You are on page 1of 26

Electrochemical Etching

Electrochemical etching was first reported by Tommasino (1970), which is an alter-


native way to develop tracks and is used widely in radiation dosimetry.

From: Handbook of Radioactivity Analysis (Third Edition), 2012

Related terms:

Nanowires, Wire, Gallium Arsenide, Oxide, Pore Size, Nanoparticles, Etching-


, Porous Silicon, Silicon Wafer

View all Topics

Learn more about Electrochemical Etching

Solid-State Nuclear Track Detectors


Shi-Lun Guo, ... S.A. Durrani, in Handbook of Radioactivity Analysis (Third Edition),
2012

B Electrochemical Etching (ECE)


Electrochemical etching was first suggested by Tommasino (1970). A plastic detector
foil with tracks of particles is used to divide a cell containing a suitable etchant, such
as NaOH solution. A stainless steel electrode is placed in each half of the cell. A
high-frequency oscillating voltage is connected to the two electrodes. The voltage
is typically equivalent to ~30−50 kV cm−1 field strength. The frequency is ranged
from several kHz to several tens of kHz. A 50 Hz 220 V sinusoidal supply can also
work. Under the joint reactions of etchant and electric voltage, the tracks start to be
etched. The forming track tip has a strong field, which makes the foil break down
near the tip as shown in Fig. 4.14. The breakdown looks like a tree. The tree is a
consequence of the charged particle track. Figure 4.15 shows microphotographs of
252Cf fission-fragment tracks in CR-39 sheets of 485 μm in thickness.
FIGURE 4.14. “Treeing” phenomenon of electrochemical etching (ECE) of charged
particle tracks in plastic foil. The sharp tips of etched cones have very strong field,
which makes the foil break down, forming tree-like tracks. The breakdown spreads
up to several hundred micrometers in width, which is easy to observe with a
low-magnification optical microscope or even by the naked eye.

(From Al-Najjar et al., 1979; printed with permission from Elsevier © 1979)
FIGURE 4.15. Microphotographs of electrochemically etched tracks of 252Cf fission
fragments recorded in CR-39 plates. The thickness of the CR-39 plate is 485 μm.
The track densities are (1) 1.5 × 103, (2) 3.0 × 103, (3) 1.2 × 104, (4) 2.4 × 104, (5)
3.6 × 104, and (6) 1.2 × 105 cm−2, respectively. The etching condition was 30 wt%
KOH, 60 °C, 90 min. The electrical field condition was 15 kVcm−1 and 40 kHz. The
track densities after ECE are the same as CE. The track size changes diminish as the
density increases.

(From Al-Najjar et al., 1979; printed with permission from Elsevier © 1979)

Many researchers have studied the “treeing” phenomenon (Tommasino et al., 1981;
Sohrabi, 1981; Sohrabi and Mahdi, 1993; Li, 1993, 2008a, 2008b).

The advantage of electrochemical etching is that the size of the etched tracks is very
large up to several hundreds of micrometers, which are easy to scan with an optical
microscope or track image analyzer, or even can be seen with the naked eye. The
disadvantage is that the track density must be lower than ~105 cm−2.

The applications of electrochemical etching are mainly in the detection of fission


fragments, -particles, and neutron-recoil tracks. The main fields of application are
neutron dosimetry (Tanner et al., 2005; Dhairyawan et al., 2003; Zainali and Afkar,
2005; Hankins et al., 1989) and Rn monitoring (Zainali and Afkar, 2005).

> Read full chapter

Modern techniques of surface geome-


try modification for the implants based
on titanium and its alloys used for im-
provement of the biomedical character-
istics
E.G. Zemtsova, ... V.M. Smirnov, in Titanium in Medical and Dental Applications,
2018

2.2.2.3 Anodization
Electrochemical etching can take place without complete dissolution of anode ma-
terial in the etchant. In this case, an oxide layer grows on the surface. This is anodiza-
tion that is widely used for the modification of titanium implant surfaces [45–47].
In Ref. [46], microphotographs of the Ti alloy surface are given after anodization in
aqueous sulfuric acid (Fig. 2.2.13).

Fig. 2.2.13. Electronic microphotograph of an anodized titanium alloy surface [46].


Reprinted from N. Masahashi, Y. Mizukoshi, S. Semboshi, K. Ohmura, S. Hanada,
Photo-induced properties of anodic oxide films on Ti-6Al-4V, Thin Solid Films 520
(2012) 4956–4964. Copyright (2012), with permission from Elsevier.

So, one can observe the presence of a nanometer-sized texture on the sample
surface. At lower magnification, a weakly developed micron-sized texture can be
revealed (Fig. 2.2.14).

Fig. 2.2.14. Electronic microphotograph of anodized titanium alloy surface [46] at


low magnification.

Reprinted from N. Masahashi, Y. Mizukoshi, S. Semboshi, K. Ohmura, S. Hanada,


Photo-induced properties of anodic oxide films on Ti-6Al-4V, Thin Solid Films 520
(2012) 4956–4964. Copyright (2012), with permission from Elsevier.

Therefore, this technique is also prospective for implant development. This method
looks even more attractive when considering the works describing the formation of
the structures, anisotropic along the normal to the surface. Under certain conditions
localization of the charge occurs during the anodization on the areas with a high
etching rate. As a result, these areas repelled and formed a distorted hexagonal pore
network [48] (Fig. 2.2.15).

Fig. 2.2.15. Electron microphotograph of anodized porous (tubular) titanium surface


(A). Electron microphotograph of a chip of the surface (B) [48].

Reprinted from M. Goudarzi, F. Batmanghelich, A. Afshar, A. Dolati, G. Mortazavi,


Development of electrophoretically deposited hydroxyapatite coatings on anodized
nanotubular TiO2 structures: Corrosion and sintering temperature, Appl. Surf. Sci.
301 (2014) 250–257. Copyright (2014), with permission from Elsevier.

Such a structure has good cellular response. So, in Ref. [49] the same nanotubes
were obtained on the originally etched surface of titanium. These structures had no
significant effect on osteoblast proliferation; however, a huge increase of alkaline
phosphatase activity was discovered, indicating a significant promotion of the
differentiation process. Subsequent studies showed that this effect on the surfaces
after the classical anodizing is less pronounced [50].

Thus, it is concluded that the anodizing is suitable for creating bioactive coatings.
The use of pile-like surfaces is especially efficient.

> Read full chapter

Nanosilicon
In Fundamentals and Applications of Nano Silicon in Plasmonics and Fullerines,
2018

9.11.1 Synthesis of Porous Silicon


This electro-chemical etching of silicon was first studied by Arthur Uhlir Jr. and
Ingeborg Uhlir at the Bell Labs in the 1950s [51]. The main purpose of their studies
is polishing and shaping silicon surface. But occasionally underlower current, it
was found that the process does not polish rather it produces an ugly looking
thick black/brow/red film. A schematic of the electrochemical etching process for
the synthesis of porous silicon is presented in Fig. 9.43A. The backside of a silicon
wafer is contacted with a metal (copper for example). It is immersed in a solution
of hydrofluoric acid (HF) in water (aqueous solution). The copper film is biased
positively against an immersed platinum electrode. A voltages applied between
the wafer backside contact and the platinum electrode, setting up a current. The
current causes etching (dissolution) of the silicon wafer. The etching results in the
progressive formation of a sponge-like structure with disordered pores. When the
silicon wafer is highly doped (large conductivity) and the current is high, the current
skims the surface resulting in very high current density. Under high-current density,
the surface becomes mechanically weak and gets eaten away the top layer of silicon,
which essentially electro polishes the surface. Another configuration which isolates
the back metal from the etching solution is shown in Fig. 9.43B.
Figure 9.43. Synthesis of porous silicon.(A) Schematic of the electro-
chemical etching process for the synthesis of porous silicon. (From
http://cibaprocess.nus.edu.sg/process/index.html) (B) Another configuration which
isolates the back metal from the etching solution.

(From Elisabet Xifré-Pérez, Josep Ferré-Borrull, Josep Pallarés, Lluís F. Marsal, ,


Methods, Properties and Applications of Porous Silicon Electrochem. Eng. Nanopor.
Mater. 37-63)

Leigh Canham

Leigh Canham is a British scientist who has pioneered the optoelectronic and
biomedical applications of porous silicon. Leigh Canham graduated from University
College London in 1979 with a BSc in Physics and completed his PhD at King’s
College London in 1983. His early work in this area took place at the Royal Signals
and Radar Establishment in Malvern, Worcestershire. Canham and his colleagues
showed that electrochemically etched silicon could be made porous. This porous
material could emit visible light when a current was passed through it (electrolu-
minescence). Later the group demonstrated the biocompatibility of porous silicon.
Canham now works as Chief Scientific Officer of psiMedica (part of pSiVida).
According to the pSiVida web site, Canham is the most cited author on porous
silicon. In a study of most cited physicists up to 1997 Canham ranked at 771.

Nearly 30 years later it was found that in some cases the etched silicon displayed
photoluminescence under UV light [5]. Those studies were conducted by Leigh
Canham at the Defense Agency in the UK. Close observations concluded that etching
results in the progressive formation of a sponge-like structure with disordered
pores (porous silicon layer). This caused the top layer of the wafer to exhibit red
luminescence when exposed to UV light. It was then conjectured that the film
consists of interconnected nanostructures, and quantum confinement effects was
invoked where photoexcited electrons and holes, such as a particle in a box model,
are behind the enhancement and blue-shift to the visible. The process is also called
now anodic etching and the cell is called anodic cell.

There are several features worth noting with regard to the material choice, reaction
equations and mechanism, electrode configuration, and material type, as well as
other configurations:

(1)
Dopings: Formation of fluorescent porous silicon could be obtained in all
dopings from n + , n−, p−, and p+ if the etching conditions produced high
density porosity.
(2) Impurities: The fluorescence was not observed in material with porous struc-
ture too large, nor was it observed in material made amorphous by defect con-
centration with ion implantation before anodization. Therefore if fluorescence
is desired, a porous silicon fabrication technique can be largely judged by the
crystallinity and size scale of the resulting structure.
(3) Reaction equation: for the etching process seems to be:

The divalent silicon species formed on the surface is then oxidized to tetravalent
silicon ions in the form of (solution) through a redox reaction in solution:

with the partial reactions being:

1. External current mechanism: Holes come to the Si surface through the bulk
where they weaken the Si–H passivating bonds allowing the fluorine ions
to break them. The Si–F bonds weaken the Si–Si backbonds allowing them
to be broken as well. Then a SiF4 molecule leaves the bulk which is now H
passivated again, and the cycle repeats, which results in dissolution of silicon
which is turned into porous silicon. In n-type silicon, holes need to be created
by illumination of the sample, where electron–hole pairs are broken apart by
the light to release more liberated holes. However, in p-type silicon, holes are
majority charge carriers, which are easily driven to reach the wafer surface to
participate in the reaction. Thus, p-type silicon is the preferred material for the
process.
2. Electrodes: The backside of the p-type silicon wafer is usually connected to a
metal plate or wire made of copper for example to achieve an Ohmic contact.
The cathode is typically a platinum grid which is inert in HF. The back metal
electrode may introduce metal impurities in the electrolyte which quenches
luminescence. Recently, inert diamond cathodes are used instead of metal
electrodes to avoid metallic impurities in the electrolyte; inert diamond anodes
provide improved electrical back plate contact to the silicon wafers. Corrosion
of the anode is produced by running electric current through the cell.
3. Both DC and pulsed current are used but the pulsed case allows formation of
silicon wafers thicker than 50 μm [4].
4. Solvent: Ethanol is usually added to the HF electrolyte as a surfactant to prevent
bubbles accumulating on the silicon surface. Usually, ethanol or methanol
is added to the HF electrolyte to prevent bubbles from accumulating on the
silicon surface and to facilitate infiltration of HF solution within the pores, as
well as to improve the uniformity of pores and thickness. It is to be noted that
during the formation of porous silicon, hydrogen gas forms. In the absence of
ethanol or methanol hydrogen bubbles stick to the surface and induce lateral
and in-depth inhomogeneity.

9.11.1.1 Stain Etching


Another method to obtain porous silicon is through what has been called stain-etch-
ing with HF acid, nitric acid, and water [54]. Porous silicon formation by stain-etch-
ing is particularly attractive because of its simplicity and the presence of readily
available corrosive reagents; namely nitric acid (HNO3) and hydrogen fluoride (HF).
Furthermore, stain-etching is useful if one needs to produce a very thin porous
Si films. R.J. Archer revealed that it is possible to create stain films as thin as
25 Å through stain-etching with HF–HNO3 solution. Finally, porous silicon can be
synthesized chemically from silicon tetrachloride, using self-forming salt byproducts
as templates for pore formation. The salt templates are later removed with water.

> Read full chapter

SiC as a Biocompatible Marker for Cell


Labeling
J.-M. Bluet, ... V. Lysenko, in Silicon Carbide Biotechnology, 2012

11.4.4 Conclusion
SiC QDs realized by either electrochemical etching or laser ablation of the initial bulk
material show intense luminescence coming from both the quantum confinement
effect and surface states. It has been shown that quantum confinement can be
evidenced in 6H-SiC NPs by dispersion in a high dielectric constant solvent and size
selection by centrifugation to filter the QD size. In the case of 3C-SiC nanoparticles,
the role of surface states on the PL spectra has been evidenced. These surface states
are most probably due to the presence of C O double bonds. From a practical point
of view, the strong luminescence from QDs made from SiC, known as a chemically
inert material, opens the way to marking applications in different domains. One
of these applications, cell labeling, using these SiC QDs, is described in the next
section.

> Read full chapter


Optical properties of porous silicon ma-
terials for biomedical applications
V. Torres-Costa, R.J. Martín-Palma, in Porous Silicon for Biomedical Applications,
2014

8.6 Multilayer structures


One major advantage of the electrochemical etching process for producing nano-PSi
is the ease of fabrication of multilayer structures. PSi is depleted of carriers due to the
widening of its bandgap and the enhancement of its resistivity. As a consequence,
PSi itself is not affected by the ongoing formation reaction. This is one of the reasons
for in-depth PSi growth. Once a PSi layer is formed, a change in the fabrication
parameters (i.e., current density) will have an effect on the new layer being formed
beneath the already existing PSi. In this way, in-depth porosity profiles can be
achieved, for example, by varying the anodization current density over time.

By alternately switching the current density between high and low values, stacks of
high and low porosity layers can be achieved. Such a structure results in a multilayer
stack of dielectric layers with low and high refractive indexes and, as a consequence,
will show interference effects. In this way, a precise correlation between formation
parameters and effective refractive index, and etch rate allows the fabrication of
nano-PSi interference devices with predefined optical behavior.

In order for such devices to show the expected optical behavior, however, lateral
and in-depth homogeneity and sharp interfaces are required. Cross-sectional SEM
images, such as the one shown in Fig. 8.14, show that these structures have indeed
a notable lateral and in-depth homogeneity, with well-defined and sharp interfaces
between adjacent layers (Beale et al., 1985; Smith and Collins, 1992).
8.14. Cross-sectional SEM image of a nano-PSi multilayer. High porosity layers
appear as dark stripes due to their lower density, whereas low porosity layers appear
as bright stripes.

However, not only geometrical homogeneity is important for optical applications of


nano-PSi multilayers, but also morphological (i.e., porosity) in-depth homogeneity,
since it is directly related to the effective refractive index of the nano-PSi layers (see
Section 8.3). Figure 8.15 shows the porosity depth profile of a nano-PSi interference
filter composed of five high-index/low-index (low porosity/high porosity) bilayers.
The porosity is calculated from the compositional profile measured by Rutherford
backscattering spectroscopy. As it can be observed, the porosity of alternating layers
remains fairly constant along the whole stack, apart from a measurement artifact at
the outermost layer and a transition layer at the substrate interface (Torres-Costa et
al., 2005) previously observed by HR-TEM (Martín-Palma et al., 2002). This result
shows that the effective optical constants determined for nano-PSi layers as a
function of their fabrication parameters can be used and exploited for the fabrication
of nano-PSi multilayer stacks.
8.15. Rutherford backscattering spectroscopy (RBS) spectrum of a nano-PSi multi-
layer stack consisting of five low porosity/high porosity pairs (a), and compositional
and porosity profile obtained from the theoretical fit of the spectrum (b).

The correlation of the optical properties and thickness of nano-PSi layers with
their formation parameters allows the prediction of the interference effects of
nano-PSi-based multilayer stacks. In this way, optical interference filters can be
designed, optimized, and later fabricated using nano-PSi multilayer stacks. One of
the more simple interference filters is the so-called distributed Bragg reflector (DBR),
which consists of a stack of alternating high-index and low-index quarter-wave
layers. This layout leads to the formation of interference filters with a reflectance
peak (stop band) centered at the wavelength where the quarter-wave condition is
met. Following this rule, nano-PSi DBRs can be fabricated with reflectance peaks
tunable across the whole visible wavelength range by appropriately designing the
structure, as shown in Fig. 8.16.

8.16. Experimental reflectance spectra of nano-PSi Bragg reflectors, formed using a


1:1 HF and ethanol electrolyte. The thickness of the high-index and low-index layers
is indicated as d(H) and d(L), respectively. (‘MC’ is a label to correlate each curve in
the graph with the legend.)

It can be demonstrated (Macleod, 1969) that the half width at half maximum Δ of
a stop band centered at wavelength 0 depends on the high nH and low nL refractive
indexes as given by Equation [8.11]:

[8.11]

As a consequence, wider stop bands can be achieved by using nano-PSi layers with
a larger index mismatch. As a comparison, Fig. 8.17 shows the spectra of nano-PSi
DBRs formed with a different electrolyte from those shown in Fig. 8.16. In both
cases the reflectance peak is fully tunable over the whole visible range, although it is
found that filters formed with 2:1 HF:ethanol electrolyte (Fig. 8.17) show narrower
peaks than those formed in 1:1 HF:ethanol solutions (Fig. 8.16) for the same current
density values, for high and low porosity layers.

8.17. Experimental reflectance spectra of nano-PSi Bragg reflectors, formed using a


2:1 HF and ethanol electrolyte. The thickness of the high-index and low-index layers
is indicated as d(H) and d(L), respectively.

As shown in Fig. 8.10, electrolytes with low HF content allow achieving wider refrac-
tive index intervals than electrolytes with high HF content. Figure 8.18 shows the
stop band widths that can be achieved with two different electrolyte concentrations
according to Equation [8.11]. Clearly, the electrolyte with a lower HF to ethanol ratio
gives a larger flexibility to design the optical filters.

8.18. Half width at half maximum (HWHM) of the reflectance peak of nano-PSi
Bragg reflectors as a function of the layer’s refractive index for two electrolyte
compositions, according to Equation [8.11].

Another feature of great significance is the peak height, since it is directly related to
the overall optical performance of the filters. From Figs 8.16 and 8.17, it is observed
that multilayers formed using lower HF concentrations show higher reflectance
peaks. This behavior is attributed to a higher average porosity. As reported elsewhere,
increasing porosity leads to a blue-shift of the absorption edge (Canham, 1990; von
Behren et al., 1998) as a consequence of the band gap widening due to the quantum
confinement effects, resulting in less absorption losses, especially for wavelengths
below 500 nm.

Besides the previously described ‘ step-like’ multilayer structures, the electrochem-


ical process allows the fabrication of graded structures, optical devices with a con-
tinuously varying refractive index profile, known as rugate filters (Bovard, 1993).
In such filters, the refractive index alternates between low and high values, but
in a smooth – usually sinusoidal – fashion. Rugate filters show a high reflectance
band like conventional DBRs, but due to the smoothly varying refractive index,
dispersion of light at the interfaces is notably reduced and high order harmonics
are suppressed. Furthermore, by using an apodized index profile, which gradually
matches the refractive index of air and the filter, and the filter and the substrate, the
sidelobes around the main reflectance peak can be suppressed (Southwell, 1983).

A typical example of a rugate index profile is shown in Fig. 8.19, along with the
reflectance spectra of the actual structure realized in nano-PSi (Ishikura et al., 2008).
A further advantage of this kind of structure is the possibility of combining several
rugate profiles at the same time. Thus, by combining rugate profiles with adjacent
stopbands, the final filter will show an extended broad band, as shown in Fig. 8.20.
These broadband filters can be used as omnidirectional reflectors, broadband filters,
or cutoff filters.
8.19. (a) Refractive index profile and (b) reflectance spectrum of a nano-PSi rugate
filter.

Source: Reprinted from Ishikura et al. (2008), Copyright 2008, with permission from
Elsevier.
8.20. Refractive index profile of a rugate filter resulting from the combination of
several single profiles (a), leading to a wide band reflector (b).

Source: Reprinted from Ishikura et al. (2008), Copyright 2008, with permission from
Elsevier.

As mentioned in Section 8.5, untreated nano-PSi is chemically unstable and prone


to oxidation even in ambient air. Thermally carbonized nano-PSi, on the other hand,
is chemically stable, the carbonization process having minor effects on the optical
constants of nano-PSi. The chemical stability of thermally carbonized nano-PSi
Bragg reflectors is shown in Fig. 8.21. In order to check the oxidation rate of the
Bragg reflectors, the reflectance peak position is plotted as a function of the days of
immersion in pure ethanol, a strong oxidizer. In this plot, oxidation is evidenced
by a shift of the peak position to shorter wavelengths (a negative % shift) as a
consequence of the decrease of the refractive index due to oxidation. As can be
observed, no significant shift is observed even after 6 months of storage of the
nano-PSi devices in ethanol. As a comparison, it is observed that untreated reflectors
suffer blue-shifts of up to 15% in the first few hours.
8.21. Evolution of the reflectance peak position of untreated and thermally car-
bonized nano-PSi Bragg reflectors, as a function of days of storage in absolute
ethanol.

> Read full chapter

Porous silicon for tumour targeting and


imaging
J-H. Park, in Porous Silicon for Biomedical Applications, 2014

17.3 Preparation of PSi particles


PSi materials are mainly fabricated by electrochemical etching of Si wafers in
an ethanolic HF solution. To use the PSi materials in vivo, their microparticle or
nanoparticle formulation should be prepared to perform their biological functions.
The microparticles travel through the bloodstream to the tumour site while the
nanoparticles enter into the tumour cells. In this section, the preparations of PSi
microparticles and nanoparticles used for tumour targeting and imaging will be
briefly described.

17.3.1 Microparticles
Many researchers have prepared PSi microparticles for tumour targeting and imag-
ing. In the microparticle formulation, the unique porous nanostructure is well
preserved so that imaging agents or nanoparticles can be easily incorporated, and
both shape and size can be easily controlled so that they are suitable for tumour
targeting. Gu and co-workers synthesized magnetic luminescent PSi microparticles
for magnetically targeting tumour cells in vitro and tracking the particles (Fig. 17.1)
(Gu et al., 2010). Electrochemical etching of highly doped p-type single-crystal Si
wafers in an aqueous HF solution containing ethanol produces a porous layer, which
is removed from the Si substrate with a secondary current pulse. In order to activate
photoluminescence and impart hydrophilicity, the porous Si layer is mildly oxidized
in an aqueous solution. The film is then fractured by ultrasonication and filtered
to remove fragments that are less than 5 μm in size. Finally, magnetic properties
were imparted to the luminescent PSi microparticles using an infusion of 13 nm
magnetite nanoparticles.

17.1. Preparation and properties of PSi microparticles. (a) Schematic depicting the
synthesis of magnetic luminescent PSi microparticles. Scanning electron micro-
scope (SEM) images of the PSi microparticles before (b) and after (c) loading of 13 nm
magnetite nanoparticles. Insets of both images reveal the porous nanostructures.
The scale bars indicate 100 μm (100 nm for the insets).

(Source: Adapted from Gu et al. (2010) Small 6(22): 2546–2552).

Tasciotti and co-workers prepared hemispherical PSi microparticles for multistage


delivery (Tasciotti et al., 2008). A heavily doped p++-type Si wafer was used as the
substrate, and a 200 nm layer of Si nitride was deposited by a low-pressure chemical
vapour deposition system. Standard photolithography was applied to pattern the
PSi. The hemispherical PSi microparticles were formed in ethanolic HF by electro-
chemical etching, and after removing the nitride layer by HF, the microparticles
were released by ultrasound. The particle was then either oxidized or modified
with (3-aminopropyl)triethoxysilane to create negative or positive surface charges.
Finally, fluorescent nanoparticles – QDs – were loaded into the pores of the PSi
microparticles for use in imaging.

17.3.2 Nanoparticles
Recent advances have led to the development of PSi nanoparticles that are
biodegradable and intrinsically photoluminescent. Nanoparticles of size between 10
and 100 nm have been shown to circulate for a relatively long time, and they accu-
mulate in the tumours via an EPR effect. Investigators at the University of California,
San Diego, prepared biodegradable luminescent PSi nanoparticles (PSiNP) for in vivo
tumour targeting and imaging (Fig. 17.2) (Park et al., 2009). The PSi nanoparticles
were prepared through a process involving electrochemical etching of single-crystal
Si wafers in ethanolic HF solution, lift-off of the PSi film, ultrasonication, filtra-
tion of the formed particles through a 0.22 μm filtration membrane and, finally,
activation of the luminescence in an aqueous solution. During the activation step,
Si oxide grows on the hydrogen-terminated PSi surface, generating significant
luminescence, which is attributed to quantum confinement effects and to defects
localized at the Si/SiO2 interface (Heinrich et al., 1992; Wilson et al., 1993; Godefroo
et al., 2008). Angelis and co-workers developed PSi nanoparticles that could be
successfully employed for in vivo tumour targeting and imaging (Angelis et al.,
2010). PSi was obtained by electrochemical anodization of boron-doped Si wafers
in an ethanolic HF. In order to obtain nanoparticles, the PSi films were sonicated in
dimethylformamide, ultra-sonicated in water at a constant temperature of 4 °C, and
finally filtered to eliminate impurities larger than 500 nm.
17.2. Preparation and properties of biodegradable luminescent PSi nanoparticles. (a)
Schematic depicting the synthesis of biodegradable luminescent PSi nanoparticles.
(b) Schematic depicting the structure and in vivo degradation process for the PSi
nanoparticles. (c) SEM image of the PSi nanoparticles. The scale bars indicate 500 nm
(50 nm for the inset).

(Source: Adapted from Park et al. (2009) Nature Mater 8: 331–336.)

Scientists in Finland prepared thermally carbonized PSi nanoparticles for tumour


targeting and imaging (Bimbo et al., 2010; Rytkonen et al., 2012). A PSi film, made
with an electrochemical etching technique, was thermally hydrocarbonized under a
continuous N2/acetylene flow at 500 °C. The partially hydrocarbon-terminated sur-
face of the PSi was modified with thermal treatment in undecylenic acid for further
bioconjugation. In order to formulate nanoparticles, the carboxylic acid-terminated
PSi was finally wet-ball milled in an ethanolic solution and centrifuged. For in vivo
targeting and imaging, these PSi nanoparticles were further functionalized with PEG
molecules, antibodies and radiotracers via the carboxyl functional group.

17.3.3 Biocompatibility and biodegradability


The most important factor in the design of any material that is to be injected into the
body is biocompatibility. In particular, materials that are being developed for in vivo
tumour targeting and imaging must be easily dispersible in a physiological solution,
so that they are suitable for intravenous administration, and able to circulate for
long enough to find their target. Furthermore, as most of the particles are inevitably
cleared by the organs of the MPS via an opsonization process – increasing the
likelihood of unintended toxicity – it is important that they degrade into non-toxic
by-products that can be eliminated from the body via the renal clearance pathway.
Shape, size, porosity and surface chemistry of the PSi particles are all relevant factors
in determining their in vivo behaviour and toxicity.

It has been demonstrated that oxidized PSi particles were completely degraded into
orthosilicic acid [Si(OH)4] in physiological conditions. The silicic acid, which is the
most available dietary form of Si, has been known to be efficiently eliminated from
the body through the urine (Popplewell et al., 1998). However, rapid degradation
of PSi particles in circulation would reduce the time available for them to carry out
their diagnostic and therapeutic functions. Thus, many researchers have developed
ways of controlling the biodegradation rate of the PSi particles. Hon and co-work-
ers tailored the biodegradability of PSi nanoparticles to improve blood circulation
(Fig. 17.3) (Hon et al., 2012). It was found that thermal oxidation and a silica coating
on the PSi nanoparticles increased their blood half-life from 10 min up to 3 and
8 h, respectively, because these techniques slowed down the biodegradation rate. A
PEG or dextran coating on the oxidized surface of PSi particles has also been used to
improve in vivo stability and circulation of the particles (Park et al., 2009; Russo et al.,
2011). Recently, a self-assembled protein coating consisting of fungal proteins (class
II hydrophobin HFBII from Trichoderma reesei) was used to increase hydrophilicity
of the thermally hydrocarbonized PSi nanoparticles (Sarparanta et al., 2012). These
methods of adjusting biodegradability allowed the temporally controlled release of
loaded diagnostic and therapeutic agents to cancerous tissues.

17.3. Characteristics of biodegradable PSi nanoparticles: (a) transmission electron


microscopy (TEM) image of PSi nanoparticles with no silica coating, 10 min silica
coating, and 30 min silica coating. (b) Degradation profiles and blood half-lives of
non-coated and silica-coated PSi nanoparticles.

(Source: Adapted from Hon et al. (2012) J Biomed Mater Res Part A 100A: 3416–3421.)

> Read full chapter

Seeing Atoms and Clusters on Surfaces


In Fundamentals and Applications of Nano Silicon in Plasmonics and Fullerines,
2018

4.6.2 Platinum–Iridium or Gold Tips


Gold tips can be obtained from wires by electrochemical etching in 0.8 M–97% KCN
solution. The wires are biased by applying a voltage in the range 8–20 V. The process
produces tip radius varying between 200 and 500 nm. Sharp edge tips may also be
prepared from these materials because these materials are rather soft. A razor blade
is placed on a 0.25 mm diameter wire at approximately a 45 degree angle. The wire
is then cut. This results in a sharp tip. One can also cut the wire with a wire cutter.
The tip quality is sufficient to obtain quality images.

> Read full chapter

Porous Silicon
Z. Gaburro, ... L. Pavesi, in Reference Module in Materials Science and Materials
Engineering, 2016

Generalities and Definitions


PSi was discovered by Uhlir in 1956 while performing electrochemical etching of
silicon. In 1990, Canham showed that certain PSi materials can have large PL
efficiency at room temperature in the visible: a surprising result, since the PL
efficiency of bulk silicon (Si) is very low, due to its indirect energy bandgap and
short nonradiative lifetime. The reason for this was the partial dissolution of silicon,
which causes (1) the formation of small silicon nanocrystals in the PSi material; (2) the
reduction of the effective refractive index of PSi with respect to silicon, and hence an
increased light extraction efficiency from PSi; and (3) the spatial confinement of the
excited carriers in small silicon regions where nonradiative recombination centers
are mostly absent. In general, PSi is an interconnected network of air holes (pores)
in Si. PSi is classified according to the pore diameter, which can vary from a few
nanometers to a few micrometers depending on the formation parameters (Figure
1). According to the general classification of porous materials, three size regimes are
defined as in Table 1.

Figure 1. Examples of PSi structures: (a) microporous, (b) mesoporous, and (c)
macroporous.

(a) Reproduced with permission from Cullis, A.G., Canham, L.T., 1991. Visible light
emission due to quantum size effects in highly porous crystalline silicon. Nature
353, 335–337; © Nature Publishing Group.

Table 1. IUPAC classification of porous materials

Dominant pore width (nm) Type of material


?2 Microporous
2–50 Mesoporous
>50 Macroporous

The word ‘nanoporous’ is sometimes used for the smallest-pore regime to empha-
size the nanometric dimension. The volumetric fraction of air of the material is called
porosity (P). The internal surface of PSi per unit volume can be very large, of the order
of 500 m2 cm-3. The enhanced PL efficiency of PSi–compared to Si–has motivated
research toward other porous semiconductor materials. For example, highly porous
SiC, GaP, Si1-xGex, and Ge structures have been investigated.

> Read full chapter

POROUS SILICON – SENSORS AND


FUTURE APPLICATIONS
James L. Gole, Stephen E. Lewis, in Nanosilicon, 2008

4.2.1 Pore Structure in PS


Several groups have studied the development of pore formation in etched silicon
surfaces. The electrochemical etching of silicon in a variety of electrolytes can
produce a diverse range of pore diameters [9–19] which can be made to vary from
the 1 to 10 nm range [11–14] (nanoporous silicon) to sizes in the 1–3 μm range
[17] (microporous silicon). Further, mixtures of two pore types are possible [12,19],
producing a hybrid micro-nanoporous structure. Additionally, “breakthrough pores”
of order 100 nm can be obtained by using a short, high potential, etch of n-type
silicon in darkness [9]. Further, “mesopores,” typically 10 nm in diameter, can be
obtained from heavily doped p+—Si and n−—Si under conditions that would produce
nanopores in moderately doped silicon.

One common form of PS which has been used in the fabrication of devices is
nanoporous silicon. This material is typically made through the electrochemical
etch of a silicon wafer, using a mixture of ethanol and aqueous HF. In producing
a nanoporous silicon film, a 1:1 (48% HF: ethanol) mixture (the etch solution) is
prepared. This solution is contained in a Teflon cell, and exposed to the polished
surface of a (100) Si wafer with a resistivity as low as 1 mΩ-cm. Electrical contact
is made to the back side of the wafer through a metal foil, as a Pt mesh electrode
(counter electrode) is also immersed in the etch solution. Anodization of the silicon
wafer is established with a constant current density of 20 mA/cm2 applied to the cell
for 20 min. Figure 4.1 shows one possible configuration of an etch cell. The outlined
etch procedure with a similar cell can result in a nanoporous sample approximately
12 μm deep with a porosity of near 75% [20].

Figure 4.1. Schematic showing a possible electrochemical etch cell configuration


used to produce a PS layer.
While the vast majority of experiments conducted thus far on PS have involved
primarily the characterization of nanoporous material. Christophersen, Carstensen,
Föll, and others have recently studied the nature of micropore formation in n-type
silicon (through backside illumination) as a function of doping, temperature, and
potential, and in p-type silicon as a function of crystal orientation and electrolyte
dependence; the former extends the studies of Lehmann and Föll [12] and Lehmann,
et al. [15,21,22] and the latter establishes the important role played by hydrogen
and oxygen in pore formation. Large up to 3 μm pores, “p-micropores,” were first
described for the non-aqueous etch of p-type silicon by Probst and Kohl [17] and
subsequently by others [23,24]. Using a hybrid approach [25,26] it is possible to
combine the microporous structure first generated by Probst and Kohl [17] with the
advantages brought by a nanoporous framework.

It is clearly established that one can fabricate a range of hybrid structures between
two limiting well-defined PS morphologies: (1) PS fabricated from aqueous elec-
trolytes that consist of highly nanoporous, branched structures and (2) PS fabricated
from non-aqueous electrolytes that is comprised of open and accessible microp-
orous structures with deep, wide, well-ordered channels that display a crystalline Si
(100) influenced pyramidal termination. The ability to control the interplay of these
two regimes of porosity provides a means to exploit both the bulk and surface prop-
erties of the supporting membrane. The hybrid microporous/nanoporous structure
etched into a silicon framework, as depicted in Fig. 4.2, represents an extrapolation
of the Probst and Kohl [17] study, and provides a useful platform for the construction
of a conductometric PS-based sensor.

Figure 4.2. Closeup views of the hybrid PS film: side and top view.

Single-crystal (100) boron-doped p-type silicon wafers with resistivities of the order
1–20 Ω-cm have been used to form the hybrid structure depicted in Fig. 4.2. Hybrid
nanopore-covered microporous p-type PS samples were fabricated in an electro-
chemical cell constructed from high-density polyethylene. A working electrode was
attached to the back of the wafer and the counter electrode corresponded to a
platinum foil placed in solution. The cell was sealed to the front of the wafer,
using a clamp, as a 1 cm2 section of the wafer made contact with the solution. A
magnetic stir-bar was used to prevent the buildup of hydrogen at the surface of
the silicon. The samples were etched in a solution of 1 M H2O,1 M HF, and 0.1 M
tetrabutylammonium percholate (TBAP), all in acetonitrile. The hybrid samples were
etched with a current density of 6 mA/cm2 for between 50 and 75 min. Using this
procedure, pores approximately 1 μm wide and 10–40 μm deep were formed, well
covered by a coating of nanoporous silicon [25,27].

> Read full chapter

Optical Properties of Nanostructured


Silicon☆
Yimin Chao, in Comprehensive Nanoscience and Nanotechnology (Second Edition),
2019

4.08.1 Introduction
Porous silicon (π-Si), typically a few micrometers-thick film, produced by electro-
chemical etching of silicon wafer in an HF solution under an anodic current, was
first discovered in 1956 by A. Uhlir at Bell’s Laboratories in USA [1], when they
were doing electrolytic shaping the surface of silicon and germanium. However, this
result did not evoke further attention from scientific communities. In 1990, Canham
reported the discovery of significant visible light emission from porous silicon
under the UV excitation [2]. This finding stimulated the interest of the scientific
community in its optical and electronic properties. It is generally considered the
luminescence owing to the presence of quantum-confined structures and is often
described as a nanocrystalline film. Porous silicon can be broken up into individual
nanoparticles by a variety of means including ultrasound [3]. It therefore provides
one of the simplest routes to silicon nanoparticles, requiring only a small power
supply (e.g., 400 mA, 40 V) for the etching and an ultrasonic horn or bath. In this
chapter the fabrication methods will be reviewed first, followed by discussions on
microstructures and quantum effect, optical spectroscopic results and applications.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like