December 7, 2006

Density of Charge Carriers in Semiconductors

What is the physical insight at the basis of the parabolic band edge approximation?

Near the band edges (extremum) the dispersion curve follows the expression below. The term <math>E_g</math> is the band gap and the effective masses are <math>m^*</math>, which is given by the curvature of the conduction and valence band).

<center>

<br>

<math>\epsilon_c ( \vec k ) = E_g + \frac

Unknown macro: {hbar^s k^2}
Unknown macro: {2m_e^*}

</math>

<br>

<math>\epsilon_v ( \vec k ) = - \frac

Unknown macro: {2m_h^*}

</math>

<br>

</center>

What is the meaning of holes?

An empty state in a band filled with electrons is called a hole. The vacant state behaves in many ways as if it were a charge carrier of positive sign and equal mass to that of the missing electron.

What role does the chemical potential play in determining the properties of a SC?

The number of charge carriers per unit volume at a given temperature is a most important property of any semiconductor. The values depend highly on the number of impurities. Impurities effect the values of <math>n_c</math>, the number of electrons in the conduction band, and <math>p_v</math>, the number of wholes in the valence band, through an effect on the chemical potential.

<center>

<br>

<math>n_c (T) = \int_

Unknown macro: {epsilon_c}

^

Unknown macro: {infty}

d \epsilon g_c (\epsilon) \frac

Unknown macro: {1}

{e^

Unknown macro: {(epsilon - mu)/k_B T}

+ 1}</math>

<br>

<math>p_v (T) = \int_{-\infty}^

Unknown macro: {epsilon_v}

d \epsilon g_v (\epsilon) \left ( 1 - \frac

{e^

Unknown macro: {(epsilon - mu)/k_B T}

+ 1 \right )</math>

<br>

<math>p_v (T) = \int_{-\infty}^

Unknown macro: {epsilon_v}

d \epsilon g_v (\epsilon) \left ( \frac

Unknown macro: {1}

{e^

Unknown macro: {(mu - epsilon)/k_B T}

+ 1 \right )</math>

<br>

</center>

Use the DOS function that is based on the 3D free electron gas.

<center>

<br>

<math>g_c(\epsilon) = \sqrt

Unknown macro: {2(epsilon - epsilon_c)}

\frac{m_c^{3/2}}

Unknown macro: {pi^2 hbar^3}

</math>

<br>

<math>g_v(\epsilon) = \sqrt

Unknown macro: {2(epsilon - epsilon_v)}

\frac{m_c^

Unknown macro: {3/2}

}

</math>

<br>

</center>

How does the position of the chemical potential relative to the band edge effect the type and density of charge carriers?

To figure out how many states are populated, multiply the Fermi function <math>f(\epsilon)</math> with the density of states <math>g(\epsilon)</math>. The product is the density of occupied states.

<p>
</p>

Based on the Fermi function, when <math>\frac

Unknown macro: {E_c - mu}
Unknown macro: {kT}

\gg 1</math> the number of electrons promoted to the conduction band is small and will occupy the lowest energy levels in that band.

<p>
</p>

Consider the following question. What is the electron and heavy hole density as a function of position of the quasi-Fermi level? Compare the exact calculation using the Fermi-Dirac distribution and the Boltzmann approximation.

<p>
</p>

The following analysis describes the electrons in the conduction band. It is also valid for the holes in the valence band replacing the electron effective mass with the hole effective mass and changing the sign of the energy difference.

<p>
</p>

In the parabolic approximation for the band edges of a bulk semiconductor, the density of states as a function of the energy is written as below.

<center>

<br>

<math>\rho_c (E) = \frac

Unknown macro: {2 pi^2}

\left ( \frac

Unknown macro: {2 m_c^*}
Unknown macro: {hbar^2}

\right ){3/2) (E - E_c)

Unknown macro: {1/2}

</math>

<br>

</center>

THe term <math>m_c^*</math> is the effective mass related to the parabolicity of the band-edge, and <math>E_c</math> is the energy of the bottom of the conduction band.

<p>
</p>

Electrons and holes inside the material (both fermions) obey to the Fermi-Dirac statistics representing the occupation of the particles over the energy levels of the system depending on the Fermi energy <math>E_F</math>.

<center>

<br>

<math>f(E) = \frac

Unknown macro: {1}

{1+ e^{\frac

Unknown macro: {E-E_F}

{kT}}</math>

<br>

</center>

For an intrinsic semiconductor at thermal equilibrium the Fermi energy lies in the middle of the band-gap and is the same for electron and holes. When a non-equilibrium situation is considered (for example under optical or electrical pumping, with carrier densities strongly exceeding the thermal values), the electron and hole distribution can be described separately through two different quasi-Fermi levels if intraband relaxation processes are much faster than interband recombination, so that quasi-thermal equilibrium within the bands can be assumed.

<p>
</p>

The product of the quasi-Fermi distribution and the density of states in the conduction band will represent the carrier distribution over the given energy levels. If we integrate this distribution over all the available energy levels we will obtain the density of electrons in the band.

<center>

<br>

<math>n = \int_

Unknown macro: {E_c}

^

\rho_c (E) f(E) dE</math>

<br>

<math>n = N_c F_

Unknown macro: {1/2}

\left (\frac

Unknown macro: {E_F^c - E_c}
Unknown macro: {kT}

\right )</math>

<br>

<math>N_c = \frac

Unknown macro: {1}
Unknown macro: {4}

\left ( \frac

Unknown macro: {2 m_c^* kT}
Unknown macro: {pi hbar^2}

\right )^

Unknown macro: {3/2}

</math>

<br>

<math>F_

(u) = \frac

Unknown macro: {2}

{\pi^{1/2}} \int_0^

Unknown macro: {infty}

\frac{x^{1/2}}{1+e^{(x-u)}}dx</math>

<br>

</center>

If the number of particles in the system is known, one can extract from the expression of <math>n</math> the value of the quasi-Fermi energy at a given temperature. The integral in the previous relation cannot be calculated analytically and requires a numerical calculation.

<p>
</p>

If the condition <math>E - E_F \gg kT</math> is valid, the Fermi distribution can be replaced by the Boltzmann distribution and there is an analytical solution of the density of particles.

<center>

<br>

<math>n = N_c e^{- \frac

Unknown macro: {E_c - E_F^c}
Unknown macro: {kT}

</math>

<br>

<math>E_F^c = E_c - kT \ln \left ( \frac

Unknown macro: {N_c}
Unknown macro: {n}

\right )</math>

<br>

</center>

How is the chemical potential engineered?

The chemical potential is engineered through the addition of impurities.

What is the Fermi Energy?

In physics and Fermi-Dirac statistics, the Fermi energy of a system of non-interacting fermions is the smallest possible increase in the ground state energy when exactly one particle is added to the system. It is equivalent to the chemical potential of the system in its ground state at absolute zero. It can also be interpreted as the maximum energy of an individual fermion in this ground state. The Fermi energy is one of the central concepts of condensed matter physics.

<p>
</p>

The position of the chemical potential is obtained from the expressions of the charge carrier density. In intrinsic semiconductors the number of electrons in the conduction band is equal to the number of holes in the valence band.

<center>

<br>

<math>n_c(T) = N_c (T) e^{-(\epsilon_c - \mu )/k_B T}</math>

<br>

<math>n_c(T) = e^{-(\mu - \epsilon_v)/k_B T}</math>

<br>

<math>n_c(T) = p_v(T)</math>

<br>

<math>n_c(T) = n_i(T)</math>

<br>

<math>n_i(T) = \sqrt

Unknown macro: {P_v(T)N_c(T)}

e^{\frac

Unknown macro: {E_g}
Unknown macro: {2k_B T}

</math>

<br>

</center>

There is then an expression of the chemical potential.

<center>

<br>

<math>\mu = \epsilon_v + \frac

Unknown macro: {1}

E_g + \frac

Unknown macro: {3}
Unknown macro: {4}

k_B T \ln \left ( \frac

Unknown macro: {m_v}
Unknown macro: {m_c}

\right )</math>

<br>

</center>

The position of the chemical potential is close to the gap center in an intrinsic semiconductor. Note that the only assumption we used was that of non-degeneracy.

What are degenerate and non-degenerate semiconductors?

In non-degenerate SC the chemical potential satisfies the relation below.

<center>

<br>

<math>\epsilon_c - \mu \gg k_B T</math>

<br>

<math>\mu - \epsilon_v \gg k_B T</math>

<br>

</center>

What is the physical significance of the exponential-like behoavior of the Fermi-function in non-degenerate semiconductors

The Fermi function shows that for <math>\frac

Unknown macro: {E_c - mu}
Unknown macro: {kT}

\gg 1</math> the number of electrons promoted to the conduction band is small and occupies the lowest energy levels in the band.

What is the density of filled states function and how does it effect the type of semiconductor

<center>

<math>n_c (T) = \int_

^

Unknown macro: {infty}

d \epsilon g_c ( \epsilon ) e^{- (\epsilon - \epsilon_c ) / k_B T} e^

Unknown macro: {(epsilon_c - mu )/k_B T}

</math>

<br>

<math>p_v (T) = \int_

^

Unknown macro: {epsilon_v}

d \epsilon g_v ( \epsilon ) e^{- (\epsilon_v - \epsilon ) / k_B T} e^

Unknown macro: {( mu - epsilon_c )/k_B T}

</math>

<br>

</center>

Because of the rapidly decaying function in the integrand only energies that within <math>k_B T</math> of the band edge contribute significantly. Assume a quadratic density of states of the form below.

<center>

<br>

<math>N_c(T) = \frac

Unknown macro: {1}
Unknown macro: {4}

\left ( \frac

Unknown macro: {2 m_c k_B T}
Unknown macro: {pi hbar^2}

\right )^

Unknown macro: {3/2}

</math>

<br>

<math>P_v(T) = \frac

Unknown macro: {4}

\left ( \frac

Unknown macro: {2 m_v k_B T}
Unknown macro: {pi hbar^2}

\right )^

Unknown macro: {3/2}

</math>

<br>

</center>

It is not possible to calculate carrier density without specific knowledge of the chemical potential. The product of the hole and electron carrier densities does not depend on the chemical potential.

What is the Law of Mass Action? How is it calculated?

<center>

<br>

<math>n_c p_v = N_c P_v e^

Unknown macro: {(epsilon_v - epsilon_c)/k_B T}

= N_c P_v e^{-E_g/k_B T}</math>

<br>

</center>

At a given temperature, the density of one carrier type can be calculated from knowledge of the other. One of the results of the law of mass action is that the product <math>np</math> is constant. By adding a large number of carriers of one type, the carrier concentration of the other type is caused to decline.

<p>
</p>

A crystal with negligible contribution of the impurities to the charge density is classified as an intrinsic semiconductor.

<center>

<br>

<math>n_x (T) = p_v (T) = n_i (T)</math>

<br>

<math>n_i = \sqrt

Unknown macro: {n_c p_v}

</math>

<br>

<math>n_i = \sqrt

Unknown macro: {N_c P_v}

e^{-E_g / 2k_BT}</math>

<br>

<math>n_c (T) = N_c (T) e^{-(\epsilon_c - \mu)/k_B T}</math>

<br>

<math>p_v (T) = P_v (T) e^{-(\mu - \epsilon_v)/k_B T}</math>

<br>

</center>

How many charge carriers does a SC have at temperature T

Calculating the position of the chemical potential in extrinsic semiconductors

<center>

Unable to render embedded object: File (Fermi_level_in_silicon_versus_T_and_doping_conc..PNG) not found.

<math>p_v - n_c + N_d^+ - N_a^- = 0</math>

<br>

<math>P_v e^

Unknown macro: {(epsilon_v - mu)/k_B T}

- N_c e^

Unknown macro: {(mu - epsilon_c)/k_B T}

+ \frac

Unknown macro: {N_d}

{1 + 2 e^{(\mu - \epsilon_d)/k_B T}} - N_c e^

+ \frac

Unknown macro: {N_d}

{1 + 2 e^{(\mu - \epsilon_d)/k_B T}}</math>

<br>

</center>

Impurity states modelled as modified hydrogen atoms

  • Consider the weakly bound 5th electron in phosphorous as a modified hydrogen atom
  • For hydrogenic donors or acceptors, think of the electron or hole, respectively, as an orbiting electron around a net fixed charge
  • Estimate the energy to free the carrier into the conduction band or valence band by using a modified expression of the energy of an electron in the hydrogen atom.

<center>

<br>

<math>E_n = \frac

Unknown macro: {me^4}
Unknown macro: {8 epsilon_0^2 h^2 n^2}

</math>

<br>

<math>E_n = - \frac

Unknown macro: {13.6}
Unknown macro: {n^2}

eV</math>

<br>

<math>E_n = \frac

Unknown macro: {m^* e^4}
Unknown macro: {8 epsilon_o^2 h^2 n^2}

\frac

Unknown macro: {1}

{\epsilon_r^2</math>

<br>

<math>E_n = - \frac

Unknown macro: {n^2}

\frac

Unknown macro: {m^*}
Unknown macro: {m}

\frac

Unknown macro: {1}
Unknown macro: {epsilon^2}

</math>

<br>

</center>

  • In the ground state <math>n=1</math>, <math>\epsilon</math> is on the order of <math>10</math>. The binding energy of the carrier to the impurity atom is <math><0.1 eV</math>.
  • Expect that many carriers are ionized at room temperature.

References

P-N Junctions

The p-n junction possesses some interesting properties which have useful applications in modern electronics. P-doped semiconductor is relatively conductive. The same is true of N-doped semiconductor, but the junction between them is a nonconductor. This nonconducting layer, called the depletion zone, occurs because the electrical charge carriers in doped n-type and p-type silicon (electrons and holes, respectively) attract and eliminate each other in a process called recombination. By manipulating this nonconductive layer, p-n junctions are commonly used as diodes: electrical switches that allow a flow of electricity in one direction but not in the other (opposite) direction. This property is explained in terms of the forward-bias and reverse-bias effects, where the term bias refers to an application of electric voltage to the p-n junction. Consider lecture 10, 11.

Doping Profile

A plot of the net doping concentration as a function of position is referred to as the doping profile.

<center>

Unable to render embedded object: File (Doping_profile.PNG) not found.

</center>

Only the doping variation in the immediate vicinity of the metallurgical junction is of prime importance. Two common idealizations are the step junction and the linearly graded junction profiles.

Poisson's Equation

Poisson's equation is a relationship from electricity and magnetism. It can be a starting point in finding quantitative solutions to electrostatic variables.

<center>

<br>

<math>\nabla \cdot E = \frac

Unknown macro: {rho}
Unknown macro: {K_s epsilon_o}

</math>

<br>

</center>

Consider when <math>E = E_x</math>.

<center>

<br>

<math>\frac

Unknown macro: {dE}
Unknown macro: {dx}

= \frac

Unknown macro: {K_s epsilon_o}

</math>

<br>

</center>

Below is the charge density inside a semiconductor assume dopants to be totally ionized.

<center>

<br>

<math>\rho = q(p - n +N_D - N_A)</math>

<br>

</center>

Qualitative Solution

There should be band bending and an internal electric field with a nonuniform doping of a pn junction diode. Assume a one-dimensional step junction. Expect regions far from the metallurgical junction behave identically to an isolated semiconductor. The fermi level is constant under equilibrium.

<center>

Unable to render embedded object: File (Step_junction_I.PNG) not found.

Unable to render embedded object: File (Step_junction_II.PNG) not found.

Unable to render embedded object: File (Step_junction_III.PNG) not found.

</center>

Deduce the functional form of the electrostatic variables. The relationship of V versus x is of the same functional form as the curve of <math>E_c</math>, <math>E_i</math>, or <math>E_v</math> upside-down. Find the <math>E</math> versus <math>x</math> relationship from the derivative of <math>E_c</math> with respect to position. Find <math>\rho</math> versus <math>x</math> from the slope of the plot of <math>E</math> versus <math>x</math>.

<center>

<br>

Unable to render embedded object: File (Step_junction_--_qVbi.PNG) not found.

<br>

Unable to render embedded object: File (Step_junction_--_electrostatic_potential.PNG) not found.

<br>

Unable to render embedded object: File (Step_junction_--_electric_field.PNG) not found.

<br>

Unable to render embedded object: File (Step_junction_--_charge_density.PNG) not found.

<br>

<math>\vec E = \int \frac

Unknown macro: {rho}
Unknown macro: {epsilon}

dx</math>

<br>

<math>V = \int \vec E dx</math>

<br>

</center>

There is a voltage drop across the junction under equilibrium conditions and there is charge near the metallurgical boundary. Consider the source of charge. In a p-material, the positive hole charges balance immobile acceptor-site charges. In an n-material, the electronic charge balances the immobile charge associated with the ionized donors. After the two materials are joined, the holes begin to diffuse from the p-side to the n-side. Electrons diffuse from the n-side to the p-side of the junction. The donors and acceptors are fixed. Unbalanced dopant site charge is left behind. There is a significant non-zero charge in the space charge region or depletion region. The built-up of charge and the associated electric field continues until the diffusion of carriers across the junction is balanced by carrier drift.

The Built-in Potential

Consider a nondegenerately-doped pn junction under equilibrium conditions. Ends of the equilibrium depletion region are at <math>-x_p</math> and <math>x_n</math>.

<center>

<br>

<math>E = - \frac

Unknown macro: {dV}
Unknown macro: {dx}

</math>

<br>

<math>- \int_{-x_p}^

Unknown macro: {x_n}

E dx = \int_

Unknown macro: {V(-x_p)}

^

Unknown macro: {V(x_n)}

dV</math>

<br>

<math>- \int_{-x_p}^

E dx = V(x_n) - V(-x_p)</math>

<br>

<math>- \int_{-x_p}^

Unknown macro: {x_n}

E dx = V_

Unknown macro: {bi}

</math>

<br>

<math>J_N = q \mu_n n E + qD_n \frac

Unknown macro: {dn}

</math>

<br>

<math>0 = q \mu_n n E + qD_n \frac

Unknown macro: {dn}
Unknown macro: {dx}

</math>

<br>

<math>E = - \frac

Unknown macro: {D_N}
Unknown macro: {mu_N}

\frac

Unknown macro: {dn/dx}
Unknown macro: {n}

</math>

<br>

<math>E = - \frac

Unknown macro: {kT}
Unknown macro: {q}

\frac

Unknown macro: {n}

</math>

<br>

<math>V_

Unknown macro: {bi}

= - \int_{-x_p}^

Unknown macro: {x_n}

E dx</math>

<br>

<math>V_

= \frac

Unknown macro: {kT}
Unknown macro: {q}

\int_

Unknown macro: {n(-x_p)}

^

Unknown macro: {n(x_n)}

\frac

Unknown macro: {n}

</math>

<br>

<math>V_

Unknown macro: {bi}

= \frac

Unknown macro: {kT}
Unknown macro: {q}

\ln \left [\frac

Unknown macro: {n(x_n)}
Unknown macro: {n(-x_p)}

\right]</math>

<br>

<math>n(x_n) = N_D</math>

<br>

<math>n(-x_p) = \frac

Unknown macro: {n_i^2}
Unknown macro: {N_A}

</math>

<br>

<math>V_

= \frac

Unknown macro: {kT}
Unknown macro: {q}

\ln \left ( \frac

Unknown macro: {N_A N_D}
Unknown macro: {n_i^2}

\right )</math>

<br>

</center>

Show that <math>qV_

Unknown macro: {bi}

</math> is the barrier to minority carrier injection.

<center>

<br>

<math>p_n = p_p e^{\frac{-qV_{bi}}

Unknown macro: {k_b T}

</math>

<br>

<math>n_p = n_n e^{\frac{-qV_

Unknown macro: {bi}

}

</math>

<br>

</center>

Depletion Approximation

The depletion approximation is a simplifying approximations used in the modeling of devices. It provides a way of obtaining approximate solutions without prior knowledge of the carrier concentrations. Two components of the approximation are below.

  • The carrier concentrations are assumed to be negligible compared to the net doping concentration in a region <math>-x_p \le x \le x_n</math> near the metallurgical junction.
  • The charge density outside the depletion region is taken to be zero

Solution of xn and xp in a step junction

<center>

<br>

<math>x_n = \left [\frac

Unknown macro: {2 K_s epsilon_0}

\frac

Unknown macro: {N_A}
Unknown macro: {N_D (N_A + N_D )}

(V_

Unknown macro: {bi}

- V_A) \right]^

Unknown macro: {1/2}

</math>

<br>

<math>x_p = \frac

Unknown macro: {N_D x_n}

</math>

<br>

<math>W = x_n + x_p</math>

<br>

</center>

Applied Bias

Depletion widths decrease under forward biasing and increase under reverse biasing. A decreased depletion width when there is a forward bias means there is less charge around the junction and a correspondingly smaller electronic field. Reverse biasing creates a large space charge region and a bigger electric field. Potential decreases at all points with a forward bias and increases at all points with a reverse bias. The potential hill shrinks in both size and extent under forward biasing, whereas reverse baising gives rise to a wider and higher potential hill. One may conceive of the diode terminals as providing direct access to the p- and n-ends of the equilibrium Fermi level. Progress from the equilibrium diagram to the forward bias diagram by moving the n-side upward by <math>qV_a</math> while holding the p-side fixed. The reverse bias is obtained from the equilibrium diagram by pulling the n-side Fermi level downward.

Forward-bias

<p>
</p>

Forward-bias occurs when the P-type block is connected to the positive terminal of a battery and the N-type block is connected to the negative terminal, as shown below.

<p>
</p>

With this set-up, the 'holes' in the P-type region and the electrons in the N-type region are pushed towards the junction. This reduces the width of the depletion zone. The positive charge applied to the P-type block repels the holes, while the negative charge applied to the N-type block repels the electrons. As electrons and holes are pushed towards the junction, the distance between them decreases. This lowers the barrier in potential. With increasing bias voltage, eventually the nonconducting depletion zone becomes so thin that the charge carriers can tunnel across the barrier, and the electrical resistance falls to a low value. The electrons which pass the junction barrier enter the P-type region (moving leftwards from one hole to the next, with reference to the above diagram).

<p>
</p>

This makes an electric current possible. An electron starts flowing around from the negative terminal to the positive terminal of the battery. It starts at the negative terminal, moving towards the N-type block. Having reached the N-type region it enters the block and makes its way towards the p-n junction. The junction barrier can no longer keep the electron in the N-type region due to the forward-bias effect (in other words, the thin depletion zone produces very little electrical resistance against the flow of electrons). The electron will therefore cross the junction and move ahead into the P-type block. Once inside the P-type region, the electron, being thermally free (from bonding)���or mobile���will move through the rest of the crystal, making its way to the positive terminal of the power supply. Please note that the electron does not jump from one hole to the next in the p-region. This actually qualifies as electron-hole recombination which immobilises both hole and electron. The electron can move freely through the crystal without needing to jump into holes which is what happens when electrons do cross the depletion layer. This process will be repeated over and over again, producing a complete circuit path through the junction.

<p>
</p>

The Shockley diode equation models the operation of a p-n junction outside the avalanche region.

<p>
</p>

Reverse Bias

<p>
</p>

Connecting the P-type region to the negative terminal of the battery and the N-type region to the positive terminal, produces the reverse-bias effect. The connections are illustrated in the following diagram:

<p>
</p>

Because the P-type region is now connected to the negative terminal of the power supply, the 'holes' in the P-type region are pulled away from the junction, causing the width of the nonconducting depletion zone to increase. Similarly, because the N-type region is connected to the positive terminal, the electrons will also be pulled away from the junction.

<p>
</p>

This effectively increases the potential barrier and greatly increases the electrical resistance against the flow of charge carriers. For this reason there will be no (or minimal) electric current across the junction.

<p>
</p>

At the middle of the junction of the p-n material, a depletion region is created to stand-off the reverse voltage. The width of the depletion region grows larger with higher voltage. The electric field grows as the reverse voltage increases. When the electric field increases beyond a critical level, the junction breaks down and current begins to flow by avalanche breakdown.

Qualitative Effect of Bias

  • Applying a potential to the ends of a diode does not increase current through drift
  • The applied voltage upsets the steady-state balance between drift and diffusion, which can unleash the flow of diffusion current
  • "Minority carrier device"

<center>

<br>

<math>n_p = n_n e^{\frac{-q(V_

Unknown macro: {bi}
  • V_a)}
    Unknown macro: {k_B T}
    </math>

<br>

<math>p_n = p_p e^{\frac{-q(V_

- V_a)}{k_B T}}</math>

</center>

  • Forward bias, which is a positive potential applied to p and a negative potential applied to n, decreases the depletion region and increases the diffusion current exponentially
  • Reverse bias, which is a negative potential applied to p and a positive potential applied to n, increases the depletion region, and no current flows ideally
  • Solve minority carrier diffusion equations on each side and determine <math>J</math> at the depletion edge.

<center>

<br>

<math>J = q \left (\frac

Unknown macro: {D_e}
Unknown macro: {L_e}

\frac

Unknown macro: {n_i^2}
Unknown macro: {N_a}

+ \frac

Unknown macro: {D_h}
Unknown macro: {L_h}

\frac

\right ) \left (e^{\frac

Unknown macro: {qV_a}

{k_B T}} - 1 \right )</math>

<br>

<math>J = J_o \left (e^{\frac

{k_B T}} - 1 \right )</math>

<br>

</center>

I-V curve

Consider the equilibrium band diagram for a pn junction. On the quasineutral n-side of the junction there are a large number of electrons a few holes. On the quasineutral p-side of the junction there are a high concentration of holes and a small number of electrons. Most carriers are of insufficient energy to "climb" the potential hill. There are some high-energy electrons that can surmount the hill and travel over to the p-side of the junction. This is the diffusion from the high-electron population n-side of the junction to the low-electron population p-side of the junction.

<p>
</p>

Electrons on the p-side are not restricted. If an electron on the p-side travels into the depletion region, it is rapidly swept over to the other side of the junction. This drift current balances the diffusion current under equilibrium conditions. A significant change resulting from a forward bias is a lowering of the potential hill between the p- and n- sides of the junction. With the potential hill decreased, more n-side electrons and p-side holes can surmount the hill and travel to the opposite side of the junction. Because the potential hill decreases linearly with the applied forward bias and the carrier concentrations vary exponentially as one progresses away from the band edges, the number of carriers of sufficient energy to surmount the potential barrier goes up exponentially with <math>V_A</math>.

<p>
</p>

A major effect of the reverse bias is to increase the potential hill between the p- and n-sides of the junction. Even a small reverse bias, anything greater than a few <math>\frac

Unknown macro: {kT}
Unknown macro: {q}

</math> in magnitude, reduces the majority carrier diffusion across the junction to a negligible level. Reverse biasing gives rise to a current flow directed from the n-side to the p-side of the junction. The reverse current is expected to saturate once the majority carrier diffusion currents are reduced to a negligible level at a small reverse bias. The overall I-V dependence is of the form below.

<center>

<br>

<math>I - I_0 \left ( e^{V_A / V_{ref}} - 1 \right )</math>

<br>

</center>

The above equation is identical to the ideal diode equation if V_

Unknown macro: {ref}

is set equal to <math>kT/q</math>.

Notes

Memorize portion of periodic table to learn what elements to use to dope

References

Initial materials selection considerations in interface light emitting devices

  • Bandgap: wavelength of emission
    • Larger atoms, weaker bonds, smaller <math>U</math>, smaller <math>E_g</math>, higher <math>\mu</math>, more costly
  • Nature of gap (direct or indirect): Radiative recombination
  • Lattice matching: Efficiency

Homojunctions and Heterojunctions

Homojunctions

  • Same material on both sides of the junction – different dopant levels
  • After junction is formed, there is constant chemical potential and continuous bands
  • Example: p-n junction in doped <math>Si</math>
  • Lattice matched by definition
  • Impurity scattering

Heterojunctions

  • Different materials on both sides of the junction – different bandgaps and work functions
  • After junction is formed there is constant chemical potential. However bands are discontinuous
  • Example: heterojunction in <math>GaAs/AlGaAs</math>
  • Need to choose lattice matched materials
  • Heterojunctions can be used as a means to create internal potentials
    • Potential barriers of holes and electrons can be created inside a material
    • There are different band gaps and electron affinity/work functions associated with different semiconductor materials
    • Internal fields from doping p-n must be superimposed on the effects.
    • Poisson Solver:

<center>

<br>

<math>\frac

Unknown macro: {dE}
Unknown macro: {dx}

= V</math>

<br>

<math>\frac

Unknown macro: {dx}

= \frac

Unknown macro: {rho}
Unknown macro: {epsilon}

</math>

<br>

</center>

In both cases the chemical potential <math>\mu</math> plays a key role in determining junction characteristics.

Lasers

Below are conditions necessary to achieve lasing action.

  • A system with more than two energy levels
  • Radiative decay path
  • Direct band gap semiconductors
  • Population inversion
  • Forward bias leads to injection of minority carriers
  • Stimulated emission
  • Reflectors for cavity

<math>GaAs</math> p-n junction laser

  • <math>GaAs</math> diode laser is doped with <math>Zn</math> (acceptor) <math>N_A = 10^
    Unknown macro: {20}
    </math> and <math>Te</math> as donor <math>N_D = 10^
    Unknown macro: {18}
    </math>. There are degenerate doping conditions such that the chemical potentials penetrate the bands.
  • <math>t</math> recombination in <math>GaAs</math> <math>10^{-9} s D - 10 cm^2/s</math>
  • <math>l = \sqrt
    Unknown macro: {Dt}
    = 1 \mu m</math>
  • Active region is determined by the large diffusion length and not the small depletion region (~<math>100nm</math>). Light emission is not well confined.
  • Donor tail merges with the conduction band at high impurity concentrations reducing the operational frequency leading to a spectral broadening and inhibiting lasing action
  • High dopant levels increase scattering and decrease carrier mobility
  • High threshold currents lead to cryogenic temperature requirements for operation, which are not practical

The solution is double heterojunction laser structures

  • The active material is <math>GaAs</math> (834 nm) or a quaternary alloy <math>InGaAsP</math>
  • There is close lattice matching of the materials
  • The refractive index of the active layer is <math>3.6</math> vs <math>3.4</math> for <math>Al_
    Unknown macro: {0.3}
    Ga_
    Unknown macro: {0.7}
    As</math>. There is strong confinement of the emitted beam to the active layer.
  • The band gap of <math>GaAs</math> is 1.42 eV while the gap in <math>Al_xGa_
    Unknown macro: {1-x}
    As</math> is about 1.8 (eV) considering <math>x = 0.3</math>. An effect is the confining of carriers to the active layer.
  • Loss is reduced due to the transparency of the <math>AlGaAs</math> layers

Quantum Wells

Approximate a well in a heterojunction structure as containing infinite potential boundaries. Modify electronic transitions through quantum wells.

<center>

<br>

<math>k = \frac

Unknown macro: {n pi}
Unknown macro: {L}

</math>

<br>

<math>E = \frac

Unknown macro: {hbar^2 k^2}
Unknown macro: {2 m^*}

</math>

<br>

<math>E = \frac

Unknown macro: {h^2 n^2}
Unknown macro: {8 m^* L^2}

</math>

<br>

</center>

Solar Cells

When useful

Two-terminal devices:

Three-terminal devices:

Four-terminal devices:

Multi-terminal devices:

Reference

Numerical Methods

Variational Principle

Say you have a system for which you know what the energy depends on, or in other words, you know the Hamiltonian H. If you cannot solve the Schr��dinger equation to figure out the wavefunction, you can guess any normalized wavefunction whatsoever, say ��, and it turns out that the expectation value of the hamiltonian for your guessed wavefunction will be greater than the actual ground state energy. Or in other words:

<center>

<br>

<math>E_

Unknown macro: {ground}

\le \left\langle\phi|H|\phi\right\rangle </math>

<br>

</center>

This holds for any �� you could have guessed!

<p>
</p>

For a hamiltonian <math>H</math> that describes the studied system and any normalizable function <math>\Psi</math> with arguments appropriate for the unknown wave function of the system, we define the functional.

<center>

<br>

<math> \varepsilon\left[\Psi\right] = \frac{\left\langle\Psi|\hat

Unknown macro: {H}

\Psi\right\rangle}

Unknown macro: {leftlanglePsi|Psirightrangle}

</math>

<br>

</center>

The variational principle states that

  • <math>\varepsilon \geq E_0</math>, where <math>E_0</math> is the lowest energy eigenstate (ground state) of the hamiltonian
  • <math>\varepsilon = E_0</math> if and only if <math>\Psi</math> is exactly equal to the wave function of the ground state of the studied system.

<p>
</p>

The variational principle formulated above is the basis of the variational method used in quantum mechanics and quantum chemistry to find approximations to the ground state.

<p>
</p>

Your guessed wavefunction, <math>\psi</math>, can be expanded as a linear combination of the actual eigenfunctions of the Hamiltonian (which we assume to be normalized and orthogonal)

<center>

<br>

<math>\phi = \sum_

Unknown macro: {n}

c_

\psi_

Unknown macro: {n}

\,</math>

<br>

</center>

Then, to find the expectation value of the hamiltonian:

<center>

<br>

<math>\left\langle\phi|H|\phi\right\rangle = \left\langle\sum_

c_

Unknown macro: {n}

\psi_

|H|\sum_

Unknown macro: {m}

c_

\psi_

Unknown macro: {m}

\right\rangle \,</math>

<br>

<math>\left\langle\phi|H|\phi\right\rangle = \sum_

Unknown macro: {n}

\sum_

\left\langle c_

Unknown macro: {n}

\psi_

|E_

Unknown macro: {m}

c_

\psi_

Unknown macro: {m}

\right\rangle \,</math>

<br>

<math>\left\langle\phi|H|\phi\right\rangle = \sum_

Unknown macro: {n}

\sum_

c_

Unknown macro: {n}

^*c_

Unknown macro: {m}

E_

\left\langle\psi_

|\psi_

Unknown macro: {m}

\right\rangle \,</math>

<br>

<math>\left\langle\phi|H|\phi\right\rangle = \sum_

Unknown macro: {n}

c_

|^2 E_

Unknown macro: {n}

\,</math>

<br>

</center>

Now, the ground state energy is the lowest energy possible, i.e. <math>E_

\ge E_

Unknown macro: {g}

</math>. Therefore, if the guessed wave function <math>\psi</math> is normalized:

<center>

<br>

<math>\left\langle\phi|H|\phi\right\rangle \ge E_

\sum_

Unknown macro: {n}

c_

|^2 = E_

Unknown macro: {g}

\,</math>

<br>

<math>E[\phi] - E_o = \frac

Unknown macro: {sum_n |a_n|^2(E_n - E_o)}
Unknown macro: {sum_n |a_n|^2}

</math>

<br>

</center>

Energy of a Hydrogen Atom

<center>

<br>

<math>E_

Unknown macro: {alpha}

= \frac{ < \Psi_

| \hat H | \Psi_

Unknown macro: {alpha}

> }{ \Psi_

| \Psi_

Unknown macro: {alpha}

> }</math>

<br>

<math>\Psi_

= C e^{(- \alpha r)</math>

<br>

<math>< \Psi_

Unknown macro: {alpha}

\Psi_

> = \pi \frac

Unknown macro: {C^2}
Unknown macro: {alpha^3}

</math>

<br>

<math>< \Psi_

Unknown macro: {alpha}
  • \frac
    Unknown macro: {1}
    Unknown macro: {2}
    \nabla^2

\Psi_

> = \pi \frac

Unknown macro: {2 alpha}

</math>

<br>

<math>< \Psi_

Unknown macro: {alpha}
  • \frac
    Unknown macro: {1}
    Unknown macro: {r}
    \nabla^2

\Psi_

> = \pi \frac

Unknown macro: {C^2}
Unknown macro: {alpha^2}

</math>

<br>

</center>

Energy of a collection of atoms

<center>

<br>

<math>\hat H = \hat T_e + \hat V_

Unknown macro: {e-e}

+ \hat V_

Unknown macro: {e-N}

+ V_

Unknown macro: {N-N}

</math>

<br>

<math>T_e: \mbox

Unknown macro: {quantum kinetic energy of the electrons}

</math>

<br>

<math>\hat T_e = - \frac

Unknown macro: {1}
Unknown macro: {2}

\sum_i \nabla_i^2</math>

<br>

<math>V_

: \mbox

Unknown macro: {electron-electron interactions}

</math>

<br>

<math>\hat V_

Unknown macro: {e-e}

= \sum_i \sum_

Unknown macro: {j>1}

\frac

Unknown macro: {1}
Unknown macro: {|vec r_i vec r_j|}

</math>

<br>

<math>V_

Unknown macro: {e-N}

: \mbox

Unknown macro: {electrostatic electron-nucleus attraction}

</math>

<br>

<math>\hat V_

= \sum_i \left [\sum_

Unknown macro: {I}

V ( \vec R_I - \vec r_i ) \right]</math>

<br>

<math>V_

Unknown macro: {N-N}

: \mbox

Unknown macro: {electrostatic nucleus-nucleus repulsion}

</math>

<br>

</center>

Molecules and Solids: Electrons and Nuclei

<center>

<br>

<math>\hat H \Psi ( \vec r_1, ..., \vec r_n, \vec R_1, ...., \vec R_N ) = E_

Unknown macro: {tot}

\Psi ( \vec r_1, ..., \vec r_n, \vec R_1, ...., \vec R_N )</math>

<br>

</center>

Treat only the electrons as quantum particles in the field of fixed or slowly varying nuclei. This is generally called the adiabatic or Born-Oppenheimer approximation. "Adiabatic" means that there is no coupling between different electronic surfaces; "B-O" implies there is no influence of the ionic motion on one electronic surface.

Linear Combination of Atomic Orbitals

This is the most common approach to find out the ground-state solution. It allows a meaningful definition of "hybridization", "bonding" and "anti-bonding" orbitals. It is also known as LCAO, LCAO-MO (molecular orbitals), or tight-binding (solids). The trial wavefunction is a linear combination of atomic orbitals. The variational parameters are the coefficients.

<center>

<br>

<math>\Psi_

Unknown macro: {trial}

= \sum_

Unknown macro: {I, (nlm)}

= c_

Unknown macro: {(nlm)}

^I \Psi_

^I (\vec r - \vec R_I)</math>

<br>

<math>E_

Unknown macro: {LCAO}

= \mbox

Unknown macro: {min}

\frac{ < \Psi_

| \hat H | \Psi_

Unknown macro: {trial}

> }{ <\Psi_

| \Psi_

Unknown macro: {trial}

></math>

<br>

</center>

Huckel approach

Huckel: planar / quasi-planar systems with delocalized <math>\pi</math> bonding. Consider two parameters.

  • <math>\alpha</math>: matrix element between same orbital
  • <math>\beta</math>: matrix element between neighboring orbitals
  • Hamiltonian between further neighbors is zero

Consider benzene

<center>

<br>

<math>\det \begin

Unknown macro: {bmatrix}

\alpha - E & \beta & 0 & 0 & 0 & \beta
\beta & \alpha - E & \beta & 0 & 0 & 0
0 & \beta & \alpha - E & \beta & 0 & 0
0 & 0 & \beta & \alpha - E & \beta & 0
0 & 0 & 0 & \beta & \alpha - E & \beta
\beta & 0 & 0 & 0 & \beta & \alpha - E \end

= 0</math>

<br>

</center>

Reference:

Hartree

Reduce the interacting many-electron system to an individual electron problem in an effective potential. This potential should be determined self-consistently by all other electrons in the system. Neglect the antisymmetric requirement as a first step, and the total wavefunction for a system with <math>N</math> electrons can be written as the product of one-electron wavefunctions.

<center>

<br>

<math>\Psi ( \vec r_1, ..., \vec r_N ) = \Pi_

Unknown macro: {i=1}

^N \Psi_i (\vec r_1)</math>

<br>

</center>

Hartree suggested a variational calculation to minimize the energy.

<center>

<br>

<math>E = \frac

Unknown macro: { <Psi | H | Psi > }
Unknown macro: { <Psi | Psi ></math> <br> </center> If <math>Psi</math> were the exact ground state wavefunction of the system, then <math>E</math> would be the ground state energy. The variational principle states that <math>E</math> is stationary with respect to variation of <math>Psi</math>, and is an upper bound to the ground state energy. The procedure leads to a set of Hartree equations. <center> <br> <math>left [vec r - vec r'|}

\right|- \frac

Unknown macro: {h^2}
Unknown macro: {2m}

\nabla^2 + v(\vec r) + \sum'_j e^2 \int \frac

Unknown macro: {Psi_j^* (vec r') Psi_j (vec r') d vec r'}

{] \Psi_i(\vec r) = \epsilon_i \Psi_i (\vec r)</math>

<br>

</center>

The prime is used to rule out the possibility of <math>j=i</math> and <math>\epsilon_i</math> are variational parameters which look like the one-electron energy eigenvalues.

<p>
</p>

Reference

Mean-field approach

  • Independent particle model (Hartree): each electron moves in an effective potential representing the attraction of the nuclei and the average effect of the repulsive interactions of the other electrons
  • This average repulsion is the electrostatic repulsion of the average charge density of all other electrons.

Hartree Equations

  • The Hartree equations can be obtained directly from the variational principle, once the search is restricted to the many-body wavefunctions that are written as the product of spin-orbitals. Independent electrons are considered.

<center>

<br>

<math>\Psi (\vec r_1, ..., \vec r_n ) = \psi_1 (\vec r_1) \psi_2 (\vec r_2)...\psi_n (\vec r_n)</math>

<br>

<math>\left [| \psi_j (\vec r_j ||^2 \frac

Unknown macro: {1}

{|\vec r_j - \vec r_i| d \vec r_j \right|\frac

Unknown macro: {2}

\nabla_i^2 + \sum_i V (\vec R_I - \vec r_i) + \sum_

Unknown macro: {j ne i}

\int] \psi_i (\vec r_i) = \epsilon \psi_i (\vec r_i)</math>

<br>

</center>

The self-consistent field

  • The single-particle Hartree operator is self-consistent. It depends on itself on the orbitals that are the solution of all other Hartree equations.
  • There are <math>n</math> simultaneous integro-differential equations for the <math>n</math> orbitals
  • Solution is achieved iteratively

Reference

Spin Statistics

  • All elementary particles are either fermions (half-integer spins) or bosons (integer)
  • A wavefunction that is antisymmetric by exchange is associated with a set of identical (indistinguishable) fermions
    • <math>\Psi(\vec r_1, \vec r_2,...,\vec r_j,...,\vec r_k,...,\vec r_n) = \Psi(\vec r_1, \vec r_2,...,\vec r_k,...,\vec r_j,...,\vec r_n)</math>
  • For bosons, it is symmetric

Reference

Slater determinant

An antisymmetric wavefunction is constructed via a Slater determinant of the individual orbitals instead of just a product, as in the Hartree approach)

<center>

<br>

<math>\Psi(\vec r_1, \vec r_2, ..., \vec r_n) = \frac

Unknown macro: {1}

{\sqrt{n!}} \begin

Unknown macro: {Vmatrix}

\psi_

Unknown macro: {alpha}

(\vec r_1 ) & \psi_

Unknown macro: {beta}

(\vec r_1 ) & ... & \psi_

Unknown macro: {v}

(\vec r_1 )
\psi_

(\vec r_2 ) & \psi_

Unknown macro: {beta}

(\vec r_2 ) & ... & \psi_

Unknown macro: {v}

(\vec r_2 )
... & ... & ... & ...
\psi_

Unknown macro: {alpha}

(\vec r_n ) & \psi_

(\vec r_n ) & ... & \psi_

Unknown macro: {v}

(\vec r_n )\end

</math>

<br>

</center>

Consider an example of two electrons in <math>H_2</math>.

<center>

<br>

<math>\Psi(\vec r_1, \vec r_2) = \frac

{\sqrt{2}} \begin

Unknown macro: {Vmatrix}

\psi_

Unknown macro: {alpha}

(\vec r_1 ) & \psi_

Unknown macro: {beta}

(\vec r_1 )
\psi_

(\vec r_1 ) & \psi_

Unknown macro: {beta}

(\vec r_2 ) \end

</math>

<br>

<math>\Psi(\vec r_1, \vec r_2) = \frac

Unknown macro: {1}

{\sqrt{2}} \left [\psi_

Unknown macro: {alpha}

(\vec r_1 ) \psi_

Unknown macro: {beta}

(\vec r_1 ) - \psi_

(\vec r_2 ) \psi_

Unknown macro: {beta}

(\vec r_2 ) \right]</math>

<br>

</center>

Reference

Pauli Principle

If two states are identical, the determinant vanishes. There can't be two electrons in the same quantum state.

Reference

Hartree-Fock

In computational physics and computational chemistry, the Hartree-Fock (HF) method is an approximate method for the determination of the ground-state wavefunction and ground-state energy of a quantum many-body system.

<p>
</p>

The Hartree-Fock method assumes that the exact, N-body wavefunction of the system can be approximated by a single Slater determinant (in the case where the particles are fermions) or by a single permanent (in the case of bosons) of N spin-orbitals. Invoking the variational principle one can derive a set of N coupled equations for the N spin-orbitals. Solution of these equations yields the Hartree-Fock wavefunction and energy of the system, which are approximations of the exact ones.

<p>
</p>

The Hartree-Fock method finds its typical application in the solution of the electronic Schr��dinger equation of atoms, molecules and solids but it has also found widespread use in nuclear physics. See Hartree-Fock-Boloyubov for a discussion of its application in nuclear structure theory. The rest of this article will focus on applications in electronic structure theory.

<p>
</p>

The Hartree-Fock method is also called, especially in the older literature, the self-consistent field method (SCF) because the resulting equations are almost universally solved by means of an iterative, fixed-point type algorithm (see the following section for more details). This solution scheme is not the only one possible and is not specific of the Hartree-Fock method. Therefore "self-consistent field" is a potentially ambiguous denomination.

Algorithm

The Hartree-Fock method is typically used to solve the time-independent Schr��dinger equation for a multi-electron atom or molecule described in the fixed-nuclei approximation by the electronic molecular Hamiltonian. Because of the complexity of the differential equations for any but the smallest systems, the problem is usually impossible to solve analytically, and so the numerical technique of iteration is used. The method makes four major simplifications in order to deal with this task:

  • The Born-Oppenheimer approximation is inherently assumed. The true wavefunction is actually a function of the coordinates of each of the nuclei, in addition to those of the electrons.
  • Typically, relativistic effects are completely neglected. The momentum operator is assumed to be completely non-relativistic.
  • The basis set is composed of a finite number of orthogonal functions. The true wavefunction is a linear combination of functions from a complete basis set.
  • The energy eigenfunctions are assumed to be anti-symmetrized linear combinations of products of one-electron wavefunctions. The effects of electron correlation, beyond that of exchange energy resulting from the anti-symmetrization of the wavefunction, are completely neglected.

<p>
</p>

The variational theorem states that, for a time-independent Hamiltonian operator, any trial wavefunction will have an energy expectation value that is greater than or equal to the true ground state wavefunction corresponding to the given Hamiltonian. Because of this, the Hartree-Fock energy is an upper bound to the true ground state energy of a given molecule. The limit of the Hartree-Fock energy as the basis set becomes infinite is called the Hartree-Fock limit. It is a unique set of one-electron orbitals, and their eigenvalues.

<p>
</p>

The starting point for the Hartree-Fock method is a set of approximate one-electron orbitals. For an atomic calculation, these are typically the orbitals for a hydrogenic atom (an atom with only one electron, but the appropriate nuclear charge). For a molecular or crystalline calculation, the initial approximate one-electron wavefunctions are typically a linear combination of atomic orbitals. This gives a collection of one electron orbitals that, due to the fermionic nature of electrons, must be anti-symmetric. This antisymmetry is achieved through the use of a Slater determinant.

<p>
</p>

At this point, a new approximate Hamiltonian operator, called the Fock operator, is constructed. The first terms in this Hamiltonian are a sum of kinetic energy operators for each electron, the internuclear repulsion energy, and a sum of nuclear-electronic coulombic attraction terms. The final set of terms models the electronic coulombic repulsion terms between each electron with a sum. The sum is composed of a net repulsion energy for each electron in the system, which is calculated by treating all of the other electrons within the molecule as a smooth distribution of negative charge. This is the major simplification inherent in the Hartree-Fock method, and is equivalent to the fourth simplification in the above list, (see post-Hartree-Fock).

<p>
</p>

The newly constructed Fock operator is then used as the Hamiltonian in the time-independent Schr��dinger Equation. Solving the equation yields a new set of approximate one-electron orbitals. This new set of orbitals is then used to construct a new Fock operator, as in the preceding paragraph, beginning the cycle again. The procedure is stopped when the change in total electronic energy is negligible between two iterations. In this way, a set of so-called "self-consistent" one-electron orbitals are calculated. The Hartree-Fock electronic wavefunction is then equal to the Slater determinant of these approximate one-electron wavefunctions. From the Hartree-Fock wavefunction, any chemical property of the system in question can be calculated in an approximate manner.

Mathematical Formulation

Because the electron-electron repulsion term of the electronic molecular Hamiltonian involves the coordinates of two different electrons, it is necessary to reformulate it in an approximate way. Under this approximation, (outlined under Hartree-Fock algorithm), all of the terms of the exact Hamiltonian except the nuclear-nuclear repulsion term are re-expressed as the sum of one-electron operators outlined below. The "(1)" following each operator symbol simply indicates that the operator is 1-electron in nature.

<center>

<br>

<math>\hat F(1) = \hat H^

Unknown macro: {core}

(1)+\sum_

Unknown macro: {j=1}

^

Unknown macro: {n/2}

[2\hat J_j(1)-\hat K_j(1)]</math>

<br>

</center>

where:

<center>

<br>

<math>\hat F(1)</math>

<br>

</center>

is the one-electron Fock operator,

<center>

<br>

<math>\hat H^

(1)=-\frac

\nabla^2_1 - \sum_

Unknown macro: {alpha}

\frac

Unknown macro: {Z_alpha}

{r_{1\alpha}}</math>

<br>

</center>

is the one-electron core Hamiltonian,

<center>

<br>

<math>\hat J_j(1)</math>

<br>

</center>

is the Coulomb operator, defining the electron-electron repulsion energy due to the j-th electron,

<center>

<br>

<math>\hat K_j(1)</math>

<br>

</center>

is the exchange operator, defining the electron exchange energy. Finding the Hartree-Fock one-electron wavefunctions is now equivalent to solving the eigenfunction equation:

<center>

<br>

<math>\hat F(1)\phi_i(1)=\epsilon_i \phi_i(1)</math>

<br>

</center>

where <math>\phi_i(1)</math> are a set of one-electron wavefunctions, called the Hartree-Fock Molecular Orbitals.

Example-benzene

Using the Huckel model, consider only the relevant molecular orbitals around the HOMO-LUMO gap. This is given by combinations of <math>p_z</math> orbitals sitting on each carbon atom. The basis consists of six orbitals, one on each of the six carbons of benzene. The diagaonal elements of the Hamiltonian are <math>\alpha</math> and the off-diagonal elements are zero. In the case that two orbitals are considered as next neighbors, the matrix elements is given by <math>-|\beta|</math>.

<p>
</p>

There are six eigenvalues found when solving the matrix involved with the Hamiltonian. As the number of atoms increases, the bandwidth between the lowest and the highest states tends to a constant (<math>-\beta</math>) and the gap goes to zero. In the rings the Hamiltonian commutes with the rotation operator. The formalism is identical to the case of solids, and the <math>N</math> atoms in the ring are equivalent to the <math>N</math> unit cells of the Born-von Karman conditions.

Hartree-Fock Equations

The Hartree-Fock equations are, again, obtained from the variational principle: we look for the minimum of the many-electron Schrodinger equation in the class of all wavefunctions that are written as a single Slater determinant.

<center>

<br>

<math>\left [-\frac

Unknown macro: {1}
Unknown macro: {2}

\nabla_i^2 + \sum_I V (\vec R_I - \vec r_i ) \right] \psi_

Unknown macro: {lambda}

(\vec r_i) + \left [| \psi_

Unknown macro: {mu}

(\vec r_j )||^2 \frac

Unknown macro: {|vec r_j - vec r_i | }

d \vec r_j \right|\sum_

Unknown macro: {mu}

\int] \psi_

Unknown macro: {lambda}

) - \sum_

\left [\vec r_j - \vec r_i |} \psi_

Unknown macro: {mu}

(\vec r_i) \psi_

Unknown macro: {lambda}

(\vec r_j) d \vec r_j \right|\int \psi_

^* \frac

Unknown macro: {1}

{] = \epsilon \psi_

Unknown macro: {lambda}

(\vec r_i)
</math>

<br>

<math>\Psi(\vec r_i,...,\vec r_n) = ||Slater||</math>

<br>

</center>

Reference

Tight binding

In the tight binding model, it is assumed that the full Hamiltonian <math>H</math> of the system may be approximated by the Hamiltonian of an isolated atom centred at each lattice point. The atomic orbitals <math>\Psi_n</math>, which are eigenfunctions of the single atom Hamiltonian Hat, are assumed to be very small at distances exceeding the lattice constant. This is what is meant by tight-binding. It is further assumed that any corrections to the atomic potential <math>\Delta U</math>, which are required to obtain the full Hamiltonian <math>H</math> of the system, are appreciable only when the atomic orbitals are small. The solution to the time-independent single electron Schr��dinger equation <math>\psi</math> is then assumed to be a linear combination of atomic orbitals

<center>

<br>

<math>\phi(\vec

Unknown macro: {r}

) = \sum_n b_n \psi_n(\vec

)</math>

<br>

</center>

This leads to a matrix equation for the coefficients <math>b_n</math> and Bloch energies <math>\varepsilon</math> of the form below.

<center>

<br>

<math>\varepsilon(\vec

Unknown macro: {k}

) = E_m - {\beta_m + \sum_{\vec

Unknown macro: {R}

\neq 0} \gamma_m(\vec

) e^{i \vec

\cdot \vec{R}}\over b_m + \sum_{\vec

Unknown macro: {R}

\neq 0} \alpha_m(\vec

) e^{i \vec

Unknown macro: {k}

\cdot \vec

Unknown macro: {R}

}}</math>

<br>

</center>

where <math>E_m</math> is the energy of the <math>m</math>th atomic level, and below are the overlap integrals.

<center>

<br>

<math> \beta_m = -\int \psi_m^*(\vec

Unknown macro: {r}

)\Delta U(\vec

) \phi(\vec

Unknown macro: {r}

) d\vec

</math>

<br>

<math> \alpha_m(\vec

) = \int \psi_m^*(\vec

Unknown macro: {r}

) \phi(\vec

-\vec

Unknown macro: {R}

) d\vec

Unknown macro: {r}

</math>

<br>

<math> \gamma_m(\vec

) = -\int \psi_m^*(\vec

Unknown macro: {r}

) \Delta U(\vec

) \phi(\vec

Unknown macro: {r}

-\vec

Unknown macro: {R}

) d\vec

</math>

<br>

</center>

The tight binding model is typically used for calculations of electronic band structure and energy gaps in the static regime. However, in combination with other methods such as the random phase approximation (RPA) model, the dynamic response of systems may also be studied.

Koopman's Theorems

The total energy is invariant under unitary transformation. It is not the sum of the canonical MO orbital energies. The ionization energy and electron affinity are given by the eigenvalue of the respective molecular orbital in the frozen orbitals approximation.

What is missing

Correlations(by definition)

  • Dynamical correlations: the electrons get too close to each other in H.-F.
  • Static correlations: a single determinant variational class is not good enough

Spin contamination

  • Even if the energy is correct (variational, quadratic) other properties might not (e.g. the UHF spin is an equal mixture of singlet and triplet)

Reference:

Amorphous Semiconductor

Properties

Amorphous materials have attracted much attention. The first reason is potential industrial applications as suitable materials for fabricating devices, and the second reason is a lack of understanding of many properties which are very different from crystalline materials. An ideal crystal is defined as an atomic arrangement that has infinite translational symmetry in all three dimensions, whereas such a definite definition is not possible for an ideal amorphous solid. An amorphous solid is defined as one that does not maintain long-range translational symmetry or has only short-range order, it does not have the same precision in its definition, because long- or short-range order is not precisely defined. In addition to surface atoms in amorphous materials, however, there are also present other structural disorders due to different bond lengths, bond angles, and coordination numbers at individual atomic sites.

<p>
</p>

The theory of amorphous systems is relatively difficult, because some of the techniques of simplification for deriving analytical results in crystals cannot be applied to amorphous structures. One needs to depend heavily on numerical simulations using computers, which itself is a relatively new field.

<p>
</p>

It is commonly established that there are three types of structural disorders in an amorphous solid which do not exist in crystalline solids. These are stated below.

  • Different bond lengths
  • Different bond angles
  • Under- and over-coordinated sites

Bond States

In amorphous silicon, all silicon atoms are bonded covalently but it is not necessary that all atomic sites are of the same coordination number four. Some are under coordinated, which means that one or more covalent electrons on a silicon atom cannot form covalent bonds with neighboring atoms. These uncoordinated bonds are called dangling bonds. The density of dangling bonds in amorphous silicon is very high, which reduces the photoconductivity of the material and also prevents it from doping. The hydrogenation of amorphous silicon saturates many dangling bonds and makes it more suitable for fabricating devices.

<p>
</p>

The presence of strained and weak bonds gives rise in a-Si:H to band tail states, which are also found in other a-semiconductors and insulators. A crystalline semiconductor has quite well defined valence and conduction band edges, and hence a very well defined electron energy gap between the top of the valence band and bottom of the conduction band.

<p>
</p>

In amorphous semiconductors, the neutral dangling bond states lie in the middle of the energy band gap, and bonding and anti-bonding orbital of weak bonds lie above the valence band and below the conduction band edges respectively.

<p>
</p>

In addition to the neutral dangling bond states, there may also exist charged dangling bond states. If the charge carrier-phonon interaction is very strong in an amorphous solid, due to the negative <math>U</math> effect, then the positive charged dangling bond states would lie above the neutral bond state, but below the conduction band edge. The negative charged dangling bond states would lie below the neutral dangling bond state, but above the valence band edge. In the case of weak charge carrier-phonon interactions, these positions are reversed on the energy scale. These energy states found within the energy gap in amorphous solids are localized states and any charge carrier created in these states will be localized on some weak, strained or dangling bonds.

<p>
</p>

An interesting point is that as the band tail states lie above the valence and below the conduction band edges, these states are usually the highest occupied and lowest empty energy states in any amorphous semiconductor. Band tail states play the dominant role in most optical and electronic properties of a-semiconductors, particularly in the low temperature region.

<p>
</p>

The electronic energy states of amorphous semiconductors consist of delocalized states like valence and conduction bands, commonly known as the extended states and the localized states like band tails and dangling bond states. The extended states arise due to short-range order, and tail states due to disorder.

Band tail, mobility edge, and dangling bond states

In addition to the fully coordinated covalent bonds, there are also many weak, strained and even uncoordinated or dangling bonds in amorphous semiconductors. The bonding and anti-bonding states of weak and strained bonds lie close the valence and conduction band state edges, respectively, because the bond lengths are relatively larger than those giving rise to the extended states. This is clear from the fact that the energy gap between bonding and anti-bonding states reduces if the distance between the bonding atoms increases. For larger separation between atoms the energy band gap, which is similar to the separation between bonding and anti-bonding states, is smaller.

<p>
</p>

The presence of weak and strained bond states gives rise to band tail states or tail states within the energy gap and near the extended state edges in amorphous semiconductor. These states are localized states.

<p>
</p>

The edge separating the conduction extended states and tail states is called the mobility edge. As the tail states are localized energy states, no conduction is expected to occur when excited electrons occupy these states. At <math>0 K</math> only conduction can occur when excited electrons are in the extended states above the conduction tail states, and that defines the mobility edge, which is the energy above which the electronic conduction can occur at <math>0 K</math>. One can define a similar edge separating the valence extended states from valence tail states. Thus there are two mobility edges, electron mobility edge at the bottom of the conduction extended states and hole mobility edge at the top of valence extended states.

<p>
</p>

Dangling bonds contribute to non-bonded states, and therefore they can be regarded as equivalent to the states of isolated atoms, which lied in the middle of the bonding and anti-bonding energy states. The dangling bond states lie in the middle of the energy gap between the edges of the valence and conduction extended states. The energy gap in amorphous semiconductors is not well defined as that in crystalline semiconductors due to the presence of tail states. The dangling bond states are also localized energy states.

<p>
</p>

The influence of the presence of dangling bonds on the electronic and optical properties of amorphous semiconductors can be very significant, depending on their numbers, because these bonds do not facilitate electronic conduction except through hopping or tunneling.

<p>
</p>

Structure

A fundamental understanding of the properties of condensed matter, whether electronic, optical, chemical, or mechanical, requires a detailed knowledge of microscopic structure (atomic arrangement). The structure of a crystalline solid is determined by studying its structure within the unit cell. The structure of the crystal as a whole is then determined by stacking unit cells. Such a procedure is impossible in determining the structure of amorphous solids. Due to the lack of long range periodicity in amorphous solids, unlike crystalline solids, determination of structure is very difficult. There is no technique to provide atomic resolution in amorphous solids compared with that in crystals.

<p>
</p>

The diffraction measurements give the structure <math>S(\vec Q)</math> with scattering vector <math>\vec Q</math>. The Fourier transform of <math>S(\vec Q)</math> produces the radial distribution function (RDF). The RDF studies show that the structure of many amorphous solids is non-random and there is a considerable degree of local ordering despite the lack of long-range order. The RDF, <math>J(r)</math>, is defined as the number of atoms lying at distances between <math>r</math> and <math>r+dr</math> and is given by the expression below, where the density function , <math>\rho (r)</math>, is an atomic pair correlation function. The function exhibits oscillatory behavior, because peaks in the probability function represent average interatomic separations.

<center>

<br>

<math>J(r) dr = 4 \pi r^2 \rho (r) dr</math>

<br>

</center>

The RDF oscillates about the average density parabola given by the curve <math>4 \pi r^2 \rho_o</math>. The position of the first peak in the RDF produces the average nearest-neighbor bond length <math>r_1</math> and the position of the second peak gives the next-nearest-neighbor distance <math>r_2</math>. The knowledge of <math>r_1</math> and <math>r_2</math> yields the bond angle <math>\theta</math>.

<center>

<br>

<math>\theta = 2 \sin^{-1} \frac

Unknown macro: {r_2}
Unknown macro: {2 r_1}

</math>

<br>

</center>

The area under a peak gives the coordination number of the structure. The second peak is generally wider than the first for covalent amorphous solids, which can be attributed to a static variation in the bond angle <math>\theta</math>. If no bond-angle variation exists, then the width of the first two peaks should be equal.

Density of States

<center>

<br>

<math>\mbox

Unknown macro: {FE gas}

</math>

<br>

<math>\frac

Unknown macro: {dN}
Unknown macro: {d epsilon}

= g(\epsilon) = \frac

Unknown macro: {hbar^2 pi^2}

\sqrt{ \frac

Unknown macro: { 2 m epsilon }

{ \hbar^2}}</math>

<br>

<math>\mbox

Unknown macro: {Crystalline Material}

</math>

<br>

<math>g_n(\epsilon) = \frac

Unknown macro: {1}
Unknown macro: {4 pi^3}

\int_{S_{n(\epsilon)}} \frac

Unknown macro: { dS }
Unknown macro: { | vec nabla_k epsilon (k) | }

</math>

<br>

</center>

Band Diagram

<center>

Unable to render embedded object: File (Description_of_states_in_amorphous_semiconductors.PNG) not found.

</center>

Van Hove

Van Hove singularities occurr where there is a non smooth density of state. It seems there won't be Van Hove singularities in an amorphous semiconductor.

Anderson

Milestones in a-Si development

  • There is poor electronic properties of pure a-Si. A dramatic increase in photoconductivity is a result of hydrogenation (10 atomic %). This was discovered in the late '60s.
  • a-Si is typically made by glow discharge plasma decomposition of silane (<math>SiH_4</math>) gas.
  • Doping of a-Si is achieved by adding <math>PH_3</math> (n-type) or <math>B_2H_6</math> (p-type).
  • Typical carrier mobilities are ~<math>1 \frac
    Unknown macro: {cm^2}
    Unknown macro: {V-s}
    </math> versus <math>1000</math> in crystalline silicon, while minority carrier diffusion lengths are about <math>100 nm</math>.
  • By the mid 80's, <math>10%</math> conversion efficiency was achieved in a-Si cells. Depending on technology <math>5-13%</math> is possible.
  • The typical solar cell design is a <math>p-i-n</math> junction. The individual layers are about <math>50 nm</math>. They serve the purpose of establishing the <math>V_
    Unknown macro: {bi}
    </math> but do not provide photocarriers.
  • Amorphous silicon is a more efficient absorber of light compared to crystalline silicon and therefore can be significantly thinner (<math>x100</math>).
  • The first large scale a-Si commercial installation was in Davis, CA.

Cause and Effect Relations Between Atomic Structure and Electronic Properties

  • The existence of short range order results in a similar overall electronic structure of an amorphous material compared to a crystal with a similar stoichiometry.
  • Silicon is semiconducting while <math>SiO_2</math> is an insulator both in the crystalline and amorphous forms
  • The abrupt band edges are replaced by broadened tails of states extending into the forbidden gap. These originate from the deviations in bond length and angle. The band tails are very important in determining the electronic properties as most of the electronic transport in semiconductors occurs near the band edges and is thus due to these states.
  • Electronic states deep in the gap occur because of coordination defects. These defects dominate the electronic properties by acting as recombination centers.
  • The presence of localized states due to the disorder- Anderson Localization

Notes

  • There is no k-conservation in amorphous materials due to the lack of long range order and lack of discrete translational symmetry.
  • short range order and correlation exists
  • The basic description of the electronic states is in terms of the DOS function.
  • Since k is not a "good quantum number" the distinction between direct bandgap and indirect bandgap semiconductors is lost. Consequently optical translations are allowed in materials such as silicon without the participation of a phonon
  • Disorder reduces carrier mobility due to scattering
  • Conductivity is due to extended as well as localized states

The nature of eigenfunctions of amorphous semiconductors

<p>
</p>

References

Calculate Eigenfunctions

Use the variational approach and a trial wavefunction that consists of a linear combination of atomic orbitals.

<center>

<br>

<math>\overline E = <\hat H></math>

<br>

<math>\overline E = \frac{\int_{-\infty}^

Unknown macro: {infty}

\chi^* \hat H \chi dV}{\int_{-\infty}^

\chi^* \chi dV} \ge E_

Unknown macro: {ground state}

</math>

<br>

<math>\chi = \sum_

Unknown macro: {n=1}

^N c_n \phi_n</math>

<br>

</center>

Search for coefficients that minimize <math>< \hat H ></math>.

<center>

<br>

<math>\frac

Unknown macro: {partial}
Unknown macro: {partial c_i^*}

< \hat H> = 0</math>

<br>

</center>

Conductive Polymers

The Nobel Prize in Chemistry was awarded in 2000 to Heeger MacDiarmid Shirakawa for the discovery of electrically conductive polymers.

Steps leading to the 2000 Nobel Prize in Chemistry

  • The original discovery focused on the role of conductivity in a simple conjugated polymer called polyacetylene
  • Made in 1974 by Shirakawa and co-workers using the Zieglar-Natta catalyst. The material was initially non-conducting but appeared metallic.
  • In 1977 Shirakawa, Heeger and MacDiarmid discovered that oxidation with chlorine, bromine or iodine vapor made polyacetylen 9-11 orders of magnitude more conductive
  • The "doped" polyacetylene had a conductivity of <math>10^5 S/m</math>. Teflon is <math>10^{-16} S/m</math>, and silver is <math>10^8S/m</math>.

Conductivity versus temperature for conjugated polymer and metal

Conductivity of polyacetylene increases with temperature, while conductivity of silver decreases.

The bonding in conjugated polymers

  • Application of a simple Huckel theory gives a pretty good approximation to the electronic eigenfunctions of the molecule.
    • Ground state properties are adequately described by Huckel like models
    • Excited states involve lattice interactions
  • There are two types of bonds in the system characterized by an <math>sp^2</math> hybridization
  • Angular momentum 0 orbitals called <math>\sigma</math> bonds in the plane of the molecule and <math>\pi</math> bonds comprising of superposition of atomic <math>p_z</math> orbitals out of the x-y planes.
  • Electronic transport properties determined by the <math>\pi</math> bonds.
    • Given the delocalized nature of the <math>\pi</math> molecular orbitals, shouldn't polyacetylene be a metal?
    • In fact pure (undoped) polyacetylene is a semiconductor with a bandgap around <math>1.7 eV</math>.
    • Electronic distribution is predicted to be across the entire chain implying equal bond lengths
  • Focus on the HOMO(<math>\pi</math>) and LUMO(<math>\pi^*</math> ) which correspond to the valence and conduction bands discussed previously in inorganic SC.

Trans and cis polyacetylene

  • Shirakawa discovered how to produce all trans polyacetylene using the Ziegler Natta organo-metallic <math>Ti(OBu)_4</math> catalyst
  • trans polyacetylen first isolated in early 70's demonstrated a conductivity <math>10^{-2}S/m</math> in all <math>cis</math> polyacetylene

Optical absorption in doped and undoped polyacetylene

  • Like a classical semiconductor the sample is transparent to light with a photon energy smaller than the band gap (<math>1.7 eV</math>).
  • There is an emergence of a midgap state in the doped material around <math>0.7 eV</math>

Pierls instability in one dimesional chains

  • Pierls showed that a hypothetical chain of sodium atoms would undergo a metal to insulator transition because the equi-spaced lattice is unstable with respect to an alternating bond configuration (1930s).
  • This small conformation change leads to the emergence of a small bandgap

The effect of doping conjugated polymers (first published in 1977)

  • Conductivity increases with increased doping
  • The role of the dopant (halogen <math>I_2/Br_2</math>) is to remove an electron from the polymer
    • The resulting cation is not completely delocalized.
  • Radical cation ("polaron" formed by the removal of one electron. The polaron migrates
  • The radical cation is localized partly because of the interaction with a counter ion.
  • A high concentration of dopants is needed in order to facilitate polaron mobility because the anion is typically of very low mobility.

===SSH model of soliton formation in polyacetylene

  • A soliton separates A and B phases

===Properties of PPV(poly (p-phenylene))

  • Electroluminescence from conjugated polymers was first reported in 1990 by R. Friend and coworkers
  • In PPV the polymer backbone is held together by <math>\sigma</math> bonds.
    • The remaining electron in each carbon is in a <math>p_z</math> or <math>p_y</math> atomic orbital which overlaps to produce a <math>\pi</math> MO.
  • The HOMO LUMO separation in PPV is aboout 2.5 eV which corresponds to a green-yellow emission.
  • The magnitude of the gap depends on the conjugation length
    • Longer lengths correspond to smaller gaps with an asymptote reached at about ten repeat units.
  • In crystalline semiconductors the electron and hole pairs are delocalized in three-dimensions which leads to a small overlap. In organic semiconductors the electron and holes are confined to the same chain.
    • This leads to bound neutral excited states called excitons.
    • Excitons play an important role in determining the PL properties of PPV and other conjugated polymers.

Simple PLED structure (polymer light emitting diode)

  • semiconducting polymer is sandwiched between a hole injecting material (ITO) and a low work function metal such as <math>Ca</math>, <math>Al</math>, or <math>Mg</math>
  • Upon external bias, electrons are injected from the <math>Ca</math> electrodes. Holes come from the ITO and recombine in the semiconductor layer and emit light corresponding to the HOMO-LUMO separation of the polymer
  • Devices that are <math>80%</math> efficient have been produced.

Applications

  • Doped polyaniline is used as a conductor and for electromagnetic shielding of electronic circuits. Polyaniline is also manufactured as a corrosion inhibitor.
  • Poly(ethylenedioxythiophene) (PEDOT) doped with polystyrenesulfonic acid is manufactured as an antistatic coating material to prevent electrical discharge exposure on photographic emulsions and also serves as a hole injecting electrode material in polymer light-emitting devices.
  • Poly(phenylene vinyliden) derivatives have been major candidates for the active layer in pilot production of electroluminescent displays (mobile telephone displays).
  • Poly(dialkylfluorene) derivatives are used as the emissive layer in full-color video matrix displays
  • Poly(thiophene) derivatives are promising in use in field-effect transistors: they may possibly find use in a supermarket checkouts
  • Poly(pyrrole) has been tested as microwave-absorbing "stealth" (radar-invisible) screen coatings and also as the active thin layer of various sensing devices.

Magnetic Materials

Energy Terms

Exchange Energy

Minimize exchange energy when all spins are parallel

<center>

<br>

<math>E = -2J \sum_

Unknown macro: {i=j}

S_i S_j</math>

<br>

</center>

Domain walls are the interfaces between domains. There is a significant angle among the magnetic moments of the atoms at the border. The higher the exchange constant <math>A_

Unknown macro: {ex}

</math>, the higher the exchange energy.

<p>
</p>

Reduce the exchange energy by spreading the angular transition over a very large number of atoms. The angle between two adjacent moments is very small, but there is a cost in anisotropy energy. Almost all the moments that are in the transition deviate from the easy magnetization direction, and there is a cost in magnetocrystalline energy. This grows with the anisotropy constant, <math>K</math>.

<p>
</p>

The width of the domain wall is defined by a compromise. The compromise depends on the relative values of the constants <math>A_

</math> and <math>K</math>. The higher the <math>A_

Unknown macro: {ex}

/ K</math> ratio, the wider the domain walls.

Magnetostatic Energy

<center>

<br>

<math>E = -M \cdot H</math>

<br>

</center>

Demagnetization factor

Anisotropy Energy

Easy and hard direction. The angle, <math>\theta</math>, is the angle from the easy angle. Extra energy to magnetize along a particular angle. Squared due to symmetry. Energy required to rotate spins from favored direction.

<center>

<br>

<math>K = K_1 \sin^2 \theta + K_2 \sin^4 \theta + K_3 \sin^6 \theta</math>

<br>

</center>

Magnetostrictive Energy

The magnetization of a ferromagnetic material is in nearly all cases accompanied by changes in dimensions. The resulting strain is called the magnetostriction <math>\lambda</math>. From a phenomenological viewpoint there are really two main types of magnetostriction to consider: spontaneous magnetostriction arising from the ordering of magnetic moments into domains at the Curie temperature, and field-induced magnetostriction.

<p>
</p>

Spontaneous magnetostriction within domains arises from the creation of domains as the temperature of the ferromagnet passes through the Currie (or ordering) temperature. Field-induced magnetostriction arises when domains that show spontaneous magnetostriction are reoriented under action of a magnetic field.

<p>
</p>

The field-induced bulk-magnetostriction is the variation of <math>\lambda</math> with <math>\vec H</math> or <math>\vec B</math> and is often the most interesting feature of the magnetostrictive properties to the materials scientist. However, the variations <math>\lambda (\vec H)</math> or <math>\lambda (\vec B)</math> are very structure sensitive so that it is not possible to give any general formula for the relation of magnetostrictino to field.

Demagnetizing fields

In view of the fact that the magnetization <math>\vec M</math> and the magnetic field <math>\vec H</math> point in opposite directions inside a magnetized material of finite dimensions, due to the presence of the magnetic dipole moment, it is possible to define a demagnetizing field <math>\vec H_d</math> which is present whenever magnetic poles are created in a material

<p>
</p>

The demagnetizing field depends on two factors only. These are the magnetization in the material (i.e. the pole strength) and the shape of the specimen (i.e the pole separation which is determined by sample geometry). The demagnetizing field is proportional to the magnetization.

<p>
</p>

When dealing with samples of finite dimensions in an applied magnetic field <math>\vec H_

Unknown macro: {app}

</math> it is necessary to make some demagnetizing field correction to determine the exact internal field in the solid, <math>\vec H_

Unknown macro: {in}

</math>.

<center>

<br>

<math>\vec H_

= \vec H_

- N_d \vec M</math>

<br>

</center>

Examples involving energy terms

  • Spins in sphere
  • Small, long rods to store energy

References:

  • David Jiles Introduction to Magnetism and Magnetic Materials

Spintronics

In order to make a spintronic device, the primary requirement is to have a system that can generate a current of spin polarised electrons, and a system that is sensitive to the spin polarization of the electrons. Most devices also have a unit in between that changes the current of electrons depending on the spin states.

<p>
</p>

The simplest method of generating a spin polarised current is to inject the current through a ferromagnetic material. The most common application of this effect is a giant magnetoresistance (GMR) device. A typical GMR device consists of at least two layers of ferromagnetic materials separated by a spacer layer. When the two magnetization vectors of the ferromagnetic layers are aligned, then an electrical current will flow freely, whereas if the magnetization vectors are antiparrallel then the resistance of the system is higher. Two variants of GMR have been applied in devices, current-in-plane where the electric current flows parallel to the layers and current-perpendicular-to-the-plane where the electric current flows in a direction perpendicular to the layers.

<p>
</p>

The most successful spintronic device to date is the spin valve. This device utilizes a layered structure of thin films of magnetic materials, which changes electrical resistance depending on applied magnetic field direction. In a spin valve, one of the ferromagnetic layers is "pinned" so its magnetization direction remains fixed and the other ferromagnetic layer is "free" to rotate with the application of a magnetic field. When the magnetic field aligns the free layer and the pinned layer magnetization vectors, the electrical resistance of the device is at its minimum. When the magnetic field causes the free layer magnetization vector to rotate in a direction antiparallel to the pinned layer magnetization vector, the electrical resistance of the device increases due to spin dependent scattering. The magnitude of the change, (Antiparallel Resistance - Parallel Resistance) / Parallel Resistance x 100% is called the GMR ratio. Devices have been demostrated with GMR ratios as high as 200% with typical values greater than 10%. This is a vast improvement over the anisotropic magnetoresistance effect in single layer materials which is usually less than 3%. Spin valves can be designed with magnetically soft free layers which have a sensitive response to very weak fields (such as those originating from tiny magnetic bits on a computer disk), and have replaced anisotropic magnetoresistance sensors in computer hard disk drive heads since the late 1990s.

<p>
</p>

Future applications may include a spin-based transistor which requires the development of magnetic semiconductors exhibiting room temperature ferromagnetism. The operation of MRAM or magnetic random access memory is also based on spintronic principles.

<p>
</p>

Reference

GMR

The development of devices that sense and store information is driven by the demand for more capable computers. The utilization of the phenomenon of giant magneto-resistance (GMR) has led to the creation of more sensitive sensors that read bits coded on magnetized regions of disks. An understanding of GMR has led to a large increase in information storage capacity. The phenomenon of GMR has also been of interest to physicists.

<p>
</p>

The magneto-resistance effect (MR effect) describes the change in the electrical resistance of a material due to the application of a magnetic field. Research advances related to electron spin, tunneling effects, and production of ultra-thin layers have made possible the design of electrical devices based on quantum mechanical effects of the electron. The giant magneto-resistance effect is dependent on spin properties of electrons. Several other magneto-resistive effects have been discovered, and they are classified under XMR effects or XMR technology.

<p>
</p>

An example of a structure where the giant magneto-resistive effect is observed consists of a layer of copper, a normal metal, between layers of cobalt, a ferromagnetic material. When the magnetic moments of the ferromagnetic layers are parallel, there is less resistance to the flow of current. In current devices, the direction of magnetization of one layer is typically fixed in one direction, while the other is determined by an external field. The difference of resistance between parallel and anti-parallel configurations is 10-15% in current devices.

<p>
</p>

The effect of GMR is similar to phenomenon of polarization. In a polarization experiment, light cannot pass through if two polarizers are perpendicular to one another. Similarly, one magnetic layer may allow electrons of one type of spin to pass, and, if the second structure is aligned in the same direction, current can easily pass through. If the second layer is misaligned, neither spin channel can pass through the structure easily, and the electrical resistance is high.

<p>
</p>

Early studies in which GMR was first observed may have been published in 1987. The patent of the effect is owned by Peter Grunberg, who led a team at the Julich Research Centre and observed the effect in trilayers of Fe/Cr/Fe. The effect was simultaneously and independently observed in multilayers of Fe/Cr.

<p>
</p>

Three types of GMR are multilayer, granular, and spin-valve GMR. The effect was first observed in multilayer configurations and the effect in granular GMR is not as large as multilayer GMR. Spin-valve GMR is the most useful industrially, and the magnetic recording industry is researching structures that are very sensitive to magnetic fields. IBM and Seagate develop high density disk and tape playback heads. Additional applications include avionic compasses, swipe-card readers, wheel rotation sensors in ABS brakes, and current sensors for use in safety powerbreakers and electricity meters.

<p>
</p>

References:

TMR

In physics, the tunnel magnetoresistance effect, commonly abbreviated as TMR, occurs when two ferromagnets are separated by a thin (about 1 nm) insulator. Then the resistance of the tunneling current changes with the relative orientation of the two magnetic layers. The resistance is normally higher in the anti-parallel case.

<p>
</p>

It was discovered in 1975 by M. Julliere, using iron as the ferromagnet and germanium as the insulator.

<p>
</p>

Room temperature TMR was discovered in 1995 by Moodera et. al. following renewed interest in this field fueled by the discovery of the giant magnetoresistive effect. It is now the base for the magnetic random access memory (MRAM) and read sensors in hard disk drives. For more technical information see [Moodera and Mathon 1999].

<p>
</p>

References:

Organic Materials

References

Optical Properties

Maxwell's Equations

In electromagnetics, Maxwell's equations are a set of four equations, compiled by James Clerk Maxwell, that describe the behavior of both the electric and magnetic fields, as well as their interactions with matter.

<p>
</p>

Maxwell's four equations express, respectively, how electric charges produce electric fields (Gauss' law), the experimental absence of magnetic monopoles, how currents and changing electric fields produce magnetic fields (the Ampere-Maxwell law), and how changing magnetic fields produce electric fields (Faraday's law of induction). Refer to article in wikipedia to see in depth discussion.

! Name! Differential form! Integral form

Gauss's law: How electric charges produce electric fields

<math>\vec

Unknown macro: {nabla}

\cdot \vec

Unknown macro: {D}

= \rho</math> |

<math>\oint_S \vec

\cdot \mathrm

Unknown macro: {d}

\vec

Unknown macro: {A}

= \int_V \rho\, \mathrm

V</math> |

Gauss' law for magnetism <br /> (absence of magnetic monopoles):

<math>\vec

\cdot \vec

Unknown macro: {B}

= 0</math> |

<math>\oint_S \vec

\cdot \mathrm

Unknown macro: {d}

\vec

Unknown macro: {A}

= 0</math> |

Faraday's law of induction: How currents and changing electric fields produce magnetic fields

<math>\vec

Unknown macro: {nabla}

\times \vec

Unknown macro: {E}

= -\frac{\partial \vec{B}}

Unknown macro: {partial t}

</math> |

<math>\oint_C \vec

\cdot \mathrm

\vec

Unknown macro: {l}
  • \oint_C \vec
    Unknown macro: {B}
    \times \vec
    Unknown macro: {v}
    \cdot \mathrm
    Unknown macro: {d}
    \vec

= - \

{ \mathrm
Unknown macro: {d}

\over \mathrm

t }

\int_S \vec

Unknown macro: {B}

\cdot \mathrm

Unknown macro: {d}

\vec

Unknown macro: {A}

</math>

Amp��re's law<br /> (with Maxwell's extension): How changing magnetic fields produce electric fields

<math>\vec

Unknown macro: {nabla}

\times \vec

= \vec

Unknown macro: {J}

+ \frac{\partial \vec{D}}

Unknown macro: {partial t}

</math> |

<math>\oint_C \vec

Unknown macro: {H}

\cdot \mathrm

Unknown macro: {d}

\vec

Unknown macro: {l}

= \int_S \vec

\cdot \mathrm

Unknown macro: {d}

\vec

Unknown macro: {A}

+ \int_S \frac{\partial\vec{D}}

Unknown macro: {partial t}

\cdot \mathrm

\vec

Unknown macro: {A}

</math> |

Index of Refraction

The refractive index (or index of refraction) of a material is the factor by which the phase velocity of electromagnetic radiation is slowed in that material, relative to its velocity in a vacuum. It is usually given the symbol n, and defined for a material by:

<center>

<br>

<math> n=\sqrt

Unknown macro: {epsilon_rmu_r}

</math>

<br>

</center>

where ��r is the material's relative permittivity, and ��r is its relative permeability. For a non-magnetic material, <math>\mu_r</math> is very close to 1, therefore n is approximately <math>\sqrt

Unknown macro: {epsilon_r}

</math>.

<p>
</p>

The phase velocity is defined as the rate at which the crests of the waveform propagate; that is, the rate at which the phase of the waveform is moving. The group velocity is the rate that the envelope of the waveform is propagating; that is, the rate of variation of the amplitude of the waveform. It is the group velocity that (almost always) represents the rate that information (and energy) may be transmitted by the wave, for example the velocity at which a pulse of light travels down an optical fiber.

Stoke's Theorem

Divergence Theorem

Wave Equation

The electromagnetic wave equation is a second-order partial differential equation that describes the propagation of electromagnetic waves through a medium or in a vacuum. The homogeneous form of the equation, written in terms of either the electric field <math>E</math> or the magnetic field <math>H</math>, takes the form:

<center>

<br>

<math>(\nabla^2-

Unknown macro: {1over c^2}
Unknown macro: {partial^2 over partial t^2}

)\mathbf

Unknown macro: {E}

= 0</math>

<br>

<math>(\nabla^2-

Unknown macro: {1 over c^2}

) \mathbf

Unknown macro: {H}

= 0</math>

<br>

</center>

where c is the speed of light in the medium. In a vacuum, <math>c = 2.998 x 10^8 \frac

Unknown macro: {m}
Unknown macro: {s}

</math>, which is the speed of light in free space. The electromagnetic wave equation is derived from Maxwell's equations.

Derivation

To obtain electromagnetic waves in a vacuum, note that Maxwell's equations (SI units) in a vacuum are

<center>

<br>

<math> \nabla \cdot \mathbf

Unknown macro: {E}

= 0</math>

<br>

<math> \nabla \times \mathbf

= -\mu_o \frac{\partial \mathbf{H}}

Unknown macro: {partial t}

</math>

<br>

<math> \nabla \cdot \mathbf

= 0</math>

<br>

<math> \nabla \times \mathbf

Unknown macro: {H}

=\varepsilon_o \frac{ \partial \mathbf{E}}

Unknown macro: {partial t}

</math>

<br>

</center>

Take the curl of the curl equations we obtain

<center>

<br>

<math> \nabla \times \nabla \times \mathbf

Unknown macro: {E}

= -\mu_o \frac

Unknown macro: {partial }

\nabla \times \mathbf

= -\mu_o \varepsilon_o \frac{\partial^2 \mathbf

Unknown macro: {E}

}

Unknown macro: {partial t^2}

</math>

<br>

<math> \nabla \times \nabla \times \mathbf

Unknown macro: {H}

= \varepsilon_o \frac

Unknown macro: {partial }

Unknown macro: {partial t}

\nabla \times \mathbf

= -\mu_o \varepsilon_o \frac{\partial^2 \mathbf

Unknown macro: {H}

}

Unknown macro: {partial t^2}

</math>

<br>

</center>

Note the vector identity

<center>

<br>

<math>\nabla \times \left( \nabla \times \mathbf

Unknown macro: {V}

\right) = \nabla \left( \nabla \cdot \mathbf

\right) - \nabla^2 \mathbf

Unknown macro: {V}

</math>

<br>

</center>

where <math> \mathbf

</math> is any vector function of space. Recover the wave equations

<center>

<br>

<math> {\partial^2 \mathbf

Unknown macro: {E}

\over \partial t^2} \ - \ c^2 \cdot \nabla^2 \mathbf

\ \ = \ \ 0</math>

<br>

<math> {\partial^2 \mathbf

\over \partial t^2} \ - \ c^2 \cdot \nabla^2 \mathbf

Unknown macro: {H}

\ \ = \ \ 0</math>

<br>

</center>

where <math>c</math> is the speed of light in free space.

<center>

<br>

<math>c = { 1 \over \sqrt

Unknown macro: { mu_o varepsilon_o }

} = 2.998 \times 10^8 \frac

Unknown macro: {m}
Unknown macro: {s}

</math>

<br>

</center>

Solutions

See article in wikipedia.

EM Fields at Interfaces

Boundary Conditions

The tangential components of <math>\vec B / \mu</math> can be discontinuous at an interface if there is a layer of surface current confined there. In the equation below, <math>\mu_1</math> and <math>\mu_2</math> are the magnetic permeabilities of the two media and <math>\vec j_s</math> is the surface current density. If both materials behave as perfect insulators, no surface currents can exist, and the tangential components of \vec B / \mu are continuous across the interface.

<center>

<br>

<math>\frac

Unknown macro: {1}
Unknown macro: {mu_1}

\vec B_1 \times \vec n - \frac

Unknown macro: {mu_2}

\vec B_2 \times \vec n = \vec j_s</math>

<br>

</center>

The tangential components of the electric field are continuous at an interface.

<center>

<br>

<math>\vec E_1 \times \vec n = \vec E_2 \times \vec n</math>

<br>

</center>

Transfer Matrix

If the electric and magnetic fields in the <math>x-y</math> plane at <math>z=0</math> are known, use Maxwell's equations at a fixed frequency to integrate the wavefield and find electric and magnetic fields in the <math>x-y</math> plane at <math>z=c</math>. Assume that there is zero divergence of <math>\vec B</math> and <math>D</math>. It is only necessary to know two components of each field.

<center>

<br>

<math>\vec F(z = 0) = \left [E_x (z=0), E_y (z=0), H_x (z=0), E_y (z=0) \right]</math>

<br>

</center>

An equation below defines the transfer matrix, <math>\vec T</math>.

<center>

<br>

<math>\vec F (z = c) = \vec T (c, 0) \vec F(z=0)</math>

<br>

</center>

The transfer matrix defines how waves cross a slab of material. It is closely related to the transmission and reflection coefficients of the slab. If the slab is infinitely repeated thourout the material in a periodic array, it is possible to impose the Bloch condition and calculate the band structure, <math>K(\omega)</math>, from the eigenvalues of <math>T</math>.

<center>

<br>

<math>e^

Unknown macro: {iKc}

\vec

Unknown macro: {F}

(z=0) = \vec F (z=c)</math>

<br>

<math>e^

\vec

Unknown macro: {F}

(z=0) = \vec T (c,0) \vec F (z=0)</math>

<br>

</center>

Index of Refraction, Band Gap, and Wavelength

Index of refraction measure of how much light interacts with material. Closer to one, more transparent.

<p>
</p>

Check graph in notes. <math>Ge</math> above <math>Si</math> above <math>SiO</math>. Index of refraction of oxide less than index of non-oxide. The insulator lies below semiconductor materials. <math>Ge</math> smaller band gap. Band gap goes down with increasing index of refraction.

Definition and Examples of Isotropic, Uniaxial and Biaxial Optical Materials

Define coordinate axes (principal axes) such that the dielectric tensor is diagonal.

<center>

<br>

<math>\epsilon = \begin

Unknown macro: {bmatrix}

\epsilon_1& 0& 0
0& \epsilon_2& 0
0& 0& \epsilon_3 \end

</math>

<br>

</center>

Isotropic

  • Cubic or amorphous materials
  • <math>\epsilon_1 = \epsilon_2 = \epsilon_3</math>
  • <math>Y_3Al_5O_12</math> (Yttrium Aluminum Garnet)

Uniaxial

  • Trigonal, tetragonal, hexagonal
  • <math>\epsilon_1 \ne \epsilon_2 = \epsilon_3</math>
  • <math>LiNbO_3</math>

Biaxial

  • Triclinic, monoclinic, orthorhombic
  • <math>\epsilon_1 \ne \epsilon_2 \ne \epsilon_3</math>

Maxwell's Equations and the Constitutive Equation

! Name! Differential form! Integral form

Gauss's law: How electric charges produce electric fields

<math>\vec

Unknown macro: {nabla}

\cdot \vec

Unknown macro: {D}

= \rho</math> |

<math>\oint_S \vec

\cdot \mathrm

Unknown macro: {d}

\vec

= \int_V \rho\, \mathrm

Unknown macro: {d}

V</math> |

Gauss' law for magnetism <br /> (absence of magnetic monopoles):

<math>\vec

Unknown macro: {nabla}

\cdot \vec

Unknown macro: {B}

= 0</math> |

<math>\oint_S \vec

\cdot \mathrm

\vec

Unknown macro: {A}

= 0</math> |

Faraday's law of induction: How changing magnetic fields produce electric fields

<math>\vec

Unknown macro: {nabla}

\times \vec

Unknown macro: {E}

= -\frac{\partial \vec{B}}

Unknown macro: {partial t}

</math> |

<math>\oint_C \vec

\cdot \mathrm

Unknown macro: {d}

\vec

Unknown macro: {l}

- \oint_C \vec

Unknown macro: {B}

\times \vec

Unknown macro: {v}

\cdot \mathrm

\vec

Unknown macro: {l}

= - \

{ \mathrm
Unknown macro: {d}

\over \mathrm

t }

\int_S \vec

Unknown macro: {B}

\cdot \mathrm

Unknown macro: {d}

\vec

</math> |

Amp��re's law<br /> (with Maxwell's extension): How currents and changing electric fields produce magnetic fields

<math>\vec

Unknown macro: {nabla}

\times \vec

Unknown macro: {H}

= \vec

Unknown macro: {J}

+ \frac{\partial \vec{D}}

Unknown macro: {partial t}

</math> |

<math>\oint_C \vec

\cdot \mathrm

Unknown macro: {d}

\vec

Unknown macro: {l}

= \int_S \vec

Unknown macro: {J}

\cdot \mathrm

\vec

Unknown macro: {A}

+ \int_S \frac{\partial\vec{D}}

Unknown macro: {partial t}

\cdot \mathrm

Unknown macro: {d}

\vec

</math>

<center>

<br>

<math>\vec D = \epsilon \vec E = \vec E + 4 \pi \vec P</math>

<br>

<math>\vec D = \mu \vec H = \vec H + 4 \pi \vec M</math>

<br>

<math>\vec J = \sigma \vec E</math>

<br>

</center>

The Wave Equation and the Conditions for its Applicability

<center>

<br>

<math> {\partial^2 \mathbf

Unknown macro: {E}

\over \partial t^2} \ - \ c^2 \cdot \nabla^2 \mathbf

=0</math>

<br>

<math> {\partial^2 \mathbf

Unknown macro: {H}

\over \partial t^2} \ - \ c^2 \cdot \nabla^2 \mathbf

=0</math>

<br>

</center>

Solutions to the wave equations can take on different forms. Examples include plane waves, cylindrical waves, or spherical waves. These vector equations are satisfied by the following solutions.

<center>

<br>

<math>\vec E(x,y,z) = \vec E_o e^

Unknown macro: {i omega t - vec k r}

</math>

<br>

<math>\frac

Unknown macro: {c}

{\sqrt{\mu \epsilon}} |\vec k| = \omega</math>

<br>

</center>

The Properties of Plane Wave Solutions, Wave Vector, Transversality

Phase velocity

The solution is a complex valued function. An expression of the phase is below. Consider the fronts where the phase is constant.

<center>

<br>

<math>\vec E (x,y,z) = \vec E_o e^

</math>

<br>

<math>\phi = \omega t - \vec k \cdot \vec r</math>

<br>

<math>\omega t - \vec k \cdot \vec r = const</math>

<br>

</center>

Below is an expression of velocity of the constant fronts.

<center>

<br>

<math>\vec v_

Unknown macro: {phase}

= \frac

Unknown macro: {omega}
Unknown macro: {|k|}

\hat k</math>

<br>

<math>\vec v_

= \frac

Unknown macro: {c}

{\sqrt{\epsilon \mu}}</math>

<br>

</center>

The direction is defined by the wavevector's direction. The phase velocity is changed in a material compared with its value in vacuum.

Transversality of E-M Fields

There are three Cartesian terms of the solutions of the wave equations.

<center>

<br>

<math>\vec E(x,y,z) = E_x \hat x + E_y \hat y + E_z \hat z</math>

<br>

</center>

Consider a medium that is homogeneous and charge free. Two equations below imply the transversality of E-M fields.

<center>

<br>

<math>\vec \nabla \cdot \vec E = 0</math>

<br>

<math>\vec \nabla \cdot \vec H = 0</math>

<br>

<math>\vec E \cdot \vec k = 0</math>

<br>

<math>\vec H \cdot \vec k = 0</math>

<br>

</center>

Dispersion Relations

<center>

<br>

<math>\epsilon ( \omega ) \omega^2 = c^2 k^2</math>

<br>

</center>

The term <math>k</math> is the wavevector or propagation vector and it can be related to the wavelength, as shown below.

<center>

<br>

<math>|k| = \frac

Unknown macro: {2 pi}
Unknown macro: {lambda}

</math>

<br>

<math>|k| = \frac

Unknown macro: {n}

\lambda</math>

<br>

</center>

How are Fields on one Side of an Interface Related to the Fields on the Other Side

To calculate the behavior of the normal components of the field with respect to the interface unit vector Gauss' theorem

<center>

<br>

<math>\int_

Unknown macro: {volume}

\vec \nabla \cdot \vec B dv = \int_

Unknown macro: {surface}

\vec B \cdot \hat n dS = 0</math>

<br>

<math>\int_

\vec \nabla \cdot \vec D dv = \int_

Unknown macro: {surface}

\vec D \cdot \hat n dS = 4 \pi \int_

Unknown macro: {volume}

\rho dv</math>

<br>

</center>

  • <math>\hat n \cdot ( \vec B_2 - \vec B_1 ) = 0</math>
    • The normal component of the magnetic induction is continuous across the surface of discontinuity
  • <math>\hat n \cdot ( \vec D_2 - \vec D_1 ) = 4 \pi \rho</math>
    • In the presence of a layer of surface charge, the normal component of the electric displacement changes abruptly by an amount equal to <math>4 \pi \sigma</math>. The tangential components are found through the application of Stokes Theorem.

<center>

<br>

<math>\int_

\vec \nabla \times \vec E \cdot \hat n dS = \int_

Unknown macro: {line}

\vec E \cdot d \vec r</math>

<br>

</center>

  • <math>\hat n \times (\vec E_2 - \vec E_1) = 0</math>
    • The tangential component of the electric field is continuous across the interface.
  • <math>\vec n \times (\vec H_2 - \vec H_1 = \frac
    Unknown macro: {4 pi}
    Unknown macro: {c}

    \vec K</math>

      • In the presence of a surface current the tangential component of the magnetic induction changes abruptly.

    What is the Energy Law of EM Fields, Significance of Poynting Vector

    Energy Law of Electromagnetic Fields

    Electromagnetic theory interprets the light intensity as the energy flux of the field. The presence of a field implies the existence of energy associated with that field.

    <center>

    <br>

    <math>\vec E \cdot \vec \nabla \times \vec H - \vec H \cdot \vec \nabla \times \vec E = \frac

    Unknown macro: {c}

    \vec J \cdot \vec E + \frac

    Unknown macro: {1}
    \vec E \cdot \frac
    Unknown macro: {partial vec D}
    Unknown macro: {partial t}

    + \frac

    Unknown macro: {1}
    Unknown macro: {c}

    \vec H \cdot \frac

    Unknown macro: {partial vec B}

    </math>

    <br>

    <math>\int \vec J \cdot \vec E dv + \frac

    Unknown macro: {partial}
    Unknown macro: {partial t}

    \int ( \vec E \cdot \vec D + \vec H \cdot \vec B ) dv + \int ( \vec E \times \vec H ) \cdot \hat n dS = 0</math>

    <br>

    <math>\int \vec J \cdot \vec E dv + \frac

    Unknown macro: {partial t}

    \int ( \vec E \cdot \vec D + \vec H \cdot \vec B ) dv + \int ( \vec E \times \vec H ) \cdot \hat n dS = - \vec \nabla \cdot ( \vec E \times \vec H)</math>

    <br>

    <math>\frac

    Unknown macro: {4 pi}
    Unknown macro: {c}

    \vec J \cdot \vec E + \frac

    Unknown macro: {1}

    \vec E \cdot \frac

    Unknown macro: {partial t}

    + \frac

    Unknown macro: {1}
    Unknown macro: {c}

    \vec H \cdot \frac

    Unknown macro: {partial vec B}
    + \vec \nabla \cdot ( \vec E \times \vec H) = 0</math>

<br>

</center>

Apply Gauss's theorem.

<center>

<br>

<math>\int \frac

Unknown macro: {4 pi}
Unknown macro: {c}

\vec J \cdot \vec E dv + \int \left ( \frac

Unknown macro: {1}

\vec E \cdot \frac

Unknown macro: {partial vec D}
Unknown macro: {partial t}

\right ) dv + \frac

Unknown macro: {1}
Unknown macro: {c}

\vec H \cdot \frac

Unknown macro: {partial vec B}

+ \int \left ( \vec E \times \vec H \right ) \cdot \hat n dS = 0</math>

<br>

</center>

Use materials equations

<center>

<br>

<math>\frac

Unknown macro: {1}

\vec E \cdot \frac

Unknown macro: {partial vec D}
Unknown macro: {partial t}

= \frac

Unknown macro: {1}
Unknown macro: {4 pi}

\vec E \cdot \frac

Unknown macro: {partial epsilon vec E}

</math>

<br>

<math>\frac

Unknown macro: {1}
Unknown macro: {4 pi}

\vec E \cdot \frac

Unknown macro: {partial t}

= \frac

Unknown macro: {1}
Unknown macro: {8 pi}

\frac

Unknown macro: {partial epsilon vec E^2}

</math>

<br>

<math>\frac

Unknown macro: {1}
Unknown macro: {4 pi}

\vec E \cdot \frac

Unknown macro: {partial vec D}
Unknown macro: {partial t}

= \frac

Unknown macro: {8 pi}

\frac

Unknown macro: {partial (vec E cdot vec D)}
Unknown macro: {partial t}

</math>

<br>

<math>\frac

Unknown macro: {1}
Unknown macro: {4 pi}

\cdot \frac

Unknown macro: {partial vec B}

= \frac

Unknown macro: {1}

\frac

Unknown macro: {partial (vec H cdot vec B)}
Unknown macro: {partial t}

</math>

<br>

</center>

Poynting Vector

The Poynting vector describes the energy flux <math>\left ( \frac

Unknown macro: {J}
Unknown macro: {m^2s}

\right )</math> of an electromagnetic field. It is named after its inventor John Henry Poynting. Oliver Heaviside independently co-discovered the Poynting vector.

<p>
</p>

It points in the direction of energy flow and its magnitude is the power per unit area crossing a surface which is normal to it. (The fact that it points perhaps contributes to the frequency with which its name is misspelled.) It is derived by considering the conservation of energy and taking into account that the magnetic field can do no work. It is given the symbol <math>S</math> (in bold because it is a vector) and, in SI units, is given by:

<center>

<br>

<math>\mathbf

Unknown macro: {S}

= \vec

Unknown macro: {E}

\times \vec

Unknown macro: {H}

= \frac

Unknown macro: {1}
Unknown macro: {mu}

\vec

\times \vec

Unknown macro: {B}

</math>

<br>

</center>

where <math>E</math> is the electric field, <math>H</math> and <math>B</math> are the magnetic field and magnetic flux density respectively, and <math>\mu</math> is the permeability of the surrounding medium.

<p>
</p>

For example, the Poynting vector near an ideally conducting wire is parallel to the wire axis - so electric energy is flowing in space outside of the wire. The Poynting vector becomes tilted toward wire for a resistive wire, indicating that energy flows from the e/m field into the wire, producing resistive Joule heating in the wire.

Snell's Law and its Derivation

See below

Conserved E-M Quantities at Interfaces (Parallel Component of Wave Vector, Frequency) and the Consequence of their Conservation

Consider a wave incident from the left onto an interface.

<center>

<br>

<math>\mbox

Unknown macro: {Incident Wave}

</math>

<br>

<math>\vec E_i e^{\omega t - i \vec k_i \cdot \vec r</math>

<br>

<math>\mbox

Unknown macro: {Reflected Wave}

</math>

<br>

<math>\vec E_r e^{\omega t - i \vec k_r \cdot \vec r</math>

<br>

<math>\mbox

Unknown macro: {Transmitted Wave}

</math>

<br>

<math>\vec E_t e^{\omega t - i \vec k_t \cdot \vec r</math>

<br>

</center>

Consider a homogeneous medium and dispersion relations.

<center>

<br>

<math>|\vec k_i| = |\vec k_r| = \frac

Unknown macro: {omega n_1}
Unknown macro: {c}

</math>

<br>

<math>\vec k_t = \frac

Unknown macro: {omega n_2}

</math>

<br>

</center>

Focus on the interface plane, <math>x=0</math>. Boundary conditions dictate relations between the field components on both sides of the interface. The phase of the fields must be equal.

<center>

<br>

<math>( \vec k_1 \cdot \vec r )_

Unknown macro: {x=0}

= ( \vec k'1 \cdot \vec r )

= ( \vec k_2 \cdot \vec r )_

Unknown macro: {x=0}

</math>

<br>

<math>( \vec k_

Unknown macro: {1y}

y + \vec k_

Unknown macro: {1z}

z) = ( \vec k'_

y + \vec k'_

Unknown macro: {1z}

z) = ( \vec k_

Unknown macro: {2y}

y + \vec k_

Unknown macro: {2z}

z)</math>

<br>

<math>k_

Unknown macro: {1y}

= k'_

= k_

</math>

<br>

<math>k_

= k'_

Unknown macro: {1z}

= k_

Unknown macro: {2z}

</math>

<br>

</center>

Consider an arbitrary vector that lies in the <math>z-y</math> plane. in this case, the following is true.

<center>

<br>

<math>\vec r = (x=0,y,z) = \vec r_

Unknown macro: {z-y}

</math>

<br>

<math>(\vec k_

Unknown macro: {1t}

\cdot \vec r_t ) = (\vec k'_

\cdot \vec r_t ) = (\vec k_

Unknown macro: {2t}

\cdot \vec r_t )</math>

<br>

</center>

The vectors <math>\vec k_1</math>, <math>\vec k'_1</math>, <math>\vec k_2</math> all lie in a place called the plane of incidence. Orient the coordinate system so that the plane of incidence coincides with the <math>x-z</math> plane.

<center>

<br>

<math>\vec E = \vec E e^

Unknown macro: {i(omega t - k_x x - k_z z)}

</math>

<br>

</center>

The tangential components of the wavevector are all identical regardless of the medium that they are in and regardless of whether a medium is lossless or absorbing.

<center>

<br>

<math>k_

= k_

Unknown macro: {1'z}

= k_

Unknown macro: {2z}

= \beta</math>

<br>

<math>|\vec k_1| = |\vec k'_1| = n_1 \vec

Unknown macro: {omega}
Unknown macro: {c}

</math>

<br>

<math>|\vec k_2| = n_2 \vec

Unknown macro: {c}

</math>

<br>

<math>|\vec k_i| \sin \theta_1 = |\vec k_t| \sin \theta_2</math>

<br>

<math>\frac

Unknown macro: {c}

\sin \theta_1 = \frac

Unknown macro: {omega n_2}

\sin \theta_2</math>

<br>

</center>

This leads to two important consequences.

  • Angle of reflection is equal to the angle of incidence

<center>

<br>

<math>\theta_i = \theta_r</math>

<br>

</center>

  • Snell's law:

<center>

<br>

<math>k_

Unknown macro: {1z}

= |\vec k_1| \sin \theta_1 = n_1 \frac

Unknown macro: {omega}
Unknown macro: {c}

\sin \theta_1</math>

<br>

<math>k_

Unknown macro: {2z}

= |\vec k_2| \sin \theta_2 = n_2 \frac

Unknown macro: {c}

\sin \theta_2</math>

<br>

<math>n_1 \sin \theta_1 = n_2 \sin \theta_2</math>

<br>

</center>

Analysis of the Reflection from a Dielectric Slab (Reflectivity for s and p Polarized Waves, Field Directions, Poynting Vector, Magnitude of Wavevectors)

Find out the direction of fields from the relation below.

<center>

<br>

<math>\vec S = \frac

Unknown macro: {4 pi}

\vec E \times \vec H</math>

<br>

</center>

Express the fields on both sides of the interface in the following way.

<center>

<br>

<math>\mbox

Unknown macro: {x<0}

</math>

<br>

<math>\vec E = \vec E_1 e^{-i \vec k_1 \vec r} + \vec E'_1 e^{-i \vec k'_1 \vec r}</math>

<br>

<math>\mbox

Unknown macro: {x>0}

</math>

<br>

<math>\vec E = \vec E_2 e^{-i \vec k_2 \vec r} + \vec E'_2 e^{-i \vec k'_2 \vec r}</math>

<br>

</center>

Seprate the electric field into components that are either in the plane of incidence (p) or perpendicular to it (s), in the <math>y</math> direction. Begin with the s-polarized electric field.

<center>

<br>

<math>\vec E_s (x, y, z) = (0, E_s (x, z), 0 )</math>

<br>

</center>

Find the electric field in the <math>i</math>-th medium.

<center>

<br>

<math>\vec E_

Unknown macro: {1s}

= \hat y \left ( E_

e^{- i k_

Unknown macro: {ix}

x} + E'_

Unknown macro: {1s}

e^{i k_

x} \right ) e^

Unknown macro: {i omega t - i beta z}

</math>

<br>

</center>

The dependence of the field on the variables <math>x</math> and <math>z</math> is a result of the selection of the <math>x-z</math> plane as the plane of incidence. The components of <math>k</math> are in the <math>x</math> and <math>z</math> directions. Therefore, there is no dependence on <math>y</math>.

<p>
</p>

The component is tangential to the interface and is continuous across the interface.

<center>

<br>

<math>E_

Unknown macro: {1y}

+ E'_

= E_

Unknown macro: {2y}

+ E'_

</math>

<br>

</center>

Recall the boundary condition on the components of <math>H</math> that are tangential to the interface. Consider an insulating material in whch surface currents are not supported. The tangential component is continuous.

<center>

<br>

<math>\hat n \times ( \vec H_2 - \vec H_1 ) = \frac

Unknown macro: {c}

\vec K</math>

<br>

<math>H_

+ H'_

Unknown macro: {1z}

= H_

Unknown macro: {2z}

+ H'_

</math>

<br>

<math>\vec \nabla \times \vec E + \frac

Unknown macro: {1}
Unknown macro: {c}

\frac

Unknown macro: {partial vec B}

= 0</math>

<br>

<math>\vec B = \mu \vec H</math>

<br>

</center>

Assume a harmonic time variation of <math>H</math> of the form <math>e^

Unknown macro: {i omega t}

</math>. Assume that the material is non-magnetic, <math>\mu=1</math>, and use the angle convention in the notes.

<center>

<br>

<math>\vec \nabla \times \vec E + \frac

Unknown macro: {i omega mu}
Unknown macro: {c}

\vec H = 0</math>

<br>

<math>H_z = \frac

\frac

Unknown macro: {partial}
Unknown macro: {partial x}

E_y</math>

<br>

<math>H_z = \frac

Unknown macro: {partial x}

E_1 e^{-i (k_x x + k_z z)}</math>

<br>

<math>H_z = -k_x \frac

Unknown macro: {c}
Unknown macro: {i omega mu}

E_1 e^{-i (k_x x + k_z z)}</math>

<br>

<math>k_

Unknown macro: {1x}

= |k| \cos \theta_1 = \frac

Unknown macro: {n_1 omega}

\cos \theta_1</math>

<br>

<math>H_

Unknown macro: {1z}

= -n_1 \cos \theta_1 E_1 e^{-i(k_x x + k_z z)}</math>

<br>

<math>H'_

= n_1 \cos \theta_1 E'_1 e^{-i(k'_x x + k_z z)}</math>

<br>

</center>

This leads to two equations for the unknown field amplitudes.

<center>

<br>

<math>n_1 (E_

Unknown macro: {1y}
  • E'_

\cos \theta_1 )= n_2 (E_

Unknown macro: {2y}
  • E'_

\cos \theta_2 )</math>

<br>

</center>

Write in matrix form. These matrices are dynamic matrices of s polarized wave.

<center>

<br>

<math>\begin

Unknown macro: {bmatrix}

1 & 1
n_1 \cos \theta_1 & -n_1 \cos \theta_1\end

\begin

Unknown macro: {bmatrix}

E_

Unknown macro: {1y}


E'_

\end

= \begin

Unknown macro: {bmatrix}

1 & 1
n_2 \cos \theta_2 & -n_2 \cos \theta_2\end

\begin

Unknown macro: {bmatrix}

E_

Unknown macro: {2y}


E'_

\end

</math>

<br>

<math>D_2^{-1} D_1 = D_

Unknown macro: {12}

</math>

<br>

<math>D_

\begin

Unknown macro: {bmatrix}

E_

Unknown macro: {1y}


E'_

\end

= \begin

Unknown macro: {bmatrix}

E_

Unknown macro: {2y}


E'_

\end

</math>

<br>

</center>

Consider where the wave is incident from medium 1 moving from left to right and set <math>E'_2 = 0</math>. The reflection and transmission coefficients.

<center>

<br>

<math>r_s \left ( \frac

Unknown macro: {E'_1}
Unknown macro: {E_1}

\right ) = \frac

Unknown macro: {n_1 cos theta_1 - n_2 cos theta_2}

{n_1 \cos \theta_1 + n_2 \cos \theta_2</math>

<br>

<math>r_s \left ( \frac

Unknown macro: {E_2}

\right ) = \frac

Unknown macro: {2n_1 cos theta_1}

{n_1 \cos \theta_1 + n_2 \cos \theta_2</math>

<br>

<math>R = |r_s|^2</math>

<br>

</center>

The last term above is the reflectivity.

How to Derive the Transfer Matrix that Describes the Field Transformation at Interfaces

Consider medium with the following index of refraction. An objective is to relate any two pairs of amplitudes to any other two pairs.

<center>

<br>

<table cellpadding=10>
<tr>

<td>
<math>x<0</math>
</td>

<td>
<math>0<x<d</math>
</td>

<td>
<math>d<x</math>
</td>

</tr>

<tr>

<td>
<math>n_1</math>
</td>

<td>
<math>n_2</math>
</td>

<td>
<math>n_3</math>
</td>

</tr>

</table>

</center>

The <math>z</math> component of the wavevector does not change throughout the problem. It leads to a simple dependence on the <math>z</math> direction evolution.

<center>

<br>

<math>\vec E = \vec E e^

Unknown macro: {i(omega t - beta z)}

</math>

<br>

</center>

Define amplitudes on both sides of each interface for an s polarized field.

<center>

<br>

<math>x<0</math>

<br>

<math>E_1 e^{-ik_

Unknown macro: {1x}

x + E'1 e^{-ik

x</math>

<br>

<math>0<x<d</math>

<br>

<math>E_2 e^{-ik_

Unknown macro: {2x}

x + E'2 e^{-ik

x</math>

<br>

<math>d<x</math>

<br>

<math>E_3 e^{-ik_

Unknown macro: {3x}

x + E'3 e^{-ik

x</math>

<br>

</center>

The propagation matrix captures the effect of propagation through a medium of index <math>n_2</math> and thickness <math>d</math>.

<center>

<br>

<math>P_2 = \begin

Unknown macro: {bmatrix}

e^{ik_

Unknown macro: {2x}

d}& 0
0& e^{ik_

d}\end

</math>

<br>

<math>D_

Unknown macro: {23}

P_2 D_

Unknown macro: {12}

\begin

Unknown macro: {bmatrix}

E_1
E'_1\end

= \begin

Unknown macro: {bmatrix}

E_3
E'_3\end

</math>

<br>

</center>

Transfer Matrix Approach to Solving Periodic Systems of Dielectric Materials

<center>

<br>

<table cellpadding=10>
<tr>

<td>
<math>na<x<na+d_2</math>
</td>

<td>
<math>na-d_1<x<na</math>
</td>

</tr>

<tr>

<td>
<center>
<math>n_1</math>
</center>
</td>

<td>
<center>
<math>n_2</math>
</center>
</td>

</tr>

</table>

<br>

<math>n(x+na) = n</math>

<br>

</center>

The solution in each medium is of the form below.

<center>

<br>

<math>E = E e^

</math>

<br>

</center>

Focus on the <math>x</math> dependence of the solutions.

<center>

<br>

<math>na -d_1 <x <na</math>

<br>

<math>E = A_n e^{-ik_

Unknown macro: {1x}

(x-na)} + B_n e^{ik_

(x-na)}</math>

<br>

<math>(na-1) -d_1 <x <na - d_1</math>

<br>

<math>E = A_n e^{-ik_

Unknown macro: {2x}

(x-na+d_1)} + B_n e^{ik_

(x-na+d_1)}</math>

<br>

</center>

The field notation has been changed in order to simplify the labeling of the fields in these problems which involve a multiplicity of interfaces.

<center>

<br>

<math>k_

Unknown macro: {1x}

= \sqrt{ \left ( \frac

Unknown macro: { omega n_1}
Unknown macro: {c}

\right )^2 - k_

^2} = \sqrt{ \left ( \frac

Unknown macro: { omega n_1}
Unknown macro: {c}

\right )^2 - \beta^2} = \frac

Unknown macro: {omega n_1}

\cos \theta_1</math>

<br>

<math>k_

Unknown macro: {2x}

= \sqrt{ \left ( \frac

Unknown macro: { omega n_2}
Unknown macro: {c}

\right )^2 - \beta^2} = \frac

Unknown macro: {omega n_2}

\cos \theta_2</math>

<br>

</center>

Dynamic matrices are below.

<center>

<br>

<math>D_l^s = \begin

Unknown macro: {bmatrix}

1& 1
n_1 \cos \theta_1& -n_1 \cos \theta_1\end

</math>

<br>

<math>D_

Unknown macro: {l-1,l}

= D_l^{-1}D_

Unknown macro: {l-1}

</math>

<br>

</center>

Consider the solution form. Relate the fields across the interface.

<center>

<br>

<math>E|_{x=na^{(1)}} = A_n + B_n</math>

<br>

<math>E|{x=na-d_1^{(1)}} = A_ne^{-ik

Unknown macro: {1x}

(na-d_1-na)} + B_ne^{-ik_

(na-d_1-na)}</math>

<br>

<math>E|{x=na-d_1^{(1)}} = A_ne^{ik

Unknown macro: {1x}

d_1} + B_ne^{-ik_

d_1}</math>

<br>

<math>E|_{x=na-d_1^{(1)}} = C_n + D_n</math>

<br>

<math>P_1 = \begin

Unknown macro: {bmatrix}

e^{ik_

Unknown macro: {1x}

d_1} & 0
0 & e^{-ik_

d_1}\end

</math>

<br>

<math>D_2 \begin

Unknown macro: {bmatrix}

C_n
D_n\end

= D_1 P_1 \begin

Unknown macro: {bmatrix}

A_n
B_n\end

</math>

<br>

<math>\begin

Unknown macro: {bmatrix}

C_n
D_n\end

= D_2^{-1} D_1 P_1 \begin

Unknown macro: {bmatrix}

A_n
B_n\end

</math>

<br>

</center>

The solution brings one to the LHS of the <math>x=na-d_1</math> interface (i.e. in medium 2). Propagate the wave in medium 2 back to the RHS of interface <math>x = (n-1)a</math>.

<center>

<br>

<math>E|_{x=na-d_1^

Unknown macro: {(1)}

= C_n +D_n</math>

<br>

<math>E|_{x=(n-1)d^

Unknown macro: {(2)}

= C_ne^{ik_

d_2 +D_ne^{-ik_

Unknown macro: {2x}

d_2</math>

<br>

<math>P_2 = \begin

Unknown macro: {bmatrix}

e^{ik_

d_2} & 0
0 & e^{-ik_

Unknown macro: {2x}

d_2}\end

Unknown macro: {bmatrix}

</math>

<br>

</center>

Match fields across the interface to extract <math>A_

Unknown macro: {n-1}

</math> and <math>B_

</math>.

<center>

<br>

<math>E|_{x=na-d^

Unknown macro: {(2)}

= C_ne^{ik_

d_2 +D_ne^{-ik_

Unknown macro: {2x}

d_2</math>

<br>

<math>E|_{x=na-d^

Unknown macro: {(1)}

= A_

Unknown macro: {n-1}

+ B_

</math>

<br>

<math>D_1 \begin

Unknown macro: {bmatrix}

A_

Unknown macro: {n-1}


B_

\end

= D_2 P_2 \begin

Unknown macro: {bmatrix}

C_

Unknown macro: {n}


D_

\end

</math>

<br>

<math>\begin

Unknown macro: {bmatrix}

A_

Unknown macro: {n-1}


B_

\end

= D_1^{-1} D_2 P_2 \begin

Unknown macro: {bmatrix}

C_

Unknown macro: {n}


D_

\end

</math>

<br>

<math>\begin

Unknown macro: {bmatrix}

A_

Unknown macro: {n-1}


B_

\end

= D_1^{-1} D_2 P_2 D_2^{-1} D_1 P_1 \begin

Unknown macro: {bmatrix}

A_

Unknown macro: {n}


B_

\end

</math>

<br>

</center>

Eigenvalue Problem and its Solutions

The following is true from Bloch's Theorem.

<center>

<br>

<math>\begin

Unknown macro: {bmatrix}

A_

Unknown macro: {n-1}


B_

\end

= e^

Unknown macro: {iKa}

\begin

Unknown macro: {bmatrix}

A_

Unknown macro: {n}


B_

\end

</math>

<br>

<math>D_1^{-1} D_2 P_2 D_2^{-1} D_1 P_1 \begin

Unknown macro: {bmatrix}

A_

Unknown macro: {n}


B_

\end

= e^

\begin

Unknown macro: {bmatrix}

A_

Unknown macro: {n}


B_

\end

</math>

<br>

<math>D_1^{-1} D_2 P_2 D_2^{-1} D_1 P_1 = M</math>

<br>

<math>M = \begin

Unknown macro: {bmatrix}

M_

Unknown macro: {11}

& M_

Unknown macro: {12}


M_

Unknown macro: {21}

& M_

Unknown macro: {22}

\end

</math>

<br>

</center>

Each one of the matrix elements depends on <math>\omega</math> and <math>\beta</math>. The eigenvalues of the 2x2 matrix are below.

<center>

<br>

<math>e^

Unknown macro: {iKa}

= \frac

Unknown macro: {1}
Unknown macro: {2}

(M_

Unknown macro: {11}

+ M_

Unknown macro: {22}

) \pm \left ( \frac

(M_

Unknown macro: {11}

+ M_

Unknown macro: {22}

)2 - 1 \right )

Unknown macro: {1/2}

</math>

<br>

</center>

The equation defines the dispersion relations for the Bloch wavenumber <math>K</math> and <math>\omega</math> and <math>\beta</math>. The determinant of the <math>M</math> matrix is unity, and e^

Unknown macro: {iKa}

and e^{-iKa} are two eigenvalues.

<center>

<br>

<math>\frac{e^

+ e^{-iKa}}

Unknown macro: {2}

= \cos Ka</math>

<br>

<math>K(\beta, \omega) = \frac

Unknown macro: {1}
Unknown macro: {a}

\cos^{-1} \left ( \frac{M_

+ M_{22}}

Unknown macro: {2}

\right )</math>

<br>

</center>

The eigenvectors are below

<center>

<br>

<math>\begin

Unknown macro: {bmatrix}

A_

Unknown macro: {0}


B_

\end

= e^

Unknown macro: {iKa}

\begin

Unknown macro: {bmatrix}

M_

Unknown macro: {12}


e^

- M_

Unknown macro: {11}

\end

Unknown macro: {bmatrix}

</math>

<br>

</center>

Photonic Band Diagrams and Bandgaps

Consider the case where <math>\beta=0</math> which corresponds to normal incidence. The dispersion relation <math>\omega</math> vs. <math>K</math> can be expressed as below.

<center>

<br>

<math>\cos Ka = \frac{M_

+ M_

Unknown macro: {22}

}

= \cos k_1 d_1 \cos k_2 d_2 - \frac

Unknown macro: {1}
Unknown macro: {2}

\left ( \frac

Unknown macro: {n_2}
Unknown macro: {n_1}

+ \frac

\right ) \sin k_1 d_1 \sin k_2 d_2</math>

<br>

<math>k_1 = \frac

Unknown macro: {omega n_1}
Unknown macro: {c}

</math>

<br>

<math>k_2 = \frac

Unknown macro: {omega n_2}

</math>

<br>

</center>

The width of the normal incidence bandgap is given approximately by the expression below.

<center>

<br>

<math>\frac

Unknown macro: {Delta omega}
Unknown macro: {omega_o}

\approx \frac

Unknown macro: {pi}

\frac

Unknown macro: {Delta n}
Unknown macro: {n}

</math>

<br>

</center>

Additional Topics

  • States within the band gap
    *Organic semiconductors
  • No labels