Interpreting Defect and Energy Level Diagrams

A beginner's guide to understanding defect and energy level diagrams obtained from first-principles calculations

The formation energy of point defects is calculated using the supercell approach described here. The results of the defect calculations are typically presented in the form of "defect diagrams" and "energy level diagrams" – both are discussed below. These diagrams are useful for qualitative assessment of defect and doping properties (e.g., dopability) as well as quantitative determination of defect and electronic charge carrier (electrons, holes) concentrations. The discussion here will focus on semiconductors and insulators i.e., materials with a band gap.

Defect Diagrams

How are these defect diagrams plotted?

What does the charge state/slope of a defect signify and how is related to doping?

Charged defects create electronic carriers – electrons and holes. A neutral defect (e.g. isoelectronic doping, alloying) does not create electronic carriers. A positively-charged defect has a positive slope in a defect diagram and represent a donor-like defect. In other words, the defect ionizes to a positively-charged state by giving up (donating, hence donor) electron(s). Donors tend to make the material n-type. Similarly, negatively-charged defects are acceptor-like and they tend to make the p-type. Whether a material is actually doped n- or p-type depends on several other factors. We will return to the topic of doping once we have understood a few additional concepts.

What is the significance of a charge transition level in defect diagram?

The charge transition levels represent the approximate energetic location of defect states (a.k.a. defect levels) inside the band gap. These defect states give rise to many interesting phenomena that have implications for the electronic, optical, magnetic properties and many more. It is beyond the scope of this tutorial to delve into these effects. At this point, the mere acknowledgement of the existence of these defect states is sufficient.

What are shallow and deep defects and how to find them in defect diagrams?

In rare cases, an extremely deep donor defect will have CTLs lie inside the valence band. Analogously, an extremely deep acceptor will have CTLs inside the conduction band.

What is the effect of elemental chemical potentials on the defect diagram and how is related to the synthesis/growth conditions?

Let us consider the example of ZnO grown under oxygen-rich (O-rich) and oxygen-deficient (O-poor) conditions. Under what conditions would oxygen vacancy formation in ZnO be more favorable? We can arrive at the answer by simply thinking about the process. Oxygen vacancy formation involves the removal of O atoms from ZnO and placing it in the external O reservoir. Under O-rich conditions, this external reservoir already has a high concentration of O, so placing additional O atoms would require a lot more energy. In contrast, under O-poor conditions, where the reservoir has a lower concentration of oxygen, removal of O from ZnO and placing it in this reservoir should be energetically cheaper. Therefore, O vacancy formation is more favorable under O-poor conditions. A detailed mathematical treatment on the effect of elemental chemical potentials and growth conditions on defect energetics will be added later.

How do we calculate the defect concentrations and associated electron and hole concentrations?

The defect concentration at a given temperature is calculated from the Boltzmann probability:

The equation for calculating defect concentration using Boltzmann probability implicitly includes the configurational entropy contribution. However, it does not include the vibrational entropy contribution, which is generally negligible in most cases. The latter is an important consideration when defect formation involves elements that are gases under standard conditions. In such cases, chemical potential of the elemental (gaseous) reservoir has a strong T dependence.

The defect and electronic carrier (electron, hole) concentrations are calculated by imposing the condition of overall charge neutrality. This means, for the material to be charge neutral, the total number of positive and negative charges must be balanced. The positive charges include donor defects (+ve slope in defect diagrams) and holes, and negative charges are acceptors (-ve slope in defect diagrams) and electrons.

The equilibrium Fermi energy is also often referred to as pinned Fermi level. Strict practitioners of the field may argue that Fermi level pinning is not an appropriate term in this context.

This tutorial focuses on the equilibrium thermodynamics of defect formation. Defects can and do form under non-equilibrium condition, but determination of their concentrations is not straightforward without detailed experimental inputs. It is a specialized topic that is beyond the scope of this guide designed for beginners.

Can we guess the approximate location of the equilibrium Fermi energy? What can we learn?

How do we qualitatively interpret the defect diagrams in the context of doping?

We learned above that donor-like defects tend to donate electrons while acceptor-like defects tend to accept electrons (create holes). Rarely does a material contain only donor (or acceptor) defects; often, both donor-like and acceptor-like defects are present. Whether a material is overall n- or p-type, depends on the relative concentrations of donor and acceptor defects. Intuitively, if the concentration of donors (+ve charged defects) is much higher than acceptors (-ve charged defects), one can imagine that the material will be n-type. Mathematically, this can be understood by charge balance discussed above. If the +ve charged defects are in higher concentration, the overall charge balance is restored by creating more electrons (-ve charges) than holes. As such, the material will have a high concentration of electrons, making it n-type. Here, we have discussed the scenario of "self-doping" by the native defects. However, in practical applications, materials are often doped with extrinsic dopants e.g., n-type Si doped with P, n-type ZnO doped with Al.

Materials cannot be extrinsically doped n- or p-type at will (otherwise life would have been easy!). A lot depends on the native defect chemistry. In other words, whether a material can be doped n- or p-type depends critically on the donors and acceptors natively present in material e.g., Si vacancy in Si or O vacancy in ZnO. Native defects, as the name suggests, are inherently present in the material and are not introduced through extrinsic doping. Therefore, one cannot bypass the existence of such native defects. The formation energetics of these native defects determine if a material is n- or p-type (or both, or neither) dopable. Here, dopable means the possibility of extrinsically doping a material. The figure below showcases four different scenarios for materials that are (1) p-type dopable, (2) n-type dopable, (3) both p- and n-type dopable ("ambipolar"), and (4) neither p- nor n-type dopable (typical insulators).

Scenario 1:

Last updated