Вы находитесь на странице: 1из 15

Chapter 2

Porous silicon

500

450

400
Number of publications

350

300

250

200

150

100

50

0
56 66 76 86 96
Year
Figure 2-1 Number of publications per year regarding porous silicon since it was first
discovered

Although porous silicon was first discovered by Uhlir[1] in 1956, significant interest in this material is
more recent. Figure 2-1[2,3] illustrates this increasing interest by plotting the number of publications per
year on the subject of porous silicon since 1956. The small amount of interest shown in porous silicon
from the mid-1970’s and throughout the 1980’s relates almost exclusively to its use for device isolation
in integrated circuits[4,5]. The more noticeable interest shown from the start of this decade came with
the demonstration by Canham[6] of room temperature photoluminescence from this material. Since this
time the majority of research into porous silicon has focussed on observations of and explanations for
both photoluminescence and electroluminescence from this material, and its potential optoelectronic
applications.
The work described in this thesis uses p-type† porous silicon in the main. This chapter briefly reviews
the fabrication and structure of porous silicon this p-type silicon and possible applications of both n-
and p-type porous silicon.
The demonstration of photoluminescence from porous silicon stimulated research into the use of
porous silicon for optoelectronic circuits and forms around half of the literature currently available on
porous silicon. Several mechanisms have been proposed for this photoluminescence and a brief
review of both the observations and possible mechanisms is given in Appendix A. From a device point
of view it is electroluminescence rather than photoluminescence which is important, and this is
discussed in section 2.2.3 of this chapter.


Porous silicon fabricated from a p+ substrate is abbreviated to p+ porous silicon and likewise for
porous silicon fabricated from p–, n– and n+ substrates. Porous silicon fabricated from p-type
substrates or n-type substrates where the doping level is unspecified will be referred to p-type
porous silicon and n-type porous silicon respectively.

3
Porous Silicon

From pump
Cathode Electrode
Cell
_ +
O-ring
Silicon wafer
Metal contact
Porous silicon
To pump

a) Single-tank Cell b) Double-tank Cell

Figure 2-2 Schematic diagram of two arrangements commonly used to fabricate porous silicon
(adapted from ref [15])

2.1 The fabrication and structure of porous silicon.


The porous silicon described in this thesis was fabricated by the electrochemical anodisation of silicon
in a hydrofluoric acid (HF) based electrolyte. This is the most common method of fabricating porous
silicon though the use of an ammonium fluoride based electrolyte has also been reported[7,8,9]. The
fabrication is usually conducted in the dark to prevent photogenerated currents contributing to the
formation process.
An alternative method for fabrication is by a chemical stain etch[10,11,12] that requires dipping the silicon
substrate in a hydrofluoric acid : nitric acid : water solution for 3-15 minutes. The porous silicon
fabricated using this method is, however, inhomogeneous in both porosity and thickness due to the
fact that hydrogen gas evolved during fabrication remains on the surface of the wafer[10]. For these
reasons the use of this method is rare, although it has been reported that the physical structure of
these layers is similar to those fabricated by the anodisation method[12].
Porous silicon is composed of a silicon skeleton permeated by a network of pores. It is possible to
define the characteristics of a particular porous silicon layer in a number of ways. The methods of
identification include the average pore and silicon branch widths, porosity, pore and branch
orientation, and layer thickness. The specific nature of a layer depends upon the fabrication conditions
used, including the substrate doping and type, the hydrofluoric acid (HF) concentration and the
acidity(pH value) of the electrolyte, the anodisation current density and anodisation time. The
techniques used to assess these properties include various microscopy techniques (pore diameter,
microstructure and layer thickness), gravimetric analysis[13] (porosity and layer thickness‡) and gas
adsorption isotherms[14] (pore diameter).

2.1.1 Fabrication of porous silicon.


Schematic diagrams for the two methods used to form porous silicon using the anodisation method
are shown in Figure 2-2[15]. Both these arrangements are essentially the same; the silicon wafer to be
anodised forms the anode during anodisation and, together with an O-ring, seals the anodisation cell.
In the double tank cell of Figure 2-2b porous silicon is only formed on the substrate surface in contact
with the anodising electrolyte of the left side cell. This is because porous silicon is only formed during
an anodic reaction.
Using either of the two arrangements the porous silicon layers fabricated are usually homogenous in
both porosity and thickness, except within approximately 2mm of the O-ring[15]. For the work presented
here the area to be anodised had a radius of either 2.25cm or 4 inches; the inhomogeneity at the
edges could therefore be avoided when selecting the portion of the porous silicon wafer with which to
work.


The porosity of a layer can be determined by weighing the silicon substrate both before and after
anodisation (m1 and m2 respectively) and again after porous silicon layer has been removed (m3).
The porosity(P) and layer thickness(W) are then calculated by
P = (m1-m2)/(m1-m3) and W = (m1-m3)/(Sd)
where S is the surface area of the wafer which is anodised and d is the density of bulk silicon

4
Porous Silicon

300
19 -3
NA = 1x10 cm
17 -3
NA = 1x10 cm

Growth rate (nm/s)


200

100

0
0 100 200 300 400

Current density (mA/cm2)

Figure 2-3 Growth rate of porous silicon as a function of current density (Data taken from ref.
[16])

Good homogeneity within the porous silicon layers is obtained because the electrical contact to the
silicon substrate is made using the entire back surface of the wafer. This prevents lateral potential
variation across the wafer that would cause changes in the local current density. As discussed in
section 2.1.3 the porosity of a layer is partly determined by current density. Thus maintaining a
constant current density throughout the substrate allows constant porosity porous silicon to be
obtained, providing there are no local variations in the concentration of the hydrofluoric acid in the
electrolyte and that chemical leaching does not occur. The hydrofluoric acid concentration and
chemical leaching are also factors that determine the porosity of a porous silicon layer, as discussed
in section 2.1.3. Despite the nature of the anodic contact, however, hydrogen bubbles evolved during
the anodisation can cling to the surface of the wafer and cause variations in the local potential. For
this reason the electrolyte may be circulated during the anodisation to remove these bubbles. This is
particularly important in the double tank arrangement where the gas evolves during the cathodic
reaction[15] and can cause local variations in potential throughout the substrate.
Figure 2-3 illustrates how the growth rate of porous silicon depends upon the anodisation current
density for starting substrates that are both lightly and heavily doped. It is obvious from the graph that
the thickness of a porous silicon layer after a given time period depends upon the current density at
which it has been anodised. The thickness of a porous silicon layer is therefore uniform providing a
constant current density is maintained whilst any variation in the local current density across the wafer
would cause changes in the thickness of a layer across the wafer[6].
The electrical contact to silicon substrates with low doping levels can be improved by a high dose
back implant. A metal evaporation is also necessary for these substrates when they are to be
anodised in the single tank cell of Figure 2-2a. This is unnecessary for anodisation in the double tank
because the contact is electrolytic and not metallic. An advantage of the double tank arrangement is
that it avoids a potential source of contamination of the porous silicon in any subsequent thermal and
chemical processing[15]. Substrates with high doping levels require neither a back implant nor a metal
evaporation for either arrangement.
The choice of electrolyte is determined by the necessity for the electro-active species required for the
anodisation to be efficiently transported to the porous silicon - silicon interface where the anodisation
process primarily occurs. The hydrophobic[17] and organophillic[15] nature of porous silicon means that
ethanol is a more suitable carrier than water, hence its use in the electrolyte. It can be seen from
Figure 2-4, however, that roughness is still observed at the porous silicon – silicon interface. In the
work described in Chapter 6 it will be seen that similar roughness also occurs at the interface between
porous silicon layers of different porosity. The amplitude of this roughness does decrease with
increasing porosity. In one example[18] increasing the porosity from 65% to 85% caused the amplitude
of the roughness to be reduced from 6nm to 3nm. It is thought that this reduction is a result of the
increasing pore widths associated with increasing porosity allowing easier access of the electrolyte to
the pore tips. Ethanol also acts as a surfactant agent and assists in removing hydrogen bubbles from
the surface(s) of the silicon substrate.

5
Porous Silicon

Porous Silicon
Silicon

Figure 2-4 Interface between porous silicon and silicon substrate. The porous silicon layer has
a thickness of 5µm and a porosity of 56%.

The electrolyte used for the fabrication of the porous silicon described in chapter 6 of this thesis
consisted of a 1:1:2 ratio of HF, water and ethanol. Water forms part of the electrolyte merely because
the hydrofluoric acid was supplied in a 50% aqueous form. Details of the electrolytes used for the
porous silicon described in chapter 5 and section 6.4 are given in those sections.

2.1.2 Pore formation and interfacial roughness


The first models of porous silicon layer formation assumed that the porous layer was formed on the
silicon substrate by a deposition process that involved the reduction of divalent silicon to amorphous
silicon[19,20]. It was later shown that this did not occur and a selective etching process within the silicon
and not a deposition process formed the porous silicon layers[21].
The electrochemical anodisation of silicon will only provide porous silicon if the supply of holes to the
silicon substrate is the rate-limiting step. Anodisation where the diffusion of chemical reactants in the
electrolyte is the limiting step for dissolution causes a surface charge of holes to accumulate. If this
occurs, hills on the surface of the silicon wafer (caused by surface roughness) dissolve faster than
depressions because they are more exposed to the electrolyte. Instead of forming porous silicon, the
silicon surface is then (electro-)polished. The critical current density below which porous silicon will
form is defined as Jps[22].
The exact mechanism for pore formation in a silicon substrate is still uncertain and several
mechanisms have been proposed[20]. Figure 2-5 illustrates the chemical dissolution mechanism
suggested by Lehmann and Gösele[23] that has received some attention[20,24]. Whether this is the
correct dissolution process is unclear[20] but it does explain the hydrogen gas evolved during
anodisation[15], and the need for a hole supply for the dissolution to occur, a generally accepted
requirement[19]. Another attraction of this mechanism is that it explains the fluoride contaminated
hydride passivation layer observed immediately following anodisation. Once exposed to an air
ambient, however, this surface changes to an oxide contaminated surface, the major contaminants
being mainly those elements that occur in the air in gaseous form[25,26].
Lehmann and Gösele expanded their model by suggesting that, providing the current density remains
below Jps, the pore formation is self-limited by the availability of holes within the silicon branches. For
p-type silicon substrates under anodic bias, the limitation of the hole supply may be caused by
quantum confinement.
Figure 2-6, adapted from reference [24], shows the suggested band structure at the silicon - porous
silicon interface. It is initially assumed that the pore walls are depleted of the holes necessary for the
dissolution. If a hole in the silicon substrate has sufficient energy it can penetrate into the silicon
branch causing additional dissolution and a further increase in the band gap. Holes will continue to
penetrate into the branches until the band gap has increased sufficiently to prevent further migration of
holes into the branches, limiting the dissolution to the bulk silicon - pore interface. Increased
dissolution of the branches (increased porosity) is observed as the current density is increased due to
the additional energy the increased current density gives to the hole.

6
Porous Silicon

H H
1. In the absence of electron holes, a hydrogen saturated
silicon surface is virtually free from attack by flouride ions
Si
in the HF based electrolyte. The induced polarisation
between the hydrogen and silicon atoms is low because
Si Si
the electron affinity of hydrogen is about that of silicon.

F
H H
2. If a hole reaches the surface, nucleophillic attack on an
Si Si-H bond by a fluoride ion can occur and a Si-F bond is
formed.
Si Si

H
F H 3. The Si-F bond causes a polarisation effect allowing a
F
second fluorine ion to attack and replace the remaining
Si
hydrogen bond. Two hydrogen atoms can then combine,
injecting an electron into the substrate.
Si Si

H2
4. The polarisation induced by the Si-F bonds reduces the
F F
F F electron density of the remaining Si-Si backbonds making
Si them susceptible to attack by the HF in a manner such
H+ H+ that the remaining silicon surface atoms are bonded to the
hydrogen atoms.
Si Si

F F

+2HF 5. The silicon tretrafluoride molecule reacts with the HF to


Si 2H++SiF62-
form the highly stable SiF6¯ fluoroanion.
F F
H H
The surface returns to its ‘neutral’ state until another hole
is made available.
Si Si

Figure 2-5 Suggested mechanism for the electrochemical dissolution of silicon (after ref [23])

7
Porous Silicon

- +
HF
electrolyte

Si

HF Si
Electrolyte

+
H
Porous silicon

Top left - schematic diagram for the formation of porous


silicon
Porous silicon Silicon

Top right - silicon branch isolated by two pores. Two


+
possible ways for the hole to cross the silicon - porous
silicon interface are shown (broken and dotted arrow).

HF electrolyte Silicon

Bottom - band diagram of the silicon - porous silicon


interface and the two different energy barriers for the hole +

penetrating into the wall (broken arrow) or into the


electolyte (solid arrow)

Figure 2-6 Band diagram of the silicon - porous silicon interface where the radius of a silicon
branch is small enough to exhibit quantum confinement (adapted from ref [24])

90

80
Porosity (%)

70

60

19 -3
50 NA = 1x10 cm
17 -3
NA = 1x10 cm

40
0 100 200 300 400
Current density (mA/cm2)

Figure 2-7 Porosity - current density curve for p– and p+ porous silicon (taken from ref. [16])

2.1.3 Porosity
The factors that determine the porosity of a porous silicon layer include the substrate doping,
anodisation current density and the HF concentration and pH value of the anodising electrolyte. The
relationship between porosity and current density is shown in Figure 2-7 for the porous silicon used for
the fabrication of optoelectronic components described in chapter 6. This graph shows how the
porosity of a layer increases with increasing current density and decreasing substrate doping[22]. The
porosity also increases with decreasing HF concentrations and increasing pH values of the electrolyte.

8
Porous Silicon

c
Figure 2-8 Microstructure of porous silicon - a) Cross section of p– porous silicon (photograph
taken from ref. [20]), b) Cross section of p+ porous silicon (photograph supplied by
Berger [27]), c) Planar view of p+ porous silicon (photograph supplied by Loni [28])

The relationship between porosity and pH values is caused by chemical dissolution of the porous
silicon branches by OH– ions present in the electrolyte. The dissolution rate increases with increasing
levels of the OH– ions in the electrolyte and therefore increasing pH values. This chemical dissolution
continues for as long as the porous silicon remains in contact with the electrolyte, increasing the
porosity of a layer even after the anodisation process is completed. The dissolution rate is partially
dependent upon the surface area available for reaction, a measurement that can be determined by
gas adsorption isotherms[14]. The surface area density, defined as the surface area of the silicon
branches forming the porous silicon, varies from 200m2/cm3 for porous silicon formed from p+ silicon
(ρ = 0.01Ωcm) to 600m2/cm3 for p– silicon (ρ = 1Ωcm)[14,15], though it decreases with increasing
porosity above 50%[15].
The effect of chemical dissolution on a porous silicon skeleton is to reduce the diameter of the
individual silicon branches. At higher porosities, already thin branches may disappear weakening the
remaining structure. Drying such layers can cause cracking or complete disintegration of the branches
due to capillary tensions that occur on the branch surface at the liquid - vapour phase of drying. These
forces can be avoided by supercritical drying. The use of such a technique has enabled layers of up to
97% porosity to be fabricated[29].

2.1.4 Microstructure.
The width and orientation of the branches and pores that form a porous silicon skeleton change as the
level of doping in the original substrate is altered. Figure 2-8 shows SEM and TEM photographs of
both p– and p+ porous silicon. As can be seen from Figure 2-8a porous silicon fabricated from lightly
doped p-type substrates consist of a highly inter-connected network of fine silicon branches. These
branches are typically less than 5nm wide and separated by pores of similar dimensions[30]. Figure 2-

9
Porous Silicon

8b and Figure 2-8c illustrate how porous silicon fabricated from more heavily doped p-type substrates
produces layers with wider pores and silicon branches which run parallel to each other. The widths of
the pores and branches of the p+ porous silicon typically have widths of 10 – 25nm though widths up
to 100nm have been reported[31]. These wider pores explain the lower density of the surface area[31] of
p+ porous silicon reported in the previous section. Figure 2-8c shows how the silicon branches of
these heavily doped layers have many small ‘buds’ that are not constrained to any plane[32]. It has
been noted that the distribution of the pore widths and the average pore width both increase with
increasing current density and decreasing HF concentrations in the electrolyte[14].

2.2 Applications of porous silicon


A variety of applications for porous silicon have emerged since it was first discovered. It has already
been noted that possible applications for porous silicon have been found in dielectric isolation of
integrated circuits and various optoelectronic applications. Another area is that of micromachining[33] in
which the porous silicon acts as a sacrificial layer. These main research areas are briefly reviewed
below.

2.2.1 FIPOS process


During the 1980’s the main focus of porous silicon research lay in its potential application as an
alternative to other developing silicon on insulator (SOI) and silicon on sapphire (SOS) technologies
for device isolation in integrated circuits. These were developed as an alternative to the conventional
methods of isolation by doped channel-stops, suitably biased pn-junctions and thick dielectric layers.
Compared to these conventional methods, the FIPOS (full isolation by porous oxidised silicon), SOI
and SOS methods all offered the advantages of higher speed, lower power consumption, greater
packing density and a reduced number of fabrication steps[34]. The additional attraction of the FIPOS
process was the simplicity of processing and low leakage current[35].

The method is based on the oxidation of porous silicon to isolate pre-defined islands of crystalline
silicon from the bulk silicon substrate. Providing the porosity of the porous silicon was sufficiently high,
the expansion of the silicon branches would fill the pores and not increase the thickness of the layer
causing the silicon islands to warp. The ideal porosity was estimated to be near 56%[36]. Figure 2-10
illustrates the methods used to implement the FIPOS process. This was achieved by either the
preferential anodisation to isolate predefined islands of silicon[5,39] or the epitaxial growth of silicon on
a porous silicon layer that retains the monocrystalline character of the bulk substrate[37].
The original FIPOS method suggested by Imai was the preferential anodisation of a p-type bulk silicon
substrate over implanted islands of n– silicon[5]. The current density - voltage characteristics of
different substrates vary, as shown in Figure 2-10[15]. Limiting the potential during anodisation
facilitates the preferential anodisation of p+ substrate over p– substrates, n+ substrates over n– or p-
type substrates and p-type substrates over n– substrates.

The original structures that were fabricated displayed the advantages of SOI and FIPOS devices
already mentioned. Unfortunately devices fabricated in this manner required thick porous silicon layers
in order to fabricate silicon islands of moderate widths. This was caused by the rate of pore formation
being uniform in all directions giving rise to a layer of at least half the island width[5]. The layer
thickness was reduced by either ion implantation or epitaxial growth to define the layers that would
form the porous silicon[39]. The silicon islands were then formed by the epitaxial growth and etching of
an additional silicon layer. Another problem was that of wafer warpage that was resolved by
implementing the FIPOS method in an n/n+/n structure[38]. This also removed the remaining problems
of non-uniform porous silicon layers, and the thin wisp of silicon that remained under the silicon island
where the anodisation fronts met[39].

10
Silicon substrate
Preferential anodisation Growth of silicon on porous silicon

Ion implantation to define Epitaxial growth to define


silicon layer to be porosified silicon layer to be porosified

Epitaxial growth of silicon layer Anodisation to form porous silicon

Implantation to define silicon islands Etch to define islands Epitaxial growth of silicon layer

11
Anodisation to form porous silicon Anodisation to form porous silicon Etch to define islands

Oxidation of porous silicon Oxidation of porous silicon Oxidation of porous silicon

Figure 2-9 Methods of implementing FIPOS technique


Device processing

Si substrate Porous silicon Epitaxial, or implanted


silicon, to be anodised
Silicon islands Oxidised porous silicon
Porous Silicon
Porous Silicon

140

Current density (mA/cm2)


120 p+
p-
100

80

60 n+

40

20
n-
0
-0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4

Potential (volts)

Figure 2-10 Current density - potential graph for p+, p–, n+ and n– silicon substrates (taken from
ref [15])

2.2.2 Micromachining
The techniques employed for dielectric isolation using porous silicon can also be used for
micromachining applications. Micromachining is used to fabricate small-scale mechanical devices that
are integrated with conventional microelectronics. Examples of micromachined devices include
motors, cantilevers and a wide variety of sensors that are designed to sense temperature, IR and UV
radiation, fluid flow or gas flow. Many of these structures are fabricated on free-standing membranes,
structures that can be easily fabricated using porous silicon.
Conventional micromachining methods to form free-standing membranes include anisotropically
etching[42] the rear of a substrate. This is a well established technology whose main drawback it the
need for double sided lithography. The use of double sided lithography is avoided when surface
micromachining technology[42] is used. Instead an easily etchable sacrificial layer is deposited on to
the substrate surface followed by a second layer that will form the membrane. A second layer is then
deposited that, after defining the micromachined device and removing the sacrificial layer, forms the
free-standing membrane. The drawback of this method is the limited distance that can be obtained
between the membrane and substrate. The limiting factor is defined by the maximum thickness
obtainable for the sacrificial layer and is typically limited to several microns. Although this distance
may be sufficient for applications such as micromotors, applications such as sensing often require
thickness for the sacrificial layer to be several tens of microns to reduce heat transfer to the substrate.
Porous silicon provides a good alternative to both methods described above. It is formed without the
use of double-sided lithography and can be fabricated to thicknesses of several tens of microns. The
fabricated layers, regardless of thickness, are then easily removed using a weak potassium hydroxide
solution or even photoresist developing solution. Additionally, unlike anisotropic etching the geometry
of the porous silicon layers is not limited to certain planes and so they can be formed locally on a
wafer with controlled undercutting.
A variety of devices have been demonstrated using this fabrication method including cantilever
beams[40], bolometers for thermal measurements[41], flow channels and wires[42]. Bridges[42] have been
shown to be stable under heat treatment and gas flow though not, unfortunately, to being dropped on
the floor! It has recently been suggested[41] that the porous silicon may not need to be removed in all
applications as was originally demonstrated over a decade ago for flow sensors[43]. The low thermal
conductivity of p- porous silicon means that the porous silicon may provide sufficient thermal isolation
from the substrate. This removes the need for removing the porous silicon to provide an air gap,
providing an almost identical thermal isolation function whilst improving the mechanical robustness of
the device.

12
Porous Silicon

2.2.3 Light Emitting Diodes


The possibility of electroluminescent devices fabricated from porous silicon was soon realised soon
after the demonstration of photoluminescence from porous silicon. Electroluminescent devices usually
take the form of either light emitting diodes (LEDs) or injection lasers. Though it is not certain whether
a laser will ever be fabricated from porous silicon, LEDs emitting in the red part of the spectrum have
been successfully demonstrated[44].
Electroluminescence from porous silicon was first reported in 1991[45] and was observed during anodic
oxidation that is using a liquid contact which is not practical for device applications. The first solid state
LED was reported a short time later[46] and used a Schottky-type junction between gold and n-type
porous silicon to generate red light. Unfortunately the LED emitted light with the same intensity in both
the forward and reverse bias, had a high (200V) threshold voltage and operated with a low efficiency.
Several groups demonstrated LEDs with a rectifying behaviour in 1992 that using a variety of contacts
(gold[47], ITO[48] and n-type silicon carbide[49]). The emission spectra of each of the devices were
comparable to that of the LED of ref [46] but operated with a forward bias voltage of less than 10V and
an efficiency less than 0.001%[44].
The critical characteristics concerning an LED are those of emission wavelength selection, external
efficiency, threshold voltage, carrier lifetime, width of the emission spectra, and device lifetime.
Wavelength selection throughout the visible wavelengths has been demonstrated using either different
metallic contacts[50] or by varying the fabrication conditions[48]. The external efficiency of these LEDs
has slowly been increased with the best efficiencies reported to be in excess of 0.1% (up to 0.18%) for
CW operation[27] and 0.2% for pulsed operation[51]. The operational lifetime of the LEDs has slowly
been increased and experiments to calculate the lifetime of an integrated seven segment display of
low quantum efficiency were stopped after several hundred hours[52]. Similar results have also been
observed in a separate demonstrator device[53]. Unfortunately the main disadvantages of porous
silicon LEDs are fundamental, those being the long carrier lifetime restricting the modulation rate to
between 100kHz and 10MHz with the higher modulation rates only obtainable through a trade-off with
the quantum efficiency of the device[54].

2.2.4 Photodetectors and sensors.


To complement light emission from porous silicon a variety of MSM[55] and p-n[56] photodetectors
utilising porous silicon have also been demonstrated. These detectors have been reported with
response time as low as 2ns[57] and sensitivities in excess of 0.7A/W at 500nm[58]. The quantum
efficiency of these devices has been reported[57] as high as 97% whilst the noise equivalent power has
been reported[59] to be as low as 6 x 10-13 W Hz1/2.
Porous silicon has also been investigated as a possible AR coating in solar cells[60]. Superlattices
formed using porous silicon have also been shown to act as filters allowing for the wavelength
selection of light[61]. These structures have been shown to make photodetectors colour sensitive when
used to replace the basic porous silicon layer[62]. Though as-anodised porous silicon can be used for
colour sensitivity in the red region of the electromagnetic spectrum, blue-sensitive filters are obtained
through oxidation of the porous silicon.
The use of porous silicon/silicon substrate junctions has also been used for sensing applications. The
structures are fabricated by forming a porous silicon layer on a silicon substrate and contacting both
the porous silicon surface and the rear face of the substrate. Using these structures a gas sensor
based upon the changing current due to the dipole moment of the gas[63], and a humidity sensor based
upon the changing current with humidity[64] have both been demonstrated. Additionally applications for
porous silicon in biosensing have also been demonstrated[65], using penicillin as an example. Coating
the large surface area of the porous silicon with a penicillin sensitive enzyme causes the capacitance-
voltage curve of the junction to shift with changing concentrations of penicillin.

13
Porous Silicon

2.3 References
1
A Uhlir, Electrolytic shaping of germanium and silicon, The Bell System Technical Journal, Vol
35, pp 333-347 (1956)
2
Bath Information and Data Service, ISI and Compendex 1 databases
3
Science Citation Indexes, 1956 - 1981
4
Y Wanatabe, Y Arita, T Yokoyama and Y Igarashi, Formation and properties of porous silicon
and its applications, J Electrochem Soc: Solid-State Science and Technology, Vol 122, No 10, pp
1351-1355 (1975)
5
K Imai, A new dielectric isolation method using porous silicon, Solid-State Electronics, Vol 24,
pp 159-164 (1981)
6
LT Canham, Silicon quantum wire array fabricated by electrochemical and chemical
dissolution of wafers, Appl Phys Lett, Vol 57, No 10, pp 1046-1048 (1990)
7
MM Koltun, Nature of film on surface of silicon photocell during anodic etching, R Journal
Phys Chem, Vol 38, pp 381 (1964)
8
GM O’Halloran, M Kuhl, PM Sarro, PTJ Gennissen and PJ French, New etchant for the
fabrications of porous silicon, Meetings Abstracts, Spring meeting of the Electrochemical
Society, Vol 96-1, pp 414 (1996)
9
Th Dittrich, S Rauscher, V Yu Timoshenko, J Rappich, I Sieber, H Flietner and HJ Lewerenz,
Ultrathin luminescent nanoporous silicon on n-Si: pH dependent preparation in aqueous
NH4F solutions, Appl Phys Lett, Vol 67, No 8, pp 1134-1136 (1995)
10
KH Beckmann, Investigation of the chemical properties of stain films on silicon by means of
infrared spectroscopy, Surface Science, Vol 3, pp 324-332 (1965)
11
T Lin, ME Sixta, JN Cox and ME Delaney, Optical studies of porous silicon, Mat Res Soc Symp
Proc, Vol 298, pp 379-384 (1993)
12
S Shih, KH Jung, TY Hsieh, J Sarathy, C Tsai, K-H Li, JC Campbell and DL Kwong,
Photoluminescence and structure of chemically etched Si, Mat Res Soc Symp Proc, Vol 256,
pp 27-30 (1992)
13
D Brumhead, LT Canham, DM Seekings and PJ Tufton, Gravimetric analysis of pore nucleation
and propagation in anodised silicon, Electrochimica Acta, Vol 38, No 2/3, pp 191-197 (1993)
14
R Herino, G Bomchil, K Barla and C Bertrand, Porosity and pore size distribution of porous
silicon layers, J Electrochem Soc, Vol 134, No 8, pp 1994 -2000 (1987)
15
A Halimaoui, Porous silicon: material processing, properties and applications, in JC Vial and J
Derrien (editors), Porous silicon science and technology, Springer-Verlag (1995)
16
MG Berger, Poröses Silicium für die Mikrooptik: Herstellung, Mikrostruktur und optische
Eigenschaften von Einzelschichten und Schichtsystemen, PhD Thesis, Forschungszentrum
Jühlich GmbH (1996)
17
LT Canham, Laser dye impregnation of oxidized porous silicon on silicon wafers, Appl Phys
Lett, Vol 63, No 3, pp 337-339 (1993)
18
I Berbezier and A Halimaoui, A microstructural study of porous silicon, J Appl Phy, Vol 74, No
9, pp 5421-5425 (1993)
19
R Memming and G Schwandt, Anodic dissolution of silicon in hydrofluoric acid solutions,
Surface Science, Vol 4, pp 109-124 (1966)
20
RL Smith and SD Collins, Porous silicon formation mechanisms, J Appl Phys, Vol 71, No 8, pp
R1-R22 (1992)
21
MJJ Theunissen, Etch channel formation during anodic dissolution of n-type silicon in
aqueous hydrofluoric acid, J Electrochem Soc, Vol 119, No 11, pp 351 (1972)

14
Porous Silicon

22
V Lehmann, The physics of macropore formation in low doped n-type silicon, J Electrochem
Soc, Vol 140, No 10, pp 2836-2843 (1993)
23
V Lehmann and U Gösele, Porous silicon formation: a quantum wire effect, App Phys Lett, Vol
58, No 8, pp 856-858 (1991)
24
V Lehmann, B Jobst, T Muschik, A Kux and V Petrova-Koch, Correlation between optical properties
and crystallite size in porous silicon, Jpn J Appl Phys, Vol 32, Pt 1, No 5A, pp 2095-2099 (1993)
25
LT Canham, MR Houlton, WY Leong, C Pickering and JM Keen, Atmospheric impregnation of
porous silicon at room temperature, J Appl Phys, Vol 70, No 1, pp 422-431 (1991)
26
LT Canham and GW Blackmore, SIMS analysis of the contamination of porous silicon by
ambient air, Mat Res Soc Symp Proc, Vol 256, pp 63-68 (1992)
27
A Loni, AJ Simons, TI Cox, PDJ Calcott and LT Canham, Electroluminescent porous silicon
device with an external quantum efficiency greater than 0.1% under CW operation,
Electronics Letters, Vol 31 No 15, pp 1288-1289 (1995)
28
A Loni, Defence Evaluation and Research Agency, St Andrews Road, Malvern, Worcs
29
LT Canham, AG Cullis, C Pickering, OD Dosser, TI Cox and TP Lynch, Luminescent anodized
silicon aerocrystal networks prepared by supercritical drying, Nature, Vol 368, pp 133-135
(1994)
30
PA Badoz, D Bensahel, G Bomchil, F Ferrieu, A Halimaoui, P Perret, JI Regolini, I Sagnes and G
Vincent, Characterization of porous silicon: structural, optical and electrical properties, Mat
Res Soc Symp Proc, Vol 283, pp 97-108 (1993)
31
PC Searson, JM Macaulay and FM Ross, Pore morphology and the mechanism of pore
formation in n-type silicon, J Appl Phys, Vol 72, No 1, pp 253-258 (1992)
32
MIJ Beale, NG Chew, MJ Uren, AG Cullis and HD Benjamin, Microstructure and formation
mechanism of porous silicon, Appl Phys Lett, Vol 46, No 1, pp. 86-88 (1985)
33
XZ Tu, Fabrication of silicon microstructures based on selective formation and etching of
porous silicon, J Electrochem Soc: Solid-State Science and Technology, Vol 135, No 8, pp 2105-
2107 (1987)
34
SL Partridge, Silicon-on-insulator technology, IEE Proceedings - E, Vol 133, No 3, pp 107-116
(1986)
35
NJ Thomas, JR Davis, JM Keen, JG Castledine, D Brumhead, M Goulding, J Alderman, JPG Farr,
LG Earwaker, J L’Ecuyer, IM Stirland and JM Cole, High-performance thin-film silicon-on-
insulator CMOS transistors in porous anodized silicon, IEEE Electron device letters, Vol 10, No
3, pp 129-131 (1989)
36
K Barla, G Bomchil, R Herino and A Monroy, SOI technology using buried layers of oxidized
porous Si, IEEE Circuits and Devices Magazine, pp 11-15 (November 1987)
37
C Oules, A Halimaoui, JL Regolini, A Perio and G Bomchil, Silicon on insulator structures
obtained by epitaxial growth of silicon over porous silicon, J Electrochem Soc, Vol 139, No 12,
pp 3595-3599 (1992)
38
K Barla, G Bomchil, R Herino, A Monroy and Y Gris, Characteristics of SOI CMOS circuits made
in N/N+/N oxidised porous silicon structures, Electronics Letters, Vol 22, No 24, pp 1291-1293
(1986)
39
SS Tsao, Porous silicon techniques for SOI structures, IEEE Circuits and Devices Magazines,
pp 1-7 (November 1987)
40
R Mlcak, HL Tuller, P Greiff, J Sohn and L Niles, Photoassisted electrochemical
micromachining of silicon in HF electrolytes, Sensors and Actuators A, Vol 40, pp 49-55 (1994)
41
W Lang, P Steiner, U Schaber and A Richter, A thin film bolometer using porous silicon
technology, Sensors and Actuators A, Vol 43, pp 185-187 (1994)
42
W Lang, P Steiner, A Richter, K Marusczyk, G Weimann and H Sandmaier, Applications of
porous silicon as a sacrificial layer, Sensors and Actuators A, Vol 43, pp 239-242 (1994)

15
Porous Silicon

43
O Tabata, Fast-response silicon flow sensor with an on-chop fluid temperature sensing
element, IEEE Transactions on Electron Devices, Vol Ed-33, No 3, pp 361-365 (1986)
44
RT Collins, PM Fauchet and MA Tischler, Porous silicon: from luminescence to LEDs, Physics
Today, Vol 50, No 1, pp 24-31 (1997)
45
A Halimaoui, C Oules, G Bomchil, A Bsiesy, F Gaspard, R Herino, M Ligeon and F Muller,
Electroluminescence in the visible range during anodic-oxidation of porous silicon films,
Appl Phys Lett, Vol 59, No 3, pp 304-306 (1991)
46
A Richter, P Steiner, F Kozlowski and W Lang, Current induced light-emission from a porous
silicon device, IEEE Electron Device Letters, Vol 12, No 12, pp 691-692 (1991)
47
N Koshida and H Koyama, Visible electroluminescence from porous silicon, Appl Phys Lett, Vol
60, No 3, pp 347-349 (1992)
48
NM Kalkhoran, F Namavar and HP Maruska, NP heterojunction porous silicon light-emitting
diode, Mat Res Soc Symp Proc, Vol 256, pp 89-94 (1992)
or
F Namavar, HP Maruska and NM Kalkhoran, Visible electroluminescence from porous silicon
np heterojunction diodes, Appl Phys Lett, Vol 60, No 20, pp 2514-2516 (1992)
49
T Futagi, N Ohtani, M Karsuno, K Kawamura, Y, Ohta, H Mimura and K Kitamura, An amorphous
SiC thin-film visible light-emitting diode with a µC-SiC-H electron injector, Journal of Non-
Crystalline Solids, Vol 137, Pt 2, pp 1271-1274 (1991)
50
P Steiner, A Wiedenhofer, F Kozlowski and W Lang, Influence of different metallic contacts on
porous silicon electroluminescence, Thin Solid Films, Vol 276, pp 159-163 (1996)
51
N Lalic and J Linnros, A porous silicon light-emitting diode with a high quantum efficiency
during pulsed operation, Thin Solid Films, Vol 276, pp 155-158 (1996)
52
Erik Kreifeldt, Optics and electronics get along in silicon, Optics and Photonics News, Vol 8, No
2, pp 9 (1997)
53
S Lazarouk, P Jaguiro, S Katsouba, G Masini, S La Monica, G Asiello and A Ferrari, Stable
electroluminescence from reverse biased n-type porous silicon – aluminum Schottky
junction device, Appl Phys Lett, Vol 68, No 15, pp 2108-2110 (1996)
54
A Loni, DERA Malvern, Private Communication.
55
LZ Yu and CR Wie, Fabrication of MSM photoconductor on porous Si using micromachined
silicon mask, Electronic Letters, Vol 28, No 10, pp 911-913 (1992)
56
C Tsai, KH Li and JC Campbell, Rapid-thermal-oxidised porous Si photodetectors, Electronics
Letters, Vol 29, No 1, pp 134-136 (1993)
57
JP Zheng, KL Jiao, WP Shen, WA Anderson and HS Kwok, Highly sensitive photodetector using
porous silicon, Appl Phys Lett Vol 61, No 4, pp 459-461 (1992)
58
C Tsai, KH Li, and JC Campbell, Rapid-thermal-oxidized porous Si photodetectors, Electronics
Letters, Vol 29, No 1, pp 134-136 (1993)
59
JP Zheng, KL Jiao, WP Shen, WA Anderson and HS Kwok, Unity quantum efficiency photodiode
using porous silicon film, Mat Res Soc Symp Proc, Vol 283, pp 371-375 (1993)
60
ES Kolesar Jr, VM Bright and DM Sowders, Mid-infrared (2.5 ≤ λ ≤ 12.5µm) optical absorption
enhancement of textured silicon surfaces coated with an antireflective thin film, Thin Solid
Films, Vol 270, pp 10-15 (1995)
61
MG Berger, R Arens-Fischer, M Thönissen, M Krüger, S Billat, H Lüth, S Hilbrich, W Theiss, and P
Grosse, Dielectric filters made of PS: advanced performance by oxidation and new layer
structures, Thin Solid Films, Vol 297, No 1-2, pp 237-240 (1997)
62
M Krüger, M Marso, MG Berger, M Thönissen, S Billat, R Loo, W Reetz and H Lüth, S Hilbrich, R
Arens-Fischer and P Grosse, Color-sensitive photodetector based on porous silicon
superlattices, Thin Solid Films, Vol 297, No 1-2, pp 241-244 (1997)

16
Porous Silicon

63
I Schechter, M Ben-Chorin and A Kux, Gas sensing properties of porous silicon, Anal Chem, Vol
67, pp 3727-3732 (1995)
64
M Yamana, N Kashiwazaki, A Kinoshita, T Nakano, M Yamamoto and CW Walton, Porous silicon
oxide layer formation by the electrochemical treatment of a porous silicon layer, J
Electrochem Soc, Vol 137, No 9, pp 2925-2927 (1990)
65
M Thust, MJ Schöning, S Frohnhoff, R Arens-Fischer, P Kordos and H Lüth, Porous silicon as a
substrate material for potentiometric biosensors, Meas Sci Technol, Vol 7, pp 26-29 (1996)

17

Вам также может понравиться