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 H; where appropriate, the entries also describe subsidiary but related concepts. Refer to the Directory for a full list of all the concepts treated.

A (reduction) half-reaction is a hypothetical process which shows the reduction of the oxidized component (Ox) of a redox couple Ox/Red and the formation of the reduced component (Red):
The physical states of Ox and Red are given but not that of the electron, which is regarded as stateless. The potential of the electrode at which this process is taking place is given by a version of the Nernst equation, with the activity of the electrons ignored:
The Hamiltonian operator, H ˆ , (and commonly just ‘the Hamiltonian’) of a system is the operator for its total energy, the sum of the operators for its kinetic and potential energies. In the position representation of quantum mechanics, the form in which elementary applications are usually expressed, the Hamiltonian operator for a particle of mass m in a one-dimensional system is
and in three dimensions
(See spherical polar coordinates for the form of ∇2 in polar coordinates.) For a many-electron polynuclear molecule, the operator for the electrons is the sum of such terms and the potential energy of the interactions between the particles:
In terms of the Hamiltonian operator, the time-independent Schrödinger equation is
Thus, it is an eigenvalue equation with an energy the eigenvalue and the wavefunction the corresponding eigenfunction.
A hard-sphere potential is a model intermolecular potential energy in which there is no interaction between the two particles until they are close, when the potential energy climbs abruptly to infinity (Figure H.1):
Figure H.1

The distance-dependence of a hard-sphere potential energy.

Figure H.1

The distance-dependence of a hard-sphere potential energy.

Close modal
A harmonic oscillator is a body that obeys Hook’s law, that the restoring force is proportional to the displacement, F = −kfx, where kf is the force constant. The implication is that the potential energy of the body is
and is referred to as a ‘parabolic’ potential. The Schrödinger equation for a one-dimensional harmonic oscillator of mass m is therefore
The allowed energies (Figure H.2) and the corresponding wavefunctions (Figure H.3) and probability densities (Figure H.4) are
Figure H.2

The energy levels of a harmonic oscillator.

Figure H.2

The energy levels of a harmonic oscillator.

Close modal
Figure H.3

The wavefunctions of a harmonic oscillator.

Figure H.3

The wavefunctions of a harmonic oscillator.

Close modal
Figure H.4

The probability densities of a harmonic oscillator.

Figure H.4

The probability densities of a harmonic oscillator.

Close modal
The Hv(x/α) are Hermite polynomials. That v=0 is the lowest allowed value of the quantum number implies that there is an irremovable zero-point energy, E0=12ω. The energy levels form an evenly spaced ladder of separation ħω. Note that the greater the force constant (the stiffer the spring) and the lighter the mass the greater is the separation of the levels.
From the form of the wavefunctions it follows that
The transition dipole moments for v v ± 1 (the only nonzero values) are
The thermodynamic properties of a collection of N oscillators are derived by using statistical thermodynamics and are as follows (in each case β = 1/kT and ε = ħω); see Figure H.5−H.8:
Figure H.5

The temperature dependence of the partition function of N harmonic oscillators.

Figure H.5

The temperature dependence of the partition function of N harmonic oscillators.

Close modal
Figure H.6

The temperature dependence of the internal energy of N harmonic oscillators.

Figure H.6

The temperature dependence of the internal energy of N harmonic oscillators.

Close modal
Figure H.7

The temperature dependence of the entropy of N harmonic oscillators.

Figure H.7

The temperature dependence of the entropy of N harmonic oscillators.

Close modal
Figure H.8

The temperature dependence of the heat capacity of N harmonic oscillators.

Figure H.8

The temperature dependence of the heat capacity of N harmonic oscillators.

Close modal
The Harned cell is
The cell reaction is 1 2 H2(g) + AgCl(s) → HCl(aq) + Ag(s) and the Nernst equation for the cell potential is
In the harpoon mechanism of the collision between an alkali metal ion M and a dihalogen molecule X2, an electron jumps from M to X2 when that results in a lower energy, and then M+ and X2 are drawn together by their mutual Coulombic attraction. As a result, the reaction collision cross-section is much larger than their physical cross-section. The distance at which the electron transfers, R, is predicted by equating the difference in the ionization energy of M and the electron affinity of X2 to the Coulombic potential energy of the resulting ions:

Heat is the mode of transfer of energy that is due to a temperature difference between a system and its surroundings. The quantity of energy transferred to the system in this way is denoted q, or dq of the quantity is infinitesimal. The resulting change in internal energy, provided no other types of transfer take place, is dU = dq. Specifically, if the volume of a closed (constant composition) system is constant during the transfer, and no other processes are taking place, then ΔU = qV, the subscript denoting constant volume. If the pressure on a closed system is constant, and no other processes are taking place, then ΔH = qp.

At a molecular level, heat corresponds to the transfer of energy that makes use of or generates disorderly motion in the surroundings (Figure H.9). As such, it is distinguished from work, the transfer of energy that makes use of orderly motion in the surroundings. The term ‘heat’ should be distinguished from thermal motion, the disorderly motion of molecules, which is commonly used as a synonym.

Figure H.9

The molecular basis of energy transfer as heat.

Figure H.9

The molecular basis of energy transfer as heat.

Close modal
The heat capacity, C, of a system is a measure of the energy, q, that must be transferred to a system as heat in order to bring about a specified rise in temperature, ΔT: q = CΔT. If the heat capacity is large, then a large quantity of energy as heat is needed to bring about a given rise in temperature. This general concept of heat capacity is refined in thermodynamics by noting that the energy transferred as heat to a system of constant volume can be identified with the change in its internal energy, and if the pressure is constant, then with the change in its enthalpy. Then the isochoric heat capacity (more commonly, the constant-volume heat capacity), CV, and the isobaric heat capacity (the constant-pressure heat capacity), Cp, are defined for a system of constant composition as
See Figure H.10. They are both extensive properties; the corresponding molar quantities, CV,m = CV/n and Cp,m = Cp/n are intensive. The specific heat capacity in each case is CX,s = CX/m, where m is the mass of the sample.
Figure H.10

The definition of heat capacity.

Figure H.10

The definition of heat capacity.

Close modal
The two heat capacities differ on account of the energy that must be used to expand the system (against the external atmosphere and to overcome internal cohesive forces) when it is not constrained to be at constant volume. As that energy is transferred to the surroundings or used to change the internal structure of the material, there is less available to raise its temperature, so Cp,m > CV,m for a given substance. The general relation between them is
where α is the (isobaric) expansion coefficient and κT is the isothermal compressibility of the system. For a perfect gas, where the only contribution to the work of expansion is that arising from driving back the atmosphere, this expression simplifies to

At a molecular level, the heat capacity is large if there are many energy levels thermally accessible to the system, in which case populations can be transferred readily between them. At room temperature, many monatomic nonmetallic solids follow, approximately at least, Dulong and Petit’s law, that CV,m ≈ 3R. This law can be justified on the basis of the equipartition theorem and the oscillation of atoms in three dimensions. The heat capacities of all substance tend to zero as T → 0 because fewer energy levels become accessible. At low temperatures, nonmetallic solids obey the Debye-T 3 law, that CV,m ≈ aT 3, where a is a constant characteristic of the solid. Metallic solids show an additional contribution stemming from the possibility of exciting electrons near the top of the conduction band. For them, CV,m ≈ aT 3 + bT.

Quantum mechanical theories of heat capacities take into account the quantization of vibrations of atoms in solids. The Einstein equation for the molar heat capacity of a nonmetallic solid is based on all the atoms being able to vibrate around their locations in the solid at the frequency ν, and is
See Figure H.11. The Debye equation is more elaborate: it is based on a model in which the atoms can vibrate with all frequencies up to νmax, and is
Figure H.11

The Einstein and Debye heat capacity curves.

Figure H.11

The Einstein and Debye heat capacity curves.

Close modal

A heat engine is a cyclic device for the conversion of heat into work (Figure H.12). That is, it continuously draws energy as heat from a hot source, transfers some of that energy to the surroundings as work, and deposits the rest of the energy as heat into a cold sink. See Carnot cycle and efficiency. A heat engine in reverse is a refrigerator.

Figure H.12

The structure of a heat engine.

Figure H.12

The structure of a heat engine.

Close modal
The Helmholtz energy, A, is defined as
where U is the internal energy of the system, T is its temperature, and S is its entropy. It is a state function. A change in its value at constant temperature is equal to the maximum work that the system can achieve:
The Helmholtz energy can be pictured as the energy of the system that is stored in an orderly way and hence is available for causing orderly motion in the surroundings. The change in Helmholtz energy is proportional to the change in total entropy of the system and its surroundings when the system is maintained at constant volume:
It follows that a spontaneous change in a system at constant volume is accompanied by a decrease in Helmholtz energy.

The HermannMauguin system (or International system) is used to label symmetry groups, more commonly for crystals than for individual molecules. In the system, which is set out in Table H.1,

  • n denotes the presence of an n-fold axis

  • m denotes a mirror plane

  • / indicates that the mirror plane is perpendicular to the symmetry axis

  • – (a bar over a number) indicates that the element is combined with an inversion

Table H.1

The Hermann–Mauguin system (the entries in bold) a .

Ci                   
Cs  m                  
C1  1  C2  2  C3  3  C4  4  C6  6 
    C2v  2 mm   C3v  3 m   C4v  4 mm   C6v  6 mm  
    C2h  2/ m   C3h  6  C4h  4/ m   C6h  6/ m  
    D2  222  D3  32  D4  422  D6  622 
    D2h  mmm   D3h  6̄2 m   D4h  4/ mmm   D6h  6/ mmm  
    D2d  4̄2 m   D3d  m  S4    S6   
T  23  Td  4̄3 m   Th  m 3         
O  432  Oh  m 3 m              
Ci                   
Cs  m                  
C1  1  C2  2  C3  3  C4  4  C6  6 
    C2v  2 mm   C3v  3 m   C4v  4 mm   C6v  6 mm  
    C2h  2/ m   C3h  6  C4h  4/ m   C6h  6/ m  
    D2  222  D3  32  D4  422  D6  622 
    D2h  mmm   D3h  6̄2 m   D4h  4/ mmm   D6h  6/ mmm  
    D2d  4̄2 m   D3d  m  S4    S6   
T  23  Td  4̄3 m   Th  m 3         
O  432  Oh  m 3 m              
a

The only groups listed here are the so-called ‘crystallographic point groups’.

The Hückel approximation is a primitive scheme for calculating the π orbitals and energies of conjugated polyenes. The Hamiltonian matrix is constructed by setting all diagonal elements equal to αE, all off-diagonal elements between neighbours equal to β, all other elements to zero; overlap is ignored. The eigenvalues and eigenfunctions of the matrix are the energy levels and linear combination coefficients (of the carbon atom π orbitals), respectively, and are best found by using mathematical software. Figure H.13 shows two typical outcomes.

Figure H.13

The Hückel energy levels of butadiene and benzene.

Figure H.13

The Hückel energy levels of butadiene and benzene.

Close modal

Hund’s rules are a guide to the lowest energy term of an atom and can be extended to molecules.

  1. Electrons occupy orbitals singly before occupying them doubly.

  2. Electrons occupy orbitals that result in the maximum value of the total orbital angular momentum quantum number, L.

  3. Electrons adopt the maximum number of parallel spins compatible with their orbital occupation. This is the maximum multiplicity rule, with atoms achieving the maximum value of the total spin angular momentum, S.

  4. If the shell is less than half full, minimum J lies lowest; if it is more than half full, maximum J lies lowest.

The maximum multiplicity rule stems from the quantum mechanical property of spin correlation in which electrons with the same spin tend to stay apart from one another. This mutual avoidance permits the atom to shrink slightly with the result that electrons experience a more favourable nuclear attraction.

Hybridization is the mixing of two or more atomic orbitals on the same atom. It is a feature of valence bond theory aimed at providing equivalent, local bonds that conform to the shape of the molecule. Three characteristic hybridization schemes and their mutually orthogonal orbital compositions are:

2 Linear hybrids  sp  h1 = s + px  
h2 = s − px  
3 Trigonal planar hybrids  sp2   h1 = s + 21/2px  
h2 = s + (3/2)1/2px − (1/2)1/2py  
h3 = s − (3/2)1/2px − (1/2)1/2py  
4 Tetrahedral hybrids  sp3   h1 = s + px + py + pz  
h2 = s − px − py + pz  
h3 = s − px + py − pz  
h4 = s + px − py − pz  
2 Linear hybrids  sp  h1 = s + px  
h2 = s − px  
3 Trigonal planar hybrids  sp2   h1 = s + 21/2px  
h2 = s + (3/2)1/2px − (1/2)1/2py  
h3 = s − (3/2)1/2px − (1/2)1/2py  
4 Tetrahedral hybrids  sp3   h1 = s + px + py + pz  
h2 = s − px − py + pz  
h3 = s − px + py − pz  
h4 = s + px − py − pz  

Tetrahedral and trigonal hybrids are shown in Figure H.14.

Figure H.14

Tetrahedral and trigonal hybrids.

Figure H.14

Tetrahedral and trigonal hybrids.

Close modal

Note that N atomic orbitals give rise to N mutually orthogonal hybrid orbitals. Other symmetrical schemes are listed in Table H.2. Hybrid orbitals can be constructed for a variety of geometries by selecting combinations that give rise to hybrids that match the geometry and are mutually orthogonal.

Table H.2

Hybridization schemes.

Coordination number Arrangement Composition
Linear  sp, pd, sd 
  Angular  sd 
Trigonal planar  sp2, p2
  Unsymmetrical planar  spd 
  Trigonal pyramidal  pd2  
Tetrahedral  sp3, sd3  
  Irregular tetrahedral  spd2, p3d, pd3  
  Square planar  p2d2, sp2
Trigonal bipyramidal  sp3d, spd3  
  Tetragonal pyramidal  sp2d2, sd4, pd4, p3d2  
  Pentagonal planar  p2d3  
Octahedral  sp3d2  
  Trigonal prismatic  spd4, pd5  
  Trigonal antiprismatic  p3d3  
Coordination number Arrangement Composition
Linear  sp, pd, sd 
  Angular  sd 
Trigonal planar  sp2, p2
  Unsymmetrical planar  spd 
  Trigonal pyramidal  pd2  
Tetrahedral  sp3, sd3  
  Irregular tetrahedral  spd2, p3d, pd3  
  Square planar  p2d2, sp2
Trigonal bipyramidal  sp3d, spd3  
  Tetragonal pyramidal  sp2d2, sd4, pd4, p3d2  
  Pentagonal planar  p2d3  
Octahedral  sp3d2  
  Trigonal prismatic  spd4, pd5  
  Trigonal antiprismatic  p3d3  

A hydrogen bond is a link between atoms of two electronegative elements, A and B, one of which (B) has a lone pair of electrons, bridged by a hydrogen atom and denoted A–H⋯B. Typically, A and B are any of N, O, and F, but anions may also participate. A variety of theories have been given for its formation; all depend on the small size of the proton and its ability to lie between two nearby atoms. In the molecular orbital explanation (Figure H.15), the atoms A, H, and B each supply one atomic orbital and form a bonding, an almost nonbonding, and an antibonding molecular orbital. The first two are occupied by the four electrons provided by the A–H bond and the B lone pair. As the antibonding orbital remains unoccupied, this arrangement has a lower energy than the separate AH and B components.

Figure H.15

The molecular obitals involved in the formation of a hydrogen bond.

Figure H.15

The molecular obitals involved in the formation of a hydrogen bond.

Close modal

The strength of a hydrogen bond is approximately 20 kJ mol−1. Its formation accounts for many of the anomalous properties of water (including its low vapour pressure and the relative mass densities of the liquid and solid phases). It is of major importance in determining the secondary structure of biological macromolecules.

The description of the bond in the hydrogen molecule, H2, can be regarded as the starting point of the valence bond description of the chemical bond, and that of the hydrogen molecule-ion, H2 +, the starting point of the molecular orbital theory of bonding. Valence bond theory ascribes the bond to spin pairing of the electrons supplied by the two atoms, which enables (via the Pauli principle) the two electrons to be described by the spatial wavefunction ψA(1)ψB(2) + ψA(2)ψB(1). In that combination there is constructive interference in the internuclear region and a consequent lowering of energy of the two paired electrons. More complex molecules are described by localized bonds formed similarly. In molecular orbital theory, bonding and antibonding molecular orbitals, ψA ± ψB are constructed from the H1s atomic orbitals. The occupation of the bonding combination by even a single electron lowers the energy, once again largely due to constructive interference in the internuclear region. A second electron can join the first to strengthen the bond provided its spin is opposite to that of the first electron. The theory is extended to more complex molecules by including more atomic orbitals in the linear combinations. The Schrödinger equation for the electron in H2 + can be solved analytically, but the resulting wavefunctions (in spheroidal coordinates) are of too great a complexity to be useful.

The energies of the bonding and antibonding molecular orbitals of H2 + are obtained by inserting them into the Schrödinger equation and are found to be
They are plotted in Figure H.16.
Figure H.16

The calculated energy levels of H2 +.

Figure H.16

The calculated energy levels of H2 +.

Close modal
A hydrogenic atom is a one-electron atom or ion with atomic number Z. A hydrogen atom is a hydrogenic atom with Z = 1. Its bound-state energies (Figure H.17) are
where is the Rydberg constant (a wavenumber here) and μ is the reduced mass of the atom. The quantum numbers are as follows:
Figure H.17

The energy levels of a hydrogen atom.

Figure H.17

The energy levels of a hydrogen atom.

Close modal
The energies depend only on the principal quantum number n. The wavefunctions, which are the hydrogenic atomic orbitals, are
where Rn,l(r) is the radial wavefunction and Yl,ml (θ,ϕ) is the angular wavefunction (a spherical harmonic). The normalized radial wavefunctions are
The Ln,l are associated Laguerre polynomials, which are available in mathematical software and tables; a0 is the Bohr radius. The mean radius of a hydrogenic atomic orbital is

All orbitals of the same principal quantum number, n, have the same energy and belong to the same shell of the atom. A shell is sometimes designated by the letters K, L, M,… for n = 1, 2, 3,…. The members of a shell with the same value of the orbital angular momentum quantum number l form a subshell of the atom. There are n subshells in a shell with quantum number n. They are commonly designated by the letters s, p, d, f,… for l = 0, 1, 2, 3,…. A subshell with quantum number l consists of 2l + 1 orbitals that are distinguished by the magnetic quantum number, ml. It follows that there are n 2 degenerate orbitals in a shell with quantum number n. See quantum number for a further description of their significance and the entries for each type of orbital.

The ground state of a hydrogenic atom has the configuration 1s1 and term symbol 2S1/2. Its ionization energy is hcR̃NZ 2 ≈ 13.59Z 2 eV.

The hyperfine structure of a spectrum is the splitting of absorption lines that arises from the interaction of an electron with a nucleus other than by the Coulombic interaction. Two common sources are magnetic interaction of electrons and nuclei and the interaction of the nuclear quadrupole moment with the electric field gradient at the nucleus.

Close Modal

or Create an Account

Close Modal
Close Modal