You are on page 1of 48

Sharon Harmon Term Paper: The solid-state behind ensuring Our solid State

Dr. H. Charles Solid State Physics

INTRODUCTION Proliferation of the weapons of mass destruction such as nuclear weapons is a serious threat in the world today. Since the attack on the nation in the events of September 11, 2001, preventing the spread of nuclear weapons has reached a heightened state of urgency. One way to passively determine the presence of nuclear weapons is to detect and identify characteristic signatures of highly enriched uranium and weapons grade plutonium. Neutrons and gamma rays are signatures of these materials, however with significantly different properties that weigh heavily into the means of successful detection of these nuclear materials. A lot of engineering goes into optimizing the sensitivity, resolution, discrimination, and stand-off capability of sensors and sensor systems that are capable of determining the presence of such nuclear and radiological threats, at what distances or locations, and how quickly. This is particularly important for shielded or masked nuclear and radiological materials where detectable signatures are difficult to distinguish from naturally occurring background or intentional interference. A key research objective in threat detection is maximizing sensitivity and specificity for those threats, resulting in high confidence detection and interdiction with minimal false alarms and minimal impact to the flow of commerce. Just the presence of ionizing radiation in general poses a threat to the population, which includes any particles or photons with energies above a few electron volts (eV) regardless of their intensity. Certain naturally occurring elements are not stable but slowly decompose by throwing away a portion of their nucleus; this is called radioactivity. It was discovered in 1896 by Henry Becquerel when he found that uranium atoms (Z=92) give off radiation which fogs photographic plates.[2] In addition to radioactivity that is naturally occurring, there are many man-made nuclei which are also radioactive. This introduces a variety of sources within the

electromagnetic spectrum that people need to be protected against, to include neutrons which cannot directly cause ionization but have hazardous secondary reactions. With the varying types of ionizing radiation comes the need for an array of detection mechanisms to properly identify and detect the presence of these hazardous materials. Its the charged particles interaction with matter that makes their detection possible. The most prevalent materials used for this purpose are gases one of the earliest examples being the GeigerMller counter but also liquids (e.g. liquid Argon) and even solids. Along their path of flight or in the vicinity of their interaction point, charged particles and electromagnetic radiation will ionize atoms, thus creating free electric charge carriers that may be collected and measured directly. Secondary effects are also considered, such as the generation of light in scintillators due to a recombination of ions and electrons. In the last decades, a large variety of detectors for ionizing nuclear radiation based on ionization in gases have been developed, capable of measuring both the position and energy of the radiation. Semiconductor detectors have been used for nuclear spectroscopy for quite some time, and more recently for position measurement where their development is progressing rapidly. Positionsensitive semiconductor development was initiated by experimental particle physics, which needed detectors capable of measuring particle tracks with approximately 10m precision that at the same time could operate at high rates. This was done for the purpose of investigating the rare charmed particles that were discovered in 1976. Their relatively long lifetime (10-13 to 10-12 seconds) made a direct observation of their short decay length observable with these detectors. Development of such detectors was made possible by the adaptation of microelectronic fabrication techniques used to make silicon detectors, work that was pioneered by Josef Kemmer in the 1970s.[1] Silicon strip detectors, produced with this planar technology, were introduced

into a charm search experiment in 1979 which marked the start to a revolution in experimental techniques of particle physics to include the development of low-noise-power analog microelectronics for the readout of semiconductor detectors. Another marked event was the invention of the semiconductor drift detector by Emilio Gatti and Pavel Rehak in 1983, where the signal charge transport is parallel to the semiconductor wafer surface. This principle and the possibilities opened up by the planar process were the source of many new ideas and concepts; electronics were integrated with the detector and completely new structures were invented. The primary focus of this paper will be on radiation detection using semiconductors as the interaction material, known as solid state detectors. These materials are capable of the best energy resolution, which makes them an attractive candidate for nuclear security applications where an accurate identification of threat materials is essential. An overview of the varying types of ionizing radiation will be covered with an emphasis on gamma-ray and neutron emission, as well as an introduction to the various types of radiation detection techniques, to be followed by a closer inspection of semiconductor radiation detection in particular and a survey of current semiconductor implementations from the most prevalent to more recent advances in the development of these types of detection materials for nuclear safety. NUCLEAR RADIATION

Figure 1. Spectrum of Electromagnetic Radiation

Figure 1 shows a spectrum of the electromagnetic waves. Lower-energy radiation, such as visible light, infrared, microwaves, and radio waves, are not ionizing. On the left-hand side however, there is the high-energy (shorter-wavelength) region wherein lies high-frequency Ultraviolet, X-ray, and -radiation that are ionizing due to their composition of high-energy photons. However, these are not the shortest possible wavelengths of electromagnetic waves. In addition, a spontaneous radiation from matter is not necessarily electromagnetic; there is specifically nuclear radiation which is the emission of particles from the atomic nuclei. It can be of two types: the charged ( and particles and protons) and uncharged particles, or neutrons. X- and -rays belong to the nuclear type of electromagnetic radiation of interest. [2] These are X-rays that are emitted in the rearrangement of electron shells of atoms, and gamma rays that originate from transitions within the nucleus itself. The two types of radiation emission span an energy range of interest over six decades, ranging from about 10 eV to 20 MeV. The lower energy bound is set by the minimum energy required to produce ionization in typical materials by the radiation or the secondary products of its interaction. Any radiations with energy greater than this minimum are classified as ionizing radiations. [3] The upper bound limits the set to those of primary concern in nuclear science and technology. Of these radiations, each differs in their hardness or ability to penetrate thicknesses of material. Soft radiations, such as alpha particles or low-energy X-rays penetrate only small thicknesses of material. This requires then that radioisotope sources must be deposited in very thin layers if a large fraction of these radiations is to escape from the source itself. Physically thicker sources are subject to self-absorption, which is likely to affect both the number and the energy spectrum of the radiations that emerge from its surface, and therefore typical thicknesses for such sources are measured in micrometers. [3] -particles are generally more penetrating, and sources

up to a few tenths of a millimeter in thickness can usually be tolerated. Harder radiations, such as gamma rays or neutrons, are much less affected by self-absorption and sources can be millimeters or centimeters in dimension without seriously affecting radiation properties.

Charged Particulate Radiation Fast Electrons These include beta particles (positive or negative) emitted in nuclear decay, as well as energetic electrons produced by any other process. The most common source of fast electrons in radiation measurements is a radioisotope that decays by beta-minus emission. It can be written schematically where X and Y are the initial and final nuclear species, and

is the antineutrino. Neutrinos and antineutrinos have an extremely small interaction probability with matter and therefore are undetectable for all practical purposes. Y is the recoil nucleus and appears with a very small recoil energy, which is typically below the ionization threshold, and also cannot be detected by conventional means. Thus, the only significant ionizing radiation produced by beta decay is the fast electron or beta particle itself. High-energy Betaparticles may produce bremsstrahlung as they pass through matter, or secondary (-electrons), both can ionize in turn. Energetic Beta-particles, like those emitted by 32P, are quickly decelerated by surrounding matter. The energy lost to deceleration is emitted in the form of Xrays called Bremsstrahlung, which translates braking radiation, and is of concern when shielding beta emitters. [4] As the energy of the electrons and atomic number of the absorbing medium increase, the intensity of bremsstrahlung increases.

Internal conversion is another process by which fast electrons are created, specifically called conversion electrons, and under some circumstances are nearly monoenergetic. This is useful for example if an energy calibration needs to be carried out for an electron detector, where it is much more convenient to use a source of monoenergetic electrons. Internal conversion processes begin with an excited state, which may be formed by a preceding processoften beta decay of a parent species. The common method of de-excitation is through emission of a -ray photon. For some excited states however, gamma emission may be somewhat inhibited where internal conversion becomes significant. Here the nuclear excitation energy Eex is transferred directly to one of the orbital electrons of the atom. This electron then appears with an energy given by where Eb is its binding energy in the original electron shell. [3] The analogue of internal conversion electrons are Auger electrons, which occur when the excitation energy originates in the atom rather than in the nucleus. Electron capture for instance is a proceeding process that may leave the atom with a vacancy in a normally complete electron shell. Such a vacancy is often filled by an electron from one of the outer shells of the atom with the emission of a characteristic X-ray photon. On the other hand, the excitation energy of the atom may be transferred directly to one of the outer electrons, causing it to be ejected from the atom. This electron is an Auger electron and appears with an energy given by the difference between the original atomic excitation energy and the binding energy of shell from which the electron was ejected. It is through this process that Auger electrons produce a discrete energy spectrum, with different groups corresponding to different initial and final states. When compared with beta particles or conversion electrons, their energy is relatively low, particularly because Auger electron emission is favored only in low-Z elements for which electron binding energies are small. Typical Auger electrons with a few keV initial energies are subject to

pronounced self-absorption within the source, and are easily stopped by very thin source covers or detector entrance windows. [3] Heavy Charged Particles Heavy nuclei are energetically unstable against the spontaneous emission of an alpha particle (or 4He nucleus). The probability of decay is governed by the barrier penetration mechanism, and the half-life of useful sources varies from days to many thousands of years. The decay process is written schematically as where X and Y are the initial and final nuclear

species. [3] Most alpha particles energies are limited to between about 4 and 6 MeV. There is a strong correlation between half-life of the parent isotope and alpha particle energy, and those with the highest energies have the shortest half-life. Above 6.5 MeV, the half-life is typically less than a few days, and therefore the source is of very limited utility. Conversely, if the energy drops below 4 MeV, the barrier penetration probability becomes very small and the half-life of the isotope is very large. If half-life is exceedingly long, the specific activity achievable in a practical sample of the material becomes very small, and the source is of no interest because its intensity is too low. A common source for alpha particles is 241Am, and an example of its application is the calibration of silicon solid-state detectors. Alpha particles lose energy rapidly in materials, and because of this, sources that are to be nearly monoenergetic must be prepared in very thin layers. Typical sources are covered with a metallic foil in order to contain the radioactive material, that or another material that must also be kept very thin if the original energy and monoenergetic nature of the alpha emission are to be preserved. Heavy charged particles are also created through spontaneous fission; the only spontaneous source of energetic particles with mass greater than that of the alpha particles. Detectors

intended for general application to heavy ion measurements are often tested and calibrated with these fission fragments. The spontaneous fission process in all but the extremely heavy nuclei is inhibited by the large potential barrier that must be overcome in the distortion of the nucleus from its original near-spherical shape, and is therefore not a significant process except for some transuranic (having a higher atomic number than Uranium (92)) isotopes of very large mass number.
252

Cf is the most widely used and undergoes spontaneous fission with a half-life (if it

were the only decay process) of 85 years. However, most transuranic elements undergo alpha decay, and in 252Cf the probability for alpha emission is considerably higher than that for spontaneous fission making the actual half-life for this isotope 2.65 years, and a sample of 1 microgram of 252Cf will emit 1.92x107 alpha particles and undergo 6.14x105 spontaneous fissions per second. [3] Each fission will produce two fragments emitted in opposite directions through the law of conservation of momentum. Because the normal physical form for a spontaneous fission source is a thin deposit on a flat backing, only one fragment per fission can escape from the surface, whereas the other is lost by absorption within the backing. For 252Cf in particular, each spontaneous fission process also liberates a number of fast neutrons. The fission fragments are medium-weight positive ions with a mass distribution that is predominantly asymmetric so that the fragments are clustered into a light group and heavy group with average mass numbers of 108 and 143. Appearing initially as positive ions for which the net charge approaches that of the atomic number of the fragment, the fragment then slows down as it interacts with the matter through which it is passing, and additional electrons are picked up by the ion, reducing its effective charge. The energy shared by the two fragments averages about 185 MeV. The distribution of this energy is also asymmetric with the light fragment receiving the greater fraction. Because they also lose energy readily in solid materials,

self-absorption and energy loss of the fragments are important considerations unless the source is prepared in a very thin layer.

Uncharged Radiation There is much greater emphasis today on detecting uncharged radiation because it is this type of radiation that is emitted from threats such as highly enriched uranium and weapons grade plutonium. Characteristically, these sources emit both gamma-ray and neutron signatures, and the ability to successfully detect and identify these signatures is high priority in the development of semiconductor radiation detectors. It is important to understand how this radiation is created, and how it interacts in matter, to therefore better understand how it can be best be detected.

Figure 2. Interaction of ionizing radiation with matter [4]

10

Figure 2 illustrates the types of radiation, alpha and beta, as discussed in the previous section and introduces two types to be described below. Wavy lines are used to represent gamma-rays, while charged particles and neutrons are represented by straight lines. The little circles indicate where ionization processes occur. Electromagnetic Radiation: X-rays and Gamma-Rays Gamma-Ray Sources Gamma-rays can be created by a wide range of phenomena. -rays are most typically produced as a by-product in the decay of a parent radionuclide, such as alpha or beta decay. Gamma radiation is emitted by the excited daughter nuclei following the emission of an or particle, in its transition to a lower energy state. Emission of a gamma ray from an excited state typically only requires 10-12 seconds, whereas the beta decay for instance is a relatively slow process characterized by a half-life of hundreds of days or greater. [3,5] Therefore, the gamma rays appear with a half-life characteristic of the parent beta decay, but with an energy that reflects the energy level structure of the daughter nucleus. In certain cases, there are nuclear isomers which have half-lives that are easily measurable because the excited state is more stable than average, or metastable, with decay 100 to 1000 times longer than the average 10-12 seconds and referred to as isomeric transition. The energies of gamma-rays emitted in state-to-state transitions are very specific since nuclear states have very well-defined energies. The -rays from any one transition are nearly monoenergetic, and the inherent line width of the photon energy distribution is nearly always small compared with the energy resolution of most of todays detectors. A measurement of the detector response is therefore indicative of its own limiting resolution rather than any variation in the incident gamma-ray energy. Common gamma-ray sources based on

11

beta-decay are generally limited to energies below about 2.8 MeV. Because beta decay is accompanied by the emission of a neutrino which also carries energy away, the beta spectrum does not have sharp lines, but instead is a broad peak, and so from beta decay alone it is not possible to probe the different energy levels found in the nucleus. Gamma-ray reference sources are an essential accessory, and normally consist of samples of radioisotopes of a few microcuries (around 105 Bq) encased in plastic disks or rods, with encapsulation thickness generally large enough to stop any particulate radiation from the parent decay, and the only radiation emerging from the surface is primarily the gamma radiation produced in the daughter decay sufficiently high to permit ready energy calibration of detectors with minimal hazard. Secondary radiations such as annihilation photons or bremsstrahlung can also become significant at times as additional radiation emerging, which will be discussed shortly. Gamma radiation can also be produced when particle and antiparticle collide, and is better known as annihilation radiation. Most commonly, this refers to 511 keV gamma rays produced by an normal (negative) electron and a positron (or positive electron) colliding. The origin of this radiation lies in the fate of the positrons emitted in the primary decay process, which generally only travel a few millimeters before losing their kinetic energy where the encapsulation around the source is often sufficiently thick to fully stop the positrons. It is when their energy is very low, near the end of their range, where they combine with normal negative electrons in the absorbing materials in the process of annihilation. The original positron and electron disappear and are replaced by two oppositely directed 0.511 MeV electromagnetic photons known as annihilation radiation. This radiation is generally superimposed onto whatever additional gamma radiation that may be emitted in the subsequent decay of the daughter product.

12

Gamma rays can also be produced following nuclear reactions, and their energies are higher than those available from beta-active isotopes. One such possibility is the nuclear reaction

where the product nucleus 12C is left in an excited state, and its decay gives rise to a gamma-ray photon of 4.44 MeV energy. Another possibility is the reaction

Here the product nucleus 16O can be formed in an excited state at 6.130 MeV above the ground state and with a lifetime of about 2x10-11 seconds, which is sufficiently long to eliminate almost all Doppler effects, and the resulting gamma-ray photons are monoenergetic. Both of the above reactions can be exploited by combining a radioisotope that decays by alpha emission with the appropriate target material (either 9Be or 13C), and sources of this type are more commonly used to produce neutrons, which will also be discussed shortly. Gamma rays are also commonly emitted following the absorption or thermal neutrons by typical nuclei. These gamma rays can be produced typically from sources of thermal neutrons such as intense beams from nuclear reactors or accelerator facilities, or weaker fluxes from moderated radioisotope sources of neutrons. These neutron-capture gamma rays generally have energy ranging as high as 9 MeV, and data on the probabilities of emission per neutron capture are available and useful for detector calibrations as well. [3] There are a variety of phenomena that produce gamma rays that do not involve radioactive decay. For example, when high-energy gamma rays, electrons, or protons bombard materials, the excited atoms within emit characteristic secondary (or fluorescent) gamma rays, which are products of temporary creation of excited nuclear states in the bombarded atoms, produced in the

13

nucleus but not as a result of nuclear excitement from radioactive decay. In a terrestrial gammaray flash, a brief pulse of gamma radiation occurring high in the atmosphere of Earth, gamma rays are thought to be produced by high intensity static electric fields accelerating electrons, which then produce gamma rays by bremsstrahlung interactions with atoms in the air they collide with.[5] In astronomy, high energy gamma rays include gamma ray background produced when cosmic rays (either high speed electrons or protons) interact with ordinary matter, producing both pair-production gamma rays at 511 eV, or bremsstrahlung at energies of tens of MeV or more, when cosmic ray electrons interact with nuclei of sufficiently high atomic number. Other background sources of gamma radiation that are important considerations in the calibration of detectors are those produced from inverse Compton scattering, pulsars, magnetars, quasars, and active galaxies outside the earths atmosphere within outer space. X-Ray Sources As referenced to earlier in the discussion of fast electrons, part of their energy is converted into electromagnetic radiation in the form of bremsstrahlung, and the fraction converted into this radiation increases with increasing electron energy and is largest for absorbing materials of high atomic number. This process is key in the production of X-rays form conventional X-ray tubes. With monoenergetic electrons that slow down and stop in a given material, the bremsstrahlung energy spectrum is a continuum with photon energies that extend as high as the electron energy itself. The emission of low-energy photons predominates, and the average photon energy is a small fraction of the incident electron energy. Because the spectrum is a continuum, they cannot be applied directly to the energy calibration of radiation detectors. Additionally, bremsstrahlung is also produced by other sources of fast electrons, including beta particles, and some bremsstrahlung photons are generated by any beta-active isotope encapsulated to stop the beta

14

particles, as mentioned earlier. In addition to bremsstrahlung, characteristic X-rays are also produced when fast electrons pass through an absorber. Consequently, spectra from X-ray tubes or other bremsstrahlung sources also show characteristic X-ray emission lines super-imposed on the continuous bremsstrahlung spectrum, which introduces the topic of characteristic X-rays. When the orbital electrons in an atom are disrupted from their normal configuration by some excitation process, the atom may exist in an excited state for a short period of time. The electrons then naturally tend to rearrange themselves to return the atom to its lowest energy or ground state within a time that is characteristically a nanosecond or less in a solid material. This energy that is liberated in the transition from excited state to ground state takes the form of a characteristic X-ray photon whose energy is given by the difference between the initial and final states, the difference in binding energies of the K and L shells for instance. The fact that the energy of the characteristic X-ray is unique to each individual element, they become useful in applications such as the elemental analysis of unknown samples. Atoms in an excited state will typically experience this emission, or competitive to this process, the ejection of an Auger electron. The fluorescent yield is defined as the fraction of all cases in which the excited atom emits a characteristic X-ray photon in its de-excitation, and are often tabulated as part of spectroscopic data. There are a number of various physical processes that can lead to the population of excited atomic states form which characteristic X-rays originate. Those of the most practical importance for compact sources of characteristic X-rays are excitation by radioactive decay, and excitation by external radiation.

15

Excitation by Radioactive Decay With an electron capture decay process, the nuclear charge is decreased by one unit by the capture of an orbital electron, most often a K-electron. The capture process creates a vacancy in one of the inner shells of the resulting atom, while still retaining the right number of orbital electrons. X-rays are generated when this vacancy is subsequentially filled that are characteristic of the product element. The decay may populate either the ground state or an excited state in the product nucleus, so that the characteristic X-rays may also be accompanied by gamma rays from subsequent nuclear de-excitation.[3] As discussed in an earlier section on fast electrons, internal conversion is another process by which characteristic X-rays can be created. This conversion results in the ejection of an orbital electron from the atom, again leaving behind a vacancy, and again, it is the K-electrons that are most readily converted, and therefore the K-series characteristic X-rays are predominant. With internal conversion, gamma ray de-excitation of the nuclear state is always a competing process and thus radioisotope sources of this type usually emit gamma rays in addition to characteristic X-rays. Conversion electrons may also lead to a measurable bremsstrahlung continuum, particularly when their energy is high. Excitation by External Radiation Figure 3 is the general scheme used to generate characteristic X-rays. There is an external source if radiation (X-rays, electrons, alpha particles, etc.) that is directed to strike a target, creating excited or ionized atoms in the target. The excited atoms or ions then subsequently de-excite to the ground state through the
Figure 3. General method for the generation of characteristic X-rays from a specific target element.

16

emission of characteristic X-rays, the target can serve as a localized source of these X-rays. The energy of the emitted X-rays depends on the choice of target material. Targets with low atomic number result in soft X-rays, and high-Z targets produce harder or higher energy X-rays. The incident radiation, however, must have a higher energy than the maximum photon energy expected from the target, because the excited states leading to the corresponding atomic transition must be populated by the incident radiation. As mentioned, the incident radiation can be one of a number of forms. X-rays can be used, and the process is referred to as X-ray fluorescence, and contamination from scattered photons of the incident X-ray beam can be kept to a minimum with proper choice of target and geometry. X-ray tubes can be bulky, so alternatives to this are radioisotopes that emit low-energy photons may also be used as the source of excitation. Yet another method of exciting the target is through the use of an external electron beam, where in this case the X-ray spectrum from the target is likely to be contaminated by the continuous bremsstrahlung spectrum also generated by interaction of the incident electrons in the target material. Thin targets, however, preferentially emit the bremsstrahlung photons in the forward direction, whereas the characteristic X-rays are emitted isotropically, and so by placing the exit window at a large angle (120-180) with respect to the incident electron direction, the bremsstrahlung radiation is minimized. Another alternative is the use of heavy charged particles as incident radiation. For compact and portable sources, alpha particles emitted by radioisotope sources are the most convenient source of the incident particles. Bremsstrahlung complication can be avoided with alpha particle excitation, and is therefore capable of generating a relatively clean characteristic X-ray spectrum.

17

Both gamma- and X-radiation can also be produced from whats known as synchrotron radiation. This is when a beam of energetic electrons is bent into a circular orbit, and from electromagnetic theory, a small fraction of the beam is radiated away during each cycle of the beam. When extracted from the accelerator in a tangential direction to the beam orbit, the radiation appears as an intense and highly directional beam of photons with energy that can span the range from visible light (a few eV) through 500 MeV from more recent high-energy accelerators. Despite being limited to large-scale centralized user facilities, this unique form of electromagnetic radiation is favorable because of the high intensity and tunable energy of the available source. Neutron Radiation Neutrons generated in various nuclear processes constitute the final major category, which is often further divided into slow neutron and fast neutron subcategories. Practical isotope sources of neutrons do not exist in the same sense that gamma-ray sources are available from many different nuclei populated by beta decay. Neutron sources are much more limited, and are based on either spontaneous fission or on nuclear reactions for which the incident particle is the product of a conventional decay process.[3] Just as with the production of heavy charged particles, neutrons are also produced by spontaneous fission. Several fast neutrons are promptly emitted is each fission event, so a sample of such a radionuclide can be a simple and convenient isotopic neutron source. When used specifically as a neutron source, the isotope is generally encapsulated in a sufficiently thick container so that only the fast neutrons and gamma rays emerge from the source. As mentioned, 252Cf is the most common spontaneous fission source. Its half-life of 2.65 years is long enough to be reasonably convenient, and the isotope is one of

18

the most widely produced of all the transuranics, and 2.30x106 n/s are produced per microgram of the sample on a unit mass basis. Another source of neutrons is radioisotope (,n) sources. Because energetic alpha particles are available from the direct decay of a number of convenient radionuclides, it is possible to fabricate a small self-contained neutron source by mixing an alpha-emitting isotope with a suitable target material. There are several different target materials that can lead to (,n) reactions for the alpha particle energies that are readily available in radioactive decay. Maximum neutron yield is obtained when beryllium is chosen as the target, and neutrons are produced the the reaction which has a Q-value of +5.72 MeV. The 239Pu/Be source is

probably the most widely used of the (,n) isotopic neutron sources.[3] The neutron spectra from all such alpha/Be sources are similar, and any differences reflect only the small variations in the primary alpha energies. Then there are photoneutron sources, where some radioisotope gamma-ray emitters can also be used to produce neutrons when combined with an appropriate target material. The resulting photoneutron sources are based on supplying sufficient excitation energy to a target nucleus by absorption of a gamma-ray photon to allow the emission of a free neutron. There are only two target nuclei, 9Be and 2H, that are of any practical significance for radioisotope photoneutron sources. An advantage of photoneutron sources is that if the gamma rays are monoenergetic, the neturons are also nearly monoenergetic, but the main disadvantage is the fact that very large gamma-ray activities must be used in order to produce neutron sources of attractive intensity. Reactions from accelerated charged particles are also capable of generating neutrons. Because they are not conveniently available from radioisotopes, reactions involving incident protons,

19

deuterons, and so on must rely on artificially accelerated particles. Two of the most common reactions of this type used to produce neutrons are The D-D reaction and The D-T reaction. With a relatively small coulomb barrier between the incident deuteron and the light target nucleus, the deuterons need not be accelerated to a very high energy in order to create a significant neutron yield, and these reactions are widely exploited in neutron generators in which deuterium ions are accelerated by a potential of about 100-300kV. Matter Interaction of Gammas and Neutrons For the purposes of this discussion in relation to the nuclear radiation detection of interest, only the interactions of gamma-rays and neutrons will be looked at. Unlike the charged particle alpha and beta radiation, gamma rays do not ionize along their path, but rather interact with matter in one of three ways: the photoelectric effect, the Compton effect, and electron/hole pair production. Referring back to Figure 2, there is shown an example of the Compton effect: two sequential Compton scatterings. In every scattering event, the gamma ray transfers energy to an electron, and continues on its path in a different direction with reduced energy. All the processes by which gamma rays interact lead to the partial or complete transfer of the gamma-ray photon energy to electron energy; in that the photon either disappears entirely or is scattered through a significant angle. This behavior, again, is in marked contrast to charged particles, which slow down gradually through continuous, simultaneous interactions with many absorber atoms. When a gamma ray passes through matter, the probability for absorption in a thin layer is proportional the thickness of that layer, leading to an exponential decrease of intensity with thickness.

20

Photoelectric Effect The process of photoelectric absorption involves a photon that undergoes an interaction with an absorber atom in which the photon completely disappears, and in its place, and energetic photoelectron is ejected by the atom from one of its bound shells. The most probable origin of the photoelectron is the most tightly bound, or K, shell of the atom. The kinetic energy of the resulting photoelectron is equal to the energy of the incident gamma photon minus the binding energy of the electron. The photoelectric effect is the dominant energy transfer mechanism for X-ray and gamma ray photons with energies below 50 keV, but it is much less important at higher energies. [3,5] Compton Scattering The Compton scattering interaction process takes place between the incident gamma-ray photon and an electron in the absorbing material, and it is most often the predominant interaction mechanism for gamma-ray energies typical of radioisotope sources. The incident gamma photon loses enough energy to an atomic electron to cause its ejection, with the remainder of the original photons energy being emitted as a new, lower energy gamma photon with an emission direction different from that of the incident gamma photon, as depicted in Figure 4. The probability of Compton scatter decreases with increasing
Figure 4. Schematic Diagram of the Compton Scattering Effect

photon energy. Compton scattering is through to be the principal absorption mechanism for gamma rays in the intermediate energy range 100 keV to 10 MeV. Compton scattering is relatively independent of the atomic number of the absorbing material, which is why very dense

21

metals like lead are only modestly better shields, on a per weight basis, than are less dense materials. However, the probability of Compton scattering per atom of the absorber depends on the number of electrons available as scattering targets and therefore increases linearly with Z. [5] Pair Production When the gamma-ray energy exceed twice the rest-mass energy of an electron (1.02 MeV), the process of pair production is energetically possible. This has a very low interaction probability until the gamma-ray energy approaches several MeV and therefore pair production is predominantly confined to high-energy -rays. By interaction in the coulomb field of the nucleus, the energy of the incident photon is converted into the mass of an electron-positron pair. Any gamma energy in excess of the equivalent rest mass of the two particles (1.02 MeV) appears as the kinetic energy of the pair and the recoil nucleus. At the end of the positrons range, it combines with a free electron. The entire mass of these two particles is then converted into two gamma photons each with at least 0.51 MeV energies (or higher according to the kinetic energy of the annihilated particles). The secondary electrons (and/or positrons) produced in any of these three processes frequently have enough energy to produce much ionization themselves. [5] Gamma Ray Attenuation In looking at gamma ray attenuation, a good place to start is in looking at a schematic representation. As opposed to the heavy charged particles which are readily stopped by the materials
Figure 5. Basic schematic of the interaction potential of the various radiations

Sheet of paper

Aluminum plate

22

depicted in the image, gamma rays, depending on their energy, are capable of penetrating much more material however are eventually absorbed as they penetrates a dense material. Neutron radiation consists of free neutrons that can be blocked using light elements, like hydrogen, which slow and/or capture them. Gamma ray attenuation coefficients are determined through transmission experiments, where monoenergetic gamma rays are collimated into a narrow beam and allowed to strike a detector after passing through an absorber of variable thickness. The result is a simple attenuation process, by which each of the interaction processes removes the gamma-ray photon from the beam either by absorption or by scattering away from the detector direction, and can be characterized by a fixed probability of occurrence per unit path length in the absorber. The sum of these probabilities is just the probability per unit path length that the gamma-ray photon is removed from the beam : ( ) ( ) ( ) and is called the

linear attenuation coefficient. Gamma-ray photons can also be characterized by the mean free path , defined as the average distance traveled in the absorber before an interaction takes place, and is simply the reciprocal of the linear attenuation coefficient. Typical values of range from a few mm to tens of cm in solids for common gamma-ray energies.[3] The linear attenuation coefficient is limited however in that it varies with the density of the absorber, even when the absorber material itself remains the same. In its place is the more widely used mass attenuation coefficient defined as mass attenuation coefficient = where represents the density of the medium. The product t, known as the mass thickness of the absorber, is now the significant parameter that determines its degree of attenuation, with units historically of mg/cm2, and the thickness of absorbers used in radiation measurements is often measure in mass thickness rather than physical thickness because it is a more fundamental physical quantity.

23

Neutron Interactions Neutron radiation does not ionize atoms in the same way that charged particles such as proton and electrons do (exciting an electron), and is often called indirectly ionizing radiation because neutrons have no charge and therefore cannot interact in matter by means of the coulomb force, which dominates the energy loss mechanisms for charged particles and electrons. Referring back to Figure 2 again, the neutron is shown to collide with a proton of the target material, and then becomes a fast recoil proton that ionizes in turn, and at the end of its path, the neutron is captured by a nucleus in an (n,)-reaction that leads to the emission of a neutron capture photon. Such photons always have enough energy to qualify as ionizing radiation. Neutrons can also travel through many centimeters of material without any type of interaction and thus can be totally invisible to a detector of common size. When there is neutron interaction, it is with a nucleus of the absorbing material, and as a result may either totally disappear and be replaced by one or more secondary radiations (almost always heavy charged particles), or else the energy or direction of neutron is changed significantly. The relative probabilities of the various neutron interactions change dramatically with neutron energy.[3] It was mentioned previously how neutrons are often divided into two categories: fast and slow neutrons in somewhat of an oversimplification, with a dividing line at about 0.5 eV, or about the energy of the abrupt drop in absorption cross section in cadmium (the cadmium cutoff energy). Slow Neutron Interactions Significant interactions here include elastic scattering with absorber nuclei and a large set of neutron-induced nuclear reactions. Very little energy can be imparted onto the collision nucleus

24

because of the small kinetic energy of these slow neutrons, and consequently is not an interaction on which detectors of slow neutrons can be based. These collisions are very probable, however, and tend to bring the slow neutron into thermal equilibrium with the absorber medium before a different type of interaction takes place, and therefore many neutrons in the slow energy range will be found among these thermal neutrons (0.025 eV RT). Of greater importance are the slow neutron interactions that are neutron-induced reactions that create secondary radiations of sufficient energy to be detected directly. In most materials, the radiative capture reaction [or (n,) reaction] is the most probable and plays an important part in the attenuation or shielding of neutrons. Fast Neutron Reactions With increasing energy, the probability of most neutron-induced reactions potentially useful in detectors drops off rapidly. With these fast neutrons, the importance of scattering is greater because these neutrons can transfer an appreciable amount of energy in one collision, and the secondary radiations, or recoil nuclei, have picked up a detectable amount of energy. At each scattering site, the neutron loses energy and is thereby moderated or slowed to lower energy. The most efficient moderator is hydrogen because the neutron can lose up to all its energy in a single collision with a hydrogen nucleus, however for heavier nuclei, only a partial energy transfer is possible. If the energy of the fast neutron is sufficiently high, inelastic scattering with the nucleus can occur in which the recoil nucleus undergoes an internal rearrangement into an excited state from which it eventually releases radiation, typically gamma. The total kinetic energy of the outgoing neutron and nucleus is less than the kinetic energy of the incoming neutron, where part of the original energy is used to excite the nucleus. Inelastic scattering and the subsequent secondary gamma rays play an important role in the shielding of high-energy

25

neutrons, but are an unwanted complication in the response of most fast neutron detectors based on elastic scattering.[3,6] Cross-Section Concept The likelihood of a particular event occurring between a neutron and a nucleus is expressed through the concept of the cross-section, and the probability per unit path length is a constant for any one of the interaction mechanisms for neutrons of a fixed energy. If a large number of neutrons of the same energy are directed into a thin layer of material, some may pass through with no interaction, others may have interactions that change their directions and energies, and still others may fail to emerge from the sample, and there is a probability for each of these events. As an example, the probability of a neutron not emerging from a sample (that is, of being absorbed or captured) is the ratio of the number of neutrons that do not emerge to the number originally incident on the layer. The cross section for being absorbed is the probability of neutrons being absorbed divided by the areal atomic density (the number of target atoms per unit area of the layer). The cross section thus has the dimensions of area and has traditionally been measured in units of the barn (10-28 m2), and is given the symbol . Another way to looking at the concept of cross-section is to consider the probability of a single neutron attempting to pass through a thin layer of material that has an area A and contains N target nuclei, each of cross-sectional area s. The sum of all the areas of the nuclei is Ns. The probability of a single neutron colliding with one of these nuclei is roughly the ratioof the total target area Ns to the area of the layer A. In other words, the probability of a single neutron having a collision with a nucleus is Ns/A or (N/A)s, the areal target density times s. However, on the atomic level, cross-sections for neutron interactions are not simply the geometrical crosssectional area of the target. By replacing this s by the discussed, might be thought of as an

26

effective cross-sectional area for the interaction, and the cross-section for the interaction retains the dimensions of area that s had.[6] To summarize, a neutron can have many types of interactions with a nucleus. Figure 6 below shows the various types of interactions and their cross-sections.

Figure 6. Various categories of neutron interactions. The letters separated by commas in the parentheses show the incoming and outgoing particles. [6]

RADIATION DETECTION Before looking at the different types of radiation detectors, some of the general properties that apply to all types should be noted, to include some basic definitions of detector properties, importantly efficiency and energy resolution. The function of any radiation detector depends on the manner in which the radiation interacts with the material detector itself through some mechanism as covered in the previous sections, and in order for a detector to respond at all. Interactions, or stopping times, for radiation are very small (typically a few nanoseconds in gases or a few picoseconds in solids [3]), and in most practical situations are so short that the deposition of radiation energy can be considered instantaneous. A general approach for a wide

27

category of detectors is the appearance of a given amount of electric charge within the detector active volume, followed by some collection scheme to form the basic electric signal. Typically, the charge is collected through the imposition of an electric field within the detector, causing positive and negative charges created by the radiation to flow in opposite directions. The time required for this collection varies greatly from one detector to another. For example, in ion chambers the collection time can be as long as a few milliseconds, whereas in semiconductor diode detectors, the time is a few nanoseconds. These times reflect both the mobility of the charge carriers within the detector active volume and the average distance that must be traveled before arrival at the collection electrodes. [3] Energy Resolution Energy resolution is one the key parameters taken into consideration in the development of any good radiation detector. For many applications of radiation detectors, the object is to measure the energy distribution of the incident radiation, classified under the general term radiation spectroscopy. By
Figure 7. A simple pulse-height spectrum (such a spectrum might be recorded from a scintillator for a single energy gamma-ray source) showing the definition of energy resolution R

measuring its response to a monoenergetic source of radiation, this important property can be examined. Figure 7 is an example of what is called a response function of a detector. Good resolution is defined by a tighter deviation around the peak centroid, or smaller full width at half maximum (FWHM), and will change depending on the radiation energy. An ideal detector would have as many charge carriers generated per event as possible, so that this limiting resolution

28

would be as small a percentage as possible. The popularity of solid-state detectors stems from the fact that a very large number of carriers are generated in these devices per unit energy lost by the incident radiation. Detection Efficiency Radiation detectors are inherently designed to give rise to an output puls for each quantum of radiation that interacts within its active volume. Detector efficiency is defined as the percentage of ionizing radiation hitting the detector that is successfully detected and measured. For charged particle radiation, ionization takes places immediately upon entry of the particle into the active volume, and after traveling a small fraction of its range, a typical particle will form enough ion pairs along its path to ensure that the resulting pulse is large enough to be recorded making it easy for a detector to see every alpha or beta particle for instance, that enters its active volume. Under these conditions, the detector is said to have a counting efficiency of 100%. Unlike the charged particles however, uncharged radiations such as gamma rays or neutrons must first undergo a significant interaction in the detector before detection is even possible, and detectors are often less than 100% efficient due to the long distances these radiation can travel between interactions. There are many factors that can affect the efficiency of a detector including type of detector, size and shape, distance from source, the radioisotope and type of radiation measured, backscatter, and absorption of radiation before it reaches the detector (by air and by the detector covering). It then becomes necessary to have a precise figure for the detector efficiency in order to relate the number of pulses counted to the number of neutrons or photons incident on the detector. These counting efficiencies are divided into two classes, absolute and intrinsic. Absolute efficiencies are defined as and are

29

dependent not only on detector properties, but also on the details of the counting geometry, primarily distance from the source to the detector. The intrinsic efficiency is defined as

and no longer includes the solid angle subtended by the detector as an implicit factor. The two efficiencies are related for isotropic sources by ( ), where is the solid angle of the detector seen from the actual

source position. Detector efficiencies are usually measured and quoted as absolute photopeak efficiencies for detection of gamma rays from unattenuating point sources. Thus the energy dependence is dominated by at higher energies and by at lower energies.

Detecting Neutron and Gamma-Radiation Since only the ways in which gamma rays and neutrons interact with matter were discussed at large for the purposes of this discussion, they will be the only types of radiation detection discussed. In order for a gamma ray to be detected, it must interact with matter by one or more of the mechanisms previously discussed, and that interaction must be recorded. Fortunately, the electromagnetic nature of gamma-ray photons allows them to interact strongly with the charged electrons in the atoms of all matter. The key process by which a gamma ray is detected is ionization, where it gives up part or all of its energy to an electron. The ionized electrons collide with other atoms and liberate many more electrons. The liberated charge cause electrons to move in some way, and a current is created, amplified, and collected, either directly (as with a proportional counter or a solid-state semiconductor detector) or indirectly (as with a scintillation detector), in order to register the presence of the gamma ray and measure its energy. The final result is an electrical pulse whose voltage is proportional to the energy deposited in the detecting

30

medium. [7] Gamma ray detection techniques, however, have one significant drawback. In the presence of dense, high atomic number surrounding material, gamma ray attenuation can be significant, which can mask the gamma ray signature of special nuclear materials (SNM). Neutrons, on the other hand, penetrate dense, high atomic number materials easily compared to -rays. For heterogeneous or dense materials such as samples of metal, oxides, and nuclear waste, gamma-ray attenuation is too high to permit accurate corrections of the measured signal. Under these circumstances, passive assay techniques based on neutron detection are preferable. Neutron detection is important in monitoring of special nuclear materials because neutrons are indicators of the presence of spontaneously fissioning isotopes (such as plutonium and californium) and induced fissions (as in uranium). Thus, neutron detection is an important component of the detection techniques used in identifying special nuclear materials. At present, there is a real need for an efficient, solid state detection system that allows neutron detection with an ability to discriminate gamma-events from neutron ones. Gamma discrimination is critical because gamma-rays are a very common background in neutron detection environment during SNM monitoring. [8] Neutrons are generally detected through nuclear reactions that result in prompt energetic charged particles such as protons, alpha particles, and so on. Most neutron detectors involve the combination of a target material designed to carry out this conversion together with one of the conventional radiation detectors, as outlined later in this paper. Because cross-section depends heavily on the neutrons energy, rather different techniques have been developed for neutron detection in different energy regions; the two energy types distinguished earlier, slow and fast neutrons.

31

Nuclear reactions that are useful in neutron detection are the only ones well consider. The cross-section for these reactions must be as large as possible so that efficient detectors can be built with small dimensions, particularly important when the detection material is incorporated as a gas. Also for the same reason, the target nuclide should either be of high isotopic abundance in the natural element, or alternatively, an economic source of artificially enriched samples should be available for detector fabrication. As mentioned, in many applications intense fields of gamma rays are also found with neutrons and the choice of reaction affects the capability of discriminating against these gamma rays in the detection process. An important value called the Q-value of the reaction, and it determines the energy liberated in the reaction following neutron capture. The higher the Q-value, the greater is the energy given to the reaction products, and the easier is the task of discriminating against gamma-ray events using simple amplitude discrimination. There is a common reaction that is used to detect neutrons, and they all also result in heavy charged particles :

Some of the most common reactions for the conversion of slow neutrons into directly detectable particles include the 10B(n,), the 6Li(n,), the 3He(n,p), the Gadolinium neutron capture, and neutron-induced fission reactions. [3] In the later sections where different types of detectors are discussed, we will touch upon some of these reactions that serve as the basis of operation for certain detectors. 3He in particular is useful for both fast and slow neutron detection. Fast neutron devices must employ a modified or completely different detection scheme to yield an instrument with acceptable detection efficiency. The most important additional conversion

32

process useful for fast neutrons is elastic neutron scattering, where a recoil nucleus is produced. Fast neutrons produce recoil nuclei that can be readily detected directly and are the primary basis for the operation of a fast neutron detector. By far the most popular target nucleus is hydrogen. As discussed earlier, an incident neutron can transfer up to all of its entire energy in a single collision with heavy nuclei making the resulting recoil protons relatively easy to detect. The inherently low detection efficiency for fast neutrons of any slow neutron detector can be somewhat improved by surrounding the detector with a few centimeters of hydrogen. The fast neutron incident on the detector can then lose a fraction of its initial kinetic energy in the moderator before reaching the detector as a lower-energy, or slow, neutron for which the detector is originally designed. A second factor tends to decrease the efficiency however, and with increasing moderator thickness: the probability that an incident fast neutron ever reaches the detector. By careful choice of the diameter and composition of the moderator-detector system, its overall efficiency versus energy curve can often be tailored to suit a specific application. [3]

TYPES OF RADIATION DETECTORS There are three general types of radiation detector: the scintillation detector, the gaseous detector, and the solid-state detector. Further, all detectors can be divided into two groups according to their functionality: the collision detector which merely detect the presence of a radioactive particle, and the energy detector that can measure the radiative energy. In other words, all detectors can be either quantitative or qualitative.[2] This paper will briefly introduce the first two types of detectors, with a closer inspection of the third type: the solid-state, or semiconductor, detectors.

33

Scintillating Detectors The operating principle of these detectors is based on the ability of certain materials to convert nuclear radiation into light. The sensitive volume of a scintillation detector is a luminescent material (a solid, liquid, or gas) that is viewed by a device that detects the gamma-ray-induced light emissions, usually in the form of a photomultiplier tube (PMT). The scintillation material can be either organic or inorganic, with the latter being more common. Examples of organic scintillators are anthracene, plastics, and liquids.[7] Some common inorganic scintillation materials are sodium iodide (NaI), cesium iodide (CsI), zinc sulfide (ZnS), and lithium iodide (LiI). Solid is the most common form of scintillator detector, and the most popular are the inorganic crystals NaI and CsI. No one scintillating material meets all the criteria for an ideal detector which include the following properties: [3] 1) It should convert the kinetic energy of charged particles into detectable light with a high scintillation efficiency. 2) This conversion should be linear. 3) The medium should be transparent to the wavelength of its own emission for good light collection. (Often achieved through activator impurities) 4) The decay time of the induced luminescence should be short so that fast signal pulses can be generated. 5) The material should be of good optical quality and subject to manufacture in sizes large enough to be of interest as a practical detector. 6) Its index of refraction should be near that of glass (~1.5) to permit efficient coupling of the scintillation light to a photomultiplier tube or other light sensor.

34

The choice of a particular scintillator is always a compromise among these and other factors. The inorganics usually have the best light output and linearity, but are relatively slow in response time; whereas organic scintillators are typically faster, but yield less light. The high Z-value of the constituents and high density of inorganic crystals favor their choice for gamma-ray spectroscopy, whereas organics are often preferred for fast neutron detection because of their hydrogen content. Scintillation light is emitted isotropically, so the scintillator is typically surrounded with reflective material (such as MgO) to minimize the loss of light and then is optically coupled to the photocathode of the PMT. As shown in Figure 8, photons incident on this photocathode liberate electrons through the photoelectric effect, and these photoelectrons are then accelerated by a strong electric field in the PMT, and multiplied (by a factor of 104 or more from its initial value at the photocathode surface). [7]

Figure 8. Typical arrangement of components in a scintillation detector [7]

PMTs have the disadvantage however of larger size constraints making them a less favorable candidate for certain applications and this is where solid-state alternatives are looked at to replace the PMT, such a silicon photodiodes or silicon drift detectors.

35

Ionization/Gaseous Detectors These detectors rely on the ability of some gaseous and solid materials to produce ion pairs in response to the ionization radiation, where then positive and negative ions can be separated in an electrostatic field and measured. In other words, these detectors rely heavily on the mechanisms described earlier in this paper, when ionizing radiation interacts with matter resulting in pair production along its path through the detector. These types of detectors are the oldest and most widely used.[2]

Figure 9. The equivalent circuit for a gas-filled detector. [7]

Gas counters consist of a sensitive volume of gas between two electrodes, as depicted in Figure 9. The ionizing particle causes ionization and excitation of gas molecules along its passing track, and at a minimum, the particle must transfer an amount of energy equal to the ionization energy of the gas molecule to permit the ionization process to occur. In most designs the outer electrode is the cylindrical wall of the gas pressure vessel, and the inner (positive) electrode is a thin wire positioned at the center of the cylinder. In some designs, especially ionization chambers (where the voltage between electrodes is low enough that only the primary ionization charge is collected), both electrodes can be positioned in the gas separate from the gas pressure vessel. If the voltage is increased from that of an ionization chamber, to where ionized electrons attain

36

enough kinetic energy to cause further ionizations, one then has a proportional chamber where the output signal is proportional to the energy deposited in the gas by the incident gamma-ray photon, and the energy resolution is intermediate between NaI scintillation counters and germanium (Ge) solid-state detectors. Due to the gas multiplication, the output pulses of proportional chambers are much stronger than in conventional ion chambers. These counters are generally employed in the detection and spectroscopy of low-energy X-radiation and for the detection of neutrons. If the operation voltage is increased even further, charge multiplication in the gas volume increases (avalanches) until the space charge produced by the residual ions inhibits further ionization. As a result, the amount of ionization saturates and becomes independent of the initial energy deposited in the gas, and this type of detector is commonly known as the GeigerMller (GM) detector, first invented in 1928 and is still in use because of its simplicity, low cost, and ease of operation. This counter does not differentiate among the kinds of particles it detects, or their energies; it counts only the number of particles entering the detector the quantitative type of detector.

Solid-State/Semiconductor Detectors The best energy resolution can be provided by semiconductor materials, where a comparatively large number of carriers for given incident radiation event occurs. The incident ionizing
Figure 10. Generation of electron holes by absorption of photons of energies E = Eg, E > Eg, and E < Eg [1]

energy provides a photon that is absorbed, and its energy is used to lift the electron from the valence band into the conduction band, leaving behind a hole in the valence band, as depicted in Figure 10. The charged particles can be either primary radiation or a secondary particle, and the

37

overall significant effect is the production of many electron-hole pairs along the incident particle track. The electron-hole pairs in some respects are analogous to the ion pairs produced in the gas-filled detectors. When an external electric field is applied to the semiconductive material, the created carriers form a measurable electric current, these are called solid-state or semiconductor diode detectors. Another major advantage of semiconductor detectors lies in the smallness of ionization energy, its values for Si or Ge is about 3 eV, compared with 30 eV required to create an ion pair in typical gas-filled detectors. When gamma-rays enter a detector, any of the three primary interactions discussed earlier, and shown in Figure 11, may take place between the gamma rays and electrons. The photoelectric effect produces an amount of ionization corresponding directly to gamma-ray energy, whereas Compton events produce a variable amount of ionization.[9] Only if the degraded (less energetic) secondary gamma ray is fully absorbed can we get useful information from Compton interactions about the distribution of gamma-ray energies. Unlike with heavy charged particles where the amount of ionization is almost the same for each member of a monoenergetic group of particles, the ionization produced by monoenergetic gamma rays is not constant, but instead a statistical distribution of pulses of ionization is produced containing well-defined peaks that can be used to determine gamma-ray energies. The rest of the ionization pulses constitute an undesirable
Figure 11. The interaction mechanisms between gamma rays and the electrons in a material. [9]

38

background which can be reduced only when a larger detecting volume is used to prevent the escape of degraded gamma rays. [9] The ideal semiconductor detector consists of a thick slice of very pure material having almost no free carriers of either type, with a positive electrode containing no free positive holes and a negative electrode containing no free electrons. Figure 12 is a close approximation of such a structure. Heavy doping of the n-layer almost totally suppresses its hole population; doping the p-layer suppresses its free electrons. When voltage is applied in the direction as shown in Figure 12, and electric field is produced throughout the bulk of the material, with direction as so to oppose injection of majority carriers form either contact into the bulk, although aiding injection of the minority (almost non-existant) carriers. The use of homogenous Ge or Si would be totally impractical due to an excessively high leakage current caused by the materials relatively low resistivity (50kcm for Si), and voltage applied to terminals of such a detector may cause a current which is three to five orders of magnitude greater than a minute radiationinduced electric current. [2] In typical detectors, the main source of leakage current is charge generation and injection at the side surfaces of the material where the lattice suddenly terminates, and generation of holes and electrons in the main bulk of the material by thermal lattice vibrations that excite carriers from the valence into the conduction band (usually through the intermediary of traps).[9] Cooling the device is usually implemented for reducing the current due to thermal excitation, although surface effects may still be present.
Figure 12. An ideal, fully depleted detector with heavily doped surface layers of opposite types. [9]

39

Several configurations of silicon diodes are currently produced; among them are diffused junction diodes, surface-barrier diodes, ion-implanted detectors, epitaxial layer detectors, and others. The diffused junction and surface-barrier detectors find widespread applications for the detection of -particles and other short-range radiation. To summarize, a good solid-state radiation detector should possess the following properties: 1) Excellent charge transport 2) Linearity between the energy of the incident radiation and the number of electron-hole pairs 3) Absence of free charges (low leakage current) 4) Production of a maximum number of electronhole pairs per unit of radiation 5) High detection efficiency 6) Fast response speed 7) Large collection area 8) Low cost. [2] In addition to the planar configuration, the most commonly encountered detector configuration is the coaxial, depicted in Figure 13. Coaxial detectors are produced either with open-ended (the so-called true coaxial)
Figure 13. (a) open-ended, true coaxial (b) closed-ended cylindrical (c) planar [9]

or close-ended crystals [Figure 13(a-b)]. In both cases, the electric field for charge collection is primarily radial, with some axial component present in the closed-ended configuration. They can be produced with large sensitive volumes and therefore with large detection efficiencies at high gamma-ray energies.

SOLID-STATE DETECTOR MATERIALS Silicon and Germanium are by far the most popular and widely used detector materials, however they are not the ideal from certain standpoints. Silicon for one, is not efficient for the detection of

40

gamma-rays. [2] Germanium, or high-purity Ge (HPGe), possesses the most ideal electronic characteristics in this regard and is the most widely used semiconductor material in solid-state detectors. It however, must always be operated at cryogenic temperatures to reduce thermally generated leakage current, which requires a detector package that must include capacity for cooling; and this usually involves a dewar for containing the liquid coolant. Other solid-state detection media besides Ge (bandgap 0.74 eV) and Si (bandgap 1.12 eV) have been applied to gamma-ray spectroscopy. It is advantageous to have high-resolution detectors operating at room temperature, thereby eliminating the cumbersome apparatus necessary for cooling the detector crystal. Semiconductor materials that can operate at room temperature have been extensively researched and include such materials as cadmium telluride (CdTe bangap 1.47 eV), mercuric iodine (HgI2 bandgap 2.13 eV), gallium arsenide (GaAs bandgap 1.43 eV), bismuth trisulfide (Bi2S3), and gallium selenide (GaSe). The higher average atomic numbers provide greater photoelectric efficiency per unit volume of material. Below, Table 1 summarizes some of their performance characteristics.
Table 1. Comparison of several semiconductor detector materials

41

Oftentimes these detector materials are limited in their application largely because of the ability to grow crystals sufficiently large for the total detection efficiencies needed for numerous nuclear security applications, but crystal-growth technology is always improving and these and other new compound semiconductor materials may become more attractive as convenient, highresolution room-temperature detectors for gamma-ray spectroscopy of nuclear materials. CdTe Probably the most popular of the materials mentioned above, CdTe which combines a relatively high Z-value (48 and 52) and band-gap energy large enough (1.47 eV) to permit roomtemperature operation. Crystals of high quality can be grown from CdTe to fabricate the intrinsic detector. Chlorine doping is occasionally used to compensate for the excess acceptors and to make the material a near-intrinsic type. Commercially available CdTe detectors range in size from 1 to 50 mm in diameter, and can be routinely operated at temperatures up to 50C without an excessive increase in noise. Thus, there are two types of CdTe detector available: the pure intrinsic, and the doped type. The former has a high-volume resistivity up to 1010-cm, however, its energy resolution is not that high. The doped type has significantly better energy resolution, however, its lower resistivity (108-cm) leads to a higher leakage current. [2]

MORE RECENT ADVANCES CZT A more recent candidate for room temperature radiation spectrometers has been introduced within the last decade, namely the various compositions of the ternary compound Cd1-XZnXTe,

42

where the x is the blending fraction of the ZnTe in CdTe. The acronym CZT is commonly used for this material. The gamma-ray absorption efficiency is similar to that of CdTe, yet the increased bandgap reduces the intrinsic free carrier concentration and the operating leakage current below that generally experienced with CdTe devices. High resistivity CdZnTe is most often grown through the unseeded high-pressure vertical Bridgman (HPVB) technique. TlBr Thallium bromide has many properties that make it very competitive to these existing technologies. First, it is a wide bandgap semiconductor (2.7 eV), and is operable at room temperature in the same manner as CdTe or CZT. At 20C its resistivity is large enough (>1010 -cm) to minimize dark current, and therefore, electronic noise (a source of peak broadening). Second, it intrinsically has tremendous radiation stopping power via a high density (7.5 g/cm3) and high atomic number constituents (81,35). Figure 1 compares -ray stopping power amongst various detector materials and confirms the superiority of TlBr. TlBr also has some very beneficial properties that are not readily evident in its performance as a detector but can possibly lead to TlBr meeting the ERT-01 desire for cost reductive technologies. The material requires much lower processing temperatures than those required by the competing detector materials. This can greatly impact production costs and material yields. TlBr melts congruently at a modest 480 C that allows melt-based crystal growth approaches such as Bridgman and Stockbarger to produce large volume crystals. Low processing temperatures help to lessen the introduction of impurities from containment ampoules during growth and purification. It also has a cubic crystal structure that
Figure 14. Relative detector thickness needed to stop equal numbers of 662 keV -rays.

43

simplifies crystal growth and device processing. The idea is to optimize the material to be of lower cost and higher performance than CZT. [10] Boron-filled Si Pillars He-3 tube based neutron detectors and planar solid-state neutron detectors are the two currently available technologies for thermal neutron detection. Neither is well suited to current threats. Although He-3 tube based neutron detectors have efficiencies up to 70%, they are inherently costly, fragile, contain high pressure gas, and require high voltages to operate, thus restricting the environment in which they can be deployed. Also, they are not suitable as handheld devices. Lastly, there is an increased demand for He-3 alternatives with its dwindling supply, as a byproduct of Titrium, a material used by the U.S. nuclear weapons program to make warheads. On the other hand, current solid-state detectors are compact, rugged, and operate at low voltages, but their neutron detection efficiencies are less than 5%. So, high performance, low cost and portable neutron detector are not currently available but are highly desirable. Agiltron proposes a new category of solid state neutron detector which will solve the efficiency problem and offer compactness, ruggedness, low operation voltage, low cost, and radiation hardness (Boron is immune to radiation damage). In collaboration with Lawrence Livermore National Laboratories (LLNL), the development will leverage LLNLs pioneer work in solid state neutron detectors with Agiltron's expertise in sensor fabrication, system integration and technology commercialization. The new detector will have high neutron detection efficiency, integrated electronics, large array scalability, low operating voltage, low cost, ruggedness, and small size. Theoretical simulation shows the ultimate neutron detection efficiencies will exceed 85%, while keeping the overall cost of the system competitive with current He-3 tube detectors. Figure 15 is a schematic of a proposed detector structure using this material configuration. [11]

44

Figure 15. Detector structure for B-filled Si pillars neutron detector

B6 P There is an absence of semiconductor crystals and devices that respond primarily to neutrons and not gamma rays. The key reactions of interest for neutron detection are: 3He(n,p) 3H, 6Li(n,) 3H,
10

B(n,) 7Li and the fission of uranium. In these reactions, He-3, Li-6 and B-10all become

chemically altered by the nuclear reaction, making the lifetime of any such nuclear detectors limited. Hence, it is important to have available a large quantity of low cost detector material. B6P is a most promising semiconductor material for detection of thermal neutrons. It is believed that natural boron and B-10 based detectors have the potential for replacing (i) the large volume
3

He detectors and (ii) the small handheld/backpack configurations. Recently, efforts have been

taken to re-develop and optimize the process developed by the co-PI at General Electric Laboratories in 1983, to grow from the melt crystals of B6P. B6P is III-V compound semiconductor having an energy band gap of 3.3 eV. The natural abundance of B-10 is high,

45

19.8 %. The thermal neutron absorption cross-section for natural boron, B-10 and B-11 is respectively 780 barns, 3,800 barns, and 0.0001 barn. 11B is immune to radiation damage. These properties makes boron based crystals superb candidates for semiconductor neutron detectors. P has a low stopping power for gamma rays, and low absorption cross section for neutrons (0.172 barns). In contrast to He-3, the separated B-10 isotope is commercially available. The B6P crystals are transparent allowing visible inspection of the crystal quality, and perhaps detection of neutrons through scintillation. B6P can be grown from the melt under moderate gas pressures (100 atm), and can be made with high electrical resistivity, and high optical transparency. The growth rate can be ~ 5 mm/hr, so that large quantities of B6P can be produced at a low price. The proposed work will provide high mobility semiconductor crystals that have a high reaction crosssection for neutrons, but are composed of elements that are light enough to be relatively insensitive to gamma rays. [12]

In general compound semiconductors are derived from elements in groups II to VI of the periodic table. They are so useful because of their shear range of compounds available, compared to elemental semiconductors, Sn-, C, Si, and Ge. Most elements in these groups are soluble within each other, forming homogenous solid solutions. These solutions occur when atoms of a particular element are able to substitute a given constituent of a different material without altering its crystal structure. In order that atoms can form solid solutions over large ranges of miscibility, they should satisfy the Hume-Rothery rules : the two species have similar valencies, they have comparable atomic radii allowing substitution without large mechanical distortion, their electronegativities are similar to avoid the creation of inter-metallic compounds, and individually their crystal structures are the same. In addition to binary materials, such as

46

GaAs, most compounds are also soluble within each other, making it possible to synthesize ternary, quaternary and higher-order solutions, simply by alloying binary compounds together. Figure 16 illustrates this graphically. [13] It is these factors that one must consider when attempting to develop a new compound semiconductor with the desired properties to potentially make it a competitor to previously developed solid-state detector materials.

Figure 16. Diagram illustrating the relationship of the elemental and compound semiconductors.

SUMMARY The successful development of a radiation detector relies first on understanding the science behind the sources of radiation and how they interact with matter, and then on understanding the specific materials and detection mechanisms that would be most successful to the detection application at hand, whether it be heavy charged particles, X-ray, gamma, or neutron radiation detection. Even outside the field of nuclear physics, the influence of these detectors is being

47

strongly felt. In archaeology, because of semiconductor radiation detectors, specimens can now be analyzed in fine detail by observing the gamma rays emitted from a sample after neutron bombardment in a reactor, or by observing the characteristic fluorescent X-rays produced when a samples is exposed to an X-ray or gamma-ray source. Biology, geology, mining, criminology, and many types of industrial processing have taken advantage of these solid-state semiconductor radiation detectors.

REFERENCES [1] Lutz, Gerhard. Semiconductor Radiation Detectors: Device Physics. New York: Springer, 2007. [2] Fraden, J. Handbook of Modern Sensors: Physics, Designs, and Applications. 3rd ed. New York: Springer-Verlag, 2004. [3] Knoll, G.F. Radiation Detection and Measurment. 3rd ed., John Wiley & Sons, New York, 1999. [4] Ionizing Radiation. Wikipedia. 30 November 2011. Wikimedia Foundation, Inc.. 3 December 2011 (<http://en.wikipedia.org/wiki/ionizing_radiation>). [5] Gamma Ray. Wikipedia. 28 November 2011. Wikimedia Foundation, Inc.. 4 December 2011 (<http://en.wikipedia.org/wiki/gamma_ray>). [6] Rinard, P. Neutron Interactions with Matter. Passive Nondestructive Assay of Nuclear Materials. Los Alamos, NM: Los Alamos National Laboratory, 1991. pp 357-377. [7] Smith, Hastings A. Jr., and Marcia Lucas. Gamma-Ray Detectors. Passive Nondestructive Assay of Nuclear Materials. Los Alamos, NM: Los Alamos National Laboratory, 1991. pp 43-64. [8] Shah, K., Glodo, J. et al. Low-rate Production of CLYC Scintillators, Proposal for 09-101 Broad Agency Announcement for Exploratory Research for Nuclear Detection Technology. MA: Radiation Monitoring Devices, 2009. [9] Goulding, Fred S., and Yvonne Stone. Semiconductor Radiation Detectors, Science, New Series, Vol. 170, No. 3955 (Oct. 16, 1970), pp. 280-289.

48

[10] Shah, K., et al. High Sensitivity, High Resolution Radioisotopic Detection, Proposal for 06-01 - Broad Agency Announcement for Exploratory Research for Nuclear Detection Technology. MA: Radiation Monitoring Devices, 2006. [11] Shah, K., et al. High Performance Portable Neutron Detector, Proposal for 10.12 - Broad Agency Announcement for SBIR for Nuclear Detection Technology. MA: Agiltron, Inc., 2010. [12] Ostrogorsky, A., and Glen A. Slack. B6P Semiconductor Crystals for Neutron Detection, Proposal for BAA10-02 - Broad Agency Announcement for Exploratory Research for Nuclear Detection Technology. IL: Illinois Institute of Technology, 2010. [13] Owens, A., and A. Peacock. Compound semiconductor radiation detectors, Nuclear Instruments and Methods in Physics Research. A 531 (2004), pp18-37.

You might also like