Neutron detection

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by 109.154.88.88 (talk) at 21:19, 3 June 2011 (→‎Signatures by which a neutron may be detected). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

Neutron detection is the effective detection of neutrons entering a well-positioned detector. There are two key aspects to effective neutron detection: hardware and software. Detection hardware refers to the kind of neutron detector used (the most common today is the scintillation detector) and to the electronics used in the detection setup. Further, the hardware setup also defines key experimental parameters, such as source-detector distance, solid angle and detector shielding. Detection software consists of analysis tools that perform tasks such as graphical analysis to measure the number and energies of neutrons striking the detector.

Basic physics of neutron detection

Signatures by which a neutron may be detected

Atomic and subatomic particles are detected by the signature which they produce through interaction with their surroundings. The interactions result from the particles' fundamental characteristics:

  • Charge: Neutrons are neutral particles and do not ionize directly; hence they are harder than charged particles to detect directly. Further, their paths of motion are not affected by electric and magnetic fields.
  • Mass: The neutron mass of 1.0086649156(6) u.[1] is not directly detectable, but does influence reactions through which it can be detected.
  • Reactions: Neutrons react with a number of materials through elastic scattering producing a recoiling nucleus, inelastic scattering producing an excited nucleus, or absorption with transmutation of the resulting nucleus. Most detection approaches rely on detecting the various reaction products.
  • Magnetic moment: Although neutrons have a magnetic moment of −1.9130427(5) μN, techniques for detection of the magnetic moment are too insensitive to use for neutron detection.
  • Electric dipole moment: Although the neutron is predicted to have an electric dipole moment, this has yet to be detected. Hence it is not a viable detection signature.
  • Decay: Outside the nucleus, free neutrons are unstable and have a mean lifetime of 885.7±0.8 s (about 14 minutes, 46 seconds).[1] Free neutrons decay by emission of an electron and an electron antineutrino to become a proton, a process known as beta decay:[2]

n0

p+
+
e
+
ν
e
Although the
p+
and
e
produced by neutron decay are detectable, the decay rate is too low to serve as the basis for a practical detector system.

Classic neutron detection options

As a result of these properties, detection approaches for neutrons fall into several major categories[3]:

  • Absorptive reactions with prompt reactions - Low energy neutrons are typically detected indirectly through absorption reactions. Typical absorber materials used have high cross sections for absorption of neutrons and include Helium-3, Lithium-6, Boron-10, and Uranium-235. Each of these reacts by emission of high energy ionized particles, the ionization track of which can be detected by a number of means. Commonly used reactions include 3He(n,p) 3H, 6Li(n,α) 3H, 10B(n,α) 7Li and the fission of uranium.[3]
  • Activation processes - Neutrons may be detected by reacting with absorbers in a radiative capture, spallation or similar reaction, producing reaction products which then decay at some later time, releasing beta particles or gammas. Selected materials (e.g., indium, gold, rhodium, iron (56Fe(n,p) 56Mn), aluminum (27Al(n,α)24Na), niobium (93Nb(n,2n) 92mNb), & silicon (28Si(n,p) 28Al)) have extremely large cross sections for the capture of neutrons within a very narrow band of energy. Use of multiple absorber samples allows characterization of the neutron energy spectrum. Activation also allows recreation of an historic neutron exposure (e.g., forensic recreation of neutron exposures during an accidental criticality).[3]
  • Elastic scattering reactions (also referred to as proton-recoil) - High energy neutrons are typically detected indirectly through elastic scattering reactions. Neutron collide with the nucleus of atoms in the detector, transferring energy to that nucleus and creating an ion, which is detected. Since the maximum transfer of energy occurs when the mass of the atom with which the neutron collides is comparable to the neutron mass, hydrogenous[4] materials are often the preferred medium for such detectors.[3]

Types of neutron detectors

Gas proportional detectors

Gas proportional detectors can be adapted to detect neutrons. While neutrons do not typically cause ionization, the addition of a nuclide with high neutron cross-section allows the detector to respond to neutrons. Nuclides commonly used for this purpose are boron-10, uranium-235 and helium-3. Since these materials are most likely to react with thermal neutrons (i.e., neutrons which have slowed to equilibrium with their surroundings), they are typically surrounded by moderating materials.

Further refinements are usually necessary to isolate the neutron signal from the effects of other types of radiation. Since the energy of the neutron, charged particle reaction is discrete (i.e., essentially monoenergetic) while other reactions such as gamma reactions will span a broad energy range, it is possible to discriminate among the sources.

As a class, gas ionization detectors measure the number (count rate), and not the energy of neutrons.

He3 gas-filled proportional detectors

An isotope of Helium, He3 provides for an effective neutron detector material because He3 reacts by absorbing thermal neutrons, producing a 1H1 and 1H3. Its sensitivity to gamma rays is negligible. providing a very useful neutron detector. Unfortunately the supply of He3 is limited to production as a byproduct from the decay of tritium (which has a 12.3 year half-life); tritium is produced wither as part of weapons programs as a booster for nuclear weapons or as a byproduct of reactor operation.

BF3 gas-filled proportional detectors

As elemental boron is not gaseous, neutron detectors containing boron may alternately use boron trifluoride (BF3) enriched to 96% boron-10 (natural boron is 20% B-10, 80% B-11).[5]

Boron lined proportional detectors

Alternately, boron-lined gas-filled proportional counters react similarly to BF3 gas-filled proportional detectors, with the exception that the walls are coated with 10B. In this design, since the reaction takes place on the surface, only one of the two particles will escape into the proportional counter.

Scintillation neutron detectors

Scintillation neutron detectors include liquid organic scintillators,[6] crystals,[7] plastics, and scintillation fibers.[8]

Neutron-Sensitive scintillating glass fiber detectors

Neutron-Sensitive scintillating plastic fiber detectors

Neutron-Sensitive glass fiber coated with scintillating material detectors

Semiconductor neutron detectors

Semiconductors have been used for neutron detection.[9]

Neutron activation detectors

Activation samples may be placed in a neutron field to characterize the energy spectrum and intensity of the neutrons. Activation reactions which have differing energy thresholds can be used including 56Fe(n,p) 56Mn, 27Al(n,α)24Na, 93Nb(n,2n) 92mNb, & 28Si(n,p)28Al.[10]

Fast neutron detectors

Detection of fast neutrons poses a range of special problems. A directional fast-neutron detector has been developed using multiple proton recoils in separated planes of plastic scintillator material. The paths of the recoil nuclei created by neutron collision are recorded; determination of the energy and momentum of two recoil nuclei allow calculation of the direction of travel and energy of the neutron which underwent elastic scattering with them.[11]

Applications

Neutron detection is used for varying purposes. Each application has different requirements for the detection system.

  • Reactor instrumentation: Since reactor power is essentially linearly proportional to the Neutron flux, neutron detectors provide an important measure of power in nuclear power and research reactors. Boiling water reactors may have dozens of neutron detectors, one per fuel assembly. Most neutron detectors used in thermal-spectrum nuclear reactors are optimized to detect thermal neutrons.
  • Materials science: Elastic and inelastic neutron scattering enables experimentalists to characterize the morphology of materials from scales ranging from Angstroms to about one micron.
  • Cosmic ray detection: Secondary neutrons are one component of particle showers produced in Earth's atmosphere by cosmic rays. Dedicated ground-level neutron detectors, namely neutron monitors, are employed to monitor variations in cosmic ray flux.

Experimental neutron detection

Experiments that make use of this science include scattering experiments in which neutrons directed and then scattered from a sample are to be detected. Facilities include the ISIS neutron source at the Rutherford Appleton Laboratory, the Spallation Neutron Source at the Oak Ridge National Laboratory, and the Spallation Neutron Source (SINQ) at the Paul Scherrer Institute, in which the neutrons are produced by spallation reaction, and the traditional research reactor facilities in in which neutrons are produced during fission of uranium isotopes. Noteworthy among the various neutron detection experiments is the trademark experiment of the European Muon Collaboration, first performed at CERN and now termed the "EMC experiment." The same experiment is performed today with more sophisticated equipment to obtain more definite results related to the original EMC effect.

Challenges in neutron detection in an experimental environment

Neutron detection in an experimental environment is not an easy science. The major challenges faced by modern-day neutron detection include background noise, high detection rates, neutron neutrality, and low neutron energies.

Background noise

The main components of background noise in neutron detection are high-energy photons, which aren’t easily eliminated by physical barriers. The other sources of noise, such as alpha and beta particles, can be eliminated by various shielding materials, such as lead, plastic, thermo-coal, etc. Thus, photons cause major interference in neutron detection, since it is uncertain if neutrons or photons are being detected by the neutron detector. Both register similar energies after scattering into the detector from the target or ambient light, and are thus hard to distinguish. Coincidence detection can also be used to discriminate real neutron events from photons and other radiation.

High detection rates

Since the detector lies in a region of high beam activity, it is hit continuously by neutrons and background noise at overwhelmingly high rates. This obfuscates collected data, since there is extreme overlap in measurement, and separate events are not easily distinguished from each other. Thus, part of the challenge lies in keeping detection rates as low as possible and in designing a detector that can keep up with the high rates to yield coherent data.

Neutrality of neutrons

Neutrons are neutral and thus do not respond to electric fields. This makes it hard to direct their course towards a detector to facilitate detection. Neutrons also do not ionize atoms except by direct collision, so gaseous ionization detectors are ineffective.

Varying behavior with energy

Detectors relying on neutron absorption are generally more sensitive to low-energy thermal neutrons, and are orders of magnitude less sensitive to high-energy neutrons. Scintillation detectors, on the other hand, have trouble registering the impacts of low-energy neutrons.

Experimental setup and method

Figure 1 shows the typical main components of the setup of a neutron detection unit. In principle, the diagram shows the setup as it would be in any modern particle physics lab, but the specifics describe the setup in Jefferson Lab (Newport News, Virginia).

Figure 1: The experimental setup

In this setup, the incoming particles, comprising neutrons and photons, strike the neutron detector; this is typically a scintillation detector consisting of scintillating material, a waveguide, and a photomultiplier tube (PMT), and will be connected to a data acquisition (DAQ) system to register detection details.

The detection signal from the neutron detector is connected to the scaler unit, gated delay unit, trigger unit and the oscilloscope. The scaler unit is merely used to count the number of incoming particles or events. It does so by incrementing its tally of particles every time it detects a surge in the detector signal from the zero-point. There is very little dead time in this unit, implying that no matter how fast particles are coming in, it is very unlikely for this unit to fail to count an event (e.g. incoming particle). The low dead time is due to sophisticated electronics in this unit, which take little time to recover from the relatively easy task of registering a logical high every time an event occurs. The trigger unit coordinates all the electronics of the system and gives a logical high to these units when the whole setup is ready to record an event run.

The oscilloscope registers a current pulse with every event. The pulse is merely the ionization current in the detector caused by this event plotted against time. The total energy of the incident particle can be found by integrating this current pulse with respect to time to yield the total charge deposited at the end of the PMT. This integration is carried out in the analog-digital converter (ADC). The total deposited charge is a direct measure of the energy of the ionizing particle (neutron or photon) entering the neutron detector. This signal integration technique is an established method for measuring ionization in the detector in nuclear physics.[13] The ADC has a higher dead time than the oscilloscope, which has limited memory and needs to transfer events quickly to the ADC. Thus, the ADC samples out approximately one in every 30 events from the oscilloscope for analysis. Since the typical event rate is around 106 neutrons every second,[14] this sampling will still accumulate thousands of events every second.

Separating neutrons from photons

The ADC sends its data to a DAQ unit that sorts the data in presentable form for analysis. The key to further analysis lies in the difference between the shape of the photon ionization-current pulse and that of the neutron. The photon pulse is longer at the ends (or "tails") whereas the neutron pulse is well-centered.[14] This fact can be used to identify incoming neutrons and to count the total rate of incoming neutrons. The steps leading to this separation (those that are usually performed at leading national laboratories, Jefferson Lab specifically among them) are gated pulse extraction and plotting-the-difference.

Gated pulse extraction

Ionization current signals are all pulses with a local peak in between. Using a logical AND gate in continuous time (having a stream of "1" and "0" pulses as one input and the current signal as the other), the tail portion of every current pulse signal is extracted. This gated discrimination method is used on a regular basis on liquid scintillators.[15] The gated delay unit is precisely to this end, and makes a delayed copy of the original signal in such a way that its tail section is seen alongside its main section on the oscilloscope screen.

After extracting the tail, the usual current integration is carried out on both the tail section and the complete signal. This yields two ionization values for each event, which are stored in the event table in the DAQ system.

Plotting the difference

Figure 2: Expected plot of tail energy against energy in the complete pulse plotted for all event energies. Dots represent number densities of events.

In this step lies the crucial point of the analysis: the extracted ionization values are plotted. Specifically, the graph plots energy deposition in the tail against energy deposition in the entire signal for a range of neutron energies. Typically, for a given energy, there are many events with the same tail-energy value. In this case, plotted points are simply made denser with more overlapping dots on the two-dimensional plot, and can thus be used to eyeball the number of events corresponding to each energy-deposition. A considerable random fraction (1/30) of all events is plotted on the graph.

If the tail size extracted is a fixed proportion of the total pulse, then there will be two lines on the plot, having different slopes. The line with the greater slope will correspond to photon events and the line with the lesser slope to neutron events. This is precisely because the photon energy deposition current, plotted against time, leaves a longer "tail" than does the neutron deposition plot, giving the photon tail more proportion of the total energy than neutron tails.

The effectiveness of any detection analysis can be seen by its ability to accurately count and separate the number of neutrons and photons striking the detector. Also, the effectiveness of the second and third steps reveals whether event rates in the experiment are manageable. If clear plots can be obtained in the above steps, allowing for easy neutron-photon separation, the detection can be termed effective and the rates manageable. On the other hand, smudging and indistinguishability of data points will not allow for easy separation of events.

Rate control

Detection rates can be kept low in many ways. Sampling of events can be used to choose only a few events for analysis. If the rates are so high that one event cannot be distinguished from another, physical experimental parameters (shielding, detector-target distance, solid-angle, etc.) can be manipulated to give the lowest rates possible and thus distinguishable events.

Finer detection points

It is important here to observe precisely those variables that matter, since there may be false indicators along the way. For example, ionization currents might get periodic high surges, which do not imply high rates but just high energy depositions for stray events. These surges will be tabulated and viewed with cynicism if unjustifiable, especially since there is so much background noise in the setup.

One might ask how experimenters can be sure that every current pulse in the oscilloscope corresponds to exactly one event. This is true because the pulse lasts about 50 ns, allowing for a maximum of 2×107 events every second. This number is much higher than the actual typical rate, which is usually an order of magnitude less, as mentioned above.[14] This means that is it highly unlikely for there to be two particles generating one current pulse. The current pulses last 50 ns each, and start to register the next event after a gap from the previous event.

Although sometimes facilitated by higher incoming neutron energies, neutron detection is generally a difficult task, for all the reasons stated earlier. Thus, better scintillator design is also in the foreground and has been the topic of pursuit ever since the invention of scintillation detectors. Scintillation detectors were invented in 1903 by Crookes but were not very efficient until the PMT (photomultiplier tube) was developed by Curran and Baker in 1944.[13] The PMT gives a reliable and efficient method of detection since it multiplies the detection signal tenfold. Even so, scintillation design has room for improvement as do other options for neutron detection besides scintillation.

See also

References

  1. ^ a b Particle Data Group's Review of Particle Physics 2006
  2. ^ Particle Data Group Summary Data Table on Baryons
  3. ^ a b c d Tsoulfanidis, Nicholas (1995 (2nd Edition)). Measurement and Detection of Radiation. Washington, D.C.: Taylor & Francis. pp. 467–501. ISBN 1-56032-317-5. {{cite book}}: Check date values in: |date= (help); Cite has empty unknown parameter: |coauthors= (help)
  4. ^ Materials with a high hydrogen content such as water or plastic
  5. ^ Boron Trifluoride (BF3) Neutron Detectors
  6. ^ Yousuke, I. (2000). "Deterioration of pulse-shape discrimination in liquid organic scintillator at high energies". Nuclear Science Symposium Conference Record, Volume: 1. 1. IEEE: 6/219–6/221 vol.1. doi:10.1109/NSSMIC.2000.949173. {{cite journal}}: |access-date= requires |url= (help); Unknown parameter |coauthors= ignored (|author= suggested) (help)
  7. ^ Kawaguchi, N. (2009). "Study of crystal growth and scintillation properties as a neutron detector of 2-inch diameter eu doped LiCaAlF6 single crystal". Nuclear Science Symposium Conference Record (NSS/MIC). IEEE: 1493–1495. doi:10.1109/NSSMIC.2009.5402299. {{cite journal}}: |access-date= requires |url= (help); Unknown parameter |coauthors= ignored (|author= suggested) (help)
  8. ^ Miyanaga, N. (1997). "Fiber scintillator/streak camera detector for burn history measurement in inertial confinement fusion experiment". Review of Scientific Instruments. 68 (1): 621–623. Bibcode:1997RScI...68..621M. doi:10.1063/1.1147667. {{cite journal}}: |access-date= requires |url= (help); Unknown parameter |coauthors= ignored (|author= suggested) (help)
  9. ^ Mireshghi, A. (1994). "High efficiency neutron sensitive amorphous silicon pixel detectors". Nuclear Science. 41 (4, Part: 1–2). IEEE: 915–921. Bibcode:1994ITNS...41..915M. doi:10.1109/23.322831. {{cite journal}}: |access-date= requires |url= (help); Unknown parameter |coauthors= ignored (|author= suggested) (help)
  10. ^ van Eijk, C.W.E. (2005). "Development of elpasolite and monoclinic thermal neutron scintillators". Nuclear Science Symposium Conference Record. 1. IEEE: 239–243. doi:10.1109/NSSMIC.2005.1596245. {{cite journal}}: |access-date= requires |url= (help); Unknown parameter |coauthors= ignored (|author= suggested) (help)
  11. ^ Vanier, P.E (2007). "Calibration and testing of a large-area fast-neutron directional detector". Nuclear Science Symposium Conference Record. IEEE: 179–184. doi:10.1109/NSSMIC.2007.4436312. {{cite journal}}: |access-date= requires |url= (help); Unknown parameter |coauthors= ignored (|author= suggested) (help)
  12. ^ John F. Beacom and Mark R. Vagins (2004). "Antineutrino Spectroscopy with Large Water Čerenkov Detectors". Physical Review Letters. 93 (17): 171101. arXiv:hep-ph/0309300. Bibcode:2004PhRvL..93q1101B. doi:10.1103/PhysRevLett.93.171101.
  13. ^ a b Leo, W.R. (1994). Techniques for Nuclear and Particle Physics Experiments. Springer.
  14. ^ a b c Cerny, J.C., Dolemal, Z., Ivanov, M.P., Kuzmin, E.P., Svejda, J., Wilhelm, I. (2003). "Study of neutron response and n–γ discrimination by charge comparison method for small liquid scintillation detector". Nuclear Instruments and Methods in Physics Research A. 527 (3): 512–518. arXiv:nucl-ex/0311022. Bibcode:2004NIMPA.527..512C. doi:10.1016/j.nima.2004.03.179.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  15. ^ Jastaniah, S.D., Sellin, P.J. (2003). "Digital techniques for n–γ pulse shape discrimination capture-gated neutron spectroscopy using liquid". Nuclear Instruments and Methods in Physics Research A. 517: 202–210. Bibcode:2004NIMPA.517..202J. doi:10.1016/j.nima.2003.08.178.{{cite journal}}: CS1 maint: multiple names: authors list (link)