Jump back to chapter selection.


Table of Contents

1.1 Microscopic Form of Maxwell's Equations in Vacuum
1.2 Maxwell's Equations in a Medium
1.3 The Material Equations
1.4 Macroscopic Approximation
1.5 Wave Equation
1.6 Solutions to the wave equation
1.7 Polarisation
1.8 Poynting Vector and Poynting's Theorem
1.9 Timescales
1.10 Momentum of Light


1 Electromagnetic Theory of Light

Light is an electromagnetic wave governed by the same theoretical principles that describe all forms of electromagnetic radiation. It consists of coupled oscillating electric and magnetic fields.


1.1 Microscopic Form of Maxwell's Equations in Vacuum

We begin with the simplest case by considering the electric and magnetic fields in free space, meaning there are no charges or currents present. The governing equations are Maxwell's equations:

E=0,B=0,×E=Bt,×B=ε0μ0Et.

Here, ε0 is the permittivity of free space, and μ0 is the permeability of free space. These equations describe how the electric field E and the magnetic field B evolve in time and space, with both fields being functions of position and time, so that E(r,t) and B(r,t).

A key property of Maxwell's equations is their linearity: any linear combination of solutions remains a valid solution. This has important implications, as we will see later. These equations have been experimentally confirmed for over a century and are fundamental to classical electrodynamics.


1.2 Maxwell's Equations in a Medium

To describe electromagnetic waves in a medium, we need a framework that accounts for the charge densities and currents at the atomic scale. The microscopic form of Maxwell's equations in a medium is given by:

1)E=ρε0,2)B=0,3)×E=Bt,4)×B=μ0(j+ε0Et),

where j is the total microscopic current density and ρ is the total microscopic charge density. It is worth giving a meaning to each equation:

  1. Gauss' Law: The electric field originates from charges. Positive charges act as sources, and negative charges act as sinks. The flux of E through a closed surface is proportional to the enclosed charge.
  2. Gauss' Law for Magnetism: There are no magnetic monopoles; magnetic field lines always form closed loops. This distinguishes magnetic fields from electric fields, which can have isolated point sources (charges).
  3. Faraday's Law of Induction: A time-dependent magnetic field creates a circulating electric field. This principle underlies electromagnetic induction, which is the basis of electrical generators, transformers, and inductors.
  4. Ampère-Maxwell Law: Magnetic fields are produced both by electric currents and by changing electric fields. The latter term, ε0Et, is known as the displacement current density and allows electromagnetic waves to propagate even in the absence of actual charge flow.

While these equations describe the fundamental behaviour of electric and magnetic fields, solving them exactly in a material by tracking every individual charge is impractical. Instead, we often work with macroscopic versions of Maxwell’s equations. To achieve this, we introduce two auxiliary fields: the electric displacement field D and the magnetic field H (sometimes called magnetic field intensity). These quantities result from an effective spatial averaging of the microscopic fields over volumes that are large compared to atomic dimensions but small compared to the wavelength of the electromagnetic fields. This averaging is justified since atomic-scale structures are typically on the order of Angstroms or nanometres, whereas relevant optical wavelengths are often hundreds of nanometres or larger.

The macroscopic fields are defined as:

D=ε0E+P,H=1μ0BM,

where P is the electric polarisation density (electric dipole moment per unit volume), and M is the magnetisation density (magnetic dipole moment per unit volume). These definitions allow us to describe the response of the medium without explicitly tracking all individual microscopic charges and currents. In a dielectric medium, the polarisation P is the macroscopic sum of the electric dipole moments induced by the electric field. The magnetisation M is defined analogously for magnetic materials. Both the displacement field D and the magnetic field H are often referred to as auxiliary fields. The polarisation and magnetisation are related to the electric field E and magnetic induction B through material-dependent relations called constitutive relations. In free space, both polarisation P and magnetisation M are zero, so D=ε0E and H=B/μ0.

In this course, we will be mainly concerned with isotropic media, meaning that the material response is independent of direction. This implies that the dielectric function εr(r,t) (relative permittivity) and the relative permeability μr(r,t) are scalars (or tensors that reduce to scalars). Furthermore, in many practical optical problems, it is sufficient to solve for the electric field alone. This is because in the non-relativistic regime, the force exerted by the magnetic component of light on charges is often much weaker than that of the electric component for many interactions. This assumption is further justified by the fact that most materials relevant to optics are non-magnetic at optical frequencies. However, one must always keep in mind the presence and role of the magnetic field.


1.3 The Material Equations

Solving Maxwell's equations in a medium requires explicit relationships, known as material or constitutive equations, which describe how the medium responds to the fields. As mentioned earlier, these relationships depend on the material properties. To establish the macroscopic Maxwell's equations, we begin by separating both the total charge density ρ and the total current density j into free and bound contributions:

ρ(r,t)=ρf(r,t)+ρb(r,t),j(r,t)=jf(r,t)+jb(r,t).

Free charges and currents are typically those that can move over macroscopic distances (like conduction electrons in a metal), while bound charges and currents are associated with localised atomic or molecular dipoles.

Our goal is to reformulate Maxwell's equations so that only free charges and currents appear explicitly as sources.
Starting from Gauss' Law in a medium, E=(ρf+ρb)/ε0, we can write ε0Eρb=ρf.
By defining the polarisation density P such that the bound charge density is given by

ρb=P,

we substitute this into Gauss' Law: ε0E+P=ρf. This can be rewritten using the electric displacement D=ε0E+P as

D=ρf.

At this point, we have successfully removed explicit dependence on the bound charges.
A similar approach applies to the Ampère-Maxwell Law. The total current j includes jf and jb. The bound current density jb can be expressed in terms of polarisation P and magnetisation M as

jb=Pt+×M.

Substituting this into the microscopic Ampère-Maxwell Law ×B=μ0(jf+jb+ε0Et):

×B=μ0(jf+Pt+×M+ε0Et).

Rearranging gives ×(Bμ0M)=jf+(ε0E+P)t.
Using the definitions of H=1μ0BM and D=ε0E+P, this becomes

×H=jf+Dt.

We can summarise the microscopic Maxwell's equations (which are universally valid) and the macroscopic Maxwell's equations (useful for describing fields in media):

Name Microscopic Maxwell's equations (in medium) Macroscopic Maxwell's equations
Gauss' Law E=ρtotalε0 D=ρf
Gauss' Law for Magnetism B=0 B=0
Faraday's Law of Induction ×E=Bt ×E=Bt
Ampère-Maxwell Law ×B=μ0(jtotal+ε0Et) ×H=jf+Dt

Additionally, the auxiliary relations defining D and H are:

D=ε0E+PandH=1μ0BM.

And the definitions relating bound sources to P and M are:

ρb=Pandjb=Pt+×M.

1.4 Macroscopic Approximation

The macroscopic quantities P and M (and thus D and H) are obtained by averaging microscopic properties over physically infinitesimal volumes that are nevertheless large enough to contain many atoms or molecules. The total charge in a macroscopic volume V at position R is

qR=Vρ(r)d3r,

while the total current through a surface element associated with this volume is related to

iR=Vj(r)d3r.

The electric dipole moment of the volume V is defined as

ptotal,R=V(rR)ρ(r)d3r,

while the magnetic dipole moment is

mtotal,R=12V(rR)×j(r)d3r.

The free charge and free current densities are then given by averages:

ρf(R,t)=qf,RVandjf(R,t)=if,RV,

while the macroscopic polarisation and magnetisation are dipole moments per unit volume:

P(R,t)=ptotal bound,RVandM(R,t)=mtotal bound,RV.

The macroscopic Maxwell equations effectively describe the fields averaged over these volumes. This approximation is valid as long as the fields do not vary significantly over the scale of the averaging volume.

To proceed further with solving problems, we require constitutive relations that link P to E and M to H (or B). These relations depend on the material's properties and are often established through a set of approximations:

  1. Electric and magnetic field dependence: It is often assumed that P depends primarily on E and not on B, while M depends primarily on B (or H) and not on E. This is a good approximation for many materials at optical frequencies, although magneto-optic effects do exist where fields cross-couple.
  2. Locality: P(r) and M(r) are assumed to depend only on the fields E(r) and B(r) at the same position r. This implies that the response is local and bound charges/currents do not move significantly relative to the scale over which the fields change. This is part of the long-wavelength approximation (wavelength much larger than atomic scales). Spatial dispersion occurs when this is not true.
  3. Homogeneity: The functional dependence of P and M on E and B respectively, does not vary with position r in the medium, implying the medium is optically homogeneous.
  4. Instantaneous Response (No Temporal Dispersion): P and M at time t are assumed to depend only on the values of E and B at the same time t, eliminating time integrals (convolutions) in the time domain. This assumption is only valid for optically transparent materials far from any absorption resonances. In reality, this is often a poor approximation for many materials over a broad range of frequencies, and temporal dispersion (frequency dependence of material parameters like εr(ω)) is crucial. This will be refined later.
  5. Linearity: P and M are assumed to be linear functions of E and B (or H), respectively. This is the domain of linear optics. Non-linear optics deals with higher-order dependencies.
  6. Isotropy: The response of the medium is assumed to be independent of the direction of the applied fields. For isotropic media, P is parallel to E, and M is parallel to H (or B). This means susceptibilities and permittivities can be treated as scalars. This assumption is violated in anisotropic materials like many crystals.

Let us examine the effect of these assumptions. With assumption 1, we may generally write the response as a functional of the field history. For instance:

P(r,t)=tfP(r,E(r,t),tt)dtandM(r,t)=tfM(r,H(r,t),tt)dt.

With assumptions 2, 3, and 4 (locality, homogeneity, and instantaneous response), these simplify to:

P(r,t)=fP(E(r,t))M(r,t)=fM(H(r,t)).

With assumption 5 (linearity), we can expand P and M in a Taylor series and keep only the linear terms (assuming no permanent dipole moments P(0),M(0), or that they are zero for symmetry reasons):

Pi(r,t)ε0jχij(1)Ej(r,t)Mi(r,t)jχij(m1)Hj(r,t).

Here, χij(1) is the linear electric susceptibility tensor and χij(m1) is the linear magnetic susceptibility tensor. Higher-order terms (like χ(2),χ(3)) are the focus of non-linear optics, but for this course, low-intensity light and linear responses are generally assumed unless stated otherwise.
Finally, with assumption 6 (isotropy), the susceptibility tensors reduce to scalars multiplied by the identity tensor, so χij(1)=χδij and χij(m1)=χmδij. This leads to:

P(r,t)ε0χE(r,t)andM(r,t)χmH(r,t).

The relative permittivity (dielectric constant) εr and relative permeability μr are then defined as:

εr=1+χandμr=1+χm.

Therefore, we can write the macroscopic constitutive relations as:

D=ε0εrEandB=μ0μrH.

In optics, we generally deal with non-magnetic media, so M0, which implies χm0 and thus μr1. In such cases, Bμ0H.


1.5 Wave Equation

To describe the propagation of light, we seek an equation that relates the temporal evolution of the fields to their spatial variation. We derive this for the case of a homogeneous, isotropic, linear, and non-magnetic (μr=1) medium, with no free charges (ρf=0) or free currents (jf=0), and initially assuming no dispersion (so εr is constant).
Consider the macroscopic curl equation (Faraday's Law):

×E=Bt.

Apply the curl operator to both sides:

×(×E)=(×B)t.

Using the vector identity ×(×A)=(A)2A, the left side becomes (E)2E.
Since ρf=0 and the medium is homogeneous, D=ε0εrE=0, which implies E=0.
Thus, ×(×E)=2E.
For the right side, we use the Ampère-Maxwell Law (macroscopic, no free currents): ×H=Dt.
Since B=μ0H (for μr=1) and D=ε0εrE, we have ×B=μ0(ε0εrE)t=μ0ε0εrEt.
Substituting these into the curled Faraday's Law:

2E=t(μ0ε0εrEt)=μ0ε0εr2Et2.

This yields the wave equation for E:

2E=μ0ε0εr2Et2.

A similar derivation yields the wave equation for H (or B).
Both equations have the form of a generic linear wave equation. The wave propagation speed, or phase velocity vp, in the medium is given by vp2=1/(μ0ε0εr).
The speed of light c in vacuum is given by c2=1/(ε0μ0), so c=1/ε0μ0.
The refractive index n of the medium is defined as the ratio of the speed of light in vacuum to the phase velocity in the medium:

n=cvp=ε0μ0εrε0μ0=εr.

This uses the assumption of a non-magnetic medium (μr=1). Therefore, the wave equation can be written as:

2E=n2c22Et2=1vp22Et2.

1.6 Solutions to the wave equation

One fundamental solution of the wave equation is the monochromatic plane wave:

E(r,t)=E0cos(ωtkr+ϕ),

where E0 is the constant amplitude vector, ω is the angular frequency, k is the wavevector, and ϕ is a phase constant. The angular frequency and the magnitude of the wavevector, k=|k|, are related by the dispersion relation:

ω=vpk=cnk.

The relation between the wavenumber k and the wavelength in the medium λn is k=2π/λn. The wavelength in vacuum is λ0=nλn.

The wave equation is linear (since derivatives are linear operators), and thus any superposition of solutions is also a solution. While plane waves are simple solutions, they form a complete basis, meaning any solution to the wave equation can be expressed as a linear combination (or integral) of plane waves (Fourier decomposition).
From E=0 (for a uniform medium with no free charges), it follows that plane waves are transverse:

E=(E0cos(ωtkr+ϕ))=E0(k)(sin(ωtkr+ϕ))=(E0k)sin(ωtkr+ϕ).

For this to be zero at all times and positions, we require E0k=0, meaning the electric field vector is perpendicular to the direction of propagation.
Similarly for the magnetic field, a solution is:

B(r,t)=B0cos(ωtkr+ϕm).

The electric and magnetic fields of a plane wave are not independent. From Maxwell's equation ×E=Bt, by substituting the plane wave solutions, we find the relation:

k×E0=ωB0.

This implies that B0 is perpendicular to both k and E0. Therefore, for a plane wave in an isotropic medium, E, B, and k form a mutually orthogonal triad. This is illustrated in the next figure:

Attachments/Notes/2025 Quantum Electronics/1 Electromagnetic Theory of Light/01_Introduction_and_theoretical_foundation.webp|700

The relationship between the amplitudes can also be expressed using the wave impedance of the medium, Z=μ/ε=μ0μr/ε0εr. For non-magnetic media (μr=1), Z=μ0/(ε0εr)=Z0/n, where Z0=μ0/ε0377Ω is the impedance of free space. Then |E0|=Z|H0|, and H0=1Z(k^×E0), where k^=k/k.
Because the electric and magnetic fields are orthogonal to the direction of propagation, these waves are also called transverse electro-magnetic (TEM) waves.
It is often more convenient to use complex notation:

E(r,t)=Re[E~0ei(krωt)]or simplyE(r,t)=E~(r)eiωt,

where E~(r)=E~0eikr is the complex amplitude (phasor), and E~0 may itself be complex to include the phase constant ϕ. The physical field is obtained by taking the real part. Often, the Re[] is dropped for brevity in intermediate calculations, but it must be reinstated when calculating real physical quantities, especially those that depend nonlinearly on the fields, such as intensity or the Poynting vector.


1.7 Polarisation

The polarisation of light describes the orientation of the electric field vector oscillation. For a plane wave propagating in the z-direction (k=kz^), the electric field vector E0 lies in the xy plane.
Linear polarisation means that the electric field vector oscillates along a fixed straight line in the xy plane:

E(z,t)=E0(x^cosαp+y^sinαp)cos(ωtkz+ϕ),

where αp is the angle of the polarisation direction with respect to the x-axis.

The underlying physics of how materials respond to polarised light relates to how their constituent charges (and thus dipole moments) interact with the electric field. A microscopic electric dipole moment is p=qL, where q is charge and L is the vector separating charges. The macroscopic polarisation P is the vector sum of these microscopic dipole moments per unit volume: if there are N dipoles per unit volume, and p is the average dipole moment, then P=Np.

Polarisation is important in the interaction of light with matter: the amount of light reflected from or transmitted through a surface depends on it (Fresnel equations), as does the amount of light absorbed in many materials. This is even more general - light scattering is often polarisation dependent. The refractive index itself can be polarisation dependent in anisotropic materials.
Light does not have to be linearly polarised. The general case is elliptical polarisation, where the tip of the electric field vector traces an ellipse in the xy plane over one optical cycle.

This can be best shown graphically. In the following figures (source), a wave oscillates (red) into the z-direction. The projections onto the x- and y-axis are in green and blue, respectively.

Linear polarisation - the total electric field vector oscillates along a straight line in the xy plane:

Attachments/Notes/2025 Quantum Electronics/1 Electromagnetic Theory of Light/linear.gif|700

Elliptical polarisation - the total electric field vector traces an ellipse in the xy plane:
Attachments/Notes/2025 Quantum Electronics/1 Electromagnetic Theory of Light/elliptical.gif|700

Circular polarisation - the total electric field vector traces a circle in the xy plane:
Attachments/Notes/2025 Quantum Electronics/1 Electromagnetic Theory of Light/circular.gif|700

It becomes clear that circular polarisation is a special case of elliptical polarisation, where the x- and y-amplitudes of the electric field components are equal (|E0x|=|E0y|), and their phase difference is ±π/2. The case of a +π/2 phase difference (e.g., Ey leads Ex) can define right-hand circular polarisation (RHCP) by one convention, while a π/2 phase difference defines left-hand circular polarisation (LHCP) (conventions vary, often depending on whether viewed from source or receiver).
A linear polariser is an optical element that transmits light of a specific polarisation while blocking light of the orthogonal polarisation. If e^ is the unit vector along the transmission axis of the polariser, then for a given input electric field Ein, the output field is its projection onto this axis:

Eout=(e^Ein)e^.

If e^=cosθpx^+sinθpy^, where θp is the angle of the polariser's axis with respect to the x-axis.


1.8 Poynting Vector and Poynting's Theorem

Light carries energy. The quantity quantifying the rate and direction of electromagnetic energy flow per unit area is the Poynting vector S, defined for instantaneous real fields as:

S=E×H.

Its units are Watts per square metre (W/m2). Note that for calculating instantaneous power flow, real physical fields E(r,t) and H(r,t) must be used, not their complex representations directly, as energy and power are real, non-linear quantities in terms of fields. The Poynting vector indicates that the energy flow is orthogonal to both the electric and magnetic fields.

The Poynting theorem expresses energy conservation for electromagnetic fields. It states that the rate of decrease of electromagnetic energy stored within a volume, plus the rate of energy flowing out through the surface of that volume, equals the rate of work done by the fields on the free charges within the volume:

ut=S+jfE,orS+ut=jfE.

The electromagnetic energy density u in a linear, isotropic medium is given by:

u=12(ED+BH).

The Poynting theorem represents the conservation or balance of energy: the power flow out of a volume plus the rate of increase of stored energy within that volume equals the negative of the power delivered to free charges (Ohmic losses if jf=σE). This form is valid for media where u is well-defined as above (e.g., non-dispersive or carefully treated dispersive cases). The term jfE represents the rate of energy conversion per unit volume from electromagnetic to other forms (like heat).

To gain some intuition, the sign of the divergence of the Poynting vector indicates the local change in energy density due to flow:

Proof of Poynting's Theorem:
We use real, instantaneous fields and currents. Start with Maxwell's curl equations (macroscopic form):

  1. ×E=Bt
  2. ×H=jf+Dt

Take the dot product of (1) with H: H(×E)=HBt.
Take the dot product of (2) with E: E(×H)=Ejf+EDt.
Subtract the second result from the first:

H(×E)E(×H)=HBtEDtEjf.

Using the vector identity (E×H)=H(×E)E(×H), the left side is (E×H).
So, (E×H)=(HBt+EDt)Ejf.
For linear, non-dispersive media, D=ε0εrE and B=μ0μrH, so EDt=E(ε0εrEt)=12t(ε0εrEE)=t(12ED).
Similarly, HBt=t(12HB).
Thus, the term in parenthesis is ut, where u=12(ED+HB).
Substituting S=E×H, we get:

S=utEjf,

which rearranges to the Poynting theorem: S+ut=jfE.

Lastly, an animation to illustrate the electromagnetic wave and its Poynting vector in vacuum:

Attachments/Notes/2025 Quantum Electronics/1 Electromagnetic Theory of Light/vacuum.gif|700
Attachments/Notes/2025 Quantum Electronics/1 Electromagnetic Theory of Light/poyntingvacuum.png|700


1.9 Timescales

If one optical cycle lasts roughly T0=2π/ω0, for example 10fs for visible light, and typical laser pulse durations are Tpulse and measurement durations are Tm, then generally we have:

TpulseT0andTmT0.

The duration of one pulse is typically many optical cycles, while the duration of measurement can range from nanoseconds to milliseconds or longer. We can often separate the electric field into a slowly-varying envelope A(r,t), and a fast oscillation at the carrier frequency ω0:

E(r,t)=Re[A(r,t)eiω0t].

This is depicted in the next figure. The overall pulse shape (left box-like behaviour in the example) is captured by the slowly-varying envelope A(r,t), while the fast (second, right) oscillation is captured by the eiω0t term:

Attachments/Notes/2025 Quantum Electronics/1 Electromagnetic Theory of Light/01_Introduction_and_theoretical_foundation 8.webp|700

Let us next explicitly calculate the instantaneous Poynting vector for such fields. If we define the complex envelopes E~(r,t) and H~(r,t) such that the physical fields are Ephys=Re[E~(r,t)eiω0t] and Hphys=Re[H~(r,t)eiω0t] (where E~ and H~ are the slowly varying complex amplitudes):

S(r,t)=Ephys(r,t)×Hphys(r,t)=12Re[E~(r,t)×H~(r,t)]+12Re[E~(r,t)×H~(r,t)ei2ω0t].

The first term is slowly varying, while the second term oscillates rapidly at 2ω0. Therefore, the Poynting vector has slowly varying contributions which change over timescales of the pulse envelope, and fast contributions changing over timescales of the optical cycle (through 2ω0).

To obtain a measure for the average energy flux over a measurement time interval TmT0, we calculate

S(r,t)Tm=1TmtTm/2t+Tm/2S(r,t)dt.

If Tm is much longer than the optical period T0=2π/ω0, the fast oscillating terms (at 2ω0) average to zero over Tm. Thus, if the envelopes E~ and H~ are approximately constant over Tm (or Tm is chosen as an integer multiple of optical periods and envelopes vary slowly within Tm), only the slowly-varying term contributes significantly to the average:

S(r,t)Tm12Re[E~(r,t)×H~(r,t)].

For stationary, monochromatic fields, the complex amplitudes E~(r) and H~(r) are independent of time (the envelope is constant). The time-averaged Poynting vector is then

S(r)=12Re[E~(r)×H~(r)].

This time-averaged Poynting vector is often used to define the optical intensity I:

I(r)|S(r)|.

The intensity is therefore the magnitude of the time-averaged Poynting vector, and its units are typically Watts per square metre (W/m2) or Watts per square centimetre (W/cm2). By measuring intensity, we average over the fast varying optical cycle, and thus typically discard direct phase information of the optical field.


1.10 Momentum of Light

Light carries not only energy but also momentum. The momentum density of an electromagnetic field in a medium with refractive index n is g=S/vp2=n2S/c2. The total momentum pfield contained in a volume is VgdV.
When light is absorbed or reflected by an object, it exerts a force (radiation pressure) due to the transfer of momentum.

The momentum Δp transferred to an object that completely absorbs an incident light pulse of energy Upulse propagating in a medium of refractive index n is Δp=(Upulse/c)nk^, where k^ is the direction of light propagation.
For a light beam with time-averaged Poynting vector S (intensity I=|S|) normally incident on a perfectly absorbing surface of area A for a duration Δt, the total energy incident is U=IAΔt. The momentum transferred is:

Δp=ncU=ncIAΔt.

If the surface is inclined such that its normal makes an angle θ with the incident Poynting vector S, the projected area normal to the beam is A=Acosθ. The energy incident on area A is IAcosθΔt. Thus, the magnitude of momentum transferred to a perfectly absorbing disk is:

Δp=ncI0ΔtAcosθ,

where I0 is the incident intensity.

The linear momentum of light finds applications such as in laser cooling and trapping of atoms, optical tweezers, and proposals for light sails for spacecraft. The force exerted is the radiation pressure. For example, for a high-power ultrafast laser with pulse duration 10fs, 10PW peak power, and an intensity of 1023W/cm2, the radiation pressure Prad=I/c (for normal incidence, perfect absorption in vacuum) can be immense, Prad(1023×104W/m2)/(3×108m/s)3.3×1018Pa, which is about 3×1013 bar.