Electrical Property of Semiconductor

In subject area: Engineering

Chapters and Articles

You might find these chapters and articles relevant to this topic.

1.1.4 Electrical Properties

Silicon is a group IV element in the periodic table and is a semiconductor with a bandgap of 1.12 eV, which means that pure silicon at room temperature is almost an insulator. By doping with group III or group V elements, the resistivity of silicon can be varied over a wide range.

1.1.4.1 Introduction—Dopants and Impurities in Silicon

Semiconductors are solid materials that have electrical conductivities in between those of conductors and those of insulators. The physical reason causing a material to behave as a conductor, semiconductor, or insulator lies in the availability, or lack thereof, of free current carriers in the material. Semiconductors are characterized by the narrow bandgap between the valence bands, occupied by electrons, and the conduction band, in which electrons move freely according to applied electrical fields. Intrinsic (i.e., pure) semiconductors act as insulators at room temperatures, but their behavior changes dramatically with temperature, and, more to the point, with small impurities present in the crystal. Very small amounts of electrically active impurities can totally alter the electrical properties of semiconductors such as silicon. This is because the electrically active impurities either easily donate valence electrons (donors) or accept them, creating holes (acceptors). These electrons or holes are free (i.e., not bound to individual atoms). Their movement due to applied electrical fields carries electrical currents, giving rise to the term charge carriers used to denote them.

The electrical properties of semiconductor materials such as single-crystal silicon are thus defined by the impurity concentrations present in the silicon lattice. Impurities are introduced into the starting materials during crystal growth and modified during device processing by additional doping of the silicon material with electrically active impurities. In intentional doping of silicon, impurity atoms from group III (acceptors) and group V (donors) are used. The dopants used in crystal pulling, and the conductivities reached with specific impurity concentrations, are described in more detail in Sections 3.1 and 3.3Section 3.1Section 3.3, Dopants and impurities in silicon crystals.

Manipulation of the electrical properties in the structures created during MEMS device manufacturing follows practices established in semiconductor device manufacturing. Techniques used include both very traditional methods, such as deep diffusions of dopants, which have been abandoned in mainstream semiconductor processes, and current standard techniques such as ion implantation and epitaxial deposition. While semiconductor-grade starting materials are largely free of other electrically active impurities, the incorporation of unintentional contamination into silicon during processing can have significant effects on the electrical properties of the manufactured devices.

Unintentional doping of silicon includes the introduction of unwanted donors or acceptors to the crystal lattice from the processing environment. These impurities can be either group III/V-type atoms misplaced, or other contaminants, such as some transition metals.

The generation of electrically active donors also takes place within the single-crystal CZ silicon itself, without the introduction of additional impurities. CZ silicon always includes a few parts per million atoms of interstitial oxygen atoms, originating from the quartz crucible used to hold the melt during crystal pulling (see also Chapter 3). At certain temperatures, ranging from 400°C to 550°C, these interstitial atoms create conglomerates of several oxygen atoms within the lattice. Such silicon–oxygen microclusters are known as thermal donors (TDs), as they donate free electrons to the conduction band, influencing the electrical properties accordingly [22,23]. The concentration is, however, usually below 1015×cm−3, and thus has only very marginal effects on other than high-resistivity silicon. These donors are not stable at temperatures above 600°C, and even for applications requiring high-resistivity silicon, their effect can be suppressed by quick cooling down over the generation temperature range. This method is called thermal donor anneal (TDA), and while it is effective, it does not prevent the generation of new TD if temperatures in the critical range are used later in the device processing. For further details, see Section 3.6.

The use of semiconductors is based on the fact that the charge carrier concentrations are also influenced by any electric fields that are present. This can take place intentionally, as, for example, in transistors, but also due to electric fields generated by surface effects such as charging. These effects are more pronounced in high-resistivity silicon, but they bear consideration in general.

In the vast majority of cases, the electronics required for the realization of a MEMS-based sensor consist of the basic building blocks of semiconductor electronics used since the birth of the silicon-based semiconductor industry and described in the basic handbooks of the industry, such as Sze [24]. In MEMS applications requiring very high resistivity, in fields such as RF and optical, special considerations apply. The very high resistivity materials used are obviously strongly affected by even the smallest concentrations of unintended charge carriers. These unintended charge carriers can be introduced into the material by methods such as TD generation (described earlier), oxidation, and contamination of the surface if the donors/acceptors are not subsequently evaporated. In very high resistivity silicon, these effects can be severe, in some cases even leading to type reversal.

1.1.4.2 Piezoresistive Effect in Silicon

1.1.4.2.1 General Piezoresistive Effect

The change in the resistance of metal devices due to an applied mechanical load was first discovered by Lord Kelvin in 1857. With the large-scale use of single-crystal silicon in the making of semiconductor circuits, a much stronger piezoresistive effect was discovered in silicon [25]. This discovery forms the basis for practically all piezoresistive MEMS applications.

1.1.4.2.2 Strain

When a sample is subjected to physical force, the force yields a change in length, dL, that follows the well known Hooke’s law F=kΔL, where k is a material constant. The stress is defined as the applied force per unit area. Stress (σ) is thus given by:

(1.1)σ=F/A=kΔL/A=kLΔL/ALorkεLL/A

where εL denotes the differential deformation, ΔL/L, known as strain. Most solids exhibit elastic behavior for small stress loads. In elastic deformation the strain is proportional to the applied stress and the material follows Hooke’s law, which, in the simple case of uniaxial stress, can be written as:

(1.2)σ=YεL

where the scalar Y stands for the modulus of elasticity, also known as Young’s modulus. This material parameter for crystalline materials such as silicon is not omnidirectional, but varies according to crystal planes.

For strains above certain thresholds, Hooke’s law is no longer valid. This is caused by non-elastic, or irreversible (i.e., plastic) deformation.

The piezoresistive effect is quantified using the gauge factor G, describing the relationship between the applied strain and the change in resistivity. The piezoresistive effect, and thus also the gauge factor of semiconductor materials, is several magnitudes larger than the geometrical effect observed in metals. The effect is observed in several semiconductor materials, such as germanium, polycrystalline silicon, and single-crystal silicon.

The large piezoresistive effect in silicon is due to the fact that, instead of a stress-dependent change in geometry, the effect is first and foremost due to the stress-dependent resistivity of the material. The strain applied to the lattice causes the anisotropic band structure to deform, giving rise to anisotropic change in the mobility of the current carriers.

1.1.4.2.3 Stress in Anisotropic Materials

The elastic behavior of isotropic materials can be defined with just two elastic constants, commonly used parameters being the Young’s modulus Y and Poisson’s ratio ν, but modified constants known as bulk modulus K and shear modulus G are also sometimes used. In the general anisotropic case, the stress state of the material, is defined by a matrix known as the stress tensor, a 3×3 matrix in three dimensions.

Crystallographic symmetry allows the reduction of stress tensors in silicon to six-vector notation. In static equilibrium, the situation is further simplified to just six independent variables, and the stress tensor can be written as [26]:

(1.3)σ=[σxxσxyσxzσxyσyyσyzσxzσyzσzz]

Typically, the geometries used in MEMS applications reduce the situation, as in the case of a piezoresistive sensing layer created on a relatively thick silicon substrate. For such a layer, the vertical (z) stresses can be equated to zero, and the stress is reduced to:

(1.4)σlayer=[σxxσxy0σxyσyy0000]

This expression can be further simplified if we limit our discussion to stresses parallel, and stresses perpendicular, to the direction of current. For practical sensor systems, this limitation can easily be applied.

Thus, the forces can be denoted as σl (parallel stress) and σt (perpendicular or transverse stress).

1.1.4.2.4 Strain Effect on Resistivity

In metals the conduction/valence bands are partially filled with charge carriers, and small changes in band shape do not affect most of the carriers. In semiconductors the shape and dimension of the bandgap has a much larger effect, and shifts in energy bands due to applied stress affect the mobility of charge carriers. The result is resistivity change due to applied stress and gauge factors up to two magnitudes larger than those observed in metals. Because the relative change in resistivity is proportional to strain, it can be written as:

(1.5)Δρρ=YπLεL

For strained anisotropic material, the resistivity is thus no longer scalar and is described through tensor mathematics, similar to the mathematics described earlier for stresses in anisotropic material. Thus, the dependence of the current density J on the electric field E is denoted as:

(1.6)[ExEyEz]=[ρxxρxyρxzρxyρyyρyzρxzρzyρzz][JxJyJz]

Details of the tensor mathematics explaining the anisotropic resistivity are described in, for example, Thomsen [26].

In typical applications the MEMS devices are constructed in such a way that the current in the piezoresistors is either parallel or perpendicular to the direction of the stress. Also, the vast majority of applications are made on (100) silicon wafer substrates, with the substrate oriented in such a way that the resistors are aligned on the [110] direction on p-type and the [100] direction in n-type, as depicted in Figure 1.13. These directions yield the maximum positive and negative piezo coefficients (i.e., the maximum relative resistivity change for a given applied force).

Figure 1.13. Typical piezoresistor alignments used on MEMS-standard (100) silicon wafers, aligned to utilize maximum piezo coefficients for rectangular membranes of the type used in pressure sensors. (a) P-type silicon, Dicing along <110>, (b) P-type silicon, Dicing along <100>, (c) N-type silicon, Dicing along <100>.

For a system in which the coordinates are chosen to coincide with the crystallographic axes, the relationship between the stress tensor and the anisotropic change in resistivity simplifies due to crystallographic symmetries of silicon, and can be defined as:

(1.7)1ρ·[ρ11ρ22ρ33ρ23ρ13ρ12]=[Π11Π12Π12000Π12Π11Π12000Π12Π12Π11000000Π44000000Π44000000Π44][σ11σ22σ33σ23σ13σ12]

where Π11, Π12 and Π44 represent three independent piezo coefficients, listed in Table 1.2.

Table 1.2. Independent Piezo Coefficients for Crystalline Bulk Silicon [25,27]

10−11 1/Pan-Si ND=4×1014 [25]p-Si NA 2×1015 [25]p-Si NA 1×1018 [27]p-Si NA 1×1019 [27]
Π11−102+6,6
Π12+53−1,1
Π44−14+138+103+81

In most applications the change in resistivity is limited to stresses parallel and perpendicular to the resistive path. In such cases, the situation is greatly simplified [28]. The tensor element becomes a scalar equation, and can be written as:

(1.8)ρρ=πlσl+πtσt

where the common notions of longitudinal piezo coefficient πl and transversal piezo coefficient πt are used. As an example of a common situation, let us consider the case of a p-type piezoresistor constructed to utilize the maximum piezo sensitivity as described in Figure 1.13(a). For such a resistor, which is uniaxially stressed and lying in the (110) direction, the shear strain is zero and the longitudinal piezo coefficient is simplified to:

(1.9)πl=12[π11+π12+π44]

The corresponding transversal piezo coefficient then becomes

(1.10)πt=12[π11+π12π44]

For piezoresistors in the <100> plane of an n-silicon substrate, the maximum piezo coefficient is found along the (100) crystal axis, and the piezo coefficients are reduced to:

(1.11)πl=π11

and

(1.12)πt=π12

It is immediately observed that for both n- and p-type piezoresistors, uniaxially stressed at respective optimum locations for sensitivity, the longitudinal and transverse piezo coefficients have opposite signs. The advantages offered by this fact in a Wheatstone bridge have contributed to the dominance of bridge circuits in piezoresistive sensing based on silicon.

The alignment of piezoresistors to be constructed is governed by the piezoresistive coefficients listed in Table 1.2. Almost universally, the optimal solution is to align the sensing resistors according to maximum piezoresistive effect, as described by the examples depicted in Figure 1.13. The choices of resistor direction, dicing direction, and anisotropic etching are determined by the crystal orientations. In addition, the specification of the crystallographic location of the substrate wafer flat is based on the chosen crystal alignment. The primary flat has historically been located at the <110> direction, and dicing along and perpendicular to the primary flat is depicted in Figure 1.13(a) and (c).

1.1.4.2.5 Linearity

In comparison to many other physical phenomena, the piezoresistive effect is a fairly linear effect. To be exact, the linear equations for piezo coefficients should, however, be replaced with higher order polynomials. For resistors aligned on the <110> crystal axis, for maximum piezoresistive effect in p-type silicon, a quadratic correction to the transversal piezoresistive coefficient is sufficient for a wide range of doping concentrations and applied stresses. The quadratic correction for p-type silicon along the <110> crystal axis is of the form [28]:

(1.13)ρρ=πlσl+πtσt[16.66·1010σt1/pa]

yielding a quadratic correction of less than 1% for stresses up to 10 Mpa.

1.1.4.2.6 Effect of Temperature and Doping

In addition to the piezoresistive effect, the silicon resistors exhibit strong temperature dependence. For piezoresistors with low to moderate doping, the resistivity changes can reach one magnitude for every 100°C. Change in offset is caused by leakage currents, which are strongly dependent on temperature according to Laermer [29]. While in some cases the temperature dependence can be utilized or ignored, in the vast majority of cases this must be compensated for. Combined with the effect of doping on the piezoresistive effect, the temperature effect can be corrected with the piezoresistance correction factor P(N,T). Generally, a higher dopant concentration in silicon reduces the temperature effect, whereas the higher the temperature, the lower is the effect of doping level. Thus, designing of the resistor requires a trade-off between sensitivity and temperature stability (Figure 1.14).

Figure 1.14. Piezo correction factor P(N,T) in n-Si as a function of doping level, for various temperatures.

Redrawn and modified from Kanda [30].

The modern IC technology can provide a mathematical solution to the compensation of the piezoresistors, by introducing signal conditioning of the linearity and temperature effect to the sensing setup. This approach provides additional accuracy and flexibility at the cost of some cost and complexity. While the approach is commonly used to improve nonidealities in surface micromachined sensors, the performance of sensing elements manufactured from bulk, single-crystal silicon is usually close to ideal when temperature effects are accounted for.

Basic solutions for correcting for temperature variations include the use of a reference (i.e., a non-strained resistor which is at the same temperature). The difference of signal between the strained and non-strained sensors provides the desired signal due to strain only. The other basic technique is to utilize a temperature independent circuit—in practice, a bridge circuit.

In a Wheatstone bridge circuit application, the four resistors are doped to the silicon and arranged in such a way on the sensing membrane that they react to strain in pairs. In addition to having the advantages in sensitivity, discussed earlier, the bridge circuit can be used to compensate for temperature effects. The standard setup of a Wheatstone bridge in (100) silicon substrate consists of one pair of resistors aligned to give maximum positive change in resistivity (longitudinally stressed if p-type, transversely stressed if n-type), while the other pair yields the maximum negative change (a perpendicular alignment with respect to the first pair of resistors). This bridge circuit provides an output signal which is independent of temperature, assuming that the whole membrane is subjected to equal temperature changes. The circuit output is thus, to a first-degree approximation, directly proportional to the change in resistivity due to the applied stress. The main disadvantage of piezoresistive gauging using bridge circuits lies in the current requirements in the milliampere range, which cause limitations in the use of these circuits in power-sensitive applications.

1.1.4.2.7 Example of a Piezoresistive Sensor Design

Pressure sensing has utilized piezoresistive sensing based on MEMS technology for years and is the first widely adopted application of the technology. This is due to the fact that thin micromachined silicon membranes with a piezoresistor bridge implemented on top of them provide a rather ideal solution to the common problem of sensing pressure over a known, rather limited range of pressures, over long periods of time and countless cycles of pressure changes, for low cost and no maintenance.

A very basic sensing setup is as follows:

Starting point:

Silicon substrate, N-type, (100) orientation, device dicing along <110> direction.

Membrane defined by etching, using an alkali etchant from the backside of the wafer. The long etching is carried out in concentrated KOH or TMAH, requiring a thick protective layer of oxide, or nitride. The traditional bulk micromachining has utilized backside patterning for these cavity-etching processes.

Cavity sealed through wafer bonding. The sealing options range from glues and glass frits to fusion bonding of silicon wafers.

Piezoresistors are boron implanted on top of the wafer, at the edges of the membrane, where the stress maxims are located when the cavity pressure differs from the outside pressure. Boron implantation of the ohmic contacts can usually also be performed at this stage.

The basic device is finished with one or two layers of metals.

The dopant profile of a simple implanted and diffused piezoresistor of the type described is illustrated in Figure 1.15. The denuded zone responsible for the electrical isolation is shown without bias; with reverse bias, it can be extended to meet the requirements.

Figure 1.15. Basic diffused piezoresistor dopant depth profile. The wafer surface is to the left, and the piezoresistor is hatched. The transition zone under the piezoresistor acts as the electrical isolation. The hatched line (– · · –) on the left denotes free hole concentration, and the hatched line (— · —) on the right side of the transition denotes free electron density, approaching the starting material ND density in the depth of the wafer. The graph is calculated using IC-simulation software ICECREM 4.3 for WINDOWS™ by Fraunhofer Gesellschaft for Integrated Systems and Device Technology (IISB), Erlangen, Germany [31], and the carrier concentrations refer to zero bias.

The basic device outlined earlier follows the geometry outlined in Figure 1.16. This device has the piezoresistors located at the stress maximums of the rectangular membrane and along the crystal axis yielding greatest change in resistivity for a given stress. Also, the shear stresses of the membrane are minimized, theoretically to zero, at the locations of the resistors.

Figure 1.16. Basic MEMS-based sensor application using piezoresistors in &lt;100&gt; silicon. (a) Side view (cross-section along the &lt;110&gt; plane). (b) Top view, with the piezoresistors located in a Wheatstone bridge, along the (110) direction on the top of the silicon membrane.

1.1.4.2.8 Surface Effects

In many cases, the carrier concentration depth profile of the piezoresistor can be very important. Especially in accelerometers, the piezoresistive MEMS-sensor is often loaded so as to cause compressive strain on one side of the membrane and tensile strain on the other. In such a case the effective piezo coefficient, and thus the sensitivity, is very strongly influenced by the ratio of carrier concentrations between the membrane surfaces. Also, those piezoresistors whose stress is purely compressive or tensile are affected by the vertical carrier profiles, as the greatest strains are located at the very surfaces of the membrane. On the other hand, the carrier concentrations tend to change more rapidly the closer they are to the surface. Because the carrier concentrations near the device surfaces can also be affected by charging of the dielectric layers used to isolate and seal the micromachined structures, a careful consideration of the design principles used in analog transistor design is necessary to ensure the reliable operation of the piezoresistive element.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780323299657000014

1.1.4.1 Introduction—Dopants and Impurities in Silicon

Semiconductors are solid materials that have electrical conductivities in between those of conductors and those of insulators. The physical reason causing a material to behave as a conductor, semiconductor, or insulator lies in the availability, or lack thereof, of free current carriers in the material. Semiconductors are characterized by the narrow bandgap between the valence bands, occupied by electrons, and the conduction band, in which electrons move freely according to applied electrical fields. Intrinsic (i.e., pure) semiconductors act as insulators at room temperatures, but their behavior changes dramatically with temperature, and, more to the point, with small impurities present in the crystal. Very small amounts of electrically active impurities can totally alter the electrical properties of semiconductors such as silicon. This is because the electrically active impurities either easily donate valence electrons (donors) or accept them, creating holes (acceptors). These electrons or holes are free (i.e., not bound to individual atoms). Their movement due to applied electrical fields carries electrical currents, giving rise to the term charge carriers used to denote them.

The electrical properties of semiconductor materials such as single-crystal silicon are thus defined by the impurity concentrations present in the silicon lattice. Impurities are introduced into the starting materials during crystal growth and modified during device processing by additional doping of the silicon material with electrically active impurities. In intentional doping of silicon, impurity atoms from group III (acceptors) and group V (donors) are used. The dopants used in crystal pulling, and the conductivities reached with specific impurity concentrations, are described in more detail in Sections 3.1 and 3.3Section 3.1Section 3.3, Dopants and impurities in silicon crystals.

Manipulation of the electrical properties in the structures created during MEMS device manufacturing follows practices established in semiconductor device manufacturing. Techniques used include both very traditional methods, such as deep diffusions of dopants, which have been abandoned in mainstream semiconductor processes, and current standard techniques such as ion implantation and epitaxial deposition. While semiconductor-grade starting materials are largely free of other electrically active impurities, the incorporation of unintentional contamination into silicon during processing can have significant effects on the electrical properties of the manufactured devices.

Unintentional doping of silicon includes the introduction of unwanted donors or acceptors to the crystal lattice from the processing environment. These impurities can be either group III/V-type atoms misplaced, or other contaminants, such as some transition metals.

The generation of electrically active donors also takes place within the single-crystal CZ silicon itself, without the introduction of additional impurities. CZ silicon always includes a few parts per million atoms of interstitial oxygen atoms, originating from the quartz crucible used to hold the melt during crystal pulling (see also Chapter 3). At certain temperatures, ranging from 400°C to 550°C, these interstitial atoms create conglomerates of several oxygen atoms within the lattice. Such silicon–oxygen microclusters are known as thermal donors (TDs), as they donate free electrons to the conduction band, influencing the electrical properties accordingly [22,23]. The concentration is, however, usually below 1015×cm−3, and thus has only very marginal effects on other than high-resistivity silicon. These donors are not stable at temperatures above 600°C, and even for applications requiring high-resistivity silicon, their effect can be suppressed by quick cooling down over the generation temperature range. This method is called thermal donor anneal (TDA), and while it is effective, it does not prevent the generation of new TD if temperatures in the critical range are used later in the device processing. For further details, see Section 3.6.

The use of semiconductors is based on the fact that the charge carrier concentrations are also influenced by any electric fields that are present. This can take place intentionally, as, for example, in transistors, but also due to electric fields generated by surface effects such as charging. These effects are more pronounced in high-resistivity silicon, but they bear consideration in general.

In the vast majority of cases, the electronics required for the realization of a MEMS-based sensor consist of the basic building blocks of semiconductor electronics used since the birth of the silicon-based semiconductor industry and described in the basic handbooks of the industry, such as Sze [24]. In MEMS applications requiring very high resistivity, in fields such as RF and optical, special considerations apply. The very high resistivity materials used are obviously strongly affected by even the smallest concentrations of unintended charge carriers. These unintended charge carriers can be introduced into the material by methods such as TD generation (described earlier), oxidation, and contamination of the surface if the donors/acceptors are not subsequently evaporated. In very high resistivity silicon, these effects can be severe, in some cases even leading to type reversal.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780323299657000014

1 Elemental and Compound Semiconductors

Broadly speaking, semiconductors may be divided into two distinct groups, namely elemental or compound semiconductors. As the name suggests, elemental semiconductors are formed from single chemical elements, and important members of this family include Si and Ge. Compound semiconductors represent the largest group, however, and are formed as a result of the chemical reaction between two or more different elements (examples include GaAs, InAs, InP, and GaN). Another important group of materials is provided by semiconductor alloys, among which include AlxGa1−xAs and Si1−xGex. These materials are compound semiconductors in which the atoms are arranged in a well-defined crystal structure, but in which the different chemical species are randomly distributed throughout this crystal. While the bonding in most semiconductors is largely covalent in nature, in compound semiconductors formed from elements other than those in Group IV of the periodic table, the bonding also exhibits a small ionic nature and the material is said to be polar (Ferry 1991). Elemental semiconductors, on the other hand, are referred to as nonpolar.

Crystal structure plays an important role in determining the bandstructure, and the resulting electrical properties, of semiconductors. The vast majority of these materials exhibit either the diamond or the zincblende crystal forms (Fig. 2), the underlying lattice of which is the face-centered cubic lattice. In a compound semiconductor, such as GaAs, the two atoms in the basis are from different chemical species and this is known as the zincblende form. In either case, hybridization of the outermost occupied s- and p-orbitals gives rise to a tetrahedral bonding arrangement in the crystal (Ferry 1991) (Fig. 2), in which each atom bonds to its four nearest neighbors, forming a tetrahedral structure with the center atom located at the body of the tetrahedron.

Figure 2. Illustrations of the unit cell in the diamond (left) and zincblende (right) crystal structures. The solid lines between atoms in both figures represent chemical bonds. Note that, for the sake of clarity, chemical bonds extending outside the unit cell are not shown. In the case of the zincblende structure, the different colored balls represent atoms from different chemical species (for example, Ga and As in GaAs).

While the diamond and zincblende structures account for the majority of semiconductors encountered in nature, a less common form is the wurtzite structure. This is essentially based on the hexagonal lattice. The tetrahedral bonding arrangement remains in this crystal form, and GaN is an important example of a semiconductor that can exhibit the wurtzite structure. The crystal structure, and value of the principal energy gap (Eg), of a number of different semiconductors are summarized in Table 1. The form of the principal gap (direct versus indirect) is also shown in this table, although we defer a discussion of this issue until Sect. 2.1.

Table 1. Key semiconductors, their crystal structure, and energy-gap values. Note that the other form of C is graphite, which exhibits a hexagonal crystal structure and is a semimetal.

SemiconductorCrystal structureGap typeEnergy gap (Eg) at 300 K (eV)
SiDiamondIndirect1.124
CDiamondIndirect5.50
SiCZincblendeIndirect2.416
GeDiamondIndirect0.664
AlNWurtziteDirect6.2
AlPZincblendeIndirect2.45
AlAsZincblendeIndirect2.153
AlSbZincblendeIndirect1.615
GaNWurtziteDirect3.44
GaPZincblendeIndirect2.272
GaAsZincblendeDirect1.424
GaSbZincblendeDirect0.75
InAsZincblendeDirect0.354
InSbZincblendeDirect0.230
CdSZincblendeDirect2.50
CdSeWurtziteDirect1.751
CdTeZincblendeDirect1.475

Source: Madelung (1996), Pierret (1996).

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B0080431526018544

More than 100 000 papers quoted by the ISI Web of Knowledge have been published since 1975 solely on the scientific applications of the scanning electron microscope (SEM). Among them, more than 20 000 reported about semiconductors. As a consequence, the necessity to be concise and to present results with a scientific train of thought drove the authors to focus this chapter on the introduction of the basics of the two most used SEM techniques for the assessment of both optical and electrical properties of semiconductors, namely, the cathodoluminescence and the electron beam-induced current. Their complementarities, as well as a few updated emblematic applications to semiconductor nanostructures and devices, are presented and critically discussed. For the sake of completeness, a synthetic description of other techniques used for more specific investigations will also be presented.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780444531537001103
2019, Journal of Environmental ManagementAkansha Mehta, ... Tejraj M. Aminabhavi

1 Introduction

1.1 General background

In recent years, there has been great deal of efforts by chemists and physicists to discover novel nanomaterials for different applications (Haque et al., 2018). One of the largely moving areas of chemical research is the development of inorganic nanomaterials including metals, semiconductors and insulators in the dimension range of 1–100 nm (Haque et al., 2018). Dynamical behavior of low dimensional semiconductors has also become a crucial part of recent efforts. This has enabled intense investigations related to their novel optical, mechanical and transport phenomena. A variety of low dimensional semiconductors, which include quantum wires, quantum wells and quantum dots are made available (Xu et al., 2018). Of these, quantum dots have gained much popularity (Chen et al., 2019) Quantum dots are the zero-dimensional structures in which electrons are delocalized along all the three spatial dimensions, leading to ‘quantum confinement effect’ (Rao et al., 2019a; Chakraborty et al., 2019). Quantum confinement occurs when the size of the nanocrystal becomes smaller compared with twice the Bohr radius, which then becomes weak, when it exceeds the Bohr radius (Bergren et al., 2016).

In semiconductor nanocrystals, when a photon absorbs energy greater than that of the semiconductor band gap, it results in the generation of electron-hole pairs or excitons in which electrons and holes are bound by electrostatic attraction. The average distance between an electron and the hole into an exciton is of the order of Bohr radius. The optical and electrical properties of semiconductor nanocrystal then becomes dependent upon its physical dimensions when its size approaches near to that of the Bohr radius (Chakraborty et al., 2019).

Luminescent semiconductor quantum dots exhibit remarkable photostability, broad absorption profiles, high quantum yields and photo bleaching stability. Controlled shape and size luminescent characteristics of quantum dots arise from the quantum confinement effect, which allow accomplishment of related materials like optical visual for equivalent investigation of dissimilar analytes. The large stokes and slender emission bands shift in the luminescence spectra of quantum dots enable efficient coupling of other fluorophores or quantum dots with the emitted light (Owen and Brus, 2017). Hence, these systems have been widely explored as the potential materials for multicolored photoluminescent probes, luminescent biological labels, markers in imaging, emitters in light emitting diodes, gain media in lasers, light harvester in photovoltaics, photocatalysis and in various environmental treatment applications (Sundheep et al., 2019; Ramachandran et al., 2019; Mallick et al., 2019).

1.2 Need for sustainable development

Clean environment, water and energy are the basic needs for human survival and economic development (Gopinath et al., 2019). However, the rise in demand for various products has led the manufacturers to indulge in risky, but profitable production modes, leading to long-term environmental threats (Zandi et al., 2019). Additionally, utilization of energy through fossil fuels such as petroleum, natural gas and coal has led to pollution of air and water bodies (BusaidiBaawain et al., 2019). The limited availability of fossil fuels and their increasing year-wise unregulated consumption may lead to energy crises of the future. Since the Kyoto protocol in 2005, these issues have emerged in the global scenario in a highly debatable manner (Chen and Chen, 2002). Formulation and implementation of an integrated set of policies addressing the energy and environmental concerns simultaneously have met with several challenges in a developed nation such as the United States of America (Greening and Bernow, 2004). However, the present policies and technologies are ineffective in the realization of long-term goals (Andrews and Govil, 1995). In order to tackle fossil fuel-based energy crisis and environmental pollution, there is a greater need for greener and sustainable source, which can resolve both the energy and the environment related issues.

Nanotechnology can be a highly promising field for development of sustainability. It is an emerging field, which could contribute for the development of smarter materials capable of both generating energy and degrading the environmental toxic pollutants. It deals with designing and manipulation of materials at the molecular scale. Rapid development of novel nanomaterials can create options regarding new product innovation and high performance applications (Jyothi et al., 2019). Fabrication of such novel functional materials with tunable sizes, shapes, crystallinity, porosity and structures are of great importance for novel innovations in sustainable energy technologies (RoyDas et al., 2019). It also allows the fabrication of materials having specific functionalities capable of recognizing a particular pollutant in a mixture (Dharupaneedi et al., 2019). Thus, tremendous progress in nanotechnology may lead to the basic understanding of physics at the nanoscale in order to control the system properties and searching for new materials for energy and environmental applications (Chung et al., 2012).

Low energy solution-based synthesis of nanomaterials would allow their incorporation into the devices (Bera et al., 2019). Among the nanomaterials, photoactive metal oxide nanoparticles, quantum dots and carbon-based nanomaterials are the potential candidature since they rely upon the sunlight, which itself is a clean and sustainable energy source. Quantum confinement effect in nanomaterials such as quantum dots has given rise to many of the fascinating optoelectronic features, which make them highly eligible candidates for energy and environment-based applications. Moreover, such smart functional materials may be helpful in the development of clean and sustainable energy sources as well as for water and air purification-based devices. Carbon-based nanocomposites such as carbon quantum dots (CQDs) may play a significant role in this regard due to their distinctive optoelectronic properties. In recent years, these carbon-based eco-friendly materials have been widely investigated for photovoltaic applications, water splitting and their ability to degrade harmful pollutants (Sarkar et al., 2016; Reddy et al., 2019a, 2019b; Mishra et al., 2019a; Rao et al., 2019b).

1.3 Carbon quantum dots (CQDs)

Carbon-based nanomaterials have gained much significance owing to their outstanding optical, electronic, mechanical and optical characteristics (Sarkar et al., 2016). Among them, carbon dots (CDs) have the greatest consideration due to their biocompatibility, abundance of raw materials in the nature, low toxicity, resistance to photo-bleaching and cost-effectiveness (Wang et al., 2014). With their sizes in the range of 1–10 nm, these materials are classified as polymer dots (PDs), graphene quantum dots (GQDs), and carbon quantum dots (CQDs) (Sarkar et al., 2016). CQDs are the zero-dimensional fluorescent carbon nanomaterials possessing sp2 hybridized carbon atoms with surfaced passivity for oxygen containing functional groups such as hydroxyl and carboxylic groups. In contrast to other semiconductor CQDs are chemically inert, biocompatible, can be easily functionalized and are resistant to photo-bleaching (Linehan and Doyle, 2014). Their outstanding photo-physical properties, biocompatibility and low-cost make them ideal in the fields of bioimaging (Roy et al., 2019), sensing (Shetti et al., 2019a) and photocatalysis.

Researchers have reported several major developments in CQDs modifications and their applications. However, the issues based on optical and electronic properties are mainly due to the defects in CQDs, agglomeration of CQDs during the synthesis of CQDs affects the uniformity in size and the surface properties. Hitherto, approaches to precisely controlling the defects in CQDs are unavailable. Hence, it is required to put more efforts in atom-precise structural synthesis of CQDs. Also, modification of CQDs with controllable functionalization and doping of CQDs are the crucial issues. The majority of CQDs in photo-catalysis have shown restricted light-harvesting capacity, but only a few reports are available on photo-stability of CQD-based nanocomposites, thus signifying to develop further stable CQD-based photocatalytic systems for useful applications (Duarah and Karak, 2019; Ghosh et al., 2019; Mishra et al., 2019b). In biomedical field, many studies have been reported based on CQDs toxicity and their related mechanism. The C-dots are considered to be the competent nano-architectures for drug delivery and bioimaging, but their poor size and surface characteristics hinder their wider usage (Kaushik et al., 2019; Gulla et al., 2019). In order to overcome the biodegradability and toxicity problems, new fluorescent materials such as CQDs have been developed possessing better optical, biological, chemical inertness, and low toxicity compared to semiconductor quantum dots (Deng et al., 2019).

This timely review summarizes surface modifications and band gap tuning of CQDs by different novel metals and metal oxides. The review will address an outlook on the change in optical properties by surface modifications of CQDs. In later parts, more focus will be placed on water treatment/splitting applications such as heavy metal ion detection, photodegradation of dyes and H2 production of CQDs. Lastly, limitations and challenges of CQDs will be discussed along with the latest published reports for a quantitative discussion.

View article
Read full article
URL: https://www.sciencedirect.com/science/article/pii/S0301479719312046
1995, Handbook of Compound SemiconductorsVinod Malhotra, Carl W. Wilmsen

8.0 EPILOGUE

It is apparent that, in the last five to six years, significant progress has been made in passivation of GaAs and InP surfaces. Dramatic improvements in the electrical properties of semiconductor surfaces have been achieved using a variety of techniques. These techniques range from simple chemical treatments, using sulfur- and selenium-based solutions, to sophisticated epitaxial growth/regrowth methods which include the deposition of a few monolayers of Si. It is also very encouraging to note that excellent passivation of the surface defects can be achieved even on air-exposed surfaces. This is particularly true for the GaAs surface which has been relatively much harder to passivate than its InP counterpart. The much sought after MISFETs in these materials also exhibit significant improvements in their performance, and more work in the area will certainly be forthcoming.

The usefulness of superior M-S and I-S interfaces are obviously numerous. However, before these techniques become practically viable, the problems associated with uniformity, reproducibility, and stability of the passivated surfaces, devices and circuits will have to be carefully evaluated. The nature of surface electronic defects and the fundamental mechanism of passivation continue to remain controversial, but nevertheless, due to the amount of experimental data now available, it appears that dos and don ‘ts of III-V surface passivation may emerge. At least for now, it appears that the demon of Fermi-level pinning and the myriad of other related consequences that have plagued GaAs and InP are removed.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780815513742500090
2019, Journal of Environmental ManagementAkansha Mehta, ... Tejraj M. Aminabhavi

1.1 General background

In recent years, there has been great deal of efforts by chemists and physicists to discover novel nanomaterials for different applications (Haque et al., 2018). One of the largely moving areas of chemical research is the development of inorganic nanomaterials including metals, semiconductors and insulators in the dimension range of 1–100 nm (Haque et al., 2018). Dynamical behavior of low dimensional semiconductors has also become a crucial part of recent efforts. This has enabled intense investigations related to their novel optical, mechanical and transport phenomena. A variety of low dimensional semiconductors, which include quantum wires, quantum wells and quantum dots are made available (Xu et al., 2018). Of these, quantum dots have gained much popularity (Chen et al., 2019) Quantum dots are the zero-dimensional structures in which electrons are delocalized along all the three spatial dimensions, leading to ‘quantum confinement effect’ (Rao et al., 2019a; Chakraborty et al., 2019). Quantum confinement occurs when the size of the nanocrystal becomes smaller compared with twice the Bohr radius, which then becomes weak, when it exceeds the Bohr radius (Bergren et al., 2016).

In semiconductor nanocrystals, when a photon absorbs energy greater than that of the semiconductor band gap, it results in the generation of electron-hole pairs or excitons in which electrons and holes are bound by electrostatic attraction. The average distance between an electron and the hole into an exciton is of the order of Bohr radius. The optical and electrical properties of semiconductor nanocrystal then becomes dependent upon its physical dimensions when its size approaches near to that of the Bohr radius (Chakraborty et al., 2019).

Luminescent semiconductor quantum dots exhibit remarkable photostability, broad absorption profiles, high quantum yields and photo bleaching stability. Controlled shape and size luminescent characteristics of quantum dots arise from the quantum confinement effect, which allow accomplishment of related materials like optical visual for equivalent investigation of dissimilar analytes. The large stokes and slender emission bands shift in the luminescence spectra of quantum dots enable efficient coupling of other fluorophores or quantum dots with the emitted light (Owen and Brus, 2017). Hence, these systems have been widely explored as the potential materials for multicolored photoluminescent probes, luminescent biological labels, markers in imaging, emitters in light emitting diodes, gain media in lasers, light harvester in photovoltaics, photocatalysis and in various environmental treatment applications (Sundheep et al., 2019; Ramachandran et al., 2019; Mallick et al., 2019).

View article
Read full article
URL: https://www.sciencedirect.com/science/article/pii/S0301479719312046

1.0 INTRODUCTION

Successful development of compound semiconductor device technology relies greatly on accurate materials characterization of the very complicated structures that produce high speed optical and electrical products. Elemental or chemical analysis of GaAs and related materials for both bulk (matrix) and minor (dopant) constituents is necessary in the research, development, and production stages of transistors, lasers, and related devices. The diagnostic techniques described in this chapter are concerned with elucidating chemical composition of layered III-V materials by detecting the results of charged particle (ion) beam interactions with solids. The discussion is separated into two parts; the detection of the (back) scattered incident ions by RBS, Rutherford Backscattering Spectrometry, and the analysis of target material sputtered during ion bombardment by SIMS, secondary ion mass spectrometry, and related techniques.

Because RBS and SIMS use the production, transport and detection of ions in their characterizations, the two techniques are very expensive to implement. Both techniques require high (or ultrahigh) vacuum, high voltages and/or magnetic fields, sensitive charged particle detectors, sophisticated computing, etc., so the cost of a modern instrument may easily exceed one million dollars. The knowledge gained from their use should be commensurate with the capital and operating costs of a large and complicated experiment that often needs more than one skilled operator. Hopefully it will become clear that information provided by RBS and SIMS is unavailable from any other analytical method. The sensitivity, specificity and depth resolution of matrix and dopant detection with minimal sample preparation is the hallmark of these ion beam based characterization techniques.

1.1 RBS References

An excellent introduction to RBS is found in the surface and thin film analysis textbook by Feldman and Mayer.[1] Because virtually all III-V device substrates are prepared from single crystal material, the use of channeling in RBS requires much attention. This review draws heavily from the text by Feldman, Mayer, and Picraux[2] on this subject. Proceedings from the International Ion Beam Analysis meeting are found in Nuclear Instruments and Methods, Part B.

1.2 SIMS References

This review on SIMS characterization of III-V materials is fortunate to have a number of excellent references upon which to draw. An in depth treatise that covers the fundamentals of almost all aspects of SIMS has been written by Benninghoven, Rüdenauer and Werner.[3] A more recent work, geared mainly to the practice of semiconductor analysis, is by Wilson, Stevie, and Magee.[4] A similar review to this one, with an emphasis on InP materials has been written by Geva.[5] The textbook by Feldman and Mayer also has an excellent section on the fundamentals and application of SIMS as well as RBS. Any similarities of this work to these excellent references is purely intentional.

The proceedings of the biannual international conferences on SIMS are excellent sources for keeping up with SIMS technology and applications.[6] Proceedings of special topics conferences, such as the Molecular Beam Epitaxy Workshop, American Vacuum Society annual meetings, and others, are often published in the Journal of Vacuum Science and Technology. The journal Surface and Interface Analysis also contains articles pertinent to compound semiconductor characterization by SIMS.

The sections of this chapter that deal with sputter-based depth profiling may seem to have an exceptionally large amount of information about sputtered neutral mass spectrometry, or post-ionization techniques, because the author has experience in this field of research. In addition, many of the applications of the techniques covered here will be examples from his institution. This is only for the convenience of obtaining useful illustrations and is not meant to suggest that other institutions are not active in these fields of work.

1.3 Fundamentals of Ion-Solid Interactions

When a high velocity ion strikes the a solid, the ion will transfer some, or all, of its energy to the solid. For singly charged ions, the energy is numerically the same as the acceleration voltage, (E = qV). The manner of energy transfer depends upon many parameters, including the masses of the incident ion and collisional target nuclei and the energy and angle of incidence of the incoming projectile. Low energy ions, (< 100 eV) typically stop on or near the surface, thus building thin films. Incident, or primary ions with 1–20 keV of energy transfer energy to the sample's nuclei, which leads to sputtering, the sequential removal of surface layers, which forms the basis of analysis of the solids by secondary ion mass spectrometry (SIMS). At higher energies, (100–300 keV), ions implant into the solid, modifying its elemental composition. Some sputtering occurs, but most of the ions and their energy are implanted deeply (> 1000 Å) into the solid. Implantation damage and mixing of material hinders the analysis of layered samples with these high energy primary ions. However, ion implantation is widely used to control the electrical properties of semiconductors. MeV ions, especially He ions, penetrate very deeply into the solid (microns), loosing small amounts of energy to the solid's electrons until a scattering event occurs with a nucleus. Minimal damage to the sample occurs because most of the energy remains in the backscattered ion. Accordingly, He ion backscattering affords nondestructive surface and bulk chemical analysis.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780815513742500156

3.02.7.2 Dopant Impurities

Pure semiconductors have only limited applicability. The presence of impurities of certain categories is essential for tuning the properties of the semiconductor materials for a variety of applications. Incorporating selected impurity species in highly controlled concentration in semiconductor crystals is known as doping. The doping changes the electrical properties of semiconductor materials such as the conductivity, charge carrier type (holes or electrons), carrier density, and carrier mobility. Depending on the conductivity and net charge carrier type, semiconductors are categorized as n-type, p-type, and semi-insulating material. Table 21 lists the dopant elements used for III–V bulk crystals. The group VI elements (S, Se, and Te) are used as dopant elements for growing n-type III–V compounds. These elements occupy the group V (P, As and Sb) atomic sites in the crystal. The group II elements (Zn, Cd, Be and Mg) are used as dopant elements for growing p-type III–V compounds. These elements occupy the group III (Al, Ga, and In) atomic sites in the crystal. The group IV elements (C, Si, Ge, and Sn) are amphoteric in nature. They can either make the material n-type or p-type. Typically, they tend to self -compensation with some exceptions. For semi-insulating compounds, transition metals (Cr and Fe) are used as dopants. While these are only examples of common dopants used, growing any compound with a specific doping concentration and conductivity is an intricate process. The selection of the dopant element is done by evaluating a variety of properties. The segregation coefficient of the dopant during growth decides the concentration of the dopant incorporated in the crystal along the growth direction (refer to Equation (6)). The position of the discrete energy level (associated with the dopant) within the energy band gap of the semiconductor decides the type and the transport properties of charge carriers.

Native defects play a significant role in limiting the concentration and atomic sites of intentional impurities (dopants) that are incorporated in semiconductor materials for tuning the electrical properties. Complex defect structures are formed by the interaction of dopant impurities and the native defects. Self-compensation effects limit the dopant concentration and the electrical conductivity type of the material. One of the examples of interaction between native defects and intentional impurities is observed in semi-insulating undoped GaAs. The high-resistivity GaAs (∼107 Ω cm) can be obtained without the addition of intentional dopants, just by reducing the concentration of boron, carbon, and oxygen in the melt during growth (Rudolph and Jurisch, 2003; Eichler et al., 2008). When these species are kept below a certain threshold, a sufficiently high compensation is then provided by the native EL2 deep donor (an AsGa-related defect), which pins the Fermi level at midgap and produces high resistivity in GaAs (Kiessling et al., 2009).

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780444531537000900

10.7 Silicon Carbide Technology for Power Semiconductor Construction

It is worldwide accepted today that a real breakthrough in the power electronics field may mainly come from the development and use of wide band gap (WBG) semiconductor devices. In power electronic systems, there has been a continuous trend toward a higher system power density (i.e., W/cm3) in the last decades and due to environmental concerns and rising energy costs, also the efficiency of the systems became an important system performance criterion over the past years besides costs. For meeting these demands, the development of power semiconductors is a crucial factor. WBG semiconductors show superior material properties, which allow operation at high switching speed, high voltage, and high temperature. These unique performances provide a qualitative change in their application to energy processing.

From energy generation (carbon, oil, gas, or any renewable) to the end user (domestic, transport, industry, etc.), the electric energy undergoes a number of conversions. These conversions are currently highly inefficient to the point that it is estimated that only 20% of the whole energy involved in energy generation reaches the end user. WBG semiconductors can increase the conversion efficiency, thanks to their outstanding material properties.

In the area of power electronic converter systems there is a general trend to higher power densities and higher efficiency which is driven by cost reduction, an increased functionality, saving resources and, in some applications, by the limited weight/space requirements. To continue this trend, new devices, which enable high switching frequencies at higher power levels or show reduced losses at moderate switching frequencies, are required.

10.7.1 Characteristics of Wide Band Gap Semiconductors

Power semiconductor devices made from materials with band gap energies larger than that in Si have been touted for many decades. The potential advantages of these WBG devices include higher achievable junction temperatures and thinner drift regions (because of the associated higher critical electric field values) that can result in much lower on resistance than is possible in Si. WBG semiconductor materials allow power electronic components to be smaller, faster, more reliable, and more efficient than their silicon (Si)-based counterparts. These capabilities make it possible to reduce weight, volume, efficiency, and life-cycle costs in a wide range of power applications. Harnessing these capabilities can lead to dramatic energy savings in industrial processing and consumer appliances, accelerate widespread use of electric vehicles and fuel cells, and help integrate renewable energy onto the electric grid.

The continuous development of improved power semiconductors is a key enabling factor for propelling the constantly increasing demand for high power density and higher efficiency in many power electronic applications. The technological progress in manufacturing power devices based on WBG materials has resulted in a significant improvement of the operating voltage range, of the switching speed and on resistance compared to silicon power devices. Table 10.2 is presented with various semiconductor materials with their characteristics. From this table, it can be noted that the materials with a value of thermal conductivity close to or exceeding Si are GaN, GaP, SiC, and C (diamond). This implies higher melting temperatures. Out of these four semiconductors, GaP has much lower carrier mobility values than Si. So, one good semiconductor material for the future is C (diamond). It has the largest thermal conductivity and band gap of any of the materials from Table 10.2. Diamond also has the largest electron mobility of any material from Table 10.2 with a band gap larger than Si. However, there are some aspects of C (diamond) that make it less than ideal. First, the material and device fabrication technology is much less mature and developed than for SiC and GaN. Second, the thermal expansion coefficient (CTE) for C (diamond) is very low. So there is a clear thermomechanical mismatch which would make the package unreliable. Also, the diamond is one of the most expensive materials which make it inappropriate for integration in large scale systems. On the other hand, GaN and SiC are by comparison to C (diamond) very well suited to typical package materials, and in fact provide a better thermomechanical match than Si.

Table 10.2. Characteristics of wide band gap semiconductor materials at 300K

Semiconductor materialBand gap (eV)
Direct, D
Indirect, I
Critical or breakdown
Field Ec (V/cm)
Thermal conductivity (W/(m∗K))Coefficient of thermal expansion (ppm/K)
GaN3.44 D3,000,0001105.4–7.2
Ge0.661 I100,000585.9
Si1.12 I300,0001302.6
GaP2.26 I1,000,0001104.65
SiC (3C,β)2.36 I1,300,0007002.77
SiC (6H,α)2.86 I2,400,0007005.12
SiC (4H,α)3.25 I3,180,0007005.12
C5.6 I5,700,000600–20000.8

At present, SiC is considered to have the best trade-off between properties and commercial maturity with considerable potential for both high temperature applications and high power devices. However, the industrial interest for GaN power devices is increasing recently. For this reason, SiC and GaN are the more attractive candidates to replace Si in such applications. In fact, some SiC devices, such as Schottky diodes, are already competing in the semiconductor market with Si power diodes. Currently, it is a sort of competition between SiC and GaN in a battle of performance versus cost. Nevertheless, scientific and industrial actors agree in considering that both will find the respective application fields with a tremendous potential market. However, many of the material advantages still remain not fully exploited due to specific material quality, technology limitations, nonoptimized device designs, and reliability issues. It is worth mentioning that diamond exhibits the best properties of all the WBG semiconductors. Theoretically diamond would be ideal for bipolar device designs, particularly in operating environments of elevated temperatures. The high values of carrier mobilities, as well as the large band gap and high thermal conductivity, make diamond the ideal future material for electronic devices of all power levels and types. However, there are critical problems related with the crystal growth (small areas single crystal wafers), both p-type and n-type dopings and processing. Therefore, there is not a diamond power device in the market and it is not expected in the next decade. SiC power devices recently reported in literature include high-voltage and high-temperature diodes, junction controlled devices (like junction field effect transistors (JFETs)), MOSFETs and metal-semiconductor-field effect transistors (MESFETs). Those based on GaN include diodes, HEMTs, and MOSFETs; and advanced research on novel devices concerning low-losses digital switches based on SiC and GaN is also of main concern. These novel devices represent a real breakthrough in power devices.

10.7.2 Silicon Carbide Overview

Silicon carbide (SiC) was accidentally discovered in 1890 by Edward G. Acheson, an assistant to Thomas Edison, when he was running an experiment on the synthesis of diamonds. Acheson thought the new material was a compound of carbon and alumina present in the clay, leading him to name it carborundum, a name that is still being used on some occasions. Silicon carbide occurs naturally in meteorites, though very rarely and in very small amounts. Being the discoverer of SiC, Acheson was the first to synthesize SiC by passing an electric current through a mixture of clay and carbon. Today, SiC is still produced via a solid-state reaction between sand (silicon dioxide) and petroleum coke (carbon) at very high temperatures in an electric arc furnace.

Silicon carbide is a semiconductor material with highly suitable properties for high-power, high-frequency, and high-temperature applications. This almost worn-out opening statement may be found in many papers dealing with SiC. Yet, it cannot be left out because it really brings forward the essence of the material's potential. Silicon carbide is a WBG semiconductor material with high breakdown electric field strength, high saturated drift velocity of electrons, and a high thermal conductivity. Therefore, these properties make SiC ideally suited for a vast number of applications. Today there are high-frequency metal-semiconductor-field effect transistors offered commercially, as well as an emerging market for Schottky diodes made from SiC.

Silicon carbide is the only chemical compound of carbon and silicon. It was originally produced by a high temperature electrochemical reaction of sand and carbon. The SiC crystals consist of 50% carbon atoms covalently bonded with 50% silicon atoms. There are over 100 different crystal structures (polytypes), each SiC polytype has its own distinct set of electrical semiconductor properties. However, only few polytypes are used for semiconductor production, the cubic 3C-SiC, hexagonal 4H-SiC, and 6H-SiC.

Mechanical and chemical properties

Silicon carbide is a very hard substance. It is chemically inert and reacts poorly (if at all) with any known material at room temperature. It is practically impossible to diffuse anything into SiC. Dopants need to be implanted or grown into the material. Furthermore, it lacks a liquid phase and instead sublimes at temperatures above 1800°C. The vapor constituents during sublimation are mainly Si, Si2C, and SiC2 in specific ratios, depending on the temperature.

Band gap

The band gap varies depending on the polytype between 2.36 eV for 3C-SiC to 3.33 eV for 2H-SiC. The most commonly used polytype is 4H-SiC, which has a band gap of 3.265 eV. The WBG makes it possible to use SiC for very high temperature operation. Thermal ionization of electrons from the valence band to the conduction band, which is the primary limitation of Si-based devices during high temperature operation, is not a problem for SiC-based devices because of this WBG.

Critical field

For power device applications, perhaps the most notable and most frequently quoted property is the breakdown electric field strength, Ec(max). This property determines how high the largest field in the material may be before material breakdown occurs. This type of breakdown is obviously referred to as catastrophic breakdown. Curiously, the absolute value of Ec(max) for SiC is frequently quoted as the relative strength of the Ec(max) against that of Si. Most discussions on this subject note that Ec(max) of SiC is 10 times that of Si. As with Si, there exists a dependence of Ec(max) with doping concentration. Thus, for a doping of approximately 1016 cm−3, Ec(max) is 2.49 MV/cm. For Si, the value of Ec(max) is about 0.401 MV/cm for the same doping. As can be seen, the value for SiC is only about a factor of six higher than that of Si and not the often-claimed 10 times higher critical field strength. Why the discrepancy? It is more correct to compare the critical strengths between devices made for the same blocking voltage. Thus, a Si device constructed for a blocking voltage of 1 kV would have critical field strength of about 0.2 MV/cm, which should be compared to the 2.49 MV/cm of SiC. This is where the order of magnitude larger breakdown field spec comes from.

Saturated drift velocity

For high-frequency devices, the breakdown electric field strength is not as important as the saturated drift velocity. In SiC, this is 2 × 107 cm/s, which is twice that of Si. A high-saturated drift velocity is advantageous to obtain as high channel currents as possible for microwave devices, and clearly SiC is an ideal material for high-gain solid-state devices.

Thermal conductivity

The second most important parameter for power and high-frequency device applications is the material's thermal conductivity. An increase in temperature generally leads to a change in the physical properties of the device, which normally affects the device in a negative way. Most important is the carrier mobility, which decreases with increasing temperature. Heat generated through various resistive losses during operation must be conducted away from the device and into the package. More detailed studies have been made where the thermal conductivity in the different crystal directions have been determined for SiC. As can be seen, there is dependence on the purity of the crystal as well as on the crystal direction. High-purity semi-insulating SiC material has the highest reported thermal conductivity with a value of 4.9 W/(cm K). Lower values are measured for the doped crystals but they are all above 4 W/(cm K) at room temperature. Finally, Fig. 10.51 presents the advantageous properties of SiC against Si.

Figure 10.51. Properties of Si and SiC materials.

10.7.3 Silicon Carbide Devices

In this section, an overview of the available SiC devices is given. The dramatic quality improvement of the SiC material in combination with excellent research and development efforts on the design and fabrication of SiC devices by several research groups has recently resulted in a strong commercialization of SiC switch-mode devices. Nevertheless, the SiC device market is still in an early stage, and today, some available SiC switches are the JFET, BJT, MOSFET, and Schottky barrier diodes (SBDs). Commercially available SiC devices are still not in mass production. Finally, it is also worth mentioning the progress of the research on the SiC IGBT.

10.7.3.1 SiC Schottky Barrier

The SiC version of this p–n junction may offer many improvements in contrast to Silicon's version. Silicon carbide SBDs have been available for more than a decade but were not commercially viable until recently. The highest performance silicon power diodes are SBDs. Not only do SBDs have the lowest reverse recovery time (trr) compared to the various types of fast recovery (fast recovery epitaxial), ultrafast recovery, and superfast recovery diodes, they also have the lowest forward voltage drop (VF).

10.7.3.2 SiC MOSFET

Several years have been spent on the research and development of the SiC MOSFET. The fabrication and stability of this oxide layer has been challenging. SiC oxides are not showing the same reliability as in Si MOSFETs. They have higher threshold voltage shifts, gate leakage, and oxide failures than comparably biased silicon MOSFET's. This has detrimental impact on SiO2 electrical quality. Consequently, much longer and higher temperature maintenance (annealing) is required to improve the SiC oxide quality. A cross section of a popular commercialized vertical double implanted SiC MOSFET structure is shown in Fig. 10.52. The normally-off behavior of the SiC MOSFET makes it attractive to the designers of power electronic converters. Unfortunately, the low channel mobility cause additional conduction-state resistance of the device, and thus increases conduction-state power losses. Additionally, the reliability and the stability of the gate oxide layer, especially over long time periods and at elevated temperatures, have not been confirmed yet. Fabrication issues also contribute to the deceleration of SiC MOSFET development.

Figure 10.52. Cross section of 4H-SiC vertical double implanted MOSFET.

10.7.3.3 SiC-Based IGBT

The Si-based IGBT has shown an excellent performance for a wide range of voltage and current ratings during the last two decades. The fabrication of a Si n-type IGBT started on a p-type substrate. Such substrates are also available in SiC, but their resistivity is unacceptably high and prevents these components from being used in power electronics applications. Furthermore, the performance of the gate oxide layer is also poor, resulting in high channel resistivities. These issues have already been investigated by many highly qualified scientists, and it is believed that such SiC devices will not be commercialized within the next 10 years.

10.7.3.4 SiC Bipolar Junction Transistor

The SiC BJT is a bipolar normally-off device, which combines both a low conduction-state voltage drop (0.32 V at 100 A/cm2) and a quite fast switching performance. A cross section of this device is shown in Fig. 10.53 where it is obvious that this is an npn BJT. The low conduction-state voltage drop is obtained because of the cancellation of the base–emitter and base–collector junction voltages. The SiC BJT is a current-driven device, which means that a substantial continuous current is required as long as it conducts a collector current.

Figure 10.53. Cross section of SiC BJT.

10.7.3.5 SiC Junction Field Effect Transistor

A JFET has no SiO2–SiC interface and could, therefore, be available as a commercial power device in the next few years. The high quality of the conduction channel and good control of the channel dimensions and doping are crucial for the JFET performance. The normally-on JFET design is capable of extremely low conduction-state resistance. Normally-on JFETs are not easily accepted by the market due to system safety requirements, regardless of their excellent on resistance. Normally-off JFETs on the other hand require a narrow and relatively low-doped channel to ensure the N-off performance, and thus pay a penalty in terms of the conduction-state performance. Normally-off JFETs are also vulnerable to the electromagnetic interference (EMI) noise due to the small range of the gate control voltage. Hence, the gate control circuitry for JFETs requires special attention to ensure reliable operation. In the case of normally-on JFETs the development of inherently safe gate drivers is particularly desired to guarantee the safety of the whole system. A significant difference between the JFET and the MOSFET is that the MOSFET is normally-off, whereas the JFET can be either normally-on or normally-off. Normally-on means that the JFET conducts when no voltage is applied to the gate. Fig. 10.54 shows the JFETs symbols and the equivalent circuit of the n-type.

Figure 10.54. SiC junction field effect transistor symbols and equivalent circuit.

(a) Symbols; (b) equivalent circuit of the n-type.

Fig. 10.55 presents the iD–vGS characteristic in the saturation region of a SiC JFET. As can be seen from Fig. 10.55, the slope of the characteristic is equal to β0.5, and the value of the threshold voltage is defined as that voltage of vGS for which the drain current ID is approximately zero. Moreover, Fig. 10.56 presents the iD–vDS characteristic of a SiC JFET (Table 10.3).

Figure 10.55. Normally-on SiC junction field effect transistor iD–vGS input characteristic for a certain value of VDS.

Figure 10.56. Normally-on SiC junction field effect transistor iD–vDS output characteristic.

Table 10.3. SiC junction field effect transistor model parameters

NameParameterUnits
VTOThreshold voltage, VthV
BetaTransconductance, βA/V2
LamdaChannel length modulation parameter, λV−1
RdDrain ohmic resistanceOhm
RsSource ohmic resistanceOhm
IsGate junction saturation currentA
CgsZero-bias G–S junction capacitanceF
CgdZero-bias G–D junction capacitanceF
PBGate junction potentialV
MJunction grading coefficient
KFFlicker-noise coefficient
AFFlicker-noise exponent
FCCoefficient for forward-bias depletion capacitance formula
TNOMParameter measurement temperature°C
XTIIS temperature coefficient
VTOTCThreshold voltage temperature coefficientV/°C
BETATCTransconductance exponential temperature coefficient°C−1

Lateral channel junction field effect transistor (LCJFET)

The most successful JFET type in terms of voltage and current ratings has been the lateral channel JFET. The cross section of the LCJFET is shown in Fig. 10.57. The LCJFET allows optimal control of the channel parameters and offers the largest ease of fabrication compared to other concepts. It also offers the use of the inherent body diode as an antiparallel diode in switching applications since the buried gate is preferably connected to source. This is necessary to reduce the Miller capacitance, and thus maintain high speed operation. The original LCJFET structure uses ion-implantation for the gate and the base region, and planar epitaxial growth for the defect-free channel layer. This leads to advantages in terms of ease of fabrication, freedom of parameter choice due to a wide design window, and small fabrication tolerances. Its main disadvantage is a relative large specific on resistance, which is related to the large cell pitch due to the lateral configuration of the channel.

Figure 10.57. Cross section of SiC lateral channel junction field effect transistor.

Vertical trench junction field effect transistor (VTJFET)

The cross section of a VTJFET is shown in Fig. 10.58. The VTJFET SiC JFET can be either a normally-off (enhancement-mode VTJFET—EMVTJFET) or a normally-on (depletion-mode VTJFET—DMVTJFET) device, depending on the thickness of the vertical channel and the doping levels of the structure. As other normally-on JFET designs, a negative gate-source voltage is necessary to keep it in the cut-off mode. On the other hand, a significant gate current (approximately 200 mA for a 30-A device) is necessary for the normally-off JFET to keep it in the conduction mode. The pinch-off voltage for the DMVJFET equals approximately −6 V, whereas the positive pinch-off voltage for the normally-off one is slightly higher than 1 V. Comparing this type of SiC JFETs to the LCJFET there is no physical antiparallel body diode in this design. However, the VTJFET can conduct current in the reverse direction. When the vertical channel JFETs are used for freewheeling purpose, the reverse-recovery currents are only caused by the depletion charge of the device and no minority injection is involved.

Figure 10.58. Cross-section of vertical trench junction field effect transistor.

Fig. 10.59 presents the heat sinks of two inverter stages of the same power one implemented with Si IGBTs and the other with SiC JFETs. As can be seen from Fig. 10.59, the power semiconductor stage of the SiC inverter is exhibiting higher power density (W/cm2) due to the higher efficiency.

Figure 10.59. Heat sinks of two inverter stages of the same power but with different semiconductor devices.

(a) Using Si semiconductor devices; (b) using SiC semiconductor devices.

Fig. 10.60 shows the power circuit of a three-phase voltage source inverter implemented with SiC JFETs.

Figure 10.60. Three-phase inverter implemented using normally-on SiC junction field effect transistors.

Fig. 10.61 shows a driving circuit of a normally-on SiC JFET. Since the normally-on SiC JFET needs 0 V at its gate to turn on and −15 V to turn off the conventional pulses of +15 are converted through the totem pole MOSFETs (IXDD_609) to −15 V pulses.

Figure 10.61. Driving circuit for normally-on SiC junction field effect transistors.

Fig. 10.62 shows a driving circuit of a normally-off SiC JFET.

Figure 10.62. Driving circuit for normally-off SiC junction field effect transistors.

When an inverter is implemented with SiC JFETs, it is important to solve the Miller effect; otherwise unwanted overvoltage effects may appear across the semiconductor devices. Fig. 10.63 shows the Miller effect on one phase leg of an inverter. Specifically, when the upper SiC JFET is in the turn-on transition and the lower is in the cut-off mode, then the voltage of the lower JFET drain increases very fast from 0 to +Vin with a very high dv/dt. Since the semiconductor switch S2 is conducting the parasitic current ICGD, which is generated from Miller capacitance CGD, will flow through the resistor Rgl toward the ground increasing the voltage VGS. If VGS =  Vgg + Rgl·IGCD is increased above its threshold voltage value, then S2 might go faulty to conduction and together with S1 a short-circuit is created on the phase leg which will be catastrophic for both semiconductor switches.

Figure 10.63. SiC junction field effect transistor Miller effect on an inverter leg.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B978012811798900010X