Phonovoltaic

edit
 
The phonovoltaic cell and an energy diagram of its operation.

The phonvoltaic (pV) cell converts vibrational (phonon) energy into a direct current much like the photovoltaic effect in a photovoltaic (PV) cell converts light (photon) into power. That is, it uses a p-n junction to separate the electrons and holes generated as valence electrons absorb optical phonons more energetic than the band gap, and then collects them in the metallic contacts for use in a circuit[1]. The pV cell is an application of heat transfer physics[2] and competes with other thermal energy harvesting devices like the thermoelectric generator.

While the thermoelectric generator converts heat, a broad spectrum of phonon and electron energy, to electricity, the pV cell converts only a narrow band of phonon energy, i.e., only the most energetic optical phonon modes. A narrow band of excited optical phonons has much less entropy than a similary excited broad spectrum of phonons. Thus, the pV cell can exceed the thermoelectric efficiency. However, exciting and harvesting the optical phonon poses a challenge.

Satisfying the laws of thermodynamics

edit

By the first law of thermodynamics, the excitation driving electron generation in both photo- and phonovoltaic cells, i.e., the photon or phonon, must have more energy than the semiconductor band gap. For a PV cell, many materials are available with a band gap ( ) well matched to the solar photon spectrum, like Silicon or Gallium Arsenide. For a pV cell, however, no current semiconducting materials have a band gap smaller than the energy of their most energetic (optical) phonon modes ( ). Thus, novel materials are required with both energetic optical phonon modes (  meV, e.g., graphene, diamond, or boron nitride) and a small band gap ( , e.g., graphene).

By the second law of thermodynamics, the excitation must be "hotter" than the cell for power generation to occur. In a PV, the light comes from an outside source, for example, the sun, which is nearly 6000 Kelvin, whereas the PV is around 300 Kelvin. Thus, the second law is satisfied and energy conversion is possible. However, the crystal vibrations driving power generation in a pV are intrinsic to the material itself. As such, they can not be imported from an outside source like the sun, but must instead be excited by some other proceess until they are hotter than the cell. The temperature of the optical phonon population is extracted by comparing the number of optical phonon to the number expected at a given temperature, which comes from the Bose-Einstein statistics.

Non-equilibrium optical phonon population and the nanoscale requirent

edit

There are a number of ways to excite a population of vibrations, i.e., create a hot optical phonon population. For example, if the electron population is excited, using a laser or electric field, they will typically relax by emitting optical phonons. Additionally, a hot molecular gas can impart its vibrations to a crystal when chemisorbed. Regardless of method, the conversion efficiency is limited by the optical phonon temperature achieved as compared to the electron temperature within the device due to Carnot's theorem.

In a nanoscale device, this temperature is approximately equal to the temperature of the device itself. However, in a macroscale device the generated electrons accumulate faster than they are collected. Thus, the electron population is heated up to the optical phonon temperature and further generation is inhibited. The down-conversion is simultaneously inhibited as the acoustic phonon population is heated to the optical phonon temperature. Thus, the large pV cell develops a near-equilibrium state where it is heated. At best, it will act like a thermoelectric generator and exhibit thermoelectric effects. Such a device is called a thermovoltaic, rather than a phonovoltaic.

Entropy generation and efficiency

edit

Entropy generation and inefficiency in a PV cell is the result of photons more energetic than the band gap producing electrons with kinetic energy in addition to the potential energy provided by the band gap. Similarly, optical phonon energy in excess of the band gap generates an entropy flow in the pV cell, rather than electric power. The energy efficiency ( ) is quantified by the ratio of the band gap and optical phonon energy, that is

 

In addition this typical inefficiency, hot optical phonon populations tend to downconvert into multiple low-energy, acoustic phonon modes (whereas photons typically do not downconvert into low energy infrared waves). This efficiency ( ) is quantified by the tendency of a hot optical phonon to downconvert rather than generate an electron-hole pair, that is

 

where   is the rate of generation and   is the rate of downconversion, i.e., the rate at which an optical phonon produces multiple low-energy, acoustic phonons. This provides a second entropy flow reducing the efficiency of a pV cell.

Finally, entropy is generated in both pV and PV cells due to the inefficient separation of the generated electrons and holes. This efficiency ( ) is limited by the Carnot efficiency given by

 

where   is the temperature of the pV cell and   is the temperature of the optical phonon population, as dictated by the Bose-Einstein statistics. This efficiency is reduced the smaller the band gap is in comparison to the thermal energy ( , where   is the Boltzmann constant and   is the temperature). Indeed, the p-n junction efficiency is approximately

 
 
The efficiency of the phonovoltaic cell for variations in the figure of merit, Carnot limit, and band gap.

Thus, the overall efficiency ( ) is

 

where the temperature independent terms become the material figure of merit ( ),

 

If the band gap and optical phonon mode are resonant, and the optical phonon tends to generate electrons, the phonovoltaic cell can approach the Carnot limit.

The electron-phonon coupling

edit

The electron-phonon coupling is responsible for electron generation in the pV cell. In this phenomenon, the phonon leads to ion motion which perturbs the highest occupied valence state (HOS). This state begins to overlap with the lowest unoccupied conduction state (LUS), and the electron can switch states if energy and momentum are conserved. If it does, an electron-hole pair is generated.

 
Illustration of the electron-phonon coupling betwen valence and conduction states in real and momentum space.

Using a taylor expansion of the change in electron potential,  , due to a the ionic displacement of a phonon provides a matrix element for use in Fermi's golden rule, and the derivation of a generation rate. This Taylor expansion gives the following matrix element

 

where   is the atomic displacement due to a phonon with polarization   and |momentum  , and   is the electron wavefunction for an electron with momentum   in band i.

The phonon-phonon coupling

edit

Competing with the generation of electrons is the downconversion of optical phonons into multiple acoustic phonons. The coupling arises from the crystal Hamiltonian (H) expanded in terms of the ionic displacement ( ) from the equilibrium position ( ) of atom i in direction   in direction , i.e.,

 

where   is the ground-state Hamiltonian, the linear term vanishes (as the ground state is found by minimizing the energy in terms of the ionic position), and   and   are the second- and third-order force constants between atoms i, j, and k when moved in along coordinate  ,  , and  . The second order term is primarily responsible for the phonon dispersion, while the anharmonic (third order and higher) terms are responsible for thermal expansion as well as the phonon up- (multiple low-energy optical phonons combine to form a high-energy phonon) and downconversion (a high-energy phonon splits into multiple low-energy phonons).

Typically, up- and down-conversion is dominated by the third-order interaction. Thus, the perturbation Hamiltonian used in Fermi's golden rule for phonon up- and downconversion is

 

where   is the direction of atomic displacement for atom i due to a phonon with polarization   and momentum  .

pV materials

edit
 
The phonon energy and density of states for various semiconductors and semimetals.
 
The band gap of various semiconductors and semimetals.

As outlined above, an efficient pV cell requires a material with an optical phonon mode more energetic than the bandgap, which in turn is much more energetic than the thermal energy at the intended operating temperature  . Furthermore, the pV cell requires a material wherein a hot optical phonon prefers to produce an electron rather than multiple low energy acoustic phonons ( ).

No current material offers this combination of properties. In general, a material with a first-row element is required to have a highly energetic optical phonon, e.g., greater than 100 meV. For example, the carbon allotropes (diamond, graphene, and graphite) or boron nitride allotropes produce optical phonon modes exceeding 150 meV. However, the high electronegativity of a first-row elements tends to create a very large band gap, as in diamond and the boron nitride allotropes (band gaps greater than 3 eV), unless they are sufficiently symmetric, like graphene (no band gap)

Therefore, Tuning the band gap and manipulating the symmetry of the first-row semiconductors (periodic table) is required to reach the first requirements of a pV cell ( ). The band gap in graphene has been tuned in a variety of ways. For instance, hydrogenated graphene (graphame for partial hydrogenation and graphane for full hydrogenation) can open a band gap ranging between 0 and 3.5 eV. At 24 hydrogen per 128 carbon, the band gap is tuned to the graphene-like optical phonon, with an energy around 200 meV[3]. If hydrogenation did not significantly affect the electron-phonon or phonon-phonon coupling, this graphame composition could achieve 90% of the Carnot limit, much better than the best thermoelectric generator. However, hydrogenation is predicted to degrade the interband electron-phonon coupling substantially and proscribes its use in an efficient pV cell.

See also

edit

References

edit
  1. ^ Melnick, C.; Kaviany, M. (March 2016), "Phonovoltaic I. Harvesting hot optical phonons in a nanoscale p-n junction", Physical Review B, 93: 094302
  2. ^ Kaviany, M. (2014). Heat transfer physics (2nd ed.). Cambridge: Cambridge University Press. ISBN 978-1-107041783.
  3. ^ Melnick, C.; Kaviany, M. (March 2016), "Phonovoltaic II. Tuning band gap to optical phonon in graphite", Physical Review B, 93: 125203

Category:Heat transfer Category:Thermodynamics Category:Condensed matter physics