You are on page 1of 10

Work function

In solid-state physics, the work function (sometimes spelled workfunction) is the minimum thermodynamic
work (i.e., energy) needed to remove an electron from a solid to a point in the vacuum immediately outside
the solid surface. Here "immediately" means that the final electron position is far from the surface on the
atomic scale, but still too close to the solid to be influenced by ambient electric fields in the vacuum. The
work function is not a characteristic of a bulk material, but rather a property of the surface of the material
(depending on crystal face and contamination).

Contents
Definition
Applications
Measurement
Methods based on thermionic emission
Work function of cold electron collector
Methods based on photoemission
Kelvin probe method
Work functions of elements[12]
Physical factors that determine the work function
Surface dipole
Doping and electric field effect (semiconductors)
Theoretical models of metal work functions
Temperature dependence of the electron work function
References
Further reading
External links

Definition
The work function W for a given surface is defined by the difference[1]

where −e is the charge of an electron, ϕ is the electrostatic potential in the vacuum nearby the surface, and
EF is the Fermi level (electrochemical potential of electrons) inside the material. The term −eϕ is the
energy of an electron at rest in the vacuum nearby the surface.

In practice, one directly controls EF by the voltage applied to the material through electrodes, and the work
function is generally a fixed characteristic of the surface material. Consequently, this means that when a
voltage is applied to a material, the electrostatic potential ϕ produced in the vacuum will be somewhat
lower than the applied voltage, the difference depending on the work function of the material surface.
Rearranging the above equation, one has
where V = −EF/e is the voltage of the material (as
measured by a voltmeter, through an attached
electrode), relative to an electrical ground that is defined
as having zero Fermi level. The fact that ϕ depends on
the material surface means that the space between two
dissimilar conductors will have a built-in electric field,
when those conductors are in total equilibrium with
each other (electrically shorted to each other, and with
equal temperatures). An example of this situation is Plot of electron energy levels against position, in a
depicted in the adjacent figure. As described in the next gold-vacuum-aluminium system. The two metals
section, these built-in vacuum electric fields can have depicted here are in complete thermodynamic
important consequences in some cases. equilibrium. However, the vacuum electrostatic
potential ϕ is not flat due to a difference in work
function.
Applications
Thermionic emission
In thermionic electron guns, the work function and temperature of the hot cathode are critical
parameters in determining the amount of current that can be emitted. Tungsten, the common
choice for vacuum tube filaments, can survive to high temperatures but its emission is
somewhat limited due to its relatively high work function (approximately 4.5 eV). By coating
the tungsten with a substance of lower work function (e.g., thorium or barium oxide), the
emission can be greatly increased. This prolongs the lifetime of the filament by allowing
operation at lower temperatures (for more information, see hot cathode).
Band bending models in solid-state electronics
The behavior of a solid-state device is strongly dependent on the size of various Schottky
barriers and band offsets in the junctions of differing materials, such as metals,
semiconductors, and insulators. Some commonly used heuristic approaches to predict the
band alignment between materials, such as Anderson's rule and the Schottky-Mott rule, are
based on the thought experiment of two materials coming together in vacuum, such that the
surfaces charge up and adjust their work functions to become equal just before contact. In
reality these work function heuristics are inaccurate due to their neglect of numerous
microscopic effects. However, they provide a convenient estimate until the true value can be
determined by experiment.[2][3]
Equilibrium electric fields in vacuum chambers
Variation in work function between different surfaces causes a non-uniform electrostatic
potential in the vacuum. Even on an ostensibly uniform surface, variations in W known as
patch potentials are always present due to microscopic inhomogeneities. Patch potentials
have disrupted sensitive apparatus that rely on a perfectly uniform vacuum, such as Casimir
force experiments[4] and the Gravity Probe B experiment.[5] Critical apparatus may have
surfaces covered with molybdenum, which shows low variations in work function between
different crystal faces.[6]
Contact electrification
If two conducting surfaces are moved relative to each other, and there is potential difference
in the space between them, then an electric current will be driven. This is because the
surface charge on a conductor depends on the magnitude of the electric field, which in turn
depends on the distance between the surfaces. The externally observed electrical effects are
largest when the conductors are separated by the smallest distance without touching (once
brought into contact, the charge will instead flow internally through the junction between the
conductors). Since two conductors in equilibrium can have a built-in potential difference due
to work function differences, this means that bringing dissimilar conductors into contact, or
pulling them apart, will drive electric currents. These contact currents can damage sensitive
microelectronic circuitry and occur even when the conductors would be grounded in the
absence of motion.[7]

Measurement
Certain physical phenomena are highly sensitive to the value of the work function. The observed data from
these effects can be fitted to simplified theoretical models, allowing one to extract a value of the work
function. These phenomenologically extracted work functions may be slightly different from the
thermodynamic definition given above. For inhomogeneous surfaces, the work function varies from place to
place, and different methods will yield different values of the typical "work function" as they average or
select differently among the microscopic work functions.[8]

Many techniques have been developed based on different physical effects to measure the electronic work
function of a sample. One may distinguish between two groups of experimental methods for work function
measurements: absolute and relative.

Absolute methods employ electron emission from the sample induced by photon absorption
(photoemission), by high temperature (thermionic emission), due to an electric field (field
electron emission), or using electron tunnelling.
Relative methods make use of the contact potential difference between the sample and a
reference electrode. Experimentally, either an anode current of a diode is used or the
displacement current between the sample and reference, created by an artificial change in the
capacitance between the two, is measured (the Kelvin Probe method, Kelvin probe force
microscope). However, absolute work function values can be obtained if the tip is first
calibrated against a reference sample.[9]

Methods based on thermionic emission

The work function is important in the theory of thermionic emission, where thermal fluctuations provide
enough energy to "evaporate" electrons out of a hot material (called the 'emitter') into the vacuum. If these
electrons are absorbed by another, cooler material (called the collector) then a measurable electric current
will be observed. Thermionic emission can be used to measure the work function of both the hot emitter and
cold collector. Generally, these measurements involve fitting to Richardson's law, and so they must be
carried out in a low temperature and low current regime where space charge effects are absent.

In order to move from the hot emitter to the vacuum, an


electron's energy must exceed the emitter Fermi level by
an amount

determined simply by the thermionic work function of


the emitter. If an electric field is applied towards the
surface of the emitter, then all of the escaping electrons
will be accelerated away from the emitter and absorbed
into whichever material is applying the electric field. Energy level diagrams for thermionic diode in
According to Richardson's law the emitted current forward bias configuration, used to extract all hot
electrons coming out from the emitter's surface.
density (per unit area of emitter), Je (A/m2), is related to
The barrier is the vacuum near emitter surface.
the absolute temperature Te of the emitter by the
equation:
where k is the Boltzmann constant and the proportionality constant Ae is the Richardson's constant of the
emitter. In this case, the dependence of Je on Te can be fitted to yield We.

Work function of cold electron collector

The same setup can be used to instead measure the work


function in the collector, simply by adjusting the applied
voltage. If an electric field is applied away from the
emitter instead, then most of the electrons coming from
the emitter will simply be reflected back to the emitter.
Only the highest energy electrons will have enough
energy to reach the collector, and the height of the
potential barrier in this case depends on the collector's
work function, rather than the emitter's.
Energy level diagrams for thermionic diode in
The current is still governed by Richardson's law. retarding potential configuration. The barrier is the
vacuum near collector surface.
However, in this case the barrier height does not depend
on We. The barrier height now depends on the work
function of the collector, as well as any additional
applied voltages:[10]

where Wc is the collector's thermionic work function, ΔVce is the applied collector–emitter voltage, and ΔVS
is the Seebeck voltage in the hot emitter (the influence of ΔVS is often omitted, as it is a small contribution
of order 10 mV). The resulting current density Jc through the collector (per unit of collector area) is again
given by Richardson's Law, except now

where A is a Richardson-type constant that depends on the collector material but may also depend on the
emitter material, and the diode geometry. In this case, the dependence of Jc on Te, or on ΔVce, can be fitted
to yield Wc.

This retarding potential method is one of the simplest and oldest methods of measuring work functions,
and is advantageous since the measured material (collector) is not required to survive high temperatures.

Methods based on photoemission

The photoelectric work function is the minimum photon energy required to liberate an electron from a
substance, in the photoelectric effect. If the photon's energy is greater than the substance's work function,
photoelectric emission occurs and the electron is liberated from the surface. Similar to the thermionic case
described above, the liberated electrons can be extracted into a collector and produce a detectable current, if
an electric field is applied into the surface of the emitter. Excess photon energy results in a liberated electron
with non-zero kinetic energy. It is expected that the minimum photon energy required to liberate an
electron (and generate a current) is
where We is the work function of the emitter.

Photoelectric measurements require a great deal of care,


as an incorrectly designed experimental geometry can
result in an erroneous measurement of work function.[8]
This may be responsible for the large variation in work
function values in scientific literature. Moreover, the
minimum energy can be misleading in materials where
there are no actual electron states at the Fermi level that
are available for excitation. For example, in a Photoelectric diode in forward bias configuration,
semiconductor the minimum photon energy would used for measuring the work function We of the
actually correspond to the valence band edge rather than illuminated emitter.
work function.[11]

Of course, the photoelectric effect may be used in the retarding mode, as with the thermionic apparatus
described above. In the retarding case, the dark collector's work function is measured instead.

Kelvin probe method

The Kelvin probe technique relies on the detection of an


electric field (gradient in ϕ) between a sample material
and probe material. The electric field can be varied by
the voltage ΔVsp that is applied to the probe relative to
the sample. If the voltage is chosen such that the electric
field is eliminated (the flat vacuum condition), then

Since the experimenter controls and knows ΔVsp, then Kelvin probe energy diagram at flat vacuum
configuration, used for measuring work function
finding the flat vacuum condition gives directly the
difference between sample and probe.
work function difference between the two materials.
The only question is, how to detect the flat vacuum
condition? Typically, the electric field is detected by
varying the distance between the sample and probe. When the distance is changed but ΔVsp is held constant,
a current will flow due to the change in capacitance. This current is proportional to the vacuum electric field,
and so when the electric field is neutralized no current will flow.

Although the Kelvin probe technique only measures a work function difference, it is possible to obtain an
absolute work function by first calibrating the probe against a reference material (with known work
function) and then using the same probe to measure a desired sample.[9] The Kelvin probe technique can be
used to obtain work function maps of a surface with extremely high spatial resolution, by using a sharp tip
for the probe (see Kelvin probe force microscope).

Work functions of elements[12]


Below is a table of work function values for various elements. Note that the work function depends on the
configurations of atoms at the surface of the material. For example, on polycrystalline silver the work
function is 4.26 eV, but on silver crystals it varies for different crystal faces as (100) face: 4.64 eV, (110)
face: 4.52 eV, (111) face: 4.74 eV.[13] Ranges for typical surfaces are shown in the table below.
Work function of elements (eV)
Ag 4.26 – 4.74 Al 4.06 – 4.26 As 3.75
Au 5.10 – 5.47 B ~4.45 Ba 2.52 – 2.70
Be 4.98 Bi 4.31 C ~5
Ca 2.87 Cd 4.08 Ce 2.9
Co 5 Cr 4.5 Cs 1.95
Cu 4.53 – 5.10 Eu 2.5 Fe: 4.67 – 4.81
Ga 4.32 Gd 2.90 Hf 3.90
Hg 4.475 In 4.09 Ir 5.00 – 5.67
K 2.29 La 3.5 Li 2.9
Lu ~3.3 Mg 3.66 Mn 4.1
Mo 4.36 – 4.95 Na 2.36 Nb 3.95 – 4.87
Nd 3.2 Ni 5.04 – 5.35 Os 5.93
Pb 4.25 Pd 5.22 – 5.60 Pt 5.12 – 5.93
Rb 2.261 Re 4.72 Rh 4.98
Ru 4.71 Sb 4.55 – 4.70 Sc 3.5
Se 5.9 Si 4.60 – 4.85 Sm 2.7
Sn 4.42 Sr ~2.59 Ta 4.00 – 4.80
Tb 3.00 Te 4.95 Th 3.4
Ti 4.33 Tl ~3.84 U 3.63 – 3.90
V 4.3 W 4.32 – 5.22 Y 3.1

Yb 2.60[14] Zn 3.63 – 4.9 Zr 4.05

Physical factors that determine the work function


Due to the complications described in the modelling section below, it is difficult to theoretically predict the
work function with accuracy. Various trends have, however, been identified. The work function tends to be
smaller for metals with an open lattice, and larger for metals in which the atoms are closely packed. It is
somewhat higher on dense crystal faces than open crystal faces, also depending on surface reconstructions
for the given crystal face.

Surface dipole

The work function is not simply dependent on the "internal vacuum level" inside the material (i.e., its
average electrostatic potential), because of the formation of an atomic-scale electric double layer at the
surface.[6] This surface electric dipole gives a jump in the electrostatic potential between the material and
the vacuum.

A variety of factors are responsible for the surface electric dipole. Even with a completely clean surface, the
electrons can spread slightly into the vacuum, leaving behind a slightly positively charged layer of material.
This primarily occurs in metals, where the bound electrons do not encounter a hard wall potential at the
surface but rather a gradual ramping potential due to image charge attraction. The amount of surface dipole
depends on the detailed layout of the atoms at the surface of the material, leading to the variation in work
function for different crystal faces.

Doping and electric field effect (semiconductors)

In a semiconductor, the work function is sensitive to the doping level


at the surface of the semiconductor. Since the doping near the
surface can also be controlled by electric fields, the work function of
a semiconductor is also sensitive to the electric field in the vacuum.

The reason for the dependence is that, typically, the vacuum level
and the conduction band edge retain a fixed spacing independent of
doping. This spacing is called the electron affinity (note that this has
a different meaning than the electron affinity of chemistry); in
silicon for example the electron affinity is 4.05 eV.[15] If the electron
affinity EEA and the surface's band-referenced Fermi level EF-EC are Band diagram of semiconductor-
known, then the work function is given by vacuum interface showing electron
affinity EEA, defined as the difference
between near-surface vacuum
energy Evac, and near-surface
where EC is taken at the surface. conduction band edge EC. Also
shown: Fermi level EF, valence band
From this one might expect that by doping the bulk of the edge EV, work function W.
semiconductor, the work function can be tuned. In reality, however,
the energies of the bands near the surface are often pinned to the
Fermi level, due to the influence of surface states.[16] If there is a large density of surface states, then the
work function of the semiconductor will show a very weak dependence on doping or electric field.[17]

Theoretical models of metal work functions

Theoretical modeling of the work function is difficult, as an accurate model requires a careful treatment of
both electronic many body effects and surface chemistry; both of these topics are already complex in their
own right.

One of the earliest successful models for metal work function trends was the jellium model,[18] which
allowed for oscillations in electronic density nearby the abrupt surface (these are similar to Friedel
oscillations) as well as the tail of electron density extending outside the surface. This model showed why the
density of conduction electrons (as represented by the Wigner–Seitz radius rs) is an important parameter in
determining work function.

The jellium model is only a partial explanation, as its predictions still show significant deviation from real
work functions. More recent models have focused on including more accurate forms of electron exchange
and correlation effects, as well as including the crystal face dependence (this requires the inclusion of the
actual atomic lattice, something that is neglected in the jellium model).[6][19]

Temperature dependence of the electron work function


The electron behavior in metals varies with temperature and is largely reflected by the electron work
function. A recent theoretical model for predicting the temperature dependence of the electron work
function, developed by Reza Rahemi et al.[20] explains the underlying mechanism and predicts this
temperature dependence for various crystal structures via calculable and measurable parameters.

References
1. Kittel, Charles. Introduction to Solid State Physics (7th ed.). Wiley.
2. Herbert Kroemer, "Quasi-Electric Fields and Band Offsets: Teaching Electrons New Tricks (htt
ps://www.nobelprize.org/nobel_prizes/physics/laureates/2000/kroemer-lecture.html)" Nobel
lecture
3. "Barrier Height Correlations and Systematics" (http://academic.brooklyn.cuny.edu/physics/tun
g/Schottky/systematics.htm). academic.brooklyn.cuny.edu. Retrieved 11 April 2018.
4. Behunin, R. O.; Intravaia, F.; Dalvit, D. A. R.; Neto, P. A. M.; Reynaud, S. (2012). "Modeling
electrostatic patch effects in Casimir force measurements". Physical Review A. 85 (1): 012504.
arXiv:1108.1761 (https://arxiv.org/abs/1108.1761). Bibcode:2012PhRvA..85a2504B (https://ui.
adsabs.harvard.edu/abs/2012PhRvA..85a2504B). doi:10.1103/PhysRevA.85.012504 (https://d
oi.org/10.1103%2FPhysRevA.85.012504).
5. Will, C. M. (2011). "Finally, results from Gravity Probe B". Physics. 4 (43): 43. arXiv:1106.1198
(https://arxiv.org/abs/1106.1198). Bibcode:2011PhyOJ...4...43W (https://ui.adsabs.harvard.ed
u/abs/2011PhyOJ...4...43W). doi:10.1103/Physics.4.43 (https://doi.org/10.1103%2FPhysics.4.
43).
6. "Metal surfaces 1a" (http://venables.asu.edu/qmms/PROJ/metal1a.html). venables.asu.edu.
Retrieved 11 April 2018.
7. Thomas Iii, S. W.; Vella, S. J.; Dickey, M. D.; Kaufman, G. K.; Whitesides, G. M. (2009).
"Controlling the Kinetics of Contact Electrification with Patterned Surfaces". Journal of the
American Chemical Society. 131 (25): 8746–8747. CiteSeerX 10.1.1.670.4392 (https://citeseer
x.ist.psu.edu/viewdoc/summary?doi=10.1.1.670.4392). doi:10.1021/ja902862b (https://doi.org/
10.1021%2Fja902862b). PMID 19499916 (https://pubmed.ncbi.nlm.nih.gov/19499916).
8. Helander, M. G.; Greiner, M. T.; Wang, Z. B.; Lu, Z. H. (2010). "Pitfalls in measuring work
function using photoelectron spectroscopy". Applied Surface Science. 256 (8): 2602.
Bibcode:2010ApSS..256.2602H (https://ui.adsabs.harvard.edu/abs/2010ApSS..256.2602H).
doi:10.1016/j.apsusc.2009.11.002 (https://doi.org/10.1016%2Fj.apsusc.2009.11.002).
9. Fernández Garrillo, P. A.; Grévin, B.; Chevalier, N.; Borowik, Ł. (2018). "Calibrated work
function mapping by Kelvin probe force microscopy". Review of Scientific Instruments. 89 (4):
043702. doi:10.1063/1.5007619 (https://doi.org/10.1063%2F1.5007619). PMID 29716375 (http
s://pubmed.ncbi.nlm.nih.gov/29716375).
10. G.L. Kulcinski, "Thermionic Energy Conversion" [1] (http://fti.neep.wisc.edu/neep602/SPRING0
0/lecture9.pdf)
11. "Photoelectron Emission" (http://www.virginia.edu/ep/SurfaceScience/PEE.html).
www.virginia.edu. Retrieved 11 April 2018.
12. CRC Handbook of Chemistry and Physics version 2008, p. 12–114.
13. Dweydari, A. W.; Mee, C. H. B. (1975). "Work function measurements on (100) and (110)
surfaces of silver". Physica Status Solidi A. 27 (1): 223. Bibcode:1975PSSAR..27..223D (http
s://ui.adsabs.harvard.edu/abs/1975PSSAR..27..223D). doi:10.1002/pssa.2210270126 (https://
doi.org/10.1002%2Fpssa.2210270126).
14. Nikolic, M. V.; Radic, S. M.; Minic, V.; Ristic, M. M. (February 1996). "The dependence of the
work function of rare earth metals on their electron structure". Microelectronics Journal. 27 (1):
93–96. doi:10.1016/0026-2692(95)00097-6 (https://doi.org/10.1016%2F0026-2692%2895%29
00097-6). ISSN 0026-2692 (https://www.worldcat.org/issn/0026-2692).
15. Virginia Semiconductor (June 2002). "The General Properties of Si, Ge, SiGe, SiO2 and
Si3N4" (http://www.virginiasemi.com/pdf/generalpropertiessi62002.pdf) (PDF). Retrieved 6 Jan
2019.
16. "Semiconductor Free Surfaces" (http://academic.brooklyn.cuny.edu/physics/tung/Schottky/surf
ace.htm). academic.brooklyn.cuny.edu. Retrieved 11 April 2018.
17. Bardeen, J. (1947). "Surface States and Rectification at a Metal Semi-Conductor Contact".
Physical Review. 71 (10): 717–727. Bibcode:1947PhRv...71..717B (https://ui.adsabs.harvard.e
du/abs/1947PhRv...71..717B). doi:10.1103/PhysRev.71.717 (https://doi.org/10.1103%2FPhysR
ev.71.717).
18. Lang, N.; Kohn, W. (1971). "Theory of Metal Surfaces: Work Function". Physical Review B. 3
(4): 1215. Bibcode:1971PhRvB...3.1215L (https://ui.adsabs.harvard.edu/abs/1971PhRvB...3.1
215L). doi:10.1103/PhysRevB.3.1215 (https://doi.org/10.1103%2FPhysRevB.3.1215).
19. Kiejna, A.; Wojciechowski, K.F. (1996). Metal Surface Electron Physics. Elsevier.
ISBN 9780080536347.
20. Rahemi, Reza; Li, Dongyang (April 2015). "Variation in electron work function with temperature
and its effect on the Young's modulus of metals". Scripta Materialia. 99 (2015): 41–44.
arXiv:1503.08250 (https://arxiv.org/abs/1503.08250). doi:10.1016/j.scriptamat.2014.11.022 (htt
ps://doi.org/10.1016%2Fj.scriptamat.2014.11.022).

Further reading
Ashcroft; Mermin (1976). Solid State Physics. Thomson Learning, Inc.
Goldstein, Newbury; et al. (2003). Scanning Electron Microscopy and X-Ray Microanalysis.
New York: Springer.

For a quick reference to values of work function of the elements:

Michaelson, Herbert B. (1977). "The work function of the elements and its periodicity" (https://s
emanticscholar.org/paper/e8c0a0669e82f16e291ffe0fe19b56d9af668688). J. Appl. Phys. 48
(11): 4729. Bibcode:1977JAP....48.4729M (https://ui.adsabs.harvard.edu/abs/1977JAP....48.47
29M). doi:10.1063/1.323539 (https://doi.org/10.1063%2F1.323539).

External links
Work function of polymeric insulators (Table 2.1) (http://repositories.tdl.org/ttu-ir/bitstream/hand
le/2346/21434/Vela_Russell_Thesis.pdf?sequence=1)
Work function of diamond and doped carbon (http://www3.ntu.edu.sg/home/ecqsun/rtf/SSC-W
F.pdf)
Work functions of common metals (http://www.pulsedpower.net/Info/WorkFunctions.htm)
Work functions of various metals for the photoelectric effect (http://hyperphysics.phy-astr.gsu.e
du/hbase/tables/photoelec.html)
Physics of free surfaces of semiconductors (http://academic.brooklyn.cuny.edu/physics/tung/S
chottky/surface.htm)

*Some of the work functions listed on these sites do not agree!*

Retrieved from "https://en.wikipedia.org/w/index.php?title=Work_function&oldid=943198693"

This page was last edited on 29 February 2020, at 13:30 (UTC).


Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using this
site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

You might also like