Translation is a symmetry operation for an infinite crystal. A symmetry element is the point, line, or plane with respect to which the operation is carried out. Examples are:

Operation: Rotation Reflection Inversion
Element:  Rotation axis  Mirror plane  Centre of inversion 
Operation: Rotation Reflection Inversion
Element:  Rotation axis  Mirror plane  Centre of inversion 

A symmetry number, σ, is a factor included in the high-temperature form of a partition function to correct for the limitation on the occupation of rotational states imposed by the Pauli principle. It is equal to the order of the rotational subgroup of the molecular point group (in effect, allowing for that number of indistinguishable orientations); for a regular tetrahedron (Figure S.20), σ = 12. The high-temperature rotational partition function of a homonuclear diatomic molecule is kT/hcB̃ if the Pauli principle is ignored but kT/σhcB̃ with σ = 2 when it is taken into account.

Figure S.20

The rotational subgroup of a regular tetrahedron.

Figure S.20

The rotational subgroup of a regular tetrahedron.

Close modal

A symmetry species (in general, Γ) is the label of the irreducible representation of a molecular point group. The letters A and B are used to label one-dimensional irreducible representations: A is used if the character of the principal rotation is +1 and B is used if it is −1. The label E denotes a two-dimensional irreducible representation and T denotes a three-dimensional irreducible representation. In icosahedral point groups, G and H are used for four- and five-dimensional irreducible representations.

A symmetry-adapted linear combination (SALC) is a weighted sum of atomic orbitals that is a basis for an irreducible representation of the molecular point group (Figure S.21). Molecular orbitals are constructed from the overlap of an SALC with orbitals of the same symmetry species on other atoms. For instance, the SALC ψH1sA + ψH1sB in H2O has A1 symmetry in the C2v molecular point group and is used to form molecular orbitals with ψO2s and ψO2pz, both of which have A1 symmetry.

Figure S.21

Three SALCs for a molecule with C3v symmetry.

Figure S.21

Three SALCs for a molecule with C3v symmetry.

Close modal

Synchrotron radiation is the high-energy electromagnetic radiation generated by electrons moving at high speed in a circle (and therefore ceaselessly accelerating). The radiation extends into the ultraviolet and X-ray regions and has a higher intensity than can be achieved by conventional sources.

In thermodynamics, a system is the object of interest, such as a gas or a reaction mixture. The surroundings are the rest of the universe, and often emulated by a large water bath. The classification of systems is based on the properties of the boundary that separates them from their surroundings (Figure S.22):

Boundary Classification of the system
Permits neither matter nor energy to pass  Isolated 
Permits energy to pass but not matter  Closed 
Permits energy and matter to pass  Open 
Boundary Classification of the system
Permits neither matter nor energy to pass  Isolated 
Permits energy to pass but not matter  Closed 
Permits energy and matter to pass  Open 
Figure S.22

Varieties of systems.

Figure S.22

Varieties of systems.

Close modal

A boundary is diathermic if it allows energy to pass as heat but adiabatic if it does not.

In X-ray crystallography, a systematic absence is the absence of reflections that follow a simple rule relating to the values of the indices hkl (Figure S.23). For example, the diffraction pattern from a cubic I (bcc) lattice can be constructed from that from a cubic P lattice by striking out all reflections with odd values of h + k + l. For a cubic F (fcc) lattice, the only reflections present are those with h, k, and l all even or all odd.

Figure S.23

Systematic absences of a cubic lattice.

Figure S.23

Systematic absences of a cubic lattice.

Close modal
Close Modal

or Create an Account

Close Modal
Close Modal