28.05.2015 Views

Understanding Infrared Thermal Testing Reading 2

Understanding Infrared Thermal Testing Reading 2

Understanding Infrared Thermal Testing Reading 2

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Infrared</strong> <strong>Thermal</strong> <strong>Testing</strong><br />

<strong>Reading</strong> II<br />

My ASNT Level III<br />

Pre-Exam Preparatory<br />

Self Study Notes<br />

26th April 2015<br />

Charlie Chong/ Fion Zhang


<strong>Infrared</strong> Thermography<br />

Charlie Chong/ Fion Zhang


<strong>Infrared</strong> Thermography<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang


Fion Zhang at Shanghai<br />

27th May 2015<br />

http://meilishouxihu.blog.163.com/<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang<br />

http://greekhouseoffonts.com/


Greek letter<br />

Charlie Chong/ Fion Zhang


IVONA TTS Capable.<br />

Charlie Chong/ Fion Zhang<br />

http://www.naturalreaders.com/


<strong>Reading</strong> 1<br />

The Ultimate <strong>Infrared</strong> Handbook for R&D Professionals<br />

Contents<br />

1. IR Thermography – How It Works<br />

2. IR Detectors For Thermographic Imaging<br />

3. Getting The Most From Your IR Camera<br />

4. Filters Extend IR Camera Usefulness<br />

5. Ultra High-Speed Thermography<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Chapter 1: IR Thermography - How It Works<br />

IR Thermography Cameras<br />

Although infrared radiation (IR) is not detectable by the human eye, an IR camera can<br />

convert it to a visual image that depicts thermal variations across an object or scene.<br />

IR covers a portion of the electromagnetic spectrum from approximately 900 to 14,000<br />

nanometers (0.9 μm–14 μm). IR is emitted by all objects at temperatures above<br />

absolute zero, and the amount of radiation increases with temperature. Thermography<br />

is a type of imaging that is accomplished with an IR camera calibrated to display<br />

temperature values across an object or scene. Therefore, thermography allows one to<br />

make non-contact measurements of an object’s temperature. IR camera construction<br />

is similar to a digital video camera, see Figure 1.1.<br />

Figure 1.1. Simplified block diagram of an IR camera<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


The main components are a lens that focuses IR onto a detector, plus<br />

electronics and software for processing and displaying the signals and<br />

images. Instead of a charge coupled device that video and digital still<br />

cameras use, the IR camera detector is a focal plane array (FPA) of<br />

micrometer size pixels made of various materials sensitive to IR wavelengths.<br />

FPA resolution can range from about 160 × 120 pixels up to 1024 × 1024<br />

pixels. Certain IR cameras have built-in software that allows the user to focus<br />

on specific areas of the FPA and calculate the temperature. Other systems<br />

utilized a computer or data system with specialized software that provides<br />

temperature analysis. Both methods can supply temperature analysis with<br />

better than ±1°C precision. FPA detector technologies are broken down<br />

into two categories: thermal detectors and quantum detectors. A common<br />

type of thermal detector is an uncooled microbolometer made of a metal or<br />

semiconductor material. These typically have lower cost and a broader IR<br />

spectral response than quantum detectors.<br />

Keywords:<br />

FPA detector technologies are broken down into two categories: thermal<br />

detectors and quantum detectors.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Still, microbolometers react to incident radiant energy and are much lower<br />

and less sensitive than quantum detectors. Quantum detectors are made<br />

from materials such as InSb, InGaAs, PtSi, HgCdTe (MCT), and layered<br />

GaAs/AlGaAs for QWIP (Quantum Well <strong>Infrared</strong> Photon) detectors. The<br />

operation of a quantum detector is based on the change of state of electrons<br />

in a crystal structure reacting to incident photons. These detectors are<br />

generally faster and more sensitive than thermal detectors. However, they<br />

require cooling, sometimes down to cryogenic temperatures using liquid<br />

nitrogen or a small Stirling cycle refrigerator unit.<br />

Ga<br />

In<br />

Si<br />

As<br />

Sb<br />

Hg<br />

Cd<br />

Te<br />

Al<br />

Gallium<br />

Indium<br />

Silicon<br />

Arsenic<br />

Antimony<br />

Mercury<br />

Cadmium<br />

Tellurium<br />

Aluminum<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


IR Spectrum Considerations<br />

Typically, IR cameras are designed and calibrated for a specific range of the<br />

IR spectrum. This means that the optics and detector materials must be<br />

selected for the desired range. Figure 1.2 illustrates the spectral response<br />

regions for various detector materials. Because IR has the same properties<br />

as visible light regarding reflection, refraction, and transmission, the optics for<br />

thermal cameras are designed in a fashion similar to those of a visual<br />

wavelength camera. However, the types of glass used in optics for visible<br />

light cameras cannot be used for optics in an infrared camera, as they do not<br />

transmit IR wavelengths well enough. Conversely, materials that are<br />

transparent to IR are often opaque to visible light. IR camera lenses typically<br />

use silicon (Si) and germanium (Ge) materials. Normally Si is used for MWIR<br />

(medium wavelength IR) camera systems, whereas Ge is used in LW (long<br />

wavelength) cameras. Si and Ge have good mechanical properties, i.e., they<br />

do not break easily, they are nonhygroscopic, and they can be formed into<br />

lenses with modern turning methods. As in visible light cameras, IR camera<br />

lenses have antireflective coatings. With proper design, IR camera lenses can<br />

transmit close to 100% of incident radiation.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 1.2. Examples of detector materials and their spectral responses<br />

relative to IR mid wave (MW) and long wave (LW) bands<br />

InSb, InGaAs, PtSi, HgCdTe (MCT), and layered GaAs/AlGaAs for QWIP<br />

(Quantum Well <strong>Infrared</strong> Photon) detectors.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


<strong>Thermal</strong> Radiation Principles<br />

The intensity of the emitted energy from an object varies with temperature and<br />

radiation wavelength. If the object is colder than about 500°C, emitted radiation lies<br />

completely within IR wavelengths. In addition to emitting radiation, an object reacts to<br />

incident radiation from its surroundings by absorbing and reflecting a portion of it, or<br />

allowing some of it to pass through (as through a lens). From this physical principle,<br />

the Total Radiation Law is derived, which can be stated with the following formula:<br />

W = αW + ρW +τW,<br />

which can be simplified to: 1 = α+ ρ +τ ,<br />

The coefficients a, r, and t describe the object’s incident energy absorbtion alpha (α),<br />

reflection Rho (ρ), and transmission Tau (τ ). Each coefficient can have a value from<br />

zero to one, depending on how well an object absorbs, reflects, or transmits incident<br />

radiation. For example, if ρ = 0, τ = 0, and α = 1, then there is no reflected or<br />

transmitted radiation, and 100% of incident radiation is absorbed. This is called a<br />

perfect blackbody. In the real world there are no objects that are perfect absorbers,<br />

reflectors, or transmitters, although some may come very close to one of these<br />

properties. Nonetheless, the concept of a perfect blackbody is very important in the<br />

science of thermography, because it is the foundation for relating IR radiation to an<br />

object’s temperature.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Keywords:<br />

Object’s incident energy:<br />

• absorbtion alpha (α),<br />

• reflection Rho (ρ),<br />

• transmission Tau (τ ),<br />

• emissivity epsilon (Ɛ).<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Fundamentally, a perfect blackbody is a perfect absorber and emitter of<br />

radiant energy. This concept is stated mathematical as Kirchhoff’s Law. The<br />

radiative properties of a body are denoted by the symbol Ɛ, the emittance or<br />

emissivity of the body. Kirchhoff’s law states that α = Ɛ, and since both values<br />

vary with the radiation wavelength, the formula can take the form α(λ) = Ɛ(λ),<br />

where λ denotes the wavelength.<br />

The total radiation law can thus take the mathematical form 1 = Ɛ + ρ + τ,<br />

which for an opaque body (τ = 0) can be simplified to 1 = Ɛ + ρ or ρ = 1 – Ɛ<br />

(i.e., reflection = 1 – emissivity). Since a perfect blackbody is a perfect<br />

absorber, ρ = 0 and Ɛ = 1. The radiative properties of a perfect blackbody can<br />

also be described mathematically by Planck’s Law. Since this has a complex<br />

mathematical formula, and is a function of temperature and radiation<br />

wavelength, a blackbody’s radiative properties are usually shown as a series<br />

of curves (Figure 1.3).<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 1.3. Illustration of Planck’s Law<br />

W α T 4<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Wien’s displacement law<br />

These curves show the radiation per wavelength unit and area unit, called the<br />

spectral radiant emittance of the blackbody. The higher the temperature, the<br />

more intense the emitted radiation. However, each emittance curve has a<br />

distinct maximum value at a certain wavelength. This maximum can be<br />

calculated from Wien’s displacement law,<br />

λ max = 2898/T, or (λ max x T = 2.898 x 10 -3 m.K)<br />

where T is the absolute temperature of the blackbody, measured in Kelvin (K),<br />

and λ max is the wavelength at the maximum intensity (in μm). Using blackbody<br />

emittance curves, one can find that an object at 30°C has a maximum near<br />

10μm, whereas an object at 1000°C has a radiant intensity with a maximum<br />

of near 2.3μm. The latter has a maximum spectral radiant emittance about<br />

1,400 times higher than a blackbody at 30°C, with a considerable portion of<br />

the radiation in the visible spectrum.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdfc


Wien’s displacement law<br />

These curves show the radiation per wavelength unit and area unit, called the<br />

spectral radiant emittance of the blackbody. The higher the temperature, the<br />

more intense the emitted radiation. However, each emittance curve has a<br />

distinct maximum value at a certain wavelength. This maximum can be<br />

calculated from Wien’s displacement law,<br />

λ max = 2898/T, or (λ max x T = 2.898 x 10 -3 m.K)<br />

where T is the absolute temperature of the blackbody, measured in Kelvin (K),<br />

and λ max is the wavelength at the maximum intensity (in μm). Using blackbody<br />

emittance curves, one can find that an object at 30°C has a maximum near<br />

10μm, whereas an object at 1000°C has a radiant intensity with a maximum<br />

of near 2.3μm. The latter has a maximum spectral radiant emittance about<br />

1,400 times higher than a blackbody at 30°C, with a considerable portion of<br />

the radiation in the visible spectrum.<br />

λ max in meter, T in Kelvin.<br />

Charlie Chong/ Fion Zhang<br />

http://hyperphysics.phy-astr.gsu.edu/hbase/wien.html


Wien's Displacement Law<br />

When the temperature of a blackbody radiator increases, the overall radiated energy<br />

increases and the peak of the radiation curve moves to shorter wavelengths. When<br />

the maximum is evaluated from the Planck radiation formula, the product of the peak<br />

wavelength and the temperature is found to be a constant.<br />

Charlie Chong/ Fion Zhang<br />

http://hyperphysics.phy-astr.gsu.edu/hbase/wien.html#c3


This relationship is called Wien's displacement law and is useful for the<br />

determining the temperatures of hot radiant objects such as stars, and indeed<br />

for a determination of the temperature of any radiant object whose<br />

temperature is far above that of its surroundings.<br />

It should be noted that the peak of the radiation curve in the Wien relationship<br />

is the peak only because the intensity is plotted as a function of wavelength. If<br />

frequency or some other variable is used on the horizontal axis, the peak will<br />

be at a different wavelength.<br />

Charlie Chong/ Fion Zhang<br />

http://hyperphysics.phy-astr.gsu.edu/hbase/wien.html


Wien's displacement law states that the black body radiation curve for different<br />

temperatures peaks at a wavelength inversely proportional to the temperature. The shift of that<br />

peak is a direct consequence of the Planck radiation law which describes the spectral brightness<br />

of black body radiation as a function of wavelength at any given temperature. However it had<br />

been discovered by Wilhelm Wien several years before Max Planck developed that more general<br />

equation, and describes the entire shift of the spectrum of black body radiation toward shorter<br />

wavelengths as temperature increases. Formally, Wien's displacement law states that the<br />

spectral radiance of black body radiation per unit wavelength, peaks at the wavelength λ max<br />

given by:<br />

λ max x T = b (2.898 x 10 -3 m.K)<br />

where T is the absolute temperature in Kelvin. b is a constant of proportionality called Wien's<br />

displacement constant, equal to 2.8977721(26)×10 −3 m K. If one is considering the peak of black<br />

body emission per unit frequency or per proportional bandwidth, one must use a different<br />

proportionality constant. However the form of the law remains the same: the peak wavelength is<br />

inversely proportional to temperature (or the peak frequency is directly proportional to<br />

temperature). Wien's displacement law may be referred to as "Wien's law", a term which is also<br />

used for the Wien approximation.<br />

Charlie Chong/ Fion Zhang<br />

http://en.wikipedia.org/wiki/Wien%27s_displacement_law


Stefan-Bolzmann law<br />

From Planck’s law, the total radiated energy W from a blackbody can be<br />

calculated. This is expressed by a formula known as the Stefan-Bolzmann<br />

law:<br />

W = σ ∙T 4 (W/m 2 ), (here the emissivity Ɛ was assumed =1)<br />

where σ is the Stefan-Bolzmann’s constant (5.67 × 10 –8 W/m 2 K 4 ). As an<br />

example, a human being with a normal temperature (about 300 K) will radiate<br />

about 500W/ m 2 of effective body surface. As a rule of thumb, the effective<br />

body surface is 1m 2 , and radiates about 0.5kW - a substantial heat loss. The<br />

equations described in this section provide important relationships between<br />

emitted radiation and temperature of a perfect blackbody. Since most objects<br />

of interest to thermographers are not perfect blackbodies, there needs to be<br />

some way for an IR camera to graph the temperature of a “normal” object.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Wien's displacement law & Stefan-Bolzmann law: Blackbody Radiation,<br />

Modern Physics, Quantum Mechanics, and the Oxford Comma | Doc Physics<br />

■<br />

https://www.youtube.com/embed/GgD3Um_f0DQ<br />

Charlie Chong/ Fion Zhang<br />

https://www.youtube.com/watch?v=GgD3Um_f0DQ


Wien's displacement law & Stefan-Bolzmann law: Max Planck Solves the<br />

Ultraviolet Catastrophe for Blackbody Radiation | Doc Physics<br />

■<br />

https://www.youtube.com/embed/H-7f-3OAXm0<br />

Charlie Chong/ Fion Zhang<br />

https://www.youtube.com/watch?v=H-7f-3OAXm0


Radiation Curves (Intensity Plot)<br />

The wavelength of the peak of the blackbody radiation curve decreases in a<br />

linear fashion ? as the temperature is increased (Wien's displacement law).<br />

This linear variation is not evident in this kind of plot since the intensity<br />

increases with the fourth power of the temperature (Stefan- Boltzmann law).<br />

The nature of the peak wavelength change is made more evident by plotting<br />

the fourth root of the intensity.<br />

Charlie Chong/ Fion Zhang<br />

http://hyperphysics.phy-astr.gsu.edu/hbase/wien.html


Radiation Curves (Intensity Plot)<br />

Charlie Chong/ Fion Zhang<br />

http://en.wikipedia.org/wiki/Wien%27s_displacement_law


Summarizing<br />

How a perfect black body changes with temperatures?<br />

• The λ max changes in a linear manner ? ( inversely proportionally)<br />

according to Wien's Displacement Law<br />

• The total total radiated energy W changes in a fourth power T 4<br />

proportionally according to Stefan-Bolzmann law<br />

Charlie Chong/ Fion Zhang<br />

http://hyperphysics.phy-astr.gsu.edu/hbase/wien.html


Wien's displacement law (inverse proportional relation)<br />

λ max x T = b<br />

λ max = b∙1/T<br />

Charlie Chong/ Fion Zhang<br />

http://en.wikipedia.org/wiki/Wien%27s_displacement_law


Emissivity Ɛ<br />

The radiative properties of objects are usually described in relation to a<br />

perfect blackbody (the perfect emitter). If the emitted energy from a blackbody<br />

is denoted as W bb , and that of a normal object at the same temperature as<br />

W obj , then the ratio between these two values describes the emissivity (Ɛ) of<br />

the object, Ɛ = W obj / W bb . Thus, emissivity is a number between 0 and 1.<br />

The better the radiative properties of the object, the higher its emissivity. An<br />

object that has the same emissivity Ɛ for all wavelengths is called a greybody.<br />

Consequently, for a greybody, Stefan- Bolzmann’s law takes the form:<br />

W = Ɛ∙σ∙T 4 (W/m 2 ),<br />

which states that the total emissive power of a greybody is the same as that<br />

of a blackbody of the same temperature reduced in proportion to the value of<br />

Ɛ for the object.<br />

Keywords:<br />

Greybody<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


■ωσμ∙Ωπ∆º≠δ≤>ηθφФρ|β≠Ɛ∠ ʋ λαρτ<br />

Charlie Chong/ Fion Zhang<br />

http://www.webexhibits.org/causesofcolor/3.html


Still, most bodies are neither blackbodies nor greybodies. The emissivity<br />

varies with wavelength. As thermography operates only inside limited spectral<br />

ranges, in practice it is often possible to treat objects as greybodies. In any<br />

case, an object having emittance that varies strongly with wavelength is<br />

called a selective radiator. For example, glass is a very selective radiator,<br />

behaving almost like a blackbody for certain wavelengths, whereas it is rather<br />

the opposite for other wavelengths. Atmospheric Influence Between the<br />

object and the thermal camera is the atmosphere, which tends to attenuate<br />

radiation due to absorption by gases and scattering by particles. The amount<br />

of attenuation depends heavily on radiation wavelength. Although the<br />

atmosphere usually transmits visible light very well, fog, clouds, rain, and<br />

snow can prevent us from seeing distant objects. The same principle applies<br />

to infrared radiation.<br />

Keywords:<br />

Blackbody<br />

Greybody<br />

selective radiator<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


For thermographic measurement we must use the so-called atmospheric<br />

windows. As can be seen from Figure 1.4, they can be found between 2 -<br />

5μm, the mid-wave windows, and 7.5μm – 13.5μm, the long-wave window.<br />

Atmospheric attenuation prevents an object’s total radiation from reaching the<br />

camera. If no correction for attenuation is applied, the measured apparent<br />

temperature will be lower and lower with increased distance. IR camera<br />

software corrects for atmospheric attenuation. Typically, LW cameras in the<br />

7.5μm to13.5μm range work well anywhere that atmospheric attenuation is<br />

involved, because the atmosphere tends to act as a high-pass filter above<br />

7.5μm (Figure 4). The MW band of 3–5μm tends to be employed with highly<br />

sensitive detectors for high end R&D and military applications. When<br />

acquiring a signal through the atmosphere with MW cameras, selected<br />

transmission bands must be used where less attenuation takes place.<br />

From other reading: Short-wave systems in particular are susceptible to<br />

attenuation by the atmosphere and, as a result, correction for relative<br />

humidity and distance to object are recommended.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 1.4. Atmospheric attenuation (white areas) with a chart of the gases<br />

and water vapor causing most of it. The areas under the curve represent the<br />

highest IR transmission.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Foggy London Bridge<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Foggy London Bridge<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 4. Atmospheric attenuation (white areas) with a chart of the gases and<br />

water vapor causing most of it. The areas under the curve represent the<br />

highest IR transmission.<br />

Charlie Chong/ Fion Zhang<br />

http://www.universe-galaxies-stars.com/infrared.html


Figure 4. Atmospheric attenuation (white areas) with a chart of the gases and<br />

water vapor causing most of it. The areas under the curve represent the<br />

highest IR transmission.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Charlie Chong/ Fion Zhang


Raleigh Scattering<br />

Rayleigh scattering, named after the British physicist Lord Rayleigh is the (dominantly) elastic scattering of light<br />

or other electromagnetic radiation by particles much smaller than the wavelength of the radiation. The Rayleigh<br />

scattering does not change the state of material hence it is a parametric process. The particles may be<br />

individual atoms or molecules. It can occur when light travels through transparent solids and liquids, but is most<br />

prominently seen in gases. Rayleigh scattering results from the electric polarizability of the particles. The<br />

oscillating electric field of a light wave acts on the charges within a particle, causing them to move at the same<br />

frequency. The particle therefore becomes a small radiating dipole whose radiation we see as scattered light.<br />

Rayleigh scattering of sunlight in the atmosphere causes diffuse sky radiation, which is the reason for the blue<br />

color of the sky and the yellow tone of the sun itself.<br />

For historical reasons, Rayleigh scattering of molecular nitrogen and oxygen in the atmosphere includes elastic<br />

scattering as well as the inelastic contribution from rotational Raman scattering in air, since the changes in<br />

wave number of the scattered photon are typically smaller than 50 cm−1. This can lead to changes in the<br />

rotational state of the molecules. Furthermore the inelastic contribution has the same wavelengths dependency<br />

as the elastic part.<br />

Scattering by particles similar to, or larger than, the wavelength of light is typically treated by the Mie theory, the<br />

discrete dipole approximation and other computational techniques. Rayleigh scattering applies to particles that<br />

are small with respect to wavelengths of light, and that are optically "soft" (i.e. with a refractive index close to 1).<br />

On the other hand, Anomalous Diffraction Theory applies to optically soft but larger particles.<br />

Charlie Chong/ Fion Zhang<br />

http://en.wikipedia.org/wiki/Rayleigh_scattering


Raleigh Scattering<br />

Elastic scattering of light or other electromagnetic radiation by particles much smaller than the wavelength of<br />

the radiation.<br />

Charlie Chong/ Fion Zhang<br />

http://en.wikipedia.org/wiki/Rayleigh_scattering


Temperature Measurements<br />

The radiation that impinges on the IR camera lens comes from three different<br />

sources. The camera receives radiation from the target object, plus radiation<br />

from its surroundings that has been reflected onto the object’s surface. Both<br />

of these radiation components become attenuated when they pass through<br />

the atmosphere. Since the atmosphere absorbs part of the radiation, it will<br />

also radiate some itself (Kirchhoff’s law). Given this situation, we can derive a<br />

formula for the calculation of the object’s temperature from a calibrated<br />

camera’s output.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


1. Emission from the object = Ɛ ∙τ∙W obj ,<br />

Where Ɛ is the emissivity of the object (unit?) τ is the transmittance of the<br />

atmosphere.<br />

2. Reflected emission from ambient sources = (1 – Ɛ) ∙ τ∙W amb ,<br />

Where (1 – Ɛ) is the reflectance of the object. (It is assumed that the ambient<br />

temperature T amb is the same for all emitting surfaces within the half sphere<br />

seen from a point on the object’s surface.)<br />

3. Emission from the atmosphere = (1 – τ) ∙ W atm , where (1 – τ) is the<br />

emissivity of the atmosphere.<br />

4. The total radiation power received by the camera can now be written:<br />

W tot = Ɛ ∙τ∙W obj + (1 – Ɛ) ∙ t ∙ W amb + (1 – τ) ∙ W atm ,<br />

Where e is the object emissivity, τ is the transmission through the<br />

atmosphere, T amb is the (effective) temperature of the object’s surroundings,<br />

or the reflected ambient (background) temperature, and T atm is the<br />

temperature of the atmosphere.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


1. Emission from the object = (Ɛ obj<br />

) ∙ (τ atm<br />

∙ W obj<br />

),<br />

Where Ɛ obj<br />

is the emissivity of the object (unit?) τ atm<br />

is the transmittance of the<br />

atmosphere, W obj<br />

emitted energy fro the object.<br />

2. Reflected emission from ambient sources = (1 – Ɛ obj<br />

)∙(τ atm<br />

∙ W amb<br />

),<br />

Where (1 – Ɛ) is the reflectance of the object. (It is assumed that the ambient<br />

temperature T amb<br />

is the same for all emitting surfaces within the half sphere seen from<br />

a point on the object’s surface.), Wamb emitted energy of ambient source.<br />

3. Emission from the atmosphere = (1 – τ) ∙ W atm<br />

, where (1 – τ) is the emissivity of the<br />

atmosphere, W atm<br />

emitted energy from the atmosphere.<br />

4. The total radiation power received by the camera can now be written:<br />

W tot<br />

= Ɛ ∙τ∙W obj<br />

+ (1 – Ɛ) ∙ τ ∙ W amb<br />

+ (1 – τ) ∙ W atm<br />

,<br />

Where e is the object emissivity, τ is the transmission through the atmosphere, T amb<br />

is<br />

the (effective) temperature of the object’s surroundings, or the reflected ambient<br />

(background) temperature, and T atm<br />

is the temperature of the atmosphere.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


• absorbtion alpha (α),<br />

• reflection Rho (ρ),<br />

• transmission Tau (τ ),<br />

• emissivity epsilon (Ɛ).<br />

W amb<br />

1 = ρ obj + Ɛ obj ,where τ obj =0<br />

W obj<br />

W atm<br />

1 = τ atm + Ɛ atm ,where ρ atm =0<br />

W tot = Ɛ obj ∙τ atm ∙ W obj + (1 – Ɛ obj ) ∙ τ atm ∙ W amb + (1- τ atm ) ∙ W atm ,<br />

Emission from object<br />

Reflection from ambient<br />

Emission from atmosphere<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


To arrive at the correct target object temperature, IR camera software<br />

requires inputs for the emissivity of the object, atmospheric attenuation and<br />

temperature, and temperature of the ambient surroundings. Depending on<br />

circumstances, these factors may be measured, assumed, or found from<br />

look-up tables.<br />

Keywords:<br />

atmospheric attenuation (transmittance, τ atm ?)<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Chapter 2: IR Detectors For Thermographic Imaging<br />

IR Cameras<br />

Thermographic imaging is accomplished with a camera that converts infrared<br />

radiation (IR) into a visual image that depicts temperature variations across<br />

an object or scene. The main IR camera components are a lens, a detector in<br />

the form of a focal plane array (FPA), possibly a cooler for the detector, and<br />

the electronics and software for processing and displaying images (Figure<br />

2.1). Most detectors have a response curve that is narrower than the full IR<br />

range (900–14,000 nanometers or 0.9μm –14μm). Therefore, a detector (or<br />

camera) must be selected that has the appropriate response for the IR range<br />

of a user’s application. In addition to wavelength response, other important<br />

detector characteristics include sensitivity, the ease of creating it as a focal<br />

plane array with micrometer size pixels, and the degree of cooling required, if<br />

any.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


In most applications, the IR camera must view a radiating object through the<br />

atmosphere. Therefore, an overriding concern is matching the detector’s<br />

response curve to what is called an atmospheric window. This is the range of<br />

IR wavelengths that readily pass through the atmosphere with little<br />

attenuation. Essentially, there are two of these windows, one in the 2μm –<br />

5.6μm range, the short/ medium wavelength (SW/MW) IR band, and one in<br />

the 8μm – 14μm range, the long wavelength (LW) IR band. There are many<br />

detector materials and cameras with response curves that meet these criteria.<br />

Keywords:<br />

Atmospheric window<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.1. Simplified block diagram of an IR camera<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Quantum vs. Non-Quantum Detectors<br />

The majority of IR cameras have a microbolometer type detector, mainly<br />

because of cost considerations. Microbolometer FPAs can be created from<br />

metal or semiconductor materials, and operate according to non-quantum<br />

principles. This means that they respond to radiant energy in a way that<br />

causes a change of state in the bulk material (i.e., the bolometer effect).<br />

Generally, microbolometers do not require cooling, which allows compact<br />

camera designs that are relatively low in cost. Other characteristics of<br />

microbolometers are:<br />

• Relatively low sensitivity (detectivity)<br />

• Broad (flat) response curve<br />

• Slow response time (time constant ~12 ms)<br />

For more demanding applications, quantum detectors are used, which<br />

operate on the basis of an intrinsic photoelectric effect. These materials<br />

respond to IR by absorbing photons that elevate the material’s electrons to a<br />

higher energy state, causing a change in conductivity, voltage, or current. By<br />

cooling these detectors to cryogenic temperatures, they can be very sensitive<br />

to the IR that is focused on them.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


They also react very quickly to changes in IR levels (i.e., temperatures),<br />

having a constant response time on the order of 1μs. Therefore, a camera<br />

with this type of detector is very useful in recording transient thermal events.<br />

Still, quantum detectors have response curves with detectivity that varies<br />

strongly with wavelength (Figure 2.2). Table 2.1 lists some of the most<br />

commonly used detectors in today’s IR cameras.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.2. Detectivity (D*) curves for different detector materials<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Quantum Well <strong>Infrared</strong> Photodetector<br />

A Quantum Well <strong>Infrared</strong> Photodetector (QWIP) is an infrared photodetector,<br />

which uses electronic inter subband transitions in quantum wells to absorb<br />

photons. The basic elements of a QWIP are quantum wells, which are<br />

separated by barriers. The quantum wells are designed to have one confined<br />

state inside the well and a first excited state which aligns with the top of the<br />

barrier. The wells are n-doped such that the ground state is filled with<br />

electrons. The barriers are wide enough to prevent quantum tunneling<br />

between the quantum wells. Typical QWIPs consists of 20 to 50 quantum<br />

wells. When a bias voltage is applied to the QWIP, the entire conduction band<br />

is tilted. Without light the electrons in the quantum wells just sit in the ground<br />

state. When the QWIP is illuminated with light of the same or higher energy<br />

as the inter subband transition energy, an electron is excited.<br />

Charlie Chong/ Fion Zhang<br />

http://en.wikipedia.org/wiki/Quantum_well_infrared_photodetector


Quantum Well <strong>Infrared</strong> Photodetector<br />

Charlie Chong/ Fion Zhang<br />

http://en.wikipedia.org/wiki/Quantum_well_infrared_photodetector


Bolometer Effect<br />

A bolometer, meaning measurer of thrown things is a device for measuring the power of incident<br />

electromagnetic radiation via the heating of a material with a temperature-dependent electrical<br />

resistance. It was invented in 1878 by the American astronomer Samuel Pierpont Langley. The name<br />

comes from the Greek word bole, for something thrown, as with a ray of light<br />

Principle of operation<br />

A bolometer consists of an absorptive element, such as a thin layer of metal, connected to a thermal<br />

reservoir (a body of constant temperature) through a thermal link. The result is that any radiation<br />

impinging on the absorptive element raises its temperature above that of the reservoir — the greater<br />

the absorbed power, the higher the temperature. The intrinsic thermal time constant, which sets the<br />

speed of the detector, is equal to the ratio of the heat capacity of the absorptive element to the thermal<br />

conductance between the absorptive element and the reservoir.[2] The temperature change can be<br />

measured directly with an attached resistive thermometer, or the resistance of the absorptive element<br />

itself can be used as a thermometer. Metal bolometers usually work without cooling. They are produced<br />

from thin foils or metal films. Today, most bolometers use semiconductor or superconductor absorptive<br />

elements rather than metals. These devices can be operated at cryogenic temperatures, enabling<br />

significantly greater sensitivity.<br />

Bolometers are directly sensitive to the energy left inside the absorber. For this reason they can be<br />

used not only for ionizing particles and photons, but also for non-ionizing particles, any sort of radiation,<br />

and even to search for unknown forms of mass or energy (like dark matter); this lack of discrimination<br />

can also be a shortcoming. The most sensitive bolometers are very slow to reset (i.e., return to thermal<br />

equilibrium with the environment). On the other hand, compared to more conventional particle detectors,<br />

they are extremely efficient in energy resolution and in sensitivity. They are also known as thermal<br />

detectors.<br />

Charlie Chong/ Fion Zhang


Conceptual schematic of a bolometer.<br />

Power P from an incident signal is<br />

absorbed by the bolometer and heats<br />

up a thermal mass with heat capacity<br />

C and temperature T. The thermal<br />

mass is connected to a reservoir of<br />

constant temperature through a link<br />

with thermal conductance G. The<br />

temperature increase is ΔT = P/G.<br />

The change in temperature is read<br />

out with a resistive thermometer. The<br />

intrinsic thermal time constant is τ =<br />

C/G.<br />

Charlie Chong/ Fion Zhang


In short: Bolometer Effect<br />

A bolometer, is a device for measuring the power of incident electromagnetic<br />

radiation via the heating of a material with a temperature-dependent electrical<br />

resistance.<br />

Charlie Chong/ Fion Zhang


Bolometer:<br />

The design of cosmic microwave background experiments is a very challenging task. The greatest problems are the receivers, the<br />

telescope optics and the atmosphere. Many improved microwave amplifier technologies have been designed for microwave<br />

background applications. Some technologies used are HEMT, MMIC, SIS and bolometers. Experiments generally use elaborate<br />

cryogenic systems to keep the amplifiers cool. Often, experiments are interferometers which only measure the spatial fluctuations<br />

in signals on the sky, and are insensitive to the average 2.7 K background.<br />

Charlie Chong/ Fion Zhang


Table 2.1. Detector types and materials commonly used in IR cameras.<br />

The Kelvin is a unit of measure for temperature based upon an absolute scale. It is one of the seven base units in the International System of Units (SI) and<br />

is assigned the unit symbol K. The Kelvin scale is an absolute, thermodynamic temperature scale using as its null point absolute zero, the temperature at<br />

which all thermal motion ceases in the classical description of thermodynamics. The Kelvin is defined as the fraction 1⁄273.16 of the thermodynamic<br />

temperature of the triple point of water (exactly 0.01 °C or 32.018 °F). In other words, it is defined such that the triple point of water or approximately 0ºC is<br />

exactly 273.16 K<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


LW Photon Detector 77K.<br />

■<br />

http://irassociates.com/index.php?page=hgcdte<br />

Charlie Chong/ Fion Zhang<br />

http://www.pro-lite.uk.com/File/MCT_detectors.php


Broad Band IR- The extended spectral area of an MWIR camera simplifies<br />

the spatial assignment of gas plumes to visible objects.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


MW Photon Detector 77K.<br />

■<br />

http://irassociates.com/index.php?page=insb<br />

Charlie Chong/ Fion Zhang<br />

http://www.pro-lite.uk.com/File/InSb_detectors.php


InGaAs Detectors & Avalanche Photodiodes<br />

Large area & high sensitivity from 800 to 2200nm<br />

Pro-Lite offers Indium Gallium Arsenide (InGaAs) detectors and avalanche photodiodes (APD) from GPD Optoelectronics<br />

Corporation. InGaAs detectors operate from 800 to 1700nm and are similar to Ge but with enhanced sensitivites at low light levels.<br />

Available options include large area InGaAs (up to 5mm), custom packages and submounts, custom filters and windows,<br />

extended range InGaAs (to 2.2µm) and high speed InGaAs (up to 4GHz). GPD also produces InGaAs avalanche photodiodes<br />

(APD) with active areas of 80 and 200µm. An APD can be regarded as the semiconductor equivalent of the photomultiplier tube<br />

(PMT) detector. With the application of a reverse bias Voltage, an APD experiences a high internal current gain, making an<br />

avalanche photodiode significantly more sensitive to low light levels than a regular InGaAs detector. Our APDs can be supplied as<br />

chip/submount combinations, in TO-46 cases with flat windows or ball lenses and as fibre pigtailed packages.<br />

Charlie Chong/ Fion Zhang<br />

http://www.pro-lite.uk.com/File/InGaAs_detectors.php


Germanium Detectors & Avalanche Photodiodes<br />

Economical near-infrared detection from 800 to 1800nm<br />

Pro-Lite offers Germanium (Ge) detectors and avalanche photodiodes (APDs) from GPD Optoelectronics Corporation. Ge<br />

detectors operate from 800 to 1800nm and are similar to InGaAs but have a reduced sensitivity, are more economical and provide<br />

improved linearity at high levels of irradiance. Available options include active areas from 0.1 to 13mm diameter, two-colour<br />

silicon/germanium (Si/Ge) detectors, cryogenic cooling (TE coolers & dewars), custom packages and submounts, custom filters<br />

and windows and avalanche Ge photodiodes.<br />

Charlie Chong/ Fion Zhang<br />

http://www.pro-lite.uk.com/File/InGaAs_detectors.php


Operating Principles for Quantum Detectors<br />

In materials used for quantum detectors, at room temperature there are<br />

electrons at different energy levels. Some electrons have sufficient thermal<br />

energy that they are in the conduction band, meaning the electrons there are<br />

free to move and the material can conduct an electrical current. Most of the<br />

electrons, however, are found in the valence band, where they do not carry<br />

any current because they cannot move freely. (See left-most views of Fig 2.3.)<br />

When the material is cooled to a low enough temperature, which varies with<br />

the chosen material, the thermal energy of the electrons may be so low that<br />

there are none in the conduction band (upper center view of Figure 2.3).<br />

Hence the material cannot carry any current. When these materials are<br />

exposed to incident photons, and the photons have sufficient energy, this<br />

energy can stimulate an electron in the valence band, causing it to move up<br />

into the conduction band (upper right view of Figure 2.3).<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.3. Operating principle of quantum detectors<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


QWIP FPA mounted on a ceramic substrate and bonded to external<br />

electronics.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Thus the material (the detector) can carry a photocurrent, which is<br />

proportional to the intensity of the incident radiation. There is a very exact<br />

lowest energy of the incident photons that will allow an electron to jump from<br />

the valence band into the conduction band. This energy is related to a certain<br />

wavelength, the cutoff wavelength. Since photon energy is inversely<br />

proportional to its wavelength, the energies are higher in the SW/MW band<br />

than in the LW band. Therefore, as a rule, the operating temperatures for LW<br />

detectors are lower than for SW/MW detectors. For an InSb MW detector, the<br />

necessary temperature must be less than 173 K (–100°C), although it may be<br />

operated at a much lower temperature. An HgCdTe (MCT) LW detector must<br />

be cooled to 77 K (–196°C) or lower. A QWIP detector typically needs to<br />

operate at about 70 K (–203°C) or lower. The lower center and right views of<br />

Figure 3 depict quantum detector wavelength dependence. The incident<br />

photon wavelength and energy must be sufficient to overcome the band gap<br />

energy, ΔE.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Cooling Methods<br />

The first detectors used in infrared radiometric instruments were cooled with<br />

liquid nitrogen. The detector was attached to the Dewar flask that held the<br />

liquid nitrogen, thus keeping the detector at a very stable and low temperature<br />

(-196°C). Later, other cooling methods were developed. The first solid-state<br />

solution to the cooling problem was presented by AGEMA in 1986, when it<br />

introduced a Peltier effect cooler for a commercial IR camera. In a Peltier<br />

cooler, DC current is forced through a thermoelectric material, removing heat<br />

from one junction and creating a cold side and a hot side. The hot side is<br />

connected to a heat sink, whereas the cold side cools the component<br />

attached to it.<br />

Charlie Chong/ Fion Zhang<br />

http://www.techbriefs.com/component/content/article/124-ntb/media/supplements/3117


Cooling Methods<br />

The first detectors used in infrared radiometric instruments were cooled with liquid<br />

nitrogen. The detector was attached to the Dewar flask that held the liquid nitrogen,<br />

thus keeping the detector at a very stable and low temperature (-196°C). Later, other<br />

cooling methods were developed. The first solid-state solution to the cooling problem<br />

was presented by AGEMA in 1986, when it introduced a Peltier effect cooler for a<br />

commercial IR camera. In a Peltier cooler, DC current is forced through a<br />

thermoelectric material, removing heat from one junction and creating a cold side and<br />

a hot side. The hot side is connected to a heat sink, whereas the cold side cools the<br />

component attached to it. See Figures 2.4 and 2.5.<br />

For very demanding applications, where the highest possible sensitivity was needed,<br />

an electrical solution to cryogenic cooling was developed. This resulted in the Stirling<br />

cooler. Only in the last 15 to 20 years were manufacturers able to extend the life of<br />

Stirling coolers to 8,000 hours or more, which is sufficient for use in thermal cameras.<br />

The Stirling process removes heat from the cold finger (Figure 2.6) and dissipates it at<br />

the warm side. The efficiency of this type of cooler is relatively low, but good enough<br />

for cooling an IR camera detector. Regardless of the cooling method, the detector<br />

focal plane is attached to the cold side of the cooler in a way that allows efficient<br />

conductive heat exchange. Because focal plane arrays are small, the attachment area<br />

and the cooler itself can be relatively small.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.4. Single stage Peltier cooler<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.5. Three-stage Peltier cooler<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.6. Integrated Stirling cooler, working with to external electronics<br />

helium gas, cooling down to -196 ºC or sometimes even lower temperatures<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Focal Plane Array Assemblies<br />

Depending on the size/resolution of an FPA assembly, it has from (approximately)<br />

60,000 to more than 1,000,000 individual detectors. For the sake of simplicity, this can<br />

be described as a two-dimensional pixel matrix with each pixel (detector) having<br />

micrometer size dimensions. FPA resolutions can range from about 160 × 120 pixels<br />

up to 1024 × 1024 pixels. In reality, assemblies are a bit more complex. Depending<br />

on the detector material and its operating principle, an optical grating may be part of<br />

the FPA assembly. This is the case for QWIP detectors, in which the optical grating<br />

disperses incident radiation to take advantage of directional sensitivity in the detector<br />

material’s crystal lattice. This has the effect of increasing overall sensitivity of a QWIP<br />

detector. Furthermore, the FPA must be bonded to the IR camera readout electronics.<br />

A finished QWIP detector and IC electronics assembly is shown in Figure 8. This<br />

would be incorporated with a Dewar or Stirling cooler in an assembly similar to those<br />

shown in Figure 7. Another complexity is the fact that each individual detector in the<br />

FPA has a slightly different gain and zero offset. To create a useful thermographic<br />

image, the different gains and offsets must be corrected to a normalized value. This<br />

multi-step calibration process is performed by the camera software. See Figures 2.9 –<br />

2.11. The ultimate result is a thermographic image that accurately portrays relative<br />

temperatures across the target object or scene (Figure 2.12). Moreover, actual<br />

temperatures can be calculated to within approximately ±1°C accuracy.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.7. Examples of cooled focal plane array assemblies used in IR<br />

cameras<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.8. QWIP FPA mounted on a ceramics substrate and bonded to<br />

external electronics<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.9. To normalize different FPA detector gains and offsets, the first<br />

correction step is offset compensation. This brings each detector response<br />

within the dynamic range of the camera’s A/D converter electronics.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.10. After offset compensation, slope correction is applied.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.11. After gain factors are brought to the same value, non-uniformity<br />

correction (NUC) is applied so that all detectors have essentially the same<br />

electronic characteristics.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.12. IR image from a 1024 × 1024 InSb detector camera (MW PD)<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


High-end InSb cameras<br />

ImageIR ® 8300/9300 Z<br />

for long-range<br />

identification<br />

Charlie Chong/ Fion Zhang<br />

http://www.opli.net/opli_magazine/eo/2013/first-thermal-camera-features-superzoom-lens-and-hd-resolution-apr/


Application Criteria<br />

As indicated earlier, different types of detectors have different thermal and<br />

spectral sensitivities. In addition, they have different cost structures due to<br />

various degrees of manufacturability. Where they otherwise fit the application,<br />

photon detectors such as InSb and QWIP types offer a number of advantages:<br />

• High thermal sensitivity<br />

• High uniformity of the detectors, i.e., very low fixed pattern noise<br />

• There is a degree of selectability in their spectral sensitivity<br />

• High yield in the production process<br />

• Relatively low cost<br />

• They are resistant to high temperatures and high radiation<br />

• They produce very good image quality<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Camera electronics can handle wide variations in absolute detector<br />

sensitivities. For example, high sensitivity that might saturate a detector at<br />

high thermal intensities can be handled by aperture control and neutral<br />

density filters. Both of these solutions can reduce the radiant energy<br />

impinging on the FPA.<br />

Price aside, spectral sensitivity is often an overriding concern in selecting a<br />

detector and camera for a specific application. Once a detector is selected,<br />

lens material and filters can be selected to somewhat alter the overall<br />

response characteristics of an IR camera system. Fig 2.13 shows the system<br />

response for a number of different detectors.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 2.13. Relative response<br />

curves for a number of IR cameras<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Chapter 3: Getting The Most From Your IR Camera<br />

<strong>Understanding</strong> IR camera calibration and corrections help ensure accurate<br />

temperature measurements and thermographic mapping.<br />

Quantitative Measurements with IR Cameras<br />

For best results, IR camera users need to think carefully about the type of<br />

measurements they need to make, and then be proactive in the camera’s<br />

calibration process. Of course, the first step is selecting a camera with the<br />

appropriate features and software for the application. An understanding of the<br />

differences between (1) thermographic and (2) radiometric measurements is<br />

very helpful in this regard.<br />

Thermography is a type of infrared imaging in which IR cameras detect<br />

radiation in the electromagnetic spectrum with wavelengths from roughly 900<br />

to 14,000 nanometers (0.9 μm –14 μm ) and produce images of that radiation.<br />

Typically, this imaging is used to measure temperature variations across an<br />

object or scene, which can be expressed in degrees Celsius, Fahrenheit, or<br />

Kelvin.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Radiometry is the measurement of radiant electromagnetic energy,<br />

especially that associated with the IR spectrum. It can be more simply defined<br />

as an absolute measurement of radiant flux. The typical unit of measure for<br />

imaging radiometry is radiance, which is expressed in units of Watts/(sr-cm2).<br />

(The abbreviation “sr” stands for steradian; a non-dimensional geometric ratio<br />

expressing the solid (conical) angle that encloses a portion of the surface of a<br />

sphere equivalent to the square of the radius.)<br />

In simple terms, one can think of thermography as “how hot” an object<br />

is, whereas radiometry is “how much energy” the object is giving off.<br />

Although these two concepts are related, they are not the same thing. IR<br />

cameras inherently measure irradiance not temperature, but thermography<br />

does stem from radiance. When you thermographically calibrate an IR system<br />

you are calibrating /measuring based on effective blackbody radiance and<br />

temperature. Therefore, the emissivity of the target object you are measuring<br />

is vital to achieving accurate temperatures. (Emissivity or emittance is the<br />

radiative property of an object relative to a perfect blackbody.)<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Entry level IR cameras with microbolometer detectors operate according to<br />

non-quantum principles. The detectors respond to radiant energy in a way<br />

that causes a change of state in the bulk material (e.g., resistance or<br />

capacitance). Calibration software in these cameras is oriented toward<br />

thermographic imaging and temperature measurements. High end IR<br />

cameras with photon detectors operate according to quantum physics<br />

principles. Although they also provide high quality images, their software is<br />

typically more sophisticated, allowing accurate measurements of both<br />

radiance and temperature.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Some reasons why radiance measurements are important include:<br />

1. Given a linear sensor, measured radiance is linear with incident energy.<br />

Temperature is non-linear with raw digital image counts, even with a linear<br />

sensor.<br />

2. Given the radiance and area of an object, radiant intensity can be<br />

calculated. Knowing total radiant intensity of a target gives a radiometric<br />

analyst the ability to model the irradiance generated by the target over<br />

various geometric and atmospheric conditions.<br />

3. The relationship between spectral bands of interest can be much easier to<br />

determine if you are working within radiometric units.<br />

4. The comparison between different objects in radiometric terms tends to<br />

have less uncertainty because emissivity is not a concern. (One still needs<br />

to consider atmospheric and spectral band-pass effects.)<br />

5. One can typically convert a radiometric signature from radiance to<br />

effective blackbody temperature given a few assumptions or ancillary<br />

measurement data. It tends to be more difficult to go from temperature to<br />

radiance.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Key Physical Relationships in Camera Operation<br />

There are five basic steps in producing radiometric and thermographic<br />

measurements with an IR camera system:<br />

1. The target object has a certain energy signature that is collected by the IR<br />

camera through its lens.<br />

2. This involves the collection of photons in the case of a photon detector, or<br />

collection of heat energy with a thermal detector, such as a<br />

microbolometer.<br />

3. The collected energy causes the detector to produce a signal voltage that<br />

results in a digital count through the system’s A/D converter. (For example,<br />

a FLIR ThermoVisionR SC6000 IR camera has a 14-bit dynamic range in<br />

its A/D converter, which creates count values ranging from 0–16,383. The<br />

more IR energy incident on the camera’s detector (within its spectral band),<br />

the higher the digital count.)<br />

4. When the camera is properly calibrated, digital counts are transformed into<br />

radiance values.<br />

5. Finally, the calibrated camera‘s electronics convert radiance values to<br />

temperature using the known or measured emissivity of the target object.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Expanding on Steps 4 and 5, an effective blackbody temperature<br />

measurement can be derived from a radiance measurement by applying a<br />

radiometric calibration, temperature vs. radiance model, and emissivity of the<br />

target object or scene. Every IR camera designed for serious measurements<br />

is calibrated at the factory. In the calibration lab, the camera takes a number<br />

of blackbody measurements at known temperatures, radiance levels,<br />

emissivities, and distances. This creates a table of values based on the A/D<br />

counts from the temperature/radiance measurements.<br />

Once the counts for each blackbody temperature measurement are entered<br />

into the calibration software, the data are then passed through an in-band<br />

radiance curve fit algorithm to produce the appropriate in-band radiance vs.<br />

count values given the camera system’s normalized spectral response<br />

function. This produces a radiometric calibration of in-band radiance [W/(srcm2)]<br />

versus the digital counts obtained while viewing a blackbody over a<br />

range of temperatures. The result is a series of calibration curves. An<br />

example of how calibration points are captured is shown in Figure 3.1.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 3.1. Example of camera measurements and corresponding in-band<br />

radiance values for given black body temperatures with resulting radiance vs.<br />

measurement curve.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


The calibration curves are stored in the camera system’s memory as a series<br />

of numeric curve-fit tables that relate radiance values to blackbody<br />

temperatures. When the system makes a measurement, it takes the digital<br />

value of the signal at a given moment, goes into the appropriate calibration<br />

table, and calculates temperature. Due consideration is given to other factors<br />

like atmospheric attenuation, reflected ambient temperature, and the<br />

camera’s ambient temperature drift before the final result is presented.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Ambient Drift Compensation (ADC).<br />

Another important consideration in the calibration process is the radiation<br />

caused by the heating and cooling of the camera itself. Any swings in camera<br />

internal temperature caused by changes in environment or the heating and<br />

cooling of camera electronics will affect the radiation intensity at the detector.<br />

The radiation that results directly from the camera is called parasitic radiation<br />

and can cause inaccuracies in camera measurement output, especially with<br />

thermographically calibrated cameras. Certain IR cameras (like the FLIR<br />

ThermoVisionR product line), have internal sensors that monitor changes in<br />

camera temperature. As part of the calibration process, these cameras are<br />

placed in an environmental chamber and focused at a black body reference.<br />

The temperature of the chamber and black body are then varied and data is<br />

collected from the internal sensors. Correction factors are then created and<br />

stored in the camera. In real-time operation, the camera sensors continually<br />

monitor internal temperature and send feedback to the camera processor.<br />

The camera output is then corrected for any parasitic radiation influences.<br />

This functionality is commonly referred to as ambient drift compensation.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Ultimately, the camera must calculate at an object’s temperature based on:<br />

(1) its emission,<br />

(2) reflected emission from ambient sources, and<br />

(3) emission from the atmosphere<br />

using the Total Radiation Law. The total radiation power received by the<br />

camera can be expressed as:<br />

W tot = Ɛ ∙τ∙W obj + (1 – Ɛ) ∙ τ∙W amb + (1 – τ) ∙ W atm ,<br />

Where:<br />

Ɛ is the object emissivity, τ is the transmission through the atmosphere, T amb<br />

is the (effective) temperature of the object surroundings, or the reflected<br />

ambient (background) temperature, and T atm is the temperature of the<br />

atmosphere.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


The best results are obtained when a user is diligent in entering known values for all<br />

the pertinent variables into the camera software. Emissivity tables are available for a<br />

wide variety of common substances. However, when in doubt, measurements should<br />

be made to obtain the correct values.<br />

Calibration and analysis software tools available to users are not always contained<br />

onboard the camera. While high-end cameras have many built-in software functions,<br />

others rely on external software that runs on a PC. Even high-end cameras are<br />

connected to PCs to expand their internal calibration, correction, and analysis<br />

capabilities. For example, FLIR’s ThermaCAMR RTools software can serve a wide<br />

variety of functions from real-time image acquisition to post-acquisition analysis.<br />

Whether the software is on the camera or an external PC, the most useful packages<br />

allow a user to easily modify calibration variables. For instance, FLIR’s ThermaCAM<br />

RTools provides the ability to enter and modify emissivity, atmospheric conditions,<br />

distances, and other ancillary data needed to calculate and represent the exact<br />

temperature of the object, both live and through saved data. This software provides a<br />

post-measurement capability to further modify atmospheric conditions, spectral<br />

responsivity, atmospheric transmission changes, internal and external filters, and<br />

other important criteria as needed. The discussions that follow below are intended to<br />

represent both onboard and external camera firmware and software functions. Where<br />

these functions reside depends on the camera.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Typical Camera Measurement Functions IR cameras have various operating<br />

modes to assure correct temperature measurements under different<br />

application conditions. Typical measurement functions include:<br />

• Spotmeter<br />

•Area<br />

• Profile<br />

• Isotherm<br />

• Temperature range<br />

• Color or gray scale settings<br />

Cursor functions allow easy selection of an area of interest, such as the<br />

crosshairs of the spot readings in Figure 3.2. In addition, the cursor may be<br />

able to select circle, square, and irregularly shaped polygon areas, or create a<br />

line for a temperature profile. Once an area is selected, it can be “frozen” so<br />

that the camera can take a snapshot of that area. Alternatively, the camera<br />

image can remain live for observation of changes in temperature.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 3.2. IR image of a printed circuit board indicating three spot<br />

temperature readings. Image colors correspond to the temperature scale on<br />

the right.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


The spotmeter finds the temperature at a particular point. Depending on the<br />

camera, this function may allow ten or more movable spots, one or more of<br />

which may automatically find the hottest point in the image. The area function<br />

isolates a selected area of an object or scene and finds the maximum,<br />

minimum, and average temperatures inside that area. The isotherm function<br />

makes it possible to portray the temperature distribution of a hot area.<br />

Multiple isotherms may be allowed. The line profile is a way to visualize the<br />

temperature along some part of the object, which may also be shown as a<br />

graph (Figure 3.3).<br />

Figure 3.3. Graph of temperature along a selected area of a target object<br />

using a camera’s profile function<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


The temperature measurement range typically is selectable by the user. This<br />

is a valuable feature when a scene has a temperature range narrower than a<br />

camera’s full-scale range. Setting a narrower range allows better resolution of<br />

the images and higher accuracy in the measured temperatures. Therefore,<br />

images will better illustrate smaller temperature differences. On the other<br />

hand, a broader scale and/or higher maximum temperature range may be<br />

needed to prevent saturation of the portion of the image at the highest<br />

temperature. As an adjunct to the temperature range selection, most cameras<br />

allow a user to set up a color scale or gray scale to optimize the camera<br />

image.<br />

Figure 3.4 illustrates two gray scale possibilities. In Figure 3.2 a so-called<br />

“iron scale” was used for a color rendering. In a manner similar to the gray<br />

scale used in Figure 4, the hottest temperatures can be rendered as either<br />

lighter colors or darker colors. Another possibility is rendering images with<br />

what is known as a rainbow scale (Figure 3.5). In some color images, gray is<br />

used to indicate areas where the camera detector has become saturated (i.e.,<br />

temperatures well above the top of the scale).<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 3.4. Gray scale images of car engine; left view has white as the<br />

hottest temperature; right view shows black as the hottest<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 3.5. Rainbow scale showing lower temperatures towards the blue end<br />

of the spectrum<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


While choice of color scale is often a matter of personal preference, there<br />

may be times when one type of scale is better than another for illustrating the<br />

range of temperatures in a scene. In the case of isotherm measurements,<br />

areas with the same thermal radiance are highlighted. If we use a color scale<br />

with ten colors, we will in fact get ten isotherms in the image. Such a scale<br />

sometimes makes it easier to see the temperature distribution over an object.<br />

In Figure 3.6, the temperature scale is selected so that each color is an<br />

isotherm with a width of 2°C. Still, it is important to realize that an isothermal<br />

temperature scale rendering will not be accurate (1) unless all of the<br />

highlighted area has the same emissivity, and (2) the ambient temperatures<br />

are the same for all objects within the area. This points out common problems<br />

for IR camera users. Often, emissivity varies across an object or scene, along<br />

with variations in ambient temperatures, accompanied by atmospheric<br />

conditions that don’t match a camera’s default values. This is why IR cameras<br />

include measurement correction and calibration functions.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 3.6. Isotherm color scale with each color having an isotherm width of<br />

2°C<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Emissivity Corrections<br />

In most applications, the emissivity of an object is based on values found in a<br />

table. Although camera software may include an emissivity table, users<br />

usually have the capability of inputting emissivity values for an object ranging<br />

from 0.1 to 1.0. Many cameras also provide automatic corrections based on<br />

user input for reflected ambient temperature, viewing distance, relative<br />

humidity, atmospheric transmission, and external optics. As described earlier,<br />

the IR camera calculates a temperature based on radiance measurements<br />

and the object’s emissivity. However, when the emissivity value is unknown<br />

or uncertain, the reverse process can be applied. Knowing the object<br />

temperature, emissivity can be calculated. This is usually done when exact<br />

emissivity values are needed. There are two common methods of doing this.<br />

The first method establishes a known temperature by using an equalization<br />

box. This is essentially a tightly controlled temperature chamber with<br />

circulating hot air. The length of time in the box must be sufficient to allow the<br />

whole object to be at a uniform temperature. In addition, it is absolutely<br />

necessary that the object stabilize at a temperature different from the<br />

surroundings where the actual measurements will take place. Usually, the<br />

object is heated to a temperature at least 10°C above the surroundings to<br />

ensure that the thermodynamics of the measurements are valid.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Once the object has reached the set temperature, the lid is drawn off and a<br />

thermogram is captured of the object. The camera and/or software for<br />

processing thermograms can be used to get the emissivity value.<br />

Another (“adjacent spot”) method is much simpler, but still gives reasonably<br />

exact values of the emissivity. It uses an area of known emissivity. The idea is<br />

to determine the temperature of the object with the camera in the usual way.<br />

The object is adjusted so that the area with unknown emissivity is very close<br />

to an area of known emissivity. The distance separating these areas must be<br />

so small that it can be safely assumed they have the same temperature. From<br />

this temperature measurement the unknown emissivity can be calculated.<br />

The problem is illustrated in Figure 7, which is an image of a printed circuit<br />

board (PCB) heated to a uniform temperature of 68.7°C. However, areas of<br />

different emissivities may actually have different temperatures, as indicated in<br />

the caption of Figure 3.7a. Using the technique just described, emissivity<br />

correction proceeds by finding a reference spot where a temperature of<br />

68.7°C is indicated and calculating the emissivity at that location. By knowing<br />

the emissivity of the reference spot, the emissivity of the target spots can be<br />

calculated. The corrected temperatures are shown in Figure 3.7b.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 3.7a. PCB heated to a uniform 68.7°C, but digital readouts are<br />

incorrect.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 3.7b. PCB with emissivity correction using the “adjacent spot”<br />

technique. Digital readouts now indicate the correct temperatures at all<br />

locations.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


As illustrated in these figures, this technique can be used with a camera’s<br />

area selection function (“AR” in the figures) and using the average<br />

temperature for that area. The reason for using the average temperature in<br />

the reference area is that there is usually a spread of temperatures within the<br />

area, especially for materials with low emissivity. In that case, using a<br />

spotmeter or an area maximum value would give a less stable result. The<br />

isotherm function is not recommended either, as it is not possible to get the<br />

averaging effect with it. It may also be possible to use a contact sensor to find<br />

the temperature of an area of unknown emissivity, but such measurements<br />

pose other problems that may not be easy to overcome.<br />

Furthermore, it is never possible to measure the emissivity of an object whose<br />

temperature is the same as the reflected ambient temperature from its<br />

surroundings. Generally, a user can also input other variables that are<br />

needed to correct for ambient conditions. These include factors for ambient<br />

temperatures and atmospheric attenuation around the target object.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Using Camera Specifications When considering IR camera performance,<br />

most users are interested in how small an object or area can be detected and<br />

accurately measured at a given distance. Knowing a camera’s field of view<br />

(FOV) specifications helps determine this.<br />

Field of View (FOV). This parameter depends on the camera lens and focal<br />

plane dimensions, and is expressed in degrees, such as 35.5° ×28.7° or<br />

18.2 × 14.6°. For a given viewing distance, this determines the dimensions of<br />

the total surface area “seen” by the instrument (Figure 3.8). For example, a<br />

FLIR ThermoVision SC6000 camera with a 25mm lens has an FOV of 0.64 ×<br />

0.51 meters at a distance of one meter, and 6.4 × 5.1 meters at a distance of<br />

ten meters.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 3.8. A camera’s field of view (FOV) varies with viewing distance.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Instantaneous Field of View (IFOV).<br />

This is a measure of the spatial resolution of a camera’s focal plane array<br />

(FPA) detector. The configuration of the FPA in the FLIR ThermoVision<br />

SC6000 is 640 × 512 detectors, which makes a total of 327,680 individual<br />

picture elements (pixels). Suppose you are looking at an object at a distance<br />

of one meter with this camera. In determining the smallest detectable object,<br />

it is important to know the area’s IFOV covered by an individual pixel in the<br />

array. The total FOV is 0.64 × 0.51 meters at a distance of one meter. If we<br />

divide these FOV dimensions by the number of pixels in a line and row,<br />

respectively, we find that a pixel’s IFOV is an area approximately 1.0 ×<br />

1.0mm at that distance. Figure 3.9 illustrates this concept.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 3.9. A camera’s geometric (spatial) resolution (IFOV) is determined by<br />

its lens and FPA configuration.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


To use this information consider, the pixel IFOV relative to the target object<br />

size (Figure 3.10). In the left view of this figure, the area of the object to be<br />

measured covers the IFOV completely. Therefore, the pixel will receive<br />

radiation only from the object, and its temperature can be measured correctly.<br />

In the right view of Figure 3.10, the pixel covers more than the target object<br />

area and will pick up radiation from extraneous objects. If the object is hotter<br />

than the objects beside or behind it, the temperature reading will be too low,<br />

and vice versa. Therefore it is important to estimate the size of the target<br />

object compared to the IFOV in each measurement situation.<br />

Figure 3.10. IFOV (red squares) relative to object size.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Spot Size Ratio (SSR).<br />

At the start of a measurement session, the distance between the camera and<br />

the target object should be considered explicitly. For cameras that do not<br />

have a calibrated spot size, the spot size ratio method can be used to<br />

optimize measurement results. SSR is a number that tells how far the camera<br />

can be from a target object of a given size in order to get a good temperature<br />

measurement. A typical figure might be 1,000:1 (also written 1,000/1, or<br />

simply abbreviated as 1,000). This can be interpreted as follows: at 1000 mm<br />

distance from a target, the camera will measure a temperature averaged over<br />

a 1mm square. Note that SSR is not just for targets far away. It can be just as<br />

important for close-up work. However, the camera’s minimum focal distance<br />

must also be considered. For shorter target distances, some manufacturers<br />

offer close-up lenses.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


For any application and camera/lens combination, the following equation<br />

applies:<br />

D/S = SSR/1 (or more conveniently S = D/SSR)<br />

Where:<br />

D is the distance from the camera to the target,<br />

S is smallest target dimension of interest, and<br />

SSR is the spot size ratio.<br />

The units of D and S must be the same. When selecting a camera, keep in<br />

mind that IFOV is a good figure of merit to use. The smaller the IFOV, the<br />

better the camera for a given total field of view.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Other Tools for Camera Users<br />

As mentioned earlier, IR cameras are calibrated at the factory, and field<br />

calibration in not practical. However, some cameras have a built-in blackbody<br />

to allow a quick calibration check. These checks should be done periodically<br />

to assure valid measurements. Bundled and optional data acquisition<br />

software available for IR cameras allows easy data capture, viewing, analysis,<br />

and storage. Software functions may include real-time radiometric output of<br />

radiance, radiant intensity, temperature, target length/area, etc. Optional<br />

software modules are also available for spatial and spectral radiometric<br />

calibration 空 间 与 光 谱 辐 射 测 量 标 定 .<br />

Functions provided by these modules might include:<br />

• Instrument calibration in terms of radiance, irradiance, and temperature<br />

• Radiometric data needed to set instrument sensitivity and spectral range<br />

• Use of different transmission and/or emissivity curves or constants for<br />

calibration data points<br />

• Adjustments for atmospheric effects<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


In addition, IR camera software and firmware provide other user inputs that<br />

refine the accuracy of temperature measurements. One of the most important<br />

functions is non-uniformity correction (NUC) of the detector FPA. This type of<br />

correction is needed due to the fact that each individual detector in the<br />

camera’s FPA has a slightly different gain and zero offset. To create a useful<br />

thermographic image, the different gains and offsets must be corrected to a<br />

normalized value. This multi-step NUC process is performed by camera<br />

software. However, some software allows the user to specify the manner in<br />

which NUC is performed by selecting from a list of menu options. For<br />

example, a user may be able to specify either a one-point or a twopoint<br />

correction.<br />

■<br />

■<br />

A one-point correction only deals with pixel offset.<br />

Two-point corrections perform both gain and offset normalization of<br />

pixel- to-pixel nonuniformity.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


With regard to NUC, another important consideration is how this function<br />

deals with the imperfections that most FPAs have as a result of<br />

semiconductor wafer processing. Some of these imperfections are manifested<br />

as bad pixels that produce no output signals or as outputs far outside of a<br />

correctable range. Ideally, the NUC process identifies bad pixels and replaces<br />

them using a nearest neighbor replacement algorithm. Bad pixels are<br />

identified based on a response and/or offset level outside user-defined points<br />

from the mean response and absolute offset level. Other NUC functions may<br />

be included with this type of software, which are too numerous to mention.<br />

The same is true of many other off-the-shelf software modules that can be<br />

purchased to facilitate thermographic image display, analysis, data file<br />

storage, manipulation, and editing. Availability of compatible software is an<br />

important consideration when selecting an IR camera for a user’s application<br />

or work environment.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Conclusions<br />

Recent advances in IR cameras have made them much easier to use.<br />

Camera firmware has made setup and operation as simple as using a<br />

conventional video camera. Onboard and PC-based software provides<br />

powerful measurement and analysis tools. Nevertheless, for accurate results,<br />

the user should have an understanding of IR camera optical principals and<br />

calibration methods. At the very least, the emissivity of a target object should<br />

be entered into the camera’s database, if not already available as a table<br />

entry.<br />

Keywords:<br />

At the very least, the emissivity of a target object should be entered into the<br />

camera’s database, if not already available as a table entry.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Chapter 4: Filters Extend IR Camera Usefulness Where<br />

Filters Can Help<br />

Materials that are transparent or opaque to IR wavelengths present problems<br />

in non-contact temperature measurements with an IR camera. With<br />

transparent materials, the camera sees through them and records a<br />

temperature that is a combination of the material itself and that which is<br />

behind it. In the second case, when an IR camera needs to see through a<br />

material to measure the temperature of an object behind it, signal attenuation<br />

and ambient reflections can make accurate temperature readings difficult or<br />

impossible. In some cases, an IR filter can be placed in the camera’s optical<br />

path to overcome these problems.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


IR Window: Cordex IR Window<br />

Charlie Chong/ Fion Zhang


Spectral Response is the Key<br />

IR cameras inherently measure irradiance not temperature. However, a<br />

camera’s software coverts radiance measurements into temperatures by<br />

using the known emissivity of a target object and applying internal calibration<br />

data for the camera’s spectral response. The spectral response is determined<br />

primarily by the camera’s lens and detector. Figure 4.1 shows the spectral<br />

response of a few IR cameras with various spectral responses. The spectral<br />

performance of most cameras can be found in their user manual or technical<br />

specifications. For many objects, emissivity is a function of their radiance<br />

wavelength, and is further influenced by their temperature, the angle at which<br />

they are viewed by a camera, and other factors. An object whose emissivity<br />

varies strongly with wavelength is called a selective radiator. One that has the<br />

same emissivity for all wavelengths is called a greybody. Transparent<br />

materials, such as glass and many plastics, tend to be selective radiators. In<br />

other words, their degree of transparency varies with wavelength. There may<br />

be IR wavelengths where they are essentially opaque due to absorption.<br />

Since, according to Kirchhoff’s Law, a good absorber is also a good emitter,<br />

this opens the possibility of measuring the radiance and temperature of a<br />

selective radiator at some wavelength.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 4.1. Relative response curves for a number of IR cameras<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Spectral Adaptation<br />

Inserting a spectral filter into the camera’s optical path is called spectral<br />

adaptation. The first step of this process is to analyze the spectral properties<br />

of the semitransparent material you are trying to measure. For common<br />

materials the data may be available in published data. Otherwise, this<br />

requires analysis with a spectrophotometer. (The camera manufacturer or a<br />

consulting firm may supply this service.) In either case, the objective is to find<br />

the spectral location of a band of complete absorption that falls within the IR<br />

camera’s response curve. Microbolometer detectors have rather broad<br />

response curves so they are not likely to present a problem in this respect.<br />

However, adding a filter decreases overall sensitivity due to narrowing of the<br />

camera’s spectral range. Sensitivity is reduced approximately by the ratio of<br />

the area under the filter’s spectral curve to the area under the camera’s<br />

spectral curve.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


This could be a problem for microbolometer systems, since they have<br />

relatively low sensitivity to start with and a broad spectral curve. Using a<br />

camera with, for example, a QWIP detector will provide greater sensitivity with<br />

a narrower spectral curve. Still, this narrow range may limit the application of<br />

such cameras for spectral adaptation. Ultimately, an optical (IR) filter must be<br />

selected that blocks all wavelengths except the band where the object<br />

absorbs. This ensures that the object has high emissivity within that band.<br />

Besides semitransparent solids, selective adaptation can also be applied to<br />

gases. However, a very narrow filter might be required for selecting an<br />

absorption “spike” in a gas. Even with proper filtering, temperature<br />

measurement of gases is difficult, mainly due to unknown gas density.<br />

Selective adaptation for a gas has a better chance of success if the objective<br />

is merely gas detection, since there are less stringent requirements for<br />

quantitative accuracy. In that case sensitivity would be more important, and<br />

some gases with very high absorption might still be measurable.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Spectral adaptation could also be applied the opposite way, i.e., selection of a<br />

spectral band where the transmission through a medium is as high as<br />

possible. The purpose would be to enable measurement on an object by<br />

seeing through the medium without any interference. The medium could be<br />

ordinary atmosphere, the atmosphere of combustion gases inside a furnace,<br />

or simply a window (or other solid) through which one wants to measure.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Filter Types<br />

The simplest filters are broadband neutral density types that are used merely<br />

to reduce optical transmission and prevent detector saturation at high<br />

temperatures. While necessary sometimes, this is not spectral adaptation. In<br />

spectral adaptation, filters are used in order to suppress or transmit certain<br />

wavelengths. For discussion purposes, filters can be described as short-pass<br />

(SP), long-pass (LP), band-pass (BP), and narrow band-pass (NBP). See<br />

Figure 4.2. SP and LP filters are specified with a cut-on and a cut-off<br />

wavelength. BP and NBP filters are specified with a center wavelength and a<br />

half-width (half-power) wavelength, the latter being the width where spectral<br />

response has decreased to 50% of its maximum. For temperature<br />

measurements on transparent materials, the filter selected must provide a<br />

band of essentially complete absorption. Incomplete absorption can be used,<br />

at least theoretically, provided that both absorptance and reflectance are<br />

known and stable at the absorption band. Unfortunately, absorption often<br />

varies with both temperature and thickness of the material.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 4.2. Response curves for different types of filters<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


An example of applying a NBP filter to the measurement of polyethylene film<br />

temperature is shown in Figure 4.3. The blue curve in the figure shows the<br />

absorption band of polyethylene film. The red curve shows the transmittance<br />

of a 3.45μm NBP filter, which is designed to match polyethylene film. The<br />

green curve shows the resulting transmission through film plus the filter. This<br />

curve, running just above the zero line, indicates an excellent filter adaptation,<br />

i.e. the film appears to be opaque to the camera, and no background radiation<br />

would disturb the measurement of film temperature. Filters can also be<br />

classified according to their application temperature. Traditionally, cold filters,<br />

filters that are stabilized at or near the same temperature as the detector, are<br />

the most accurate and desired filters for thermal signatures. Warm filters,<br />

filters screwed onto the back of the optical lens outside of the detector/cooler<br />

assembly, are also commonly used but tend to provide more radiometric<br />

calibration uncertainty due to varying IR emission with ambient temperature<br />

changes. Once a filter is selected for use with a particular camera, the<br />

camera/filter combination needs to be calibrated by the camera manufacturer.<br />

Then the performance of the system should be characterized since accuracy<br />

and sensitivity will be affected due to a reduction in energy going to the<br />

detector.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 4.3. Application of an NBP filter to achieve nearly complete absorption<br />

and high emittance from polyethylene film, allowing its temperature<br />

measurement<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Transparent Material Measurement Techniques<br />

Production of sheet glass and thin plastic film requires fairly tight temperature<br />

control to maximize production quality and yield. Traditionally, temperature<br />

sensors have been embedded at the orifice of the extruder, which provides<br />

rather coarse information about sheet/film temperature. An IR machine vision<br />

system can make noncontact temperature measurements and supply more<br />

usable data about the material as it is extruded. However, as described above,<br />

an appropriate filter is needed for the IR camera to make the material appear<br />

opaque. To ensure that the proper filter was selected, spectral response<br />

curves for the camera/filter system can be created by the camera<br />

manufacturer. (See the green curve in Figure 4.3.) In fact, this is generally<br />

required for permanent cold filter installations to validate filter response.<br />

Otherwise, (with supportive spectral data) the user can proceed by checking<br />

emissivity. This is a verification of emissivity efficiency for the overall system<br />

response, including the target material and camera with installed filter.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Recalling Kirchhoff’s law,<br />

ρ λ + Ɛ λ + τ λ = 1, or Ɛ λ = 1 – τ λ<br />

– ρ λ<br />

it is clear that in order to get an emissivity value Ɛ, transmittance τ and<br />

reflectance ρ at the pass band of the filter must be known. The transmittance,<br />

τ λ , can be taken directly from a transmission diagram like the one in Figure 3<br />

(a value of about 0.02 in that example).<br />

Where:<br />

absorbtion alpha (α),<br />

reflection Rho (ρ),<br />

transmission Tau (τ ),<br />

emissivity epsilon (Ɛ).<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Reflectance is less easy to characterize and usually is a function of material<br />

thickness. However, a transmission diagram like the one in Figure 4.4<br />

provides some indication of this parameter’s value. Using the blue curve for<br />

the thinnest polyethylene material in Figure 4, which has the lowest<br />

absorption, the transmission between absorption bands is seen to be<br />

approximately 90%. If there were no absorption bands at all, we could<br />

conclude that the reflection would be 10%. Since there are some narrow<br />

absorption bands under the curve, we can estimate the reflection to be 8% in<br />

the spectral regions where absorption is very low. However, we are interested<br />

in the reflectance where the absorption is high (i.e., where the material<br />

appears to be opaque). To estimate the reflectance of this polyethylene film,<br />

we must first make the reasonable assumption that its surface reflectance<br />

stays constant over the absorption bands. Now recognize that the 8% value is<br />

the result of reflections from both sides of the film, i.e., approximately 4% per<br />

surface.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 4.4. Transmission bands for polyethylene films of three different<br />

thicknesses<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


At the absorption band, however, since the absorption in the material is<br />

almost complete, we get reflection only on one side. Thus ρ λ = 0.04.<br />

From this ρ λ , and the τ λ value obtained from the transmission graph (Figure 3<br />

in this example), emissivity can be calculated: Ɛ λ = 1 – 0.02 – 0.04 = 0.94.<br />

This value is entered into the camera’s measurement database before having<br />

it calculate the temperatures from radiance observations. Sheet and plate<br />

glass production have similar temperature measurement requirements. The<br />

most common industrial varieties are variations of soda-lime- silica glass.<br />

Although they may vary in composition and color, their spectral characteristics<br />

do not change much. Looking at the spectral transmittance of such a glass<br />

with different thicknesses (Figure 4.5), one can conclude that IR temperature<br />

measurement must be restricted to wavelengths above 4.3μm. Depending on<br />

glass thickness, this may require either a mid wavelength (MW) or long<br />

wavelength (LW) camera/detector. MW cameras cover some portion of the<br />

spectrum from 2μm – 5μm, and LW cameras cover some portion within 8μm<br />

–12μm.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 4.5. Transmission curves for a common industrial glass in five<br />

thicknesses from 0.23mm to 5.9mm<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


In selecting a filter, the temptation might be to go for an LP type with a cut-on<br />

wavelength near the point where transmittance drops to zero. However, there<br />

are other factors to consider. For example, LP filter characteristics can<br />

interfere with the negative slope of the spectral response curve of thermolectrically<br />

cooled HgCdTe (MCT) detectors, which are used in both MW and<br />

LW cameras. A better choice may be a NBP filter. In Figure 4.6, transmission<br />

characteristics of a glass, an SW camera, and two filters are superimposed.<br />

The green curve represents the LP filter response curve, whereas the NBP<br />

filter response is shown in blue. The latter was selected for the spectral<br />

location where glass becomes “black,” and has a center wavelength of 5.0μm.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 4.6. Two alternative filters for glass measurement with a SW camera<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


The reflectance of this glass is shown in Figure 4.7. Note the peak between 8<br />

and 12 μm, which must be avoided when using an LW camera to measure<br />

the glass. Another consideration is the camera’s viewing angle, because<br />

glass reflectance can change with angle of incidence. Fortunately reflectance<br />

does not change much up to an angle of about 45° relative to normal<br />

incidence (Figure 4.8). From Figure 4.8, a value 0.025 for the glass<br />

reflectance is valid when using either the 4.7μm LP or the 5.0μm NBP filter<br />

(Figure 4.6), because they both operate in the 5μm region. Consequently a<br />

proper value for the glass emissivity in those cases would be 1 – 0.025 =<br />

0.975.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 4.7. Reflectance of a common glass at normal (perpendicular)<br />

incidence<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 4.8. Glass reflectance as a function of camera viewing angle relative<br />

to normal incidence<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Transmission Band Applications<br />

For many applications, the user will need to find a spectral band where the<br />

medium through which the camera is looking has minimum influence on the<br />

measurement. The object of interest is at the end of a measurement path on<br />

the other side of the medium. The medium is in most cases ordinary<br />

atmosphere, but it could also be a gas or a mixture of gases (e.g., combustion<br />

gases or flames), a window, or a solid semitransparent material. As is the<br />

case in absorption band applications, a spectral transmission measurement<br />

of the actual medium would be the ideal starting point. The objective is to find<br />

a band within the camera’s response curve where the medium has minimum<br />

influence on IR transmission from the target object. However, it is often<br />

impractical to perform such a measurement, particularly for gases at high<br />

temperatures. In such cases it may be possible to find the spectral properties<br />

of gas constituents (or other media) in IR literature, revealing a suitable<br />

spectrum for the measurement.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


In most cases, IR camera manufacturers have anticipated the atmospheric<br />

attenuation problem. Camera manufacturers typically add a filter that reduces<br />

measurement errors due to inaccurate and/or varying atmospheric<br />

parameters by avoiding absorption bands of the constituent gases and water<br />

vapors. This is especially needed at long measurement distances and shorter<br />

wavelengths. For MW cameras, an appropriate filter utilizes the atmospheric<br />

window between the absorption bands of H 2 O+CO 2 around 3μm or CO 2 at<br />

4.2μm. Atmospheric effects on an LW camera are much less, since the<br />

atmosphere has an excellent window from 8μm to 12μm. However, cameras<br />

with a broad response curve reaching into the MW spectrum may require an<br />

LP filter. This is particularly true for high temperature measurements where<br />

the radiation is shifted towards shorter wavelengths and atmospheric<br />

influence increases. An LP filter with a cut-on at 7.4 μm blocks the lower part<br />

of the camera’s response curve.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


An interesting transmission band application is temperature measurements<br />

on a gas-fired furnace, oven, or similar heating equipment. Objectives could<br />

be the measurement of flame temperature or the measurement of internal<br />

components through the flames. In the latter case, an unfiltered IR camera<br />

will be overwhelmed by the intense radiation from the flames, making<br />

measurement of the much weaker radiation from internal objects impossible.<br />

On the other hand, any transmission through the flames from cooler internal<br />

objects will make flame temperature measurements inaccurate. The flame<br />

absorption spectrum in Figure 4.9 reveals the spectral regions where these<br />

two types of measurement could be made. There is very little radiation from<br />

the flames in the 3.9 μm area, whereas there is a lot of radiation between the<br />

4.2 and 4.4μm range. The idea is to employ filters that utilize these spectral<br />

windows for the desired measurements.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 4.9. Flame absorption spectrum of a gas-fired furnace with two types<br />

for filters for different measurement applications<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


For measurement of internal components, you need to avoid strong<br />

absorption bands because they attenuate the radiation from the target object<br />

and they emit intensely due to the high gas temperature, thus blinding the<br />

camera. Although gas-fired combustion gases consist mostly of CO 2 and<br />

water vapor, an atmospheric filter is unsuitable because gas concentrations<br />

and temperatures are much higher. This makes the absorption bands deeper<br />

and broader. A flame filter is needed for this application. See Figure 9. This is<br />

a BP filter transmitting between 3.75 μm and 4.02 μm. With this filter installed,<br />

the camera will produce an image where the flames are almost invisible and<br />

the internal structure of the furnace is presented clearly (Figure 10).<br />

To get the maximum temperature of the flames, a CO 2 filter will show they are<br />

as high as 1400°C. By comparison, the furnace walls as seen with the flame<br />

filter are a relatively cool 700°C.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 4.10. FLIR ThermaCAM?image of furnace tubes with flame filter to<br />

allow accurate temperature measurement<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Conclusions<br />

Filters can extend the application of IR cameras into areas that might<br />

otherwise restrict their use. Still, some preliminary spectrophotometer<br />

measurements may be needed on the objects and media of interest if spectral<br />

information cannot be found in IR literature. Once a filter is selected and<br />

installed, the camera/filter system should be calibrated by the camera<br />

manufacturer. Even with a well-calibrated system, it is a good idea to avoid<br />

errors by not using spectral regions of uncertain or varying absorption relative<br />

to the camera/ filter system response spectrum.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


<strong>Infrared</strong> Filter<br />

Charlie Chong/ Fion Zhang


<strong>Infrared</strong> Filter<br />

Charlie Chong/ Fion Zhang


<strong>Infrared</strong> Filter<br />

Charlie Chong/ Fion Zhang


Chapter 5: Ultra High-Speed Thermography<br />

Recent Advances in <strong>Thermal</strong> Imaging<br />

We have all seen high-speed imagery at some point in our lives, be it a video<br />

of a missile in flight or a humming bird flapping its wings in slow motion. Both<br />

scenarios are made possible by high-speed visible cameras with ultra short<br />

exposure times and triggered strobe lighting to avoid image blur, and usually<br />

require high frame rates to ensure the captured video plays back smoothly.<br />

Until recently, the ability to capture high-speed dynamic imagery has not been<br />

possible with traditional commercial IR cameras. Now, recent advances in IR<br />

camera technologies, such as fast camera detector readouts and high<br />

performance electronics, allow high-speed imagery.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Challenges prohibiting high-speed IR cameras were based primarily on readout electronic<br />

designs, camera pixel clocks, and backend data acquisition systems being too slow. Older<br />

readout designs only allowed minimum integration times down to about 10μs, which in<br />

some cases were insufficient to stop motion on a fast moving target without image blur.<br />

Similarly, targets with very fast temperature changes could not be sampled at an adequate<br />

frame rate to accurately characterize the object of interest. Even with the advent of faster<br />

IR cameras, there still remains the hurdle of how to collect high resolution, high-speed data<br />

without overwhelming your data collection system and losing frames of data. Not all<br />

challenges for high-speed IR cameras were due to technology limitations.<br />

Some were driven by additional requirements that restricted the maximum frame rates<br />

allowed. For example, cameras that required analog video output naturally restricted the<br />

maximum frame rate due to the NTSC and PAL format requirements of 30Hz or 25Hz,<br />

respectively. This is true regardless of the detector focal plane array’s (FPA) pixel rate<br />

capabilities, because the video monitor’s pixel rates are set by the NTSC or PAL timing<br />

parameters (vertical and horizontal blanking periods). However, with new improvements in<br />

high-end commercial R&D camera technologies, all these challenges have been overcome<br />

and we can begin exploring the many benefits of high-speed IR camera technology. The<br />

core benefits are the ability to capture fast moving targets without image blur, acquire<br />

enough data to properly characterize dynamic energy targets, and increase the dynamic<br />

range without compromising the number of frames per second.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Reducing Image Blur with Short Integration Times<br />

With advanced FPA Readout Integrated Circuits (ROIC), IR cameras can<br />

have integration times (analogous to exposure time or shutter speed in visible<br />

cameras) as short as 500ns. In addition, new ROIC designs maintain linearity<br />

all the way to the bottom of their integration time limits; this was not true for<br />

ROICs developed only a few years ago. The key benefit again is to avoid<br />

motion blur as the target moves or vibrates through the field of view of the<br />

camera. With sub-microsecond integration times, these new cameras are<br />

more than sufficient for fast moving targets such as missiles or in the<br />

following example, a bullet in flight.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Faster Than a Speeding Bullet<br />

In the following experiment, a high speed IR Camera was used to capture and<br />

measure the temperature of a 0.30 caliber rifle bullet in flight. At the point of image<br />

capture the bullet was traveling at supersonic speeds (800–900 meters per second)<br />

and was heated by friction within the rifle barrel, the propellant charge, and<br />

aerodynamic forces on the bullet. Due to this heat load, the IR camera could easily<br />

see the bullet even at the very short integration time of 1μs; so unlike a visible camera,<br />

no strobe source is needed. A trigger was needed to start the camera integration time<br />

to ensure the bullet was in the Field of View (FOV) of the camera at the time of frame<br />

capture. This was done by using an acoustic trigger from the rifle shot, which locates<br />

the bullet along the axis of fire to within a distance of several centimeters. Figure 5.1a<br />

shows a close-up IR image of the bullet traveling at 840m/s (~ 1900 mph); yet using<br />

the 1μs integration time, effectively reduced the image blur to about 5 pixels. Figure<br />

5.1b shows a reference image of an identical bullet imaged with a visible light camera<br />

set to operate with a 2-microsecond integration time.<br />

The orientation of the bullets in the two images is identical – they both travel from left<br />

to right. The bright glow seen on the waist of the image is a reflection of bright studio<br />

lights that were required to properly illuminate the bullet during the exposure. Unlike<br />

the thermal image, the visible image required active illumination, since the bullet was<br />

not hot enough to glow brightly in the visible region of the spectrum.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 5.1a. <strong>Infrared</strong> image of a 0.30 caliber bullet in flight with apparent<br />

temperatures<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 5.1b. Visible-light image of an identical 0.30 caliber bullet in flight<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


High-Speed Imaging for Fast Transients<br />

Short integration times and high-speed frame rates are not always paired<br />

together in IR cameras. Many cameras have fast frame rates but not fast<br />

integration times or vice versa. Still, fast frame rates are critical for properly<br />

characterizing targets whose temperatures change very quickly. An<br />

application where both short integration time and fast frame rate are required<br />

is overload testing of integrated circuits (ICs). See Figure 5.2. The objective of<br />

this test is to monitor the maximum heat load the IC experiences when biased<br />

and reverse biased with current levels outside the design limits. Without highpeed<br />

IR technology, sufficient data might not be captured to characterize the<br />

true heat transients on the IC due to under sampling. This would not only give<br />

minimal data to analyze, but could also give incorrect readings of the true<br />

maximum temperature. When the IC was sampled at a frame rate of 1000Hz,<br />

a maximum temperature of 95°C was reported. However, when sampled at<br />

only 500Hz, the true maximum temperature was missed and a false<br />

maximum of 80°C was reported (Figure 5.3). This is just one example of why<br />

high-speed IR cameras can be so valuable for even simple applications that<br />

don’t necessarily appear to benefit from high speed at first consideration.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 5.2. Integrated circuit with 800ms overcurrent pulse<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 5.3. Maximum IC temperature data – actual vs. undersampled<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Pixel Clock vs. Analog to Digital Taps<br />

High-speed IR cameras require as a prerequisite a combination of a fast pixel<br />

clock and a higher number of analog to digital (A/D) converters, commonly<br />

called channels or taps. As a frame of reference, most low performance<br />

cameras have two channels or A/D converters and run at lower than 40<br />

megapixels/second clock rates. This may sound fast, but when you consider<br />

the amount of data, that translates into around 60Hz in most cases. Highpeed<br />

IR cameras on the other hand typically have a minimum of four<br />

channels and have clock speeds of at least 50 megapixels/second. In turn<br />

they offer 14-bit digital data at frame rates of over 120Hz at 640 × 512<br />

window sizes. In order to increase frame rates further, IR cameras usually<br />

allow the user to reduce the window size or number of pixels read out from<br />

FPA. Since there is less data per frame to digitize and transfer, the overall<br />

frame rate increases. Figure 5.4 illustrates the increase in frame rates relative<br />

to user defined window sizes. Newer camera designs offer 16 channels and<br />

pixel clocks upwards of 205 mega pixels/second. This allows for very fast<br />

frame rates without sacrificing the window size and overall resolution.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 5.4. Example of FPA window sizes relative to frame rates<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Preset Sequencing Increases Dynamic Range<br />

High-speed IR cameras have an additional benefit that does not relate to<br />

highspeed targets, but rather to increasing the dynamic range of the camera.<br />

By coupling a high-speed IR camera with a data capture method known as<br />

superframing, you can effectively increase the camera’s dynamic range from<br />

14 bits to around 18 bits – 22 bits per frame. Superframing involves cycling<br />

the IR camera through up to four multiple integration times (presets),<br />

capturing one frame at each preset. This results in multiple unique data movie<br />

files, one for each preset. This data is then combined by using off-the-shelf<br />

ABATER software. The software selects the best resolved pixel from each<br />

unique frame to build a resultant frame composed of data from all the<br />

collected data movie files at varying integration times. This method is<br />

especially beneficial for those imaging scenes with both hot and cold objects<br />

in the same field of view. Typically a 14-bit camera cannot image<br />

simultaneously both hot and cold objects with a single integration time. This<br />

would result in either over exposure on the hot object or under exposure on<br />

the cold object. The results of superframing are illustrated in the Beechcraft<br />

King Air aircraft images in Figure 5.5, captured at two different integration<br />

times.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 5.5. Active aircraft engine imaged at integration rates of 2ms (left) and<br />

30μs (right)<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


While the aircraft can be clearly seen in the left image (Preset 0 = 2ms<br />

integration time), there are portions of the engine that are clearly over<br />

exposed. Conversely, the right image in Figure 5.5 (Preset 1 = 30μs<br />

integration time), shows engine intake and exhaust detail with the remainder<br />

of the aircraft underexposed. When the two images in Figure 5.5 are<br />

processed in ABATER software, the best resolved pixels are selected and<br />

used to build a single resultant superframed image with no over or under<br />

exposed pixels (Figure 5.6). As you may have figured out, the down side to<br />

this method of data collection and analysis is the reduction in the frame rate<br />

by the number of Presets cycled. By applying some simple calculations a<br />

100Hz camera with two Presets will provide an overall frame rate of 50Hz,<br />

well under the limits of our discussion of high speed IR imagery. This only<br />

reinforces the need for a high speed camera. If a 305Hz camera is<br />

superframed as in the example above, a rate of over 150Hz per preset frame<br />

rate is achieved. This rate is well within the bounds of high speed IR imaging.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Figure 5.6. Superframed image created with ABATER software from Preset 0<br />

and Preset 1 data.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Conclusions<br />

Sophisticated IR cameras are now available with advanced readout<br />

electronics and high speed pixel clocks, which open the door for high speed<br />

IR imagery. This allows us to expand the boundaries of which applications<br />

can be solved using IR camera solutions. Furthermore, it allows us to begin<br />

capturing more data and increase our accuracy for demanding applications<br />

with fast moving targets, quick temperature transients, and wide dynamic<br />

range scenes. With the release of this new technology in the commercial IR<br />

marketplace, we can now begin to realize the benefits of high speed data<br />

capture, once only available to the visible camera realm.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


A wide range of thermal imaging cameras for R&D and Science<br />

applications<br />

FLIR markets a full product range of thermal imaging cameras for R&D<br />

applications. Whether you are just discovering the benefits that thermal<br />

imaging cameras have to offer or if you are an expert thermographer, FLIR<br />

offers you the correct tool for the job. Discover our full product range and find<br />

out why FLIR is the world leader for thermal imaging cameras.<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


Software for demanding thermal imaging professionals<br />

At FLIR, we recognize that our job is to go beyond just producing the best<br />

possible infrared camera systems. We are committed to enabling all users of<br />

our thermal imaging camera systems to work more efficiently and productively<br />

by providing them with the most professional camera-software combination.<br />

Our team of committed specialists are constantly developing new, better and<br />

more user-friendly software packages to satisfy the most demanding thermal<br />

imaging professionals. All software is Windows-based, allows fast, detailed<br />

and accurate analysis and evaluation of thermal inspections.<br />

For more information on FLIR products or software please contact your<br />

nearest FLIR representative of visit www.flir.com<br />

Charlie Chong/ Fion Zhang<br />

http://www.flirmedia.com/MMC/THG/Brochures/T559243/T559243_EN.pdf


<strong>Reading</strong> 2<br />

MWIR for Remote Sniffing and Locating of Gases<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


A mid-wave infrared camera visualizes methane gas leaking from a<br />

drilling platform.<br />

Methane is a greenhouse gas that is significantly more damaging than<br />

carbon dioxide. When it leaks uncontrolled from industrial plants, it can easily<br />

escape into the atmosphere. This is seen primarily as an economic loss, but it<br />

also contributes to the destruction of the environment. Sensitive highresolution<br />

mid-wave infrared (MWIR) cameras can accurately detect such<br />

emissions without placing a direct-measurement gas-detection system in the<br />

danger zone.<br />

Raf Vandersmissen, Sinfrared, and Thomas Zimmermann, TIB <strong>Infrared</strong><br />

Solutions<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


Figure 1. The transmission spectrum of methane (CH4) shows a strong<br />

attenuation between 3.3 and 3.6 µm and at a wavelength of 7.6 µm.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


Gas leaking from a drilling platform<br />

A March 2012 incident on the Elgin Wellhead gas drilling platform in the<br />

North Sea focused public attention on the risk connected with offshore mining<br />

of oil and gas.1 Elgin Wellhead (Figure 2) is one of more than 400 rigs in the<br />

North Sea, and it is a significant part of the exploration of the Elgin/Franklin<br />

field. As was discovered later, a gas leak had formed at the cover of well G4,<br />

which had been installed a few months earlier. A sequence and combination<br />

of several individual events never seen before had led to a kind of stress<br />

corrosion, which occurred only on G4. The incident was caused mainly by a<br />

chalk layer about 1000 m above the gas reservoir.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


Figure 2. The<br />

Elgin Wellhead<br />

drilling platform is<br />

one of more than<br />

400 rigs located in<br />

the North Sea.<br />

Charlie Chong/ Fion Zhang


The Elgin Wellhead drilling platform is one of more than 400 rigs located in<br />

the North Sea.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


The Elgin Wellhead drilling platform is one of more than 400 rigs located in<br />

the North Sea.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


The Elgin Wellhead drilling platform is one of more than 400 rigs located in<br />

the North Sea.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


The Elgin Wellhead drilling platform is one of more than 400 rigs located in<br />

the North Sea.<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang<br />

http://www.theoildrum.com/node/9072


Charlie Chong/ Fion Zhang<br />

http://www.theoildrum.com/node/9072


As a consequence, a severe methane gas leak developed in this area.<br />

Production was halted, and the drilling platform crew was evacuated,<br />

because the exact position of the leak and the amount of gas emitted could<br />

not be determined. Additionally, it was not known whether the flame on top of<br />

the platform was still burning. Because of the danger of a possible explosion,<br />

a three-mile no-go area surrounding the rig was declared, and another 4000-ft<br />

exclusion zone was established in the airspace above the rig.<br />

However, these precautions complicated the initial efforts to close the leak,<br />

as appropriate countermeasures could be attempted only after the gas<br />

concentration at the accident location had fallen below a critical value. Since<br />

these exposure limits did not permit taking a direct gas probe and subsequent<br />

gas analysis, an alternative method had to be considered to determine the<br />

methane content in the environment from afar. It was found in an absorption<br />

measurement of the infrared light in hydrocarbons.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


IR absorption of hydrocarbons<br />

Hydrocarbons are chemical compounds that consist only of carbon and<br />

hydrogen. Their simplest configuration is methane, a colorless and odorless<br />

combustible gas having the notation CH4. Its molecular model is shown in<br />

Figure 1.2 Its four C-H bonds are pointing toward the corners of a<br />

tetrahedron.This molecule structure exhibits four atomic oscillation modes,<br />

differentiated by either generating a temporal dipole moment – modes (a) and<br />

(d) – or not – modes (b) and (c). Only the first two can be excited by an<br />

external infrared radiation. They have a transmission spectrum shown in<br />

Figure 1 for the wave number area reaching from 4000/cm (which relates to a<br />

wavelength of 2.5 µm) to 500/cm (20 µm).<br />

A wave number of 3019/cm (3.3-µm wavelength) relates to the absorption<br />

band of the antisymmetrical valence oscillation (a), whereas the deformation<br />

oscillation (d), whose movement is similar to the unfolding of an umbrella,<br />

causes a strong attenuation around a wavelength of 7.6 µm (at a wave<br />

number of 1306/cm). Both areas are suited for remote methane gas sniffing,<br />

with the stronger attenuation value favoring the use of the MWIR band around<br />

3.3 µm. Additionally, it would be advantageous if the image components in<br />

the visible spectrum were captured as well for a spatial assignment.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


MWIR camera with extended spectrum area<br />

In the case of the gas leak at the Elgin Wellhead drilling platform, the<br />

measurement task was defined primarily as visualizing the gas components<br />

(methane, butane and carbon dioxide), as well as the characteristics of their<br />

spreading in the form of a gas cloud close to the sea surface and in the<br />

atmosphere. An additional goal was determining the activity of the flame on the<br />

platform. These investigations were carried out by TIB <strong>Infrared</strong> Solutions, in<br />

cooperation with the laboratory for environmental measurements at the University<br />

of Applied Sciences Düsseldorf.<br />

The Onca MWIR camera from Xenics, with expanded spectral coverage, was<br />

used for this purpose (Figure-3).<br />

The Onca camera family is equipped with a Stirling-cooled detector array in<br />

mercury-cadmium-telluride (MCT) or InSb (why or ?) detector technology. It is<br />

especially well suited for research and development tasks with high-speed image<br />

capture or IR spectroscopy. The camera operates in the extended MWIR area<br />

from 3.7 to 4.8 µm, which can be further expanded to cover the<br />

1 µm - to 5 µm wavelength area. This camera layout increases its all-weather and<br />

night-vision suitability and is of specific interest for spectroscopy applications.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


Figure 3. The extended spectral area of an MWIR camera simplifies the<br />

spatial assignment of gas plumes to visible objects.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


Using the True<strong>Thermal</strong> technique, this camera system offers an accuracy of<br />

±2 °C (or ±2 percent, whichever is largest). This is made possible by a new,<br />

stable method called NUC (nonuniformity correction), which eliminates the<br />

need for frequent adjustments. Four optional radiometric temperature ranges<br />

– −20 to 120 °C, 50 to 400 °C, 300 to 1200 °C and 1000 to 2000 °C – cover a<br />

wide variety of applications.<br />

The camera platform is optimized for autonomous operation and also in<br />

combination with a PC. Among other valuable features, it offers a mode for<br />

real-time image correction. The area-array sensor is available with a standard<br />

resolution of 320 × 256 pixels (InSb) or 384 × 288 pixels (MCT), or in a highresolution<br />

version featuring 640 × 512 pixels. Both versions offer the<br />

sensitivity specification NETD (noise-equivalent temperature difference) of<br />

less than 20 mK (?) .<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


The camera output delivers image<br />

The camera output delivers image data that is 14 bits wide at various frame rates.<br />

Two speeds are available: 30-Hz standard video with 640 × 512 pixels, and 60-<br />

Hz with 320 × 256 (InSb) or 384 × 288 (MCT) pixels. Additionally, 100 Hz is<br />

available at high resolution, as well as an extremely fast version with 460 Hz and<br />

320 × 256-pixel resolution. <strong>Reading</strong> out partial images will further increase the<br />

frame rate, which is especially useful for process-control applications at high<br />

frame rates. The camera functions can be tailored to the specific application, with<br />

the parameters stored in the camera’s nonvolatile memory.<br />

To adapt the camera to a specific measurement procedure, five filter-position<br />

options have been incorporated. Unlike most filter holders, which accept just one<br />

25-mm filter of 1-mm thickness, the holders on the Onca’s filter wheel are much<br />

deeper, and each position can hold several filters up to a total thickness of 3 mm<br />

for spectral measurements.<br />

The camera is user-programmable to offer individual filter positions and<br />

integration times for each exposure. However, because of the additional time<br />

needed for changing the filter and readjusting the operational parameters, the<br />

maximum frame rate mentioned above cannot always be reached.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


This flexibility, together with a stable thermal calibration across various<br />

integration times, enables a new operational mode called SuperFraming,<br />

which delivers thermal images of a high dynamic range in real time derived<br />

from images specifically exposed to users’ specifications. Using this special<br />

operational mode, the user can locate hot and cold spots in an image for fast<br />

process-monitoring sequences.<br />

Camera control and image capture are laid out for more than 100 images per<br />

second. Compatibility with GigE Vision and Camera Link simplifies the<br />

integration in user systems. For application development support, an API is<br />

available, as well as sample codes in C++, Visual Basic and Delphi; LabView<br />

drivers; and a comprehensive DLL library.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


Remote gas measurement<br />

The solution to the complex measurement task of visualizing gas components<br />

leaking from the well close to the drilling platform and spreading as a plume<br />

over the sea surface into the atmosphere was deploying the high-resolution<br />

(640 × 512 pixels) MWIR camera with an extended wavelength area from 1<br />

to 5 µm, operating at a frame rate of 100 Hz. Using this type of camera broke<br />

new ground for the researchers. The exclusion zone forced the IR<br />

investigation to be done from a large distance. This was an operational<br />

condition, for which there were no previously established procedures or<br />

proven results.<br />

The methane clouds can be detected even under these difficult measurement<br />

conditions; this is shown in an image taken from a plane flying over the scene<br />

(Figure 4). In the upper right is the drilling platform, with flame burning. The<br />

shadows visible in the foreground and in the center are caused by the<br />

methane gas plumes. They are illuminated by light reflections on the sea<br />

surface, and they absorb the midrange IR radiation at around 3.3 µm.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


This image-capture process requires two filters, one for the bandpass around 3.3<br />

µm, for methane absorption, and a second in the short-wave IR to make physical<br />

objects such as the drilling platform visible. By analyzing this dual-band image,<br />

certain gas concentrations can be assigned to fixed objects. This makes the<br />

interpretation of such images much easier. To detect gases whose transmission<br />

spectrum is different from that of methane, other filters or filter combinations are<br />

available. The results can be presented in different colors to distinguish among<br />

several gases in one image.<br />

Taking aerial pictures of the gas-leak scene was not the only investigative<br />

method for determining the scope of the disturbance at the drilling platform.<br />

Examinations were also carried out from aboard a ship (Figure 5). They proved<br />

that the water molecules and drops suspended in the atmosphere above the sea<br />

level are causing a strong attenuation of the MWIR radiation. Despite this effect,<br />

the gas components located behind this water vapor can be determined because<br />

naturally occurring clouds appear fully saturated, whereas the gas plume appears<br />

more like a fog. Deploying the MWIR camera allowed a quick check of the<br />

working conditions so that a repair crew could be sent to the site. On May 15, the<br />

leak was plugged by discharging heavy mud into the well. After final closure with<br />

five cement plugs in October 2012, the drilling platform went back to operation in<br />

March 2013.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


Figure 4. From above, methane plumes look like dark clouds in the MWIR<br />

spectrum.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


Figure 5. The high-resolution, high-sensitivity MWIR camera Onca performs<br />

a checkup from sea level.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


Figure 6. False-color representation of MWIR images enables the detection<br />

of the gas plume behind clouds formed by water vapor.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


False-color representation of MWIR images enables the detection of the gas<br />

plume behind clouds formed by water vapor.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


False-color representation of MWIR images enables the detection of the gas<br />

plume behind clouds formed by water vapor.<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


Gas leaking from a drilling platform<br />

http://www.dailymail.co.uk/sciencetech/article-2125471/Total-gas-leak-Greenpeace-release-infra-red-image-explosive-gas-spewing-Elgin-rig.html#v-1535918156001<br />

Charlie Chong/ Fion Zhang<br />

http://www.photonics.com/Article.aspx?AID=56861


FLIR GF-Series <strong>Infrared</strong> Cameras for Safe Gas Detection<br />

Refineries require state of the art technology to achieve results that are safe for the environment and safe for business. In fact our<br />

GF-Series infrared cameras have been developed together with oil industries and the American Petroleum Institute (API) to meet<br />

their requirements for detecting and minimizing gas leaks. The use of infrared cameras has already become a standard practice in<br />

many oil and gas companies. It's a proactive way to identify sources of Volatile Organic Compound (VOC) emissions and repair<br />

leaking components before it's too late. By using the most advanced VOC detection, you will improve safety and productivity and<br />

minimize emissions.<br />

■<br />

http://www.flir.co.uk/ogi/display/?id=49559<br />

Charlie Chong/ Fion Zhang


<strong>Reading</strong> 3<br />

What is Emissivity<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htm


More <strong>Reading</strong>: What is<br />

Emissivity<br />

All objects and materials do not<br />

radiate infrared (thermal) energy<br />

equally. Emissivity is a term<br />

describing the efficiency with which a<br />

material radiates infrared energy. A<br />

blackbody has an emissivity of 1.00<br />

and no other material can radiate<br />

more thermal energy at a given<br />

temperature. An object with an<br />

emissivity of 0 emits no infrared<br />

energy. Real-world objects have<br />

emissivity values between 0 and<br />

1.00.<br />

The lower emissivity of most real-world materials reduces the intensity of radiation<br />

from the theoretical predictions of Planck’s Law.<br />

The temperature of an object and its emissivity define how much infrared energy an<br />

object will emit. The figure below shows that quartz emits less energy than a<br />

blackbody at the same temperature and therefore has an emissivity below 1.00.<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htm


Physics of <strong>Infrared</strong>’s Emissivity<br />

<strong>Infrared</strong> (thermal) energy, when incident upon matter, be it solid, liquid or gas, will exhibit the properties of<br />

absorption σ, reflection ρ, and transmission τ to varying degrees.<br />

τ<br />

Absorption σ<br />

Absorption is the degree to which infrared energy is absorbed by a material. Materials such as plastic, ceramic,<br />

and textiles are good absorbers. <strong>Thermal</strong> energy absorbed by real-world objects is generally retransferred to<br />

their surroundings by conduction, convection, or radiation.<br />

Transmission τ<br />

Transmission is the degree to which thermal energy passes through a material. There are few materials that<br />

transmit energy efficiently in the infrared region between 7µm and 14µm. Germanium is one of the few good<br />

transmitters of infrared energy and thus it is used frequently as lens material in thermal imaging systems.<br />

Reflection ρ<br />

Reflection is the degree to which infrared energy reflects off a material. Polished metals such as aluminum,<br />

gold and nickel are very good reflectors. Conservation of energy implies that the amount of incident energy is<br />

equal to the sum of the absorbed, reflected, and transmitted energy.<br />

(1) Incident Energy = Absorbed Energy W σ<br />

+ Transmitted Energy W τ<br />

+ Reflected Energy W ρ<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htm


Emitted Energy = Absorbed Energy<br />

Consider equation 1 for an object in a vacuum at a constant temperature. Because it is in<br />

a vacuum, there are no other sources of energy input to the object or output from the object.<br />

The absorbed energy by the object increases its thermal energy - the transmitted and<br />

reflected energy does not. In order for the temperature of the object to remain constant, the<br />

object must radiate the same amount of energy as it absorbs.<br />

(2) Emitted Energy = Absorbed Energy<br />

Therefore, objects that are good absorbers are good emitters and objects that are poor<br />

absorbers are poor emitters. Applying equation 2, Equation 1 can be restated as follows:<br />

(3) Incident Energy = Emitted Energy + Transmitted Energy + Reflected Energy<br />

Setting the incident energy equal to 100%, the equation 3 becomes:<br />

(4) 100% = %Emitted Energy + %Transmitted Energy + %Reflected Energy<br />

Because emissivity equals the efficiency with which a material radiates energy, equation 4<br />

can be restated as follows:<br />

(5) 100% = Emissivity + %Transmitted Energy + %Reflected Energy<br />

Applying similar terms to %Transmitted Energy and %Reflected Energy,<br />

(6) 100% = Emissivity + Transmissivity + Reflectivity<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htm


According to equation 6, there is a balance between emissivity, transmissivity, and reflectivity.<br />

Increasing the value of one of these parameters requires a decrease in the sum of the other two<br />

parameters. If the emissivity of an object increases, the sum of its transmissivity and reflectivity<br />

must decrease. Likewise, if the reflectivity of an object increases, the sum of its emissivity and<br />

trasmissivity must decrease.<br />

Most solid objects exhibit very low transmission of infrared energy - the majority of incident<br />

energy is either absorbed or reflected. By setting transmissivity equal to zero, equation 6 can be<br />

restated as follows:<br />

(7) 100% = Emissivity + Reflectivity<br />

For objects that do not transmit energy, there is a simple balance between emissivity and<br />

reflectivity. If emissivity increases, reflectivity must decrease. If reflectivity increases, emissivity<br />

must decrease. For example, a plastic material with emissivity = 0.92 has reflectivity = 0.08. A<br />

polished aluminum surface with emissivity = 0.12 has reflectivity = 0.88.<br />

The emissive and reflective behavior of most materials is similar in the visible and infrared<br />

regions of the electromagnetic spectrum. Polished metals, for example, have low emissivity and<br />

high reflectivity in both the visible and infrared. It is important to understand, however, that some<br />

materials that are good absorbers, transmitters, or reflectors in the visible, may exhibit<br />

completely different characteristics in the infrared.<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htm


<strong>Infrared</strong> Thermometer Emissivity tables<br />

http://www.scigiene.com/pdfs/428_<strong>Infrared</strong>ThermometerEmissivitytablesrev.pdf<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htm


<strong>Infrared</strong> Thermometer Emissivity tables<br />

<strong>Understanding</strong> an object's emissivity or its characteristic "radiance" is a critical<br />

component in the proper handling of infrared measurements. Concisely, emissivity<br />

is the ratio of radiation emitted by a surface or blackbody and its theoretical<br />

radiation predicted from Planck's law. [W(L,T)=C1/(L^5*(exp(C2/LT)-1)] A<br />

material's surface emissivity is measured by the amount of energy emitted when<br />

the surface is directly observed. There are many variables that affect a specific<br />

object's emissivity, such as the wavelength of interest, field of view, the geometric<br />

shape of the blackbody, and temperature. However, for the purposes and<br />

applications of the infrared thermometer user, a comprehensive table showing the<br />

emissivity at corresponding temperatures of various surfaces and objects is<br />

displayed.<br />

The table below can be used to adjust the emissivity of any of the listed I.R.<br />

thermometers:<br />

• TN408LC<br />

• TN418LD<br />

• TN40ALC<br />

• TN425LE<br />

All have variable emissivity in order to improve the accuracy of the readings.<br />

For further assistance please contact us at Scigiene Corp. www.scigiene.com<br />

416-261-4865 and we will be happy to assist.<br />

Non-Metal Emssivity table<br />

Material(Non-Metals) Temp degF(degC) Emissivity<br />

Adobe 68 (20) 0.9<br />

Asbestos<br />

Board 100 (38) 0.96<br />

Cement 32-392 (0-200) 0.96<br />

Cement, Red 2500 (1371) 0.67<br />

Cement, White 2500 (1371) 0.65<br />

Cloth 199 (93) 0.9<br />

Paper 100-700 (38-371) 0.93<br />

Slate 68 (20) 0.97<br />

Asphalt, pavement 100 (38) 0.93<br />

Asphalt, tar paper 68 (20) 0.93<br />

Basalt 68 (20) 0.72<br />

Brick<br />

Red, rough 70 (21) 0.93<br />

Gault Cream 2500-5000 (1371-2760) .26-.30<br />

Fire Clay 2500 (1371) 0.75<br />

Light Buff 1000 (538) 0.8<br />

Lime Clay 2500 (1371) 0.43<br />

Fire Brick 1832 (1000) .75-.80<br />

Magnesite, Refractory 1832 (1000) 0.38<br />

Grey Brick 2012 (1100) 0.75<br />

Silica, Glazed 2000 (1093) 0.88<br />

Silica, Unglazed 2000 (1093) 0.8<br />

Sandlime 2500-5000 (1371-2760) .59-.63<br />

Carborundum 1850 (1010) 0.92


Material(Non-Metals) Temp degF(degC) Emissivity<br />

Ceramic<br />

Alumina on Inconel 800-2000 (427-1093) .69-.45<br />

Earthenware, Glazed 70 (21) 0.9<br />

Earthenware, Matte 70 (21) 0.93<br />

Greens No. 5210-2C 200-750 (93-399) .89-.82<br />

Coating No. C20A 200-750 (93-399) .73-.67<br />

Porcelain 72 (22) 0.92<br />

White Al2O3 200 (93) 0.9<br />

Zirconia on Inconel 800-2000 (427-1093) .62-.45<br />

Clay 68 (20) 0.39<br />

Fired 158 (70) 0.91<br />

Shale 68 (20) 0.69<br />

Tiles, Light Red 2500-5000 (1371-2760) .32-.34<br />

Tiles, Red 2500-5000 (1371-2760) .40-.51<br />

Tiles,Dark Purple 2500-5000 (1371-2760) 0.78<br />

Concrete<br />

Rough 32-2000 (0-1093) 0.94<br />

Tiles, Natural 2500-5000 (1371-2760) .63-.62<br />

Brown 2500-5000 (1371-2760) .87-.83<br />

Black 2500-5000 (1371-2760) .94-.91<br />

Cotton Cloth 68 (20) 0.77<br />

Dolomite Lime 68 (20) 0.41<br />

Emery Corundum 176 (80) 0.86<br />

Glass<br />

Convex D 212 (100) 0.8<br />

Convex D 600 (316) 0.8<br />

Convex D 932 (500) 0.76<br />

Nonex 212 (100) 0.82<br />

Nonex 600 (316) 0.82<br />

Nonex 932 (500) 0.78<br />

Smooth 32-200(0-93) .92-.94<br />

Granite 70 (21) 0.45<br />

Gravel 100 (38) 0.28<br />

Gypsum 68 (20) .80-.90<br />

Ice, Smooth 32 (0) 0.97<br />

Ice, Rough 32 (0) 0.98<br />

Lacquer<br />

Black 200 (93) 0.96<br />

Blue, on Al Foil 100 (38) 0.78<br />

Clear, on Al Foil (2 coats) 200 (93) .08-.09<br />

Clear, on Bright Cu 200 (93) 0.66<br />

Clear, on Tarnished Cu 200 (93) 0.64<br />

Red, on Al Foil (2 coats) 100 (38) .60-.74<br />

White 200 (93) 0.95<br />

White, on Al Foil (2 coats) 100 (38) .69-.88<br />

Yellow, on Al Foil (2 coats) 100 (38) .57-.79<br />

Lime Mortar 100-500 (38-260) .90-.92


Material(Non-Metals) Temp degF(degC) Emissivity<br />

Lacquer - continued<br />

Limestone 100 (38) 0.95<br />

Marble, White 100 (38) 0.95<br />

Smooth, White 100 (38) 0.56<br />

Polished Grey 100 (38) 0.75<br />

Mica 100 (38) 0.75<br />

Oil on Nickel<br />

0.001 Film 72 (22) 0.27<br />

0.002 " 72 (22) 0.46<br />

0.005 " 72 (22) 0.72<br />

Thick " 72 (22) 0.82<br />

Oil, Linseed<br />

On Al Foil, uncoated 250 (121) 0.09<br />

On Al Foil, 1 coat 250 (121) 0.56<br />

On Al Foil, 2 coats 250 (121) 0.51<br />

On Polished Iron, .001 Film 100 (38) 0.22<br />

On Polished Iron, .002 Film 100 (38) 0.45<br />

On Polished Iron, .004 Film 100 (38) 0.65<br />

On Polished Iron, Thick Film 100 (38) 0.83<br />

Paints<br />

Blue, Cu2O3 75 (24) 0.94<br />

Black, CuO 75 (24) 0.96<br />

Green, Cu2O3 75 (24) 0.92<br />

Red, Fe2O3 75 (24) 0.91<br />

White, Al2O3 75 (24) 0.94<br />

White, Y2O3 75 (24) 0.9<br />

White, ZnO 75 (24) 0.95<br />

White, MgCO3 75 (24) 0.91<br />

White, ZrO2 75 (24) 0.95<br />

White, ThO2 75 (24) 0.9<br />

White, MgO 75 (24) 0.91<br />

White, PbCO3 75 (24) 0.93<br />

Yellow, PbO 75 (24) 0.9<br />

Yellow, PbCrO4 75 (24) 0.93<br />

Paints, Aluminium 100 (38) .27-.67<br />

10% Al 100 (38) 0.52<br />

26% Al 100 (38) 0.3<br />

Dow XP-310 200 (93) 0.22<br />

Paints, Bronze Low .34-.80<br />

Gum Varnish (2 coats) 70 (21) 0.53<br />

Gum Varnish (3 coats) 70 (21) 0.5<br />

Cellulose Binder (2 coats) 70 (21) 0.34<br />

Paints, Oil<br />

All colours 200 (93) .92-.96<br />

Black 200 (93) 0.92<br />

Black Gloss 70 (21) 0.9<br />

Camouflage Green 125 (52) 0.85


Material(Non-Metals) Temp degF(degC) Emissivity<br />

Paints, Oil - continued<br />

Flat Black 80 (27) 0.88<br />

Flat White 80 (27) 0.91<br />

Grey-Green 70 (21) 0.95<br />

Green 200 (93) 0.95<br />

Lamp Black 209 (98) 0.96<br />

Red 200 (93) 0.95<br />

White 200 (93) 0.94<br />

Quartz, Rough, Fused 70 (21) 0.93<br />

Glass, 1.98 mm 540 (282) 0.9<br />

Glass, 1.98 mm 1540 (838) 0.41<br />

Glass, 6.88 mm 540 (282) 0.93<br />

Glass, 6.88 mm 1540 (838) 0.47<br />

Opaque 570 (299) 0.92<br />

Opaque 1540 (838) 0.68<br />

Red Lead 212 (100) 0.93<br />

Rubber, Hard 74 (23) 0.94<br />

Rubber, Soft, Grey 76 (24) 0.86<br />

Sand 68 (20) 0.76<br />

Sandstone 100 (38) 0.67<br />

Sandstone, Red 100 (38) .60-.83<br />

Sawdust 68 (20) 0.75<br />

Shale 68 (20) 0.69<br />

Silica,Glazed 1832 (1000) 0.85<br />

Silica, Unglazed 2012 (1100) 0.75<br />

Silicon Carbide 300-1200 (149-649) .83-.96<br />

Silk Cloth 68 (20) 0.78<br />

Slate 100 (38) .67-.80<br />

Snow, Fine Particles 20 (-7) 0.82<br />

Snow, Granular 18 (-8) 0.89<br />

Soil<br />

Surface 100 (38) 0.38<br />

Black Loam 68 (20) 0.66<br />

Plowed Field 68 (20) 0.38<br />

Soot<br />

Acetylene 75 (24) 0.97<br />

Camphor 75 (24) 0.94<br />

Candle 250 (121) 0.95<br />

Coal 68 (20) 0.95<br />

Stonework 100 (38) 0.93<br />

Water 100 (38) 0.67<br />

Waterglass 68 (20) 0.96<br />

Wood Low .80-.90<br />

Beech Planed 158 (70) 0.94<br />

Oak, Planed 100 (38) 0.91<br />

Spruce, Sanded 100 (38) 0.89


Metal Emssivity table<br />

Material(metal) Temp degF (degC) Emissivity<br />

Alloys<br />

20-Ni, 24-CR, 55-FE, Oxid. 392 (200) 0.9<br />

20-Ni, 24-CR, 55-FE, Oxid. 932(500) 0.97<br />

60-Ni, 12-CR, 28-FE, Oxid. 518 (270) 0.89<br />

60-Ni, 12-CR, 28-FE, Oxid. 1040 (560) 0.82<br />

80-Ni, 20-CR, Oxidised 212 (100) 0.87<br />

80-Ni, 20-CR, Oxidised 1112 (600) 0.87<br />

80-Ni, 20-CR, Oxidised 2372 (1300) 0.89<br />

Aluminium<br />

Unoxidised 77 (25) 0.02<br />

Unoxidised 212 (100) 0.03<br />

Unoxidised 932 (500) 0.06<br />

Oxidised 390 (199) 0.11<br />

Oxidised 1110 (599) 0.19<br />

Oxidised at 599degC(1110degF) 390 (199) 0.11<br />

Oxidised at 599degC(1110degF) 1110 (599) 0.19<br />

Heavily Oxidised 200 (93) 0.2<br />

Heavily Oxidised 940 (504) 0.31<br />

Highly Polished 212 (100) 0.09<br />

Roughly Polished 212 (100) 0.18<br />

Commercial Sheet 212 (100) 0.09<br />

Highly Polished Plate 440 (227) 0.04<br />

Highly Polished Plate 1070 (577) 0.06<br />

Bright Rolled Plate 338 (170) 0.04<br />

Bright Rolled Plate 932 (500) 0.05<br />

Alloy A3003, Oxidised 600 (316) 0.4<br />

Alloy A3003, Oxidised 900 (482) 0.4<br />

Alloy 1100-0 200-800 (93-427) 0.05<br />

Alloy 24ST 75 (24) 0.09<br />

Alloy 24ST, Polished 75 (24) 0.09<br />

Alloy 75ST 75 (24) 0.11<br />

Alloy 75ST, Polished 75 (24) 0.08<br />

Bismuth, Bright 176 (80) 0.34<br />

Bismuth, Unoxidised 77 (25) 0.05<br />

Bismuth, Unoxidised 212 (100) 0.06<br />

Brass<br />

73% Cu, 27% Zn, Polished 476 (247) 0.03<br />

73% Cu, 27% Zn, Polished 674 (357) 0.03<br />

62% Cu, 37% Zn, Polished 494 (257) 0.03<br />

62% Cu, 37% Zn, Polished 710 (377) 0.04<br />

83% Cu, 17% Zn, Polished 530 (277) 0.03<br />

Matte 68 (20) 0.07<br />

Burnished to Brown Colour 68 (20) 0.4<br />

Cu-Zn, Brass Oxidised 392 (200) 0.61<br />

Cu-Zn, Brass Oxidised 752 (400) 0.6<br />

Cu-Zn, Brass Oxidised 1112 (600) 0.61


Material(metal) Temp degF (degC) Emissivity<br />

Brass - continued<br />

Unoxidised 77 (25) 0.04<br />

Unoxidised 212 (100) 0.04<br />

Cadmium 77 (25) 0.02<br />

Carbon<br />

Lampblack 77 (25) 0.95<br />

Unoxidised 77 (25) 0.81<br />

Unoxidised 212 (100) 0.81<br />

Unoxidised 932 (500) 0.79<br />

Candle Soot 250 (121) 0.95<br />

Filament 500 (260) 0.95<br />

Graphitized 212 (100) 0.76<br />

Graphitized 572 (300) 0.75<br />

Graphitized 932 (500) 0.71<br />

Chromium 100 (38) 0.08<br />

Chromium 1000 (538) 0.26<br />

Chromium, Polished 302 (150) 0.06<br />

Cobalt, Unoxidised 932 (500) 0.13<br />

Cobalt, Unoxidised 1832 (1000) 0.23<br />

Columbium, Unoxidised 1500 (816) 0.19<br />

Columbium, Unoxidised 2000 (1093) 0.24<br />

Copper<br />

Cuprous Oxide 100 (38) 0.87<br />

Cuprous Oxide 500 (260) 0.83<br />

Cuprous Oxide 1000 (538) 0.77<br />

Black, Oxidised 100 (38) 0.78<br />

Etched 100 (38) 0.09<br />

Matte 100 (38) 0.22<br />

Roughly Polished 100 (38) 0.07<br />

Polished 100 (38) 0.03<br />

Highly Polished 100 (38) 0.02<br />

Rolled 100 (38) 0.64<br />

Rough 100 (38) 0.74<br />

Molten 1000 (538) 0.15<br />

Molten 1970 (1077) 0.16<br />

Molten 2230 (1221) 0.13<br />

Nickel Plated 100-500 (38-260) 0.37<br />

Dow Metal 0.4-600 (-18-316) 0.15<br />

Gold<br />

Enamel 212 (100) 0.37<br />

Plate (.0001) ·· ··<br />

Plate on .0005 Silver 200-750 (93-399) .11-.14<br />

Plate on .0005 Nickel 200-750 (93-399) .07-.09<br />

Polished 100-500 (38-260) 0.02<br />

Polished 1000-2000 (538-1093) 0.03<br />

Haynes Alloy C,<br />

Oxidised 600-2000 (316-1093) .90-.96


Material(metal) Temp degF (degC) Emissivity<br />

Haynes Alloy 25,<br />

Oxidised 600-2000 (316-1093) .86-.89<br />

Haynes Alloy X,<br />

Oxidised 600-2000 (316-1093) .85-.88<br />

Inconel Sheet 1000 (538) 0.28<br />

Inconel Sheet 1200 (649) 0.42<br />

Inconel Sheet 1400 (760) 0.58<br />

Inconel X, Polished 75 (24) 0.19<br />

Inconel B, Polished 75 (24) 0.21<br />

Iron<br />

Oxidised 212 (100) 0.74<br />

Oxidised 930 (499) 0.84<br />

Oxidised 2190 (1199) 0.89<br />

Unoxidised 212 (100) 0.05<br />

Red Rust 77 (25) 0.7<br />

Rusted 77 (25) 0.65<br />

Liquid 2760-3220 (1516-1771) .42-.45<br />

Cast Iron<br />

Oxidised 390 (199) 0.64<br />

Oxidised 1110 (599) 0.78<br />

Unoxidised 212 (100) 0.21<br />

Strong Oxidation 40 (104) 0.95<br />

Strong Oxidation 482 (250) 0.95<br />

Liquid 2795 (1535) 0.29<br />

Wrought Iron<br />

Dull 77 (25) 0.94<br />

Dull 660 (349) 0.94<br />

Smooth 100 (38) 0.35<br />

Polished 100 (38) 0.28<br />

Lead<br />

Polished 100-500 (38-260) .06-.08<br />

Rough 100 (38) 0.43<br />

Oxidised 100 (38) 0.43<br />

Oxidised at 1100 100 (38) 0.63<br />

Gray Oxidised 100 (38) 0.28<br />

Magnesium 100-500 (38-260) .07-.13<br />

Magnesium Oxide 1880-3140 (1027-1727) .16-.20<br />

Mercury 32 (0) 0.09<br />

Mercury 77 (25) 0.1<br />

Mercury 100 (38) 0.1<br />

Mercury 212 (100) 0.12<br />

Molybdenum 100 (38) 0.06<br />

Molybdenum 500 (260) 0.08<br />

Molybdenum 1000 (538) 0.11<br />

Molybdenum 2000 (1093) 0.18<br />

Molybdenum Oxidised at 1000degF 600 (316) 0.8<br />

Molybdenum Oxidised at 1000degF 700 (371) 0.84


Material(metal) Temp degF (degC) Emissivity<br />

Lead - continued<br />

Molybdenum Oxidised at 1000degF 800 (427) 0.84<br />

Molybdenum Oxidised at 1000degF 900 (482) 0.83<br />

Molybdenum Oxidised at 1000degF 1000 (538) 0.82<br />

Monel, Ni-Cu 392 (200) 0.41<br />

Monel, Ni-Cu 752 (400) 0.44<br />

Monel, Ni-Cu 1112 (600) 0.46<br />

Monel, Ni-Cu Oxidised 68 (20) 0.43<br />

Monel, Ni-Cu Oxid. at 1110degF 1110 (599) 0.46<br />

Nickel<br />

Polished 100 (38) 0.05<br />

Oxidised 100-500 (38-260) .31-.46<br />

Unoxidised 77 (25) 0.05<br />

Unoxidised 212 (100) 0.06<br />

Unoxidised 932 (500) 0.12<br />

Unoxidised 1832 (1000) 0.19<br />

Electrolytic 100 (38) 0.04<br />

Electrolytic 500 (260) 0.06<br />

Electrolytic 1000 (538) 0.1<br />

Electrolytic 2000 (1093) 0.16<br />

Nickel Oxide 1000-2000 (538-1093) .59-.86<br />

Palladium Plate (.00005 on .0005 silver) 200-750 (93-399) .16-.17<br />

Platinum 100 (38) 0.05<br />

Platinum 500 (260) 0.05<br />

Platinum 1000 (538) 0.1<br />

Platinum, Black 100 (38) 0.93<br />

Platinum, Black 500 (260) 0.96<br />

Platinum, Black 2000 (1093) 0.97<br />

Platinum Oxidised at 1100 500 (260) 0.07<br />

Platinum Oxidised at 1100 1000 (538) 0.11<br />

Rhodium Flash (0.0002 on 0.0005 Ni) 200-700 (93-371) .10-.18<br />

Silver<br />

Plate (0.0005 on Ni) 200-700 (93-371) .06-.07<br />

Polished 100 (38) 0.01<br />

Polished 500 (260) 0.02<br />

Polished 1000 (538) 0.03<br />

Polished 2000 (1093) 0.03<br />

Steel<br />

Cold Rolled 200 (93) .75-.85<br />

Ground Sheet 1720-2010 (938-1099) .55-.61<br />

Polished Sheet 100 (38) 0.07<br />

Polished Sheet 500 (260) 0.1<br />

Polished Sheet 1000 (538) 0.14<br />

Mild Steel, Polished 75 (24) 0.1<br />

Mild Steel, Smooth 75 (24) 0.12<br />

Mild Steel,liquid 2910-3270 (1599-1793) 0.28<br />

Steel, Unoxidised 212 (100) 0.08


Material(metal) Temp degF (degC) Emissivity<br />

Steel - continued<br />

Steel, Oxidised 77 (25) 0.8<br />

Steel Alloys<br />

Type 301, Polished 75 (24) 0.27<br />

Type 301, Polished 450 (232) 0.57<br />

Type 301, Polished 1740 (949) 0.55<br />

Type 303, Oxidised 600-2000 (316-1093) .74-.87<br />

Type 310, Rolled 1500-2100 (816-1149) .56-.81<br />

Type 316, Polished 75 (24) 0.28<br />

Type 316, Polished 450 (232) 0.57<br />

Type 316, Polished 1740 (949) 0.66<br />

Type 321 200-800 (93-427) .27-.32<br />

Type 321 Polished 300-1500 (149-815) .18-.49<br />

Type 321 w/BK Oxide 200-800 (93-427) .66-.76<br />

Type 347, Oxidised 600-2000 (316-1093) .87-.91<br />

Type 350 200-800 (93-427) .18-.27<br />

Type 350 Polished 300-1800 (149-982) .11-.35<br />

Type 446, Polished 300-1500 (149-815) .15-.37<br />

Type 17-7 PH 200-600 (93-316) .44-.51<br />

Type 17-7 PH Polished 300-1500 (149-815) .09-.16<br />

Type C1020,Oxidised 600-2000 (316-1093) .87-.91<br />

Type PH-15-7 MO 300-1200 (149-649) .07-.19<br />

Stellite, Polished 68 (20) 0.18<br />

Tantalum, Unoxidised 1340 (727) 0.14<br />

Tantalum, Unoxidised 2000 (1093) 0.19<br />

Tantalum, Unoxidised 3600 (1982) 0.26<br />

Tantalum, Unoxidised 5306 (2930) 0.3<br />

Tin, Unoxidised 77 (25) 0.04<br />

Tin, Unoxidised 212 (100) 0.05<br />

Tinned Iron, Bright 76 (24) 0.05<br />

Tinned Iron, Bright 212 (100) 0.08<br />

Titanium<br />

Alloy C110M,Polished 300-1200 (149-649) .08-.19<br />

Oxidised at 538degC(1000degF) 200-800 (93-427) .51-.61<br />

Alloy Ti-95A,Oxidised at 538degC(1000degF) 200-800 (93-427) .35-.48<br />

Anodized onto SS 200-600 (93-316) .96-.82<br />

Tungsten<br />

Unoxidised 77 (25) 0.02<br />

Unoxidised 212 (100) 0.03<br />

Unoxidised 932 (500) 0.07<br />

Unoxidised 1832 (1000) 0.15<br />

Unoxidised 2732 (1500) 0.23<br />

Unoxidised 3632 (2000) 0.28<br />

Filament (Aged) 100 (38) 0.03<br />

Filament (Aged) 1000 (538) 0.11<br />

Filament (Aged) 5000 (2760) 0.35<br />

Uranium Oxide 1880 (1027) 0.79


Material(metal) Temp degF (degC) Emissivity<br />

Zinc<br />

Bright, Galvanised 100 (38) 0.23<br />

Commercial 99.1% 500 (260) 0.05<br />

Galvanised 100 (38) 0.28<br />

Oxidised 500-1000 (260-538)<br />

Polished 100 (38) 0.02<br />

Polished 500 (260) 0.03<br />

Polished 1000 (538) 0.04<br />

Polished 2000 (1093) 0.06


Effects of Emissivity<br />

<strong>Thermal</strong> imaging cameras detect and measure the sum of infrared energy<br />

over a range of wavelengths determined by the sensitivity of the camera’s<br />

detector. <strong>Thermal</strong> imagers cannot discriminate energy at 7µm from energy at<br />

14µm the way the human eye can distinguish various wavelengths of light as<br />

colors. They calculate the temperature objects by detecting and quantifying<br />

the emitted energy over the operational wavelength range of the detector.<br />

Temperature is then calculated by relating the measured energy to the<br />

temperature of a blackbody radiating an equivalent amount of energy<br />

according to Planck’s Blackbody Law.<br />

Because the emissivity of an object affects how much energy an object emits,<br />

emissivity also influences a thermal imager’s temperature calculation.<br />

Consider the case of two objects at the same temperature, one having high<br />

emissivity and the other low. Even though the two objects have the same<br />

temperature, the one with the low emissivity will radiate less energy.<br />

Consequently, the temperature calculated by the thermal imager will be lower<br />

than that calculated for the high emissivity object.<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htmc


Discussion<br />

Subject: They (<strong>Infrared</strong> Cameras) calculate the temperature of the objects by<br />

detecting and quantifying the emitted energy over the operational wavelength<br />

range of the detector. Temperature is then calculated by relating the<br />

measured energy to the temperature of a blackbody radiating an equivalent<br />

amount of energy according to Planck’s Black-body Law.<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htm


Apparent Temperature<br />

<strong>Thermal</strong> imaging cameras cannot detect the emissivity of objects in order to calculate<br />

their true temperature. They can only calculate the “apparent” temperature of objects.<br />

The apparent temperature of an object is a function of both its temperature and<br />

emissivity. Given two objects with the same true temperature but different emissivity,<br />

a higher apparent temperature will be calculated for the object with higher emissivity<br />

(see figure below). Given two objects with the same emissivity but different true<br />

temperature, a higher apparent temperature will be calculated for the object with<br />

higher true temperature. The apparent temperature of an object may be substantially<br />

different from its true temperature. Only when the emissivity of objects is known can<br />

thermal imagers compensate for emissivity and calculate true temperature.<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htm


Reflectivity<br />

Objects with high reflectivity can reflect energy radiated by other objects. For example, polished<br />

aluminum reflects about 90% of the energy incident upon its surface. Just as thermal imagers<br />

cannot detect the emissivity of objects in order to calculate their true temperature, they also can’t<br />

detect the reflectivity of objects. Therefore, when calculating the apparent temperature of an<br />

object, thermal imagers detect and quantify energy emitted from the object, as well as, energy<br />

reflected from the surface of the object. If an object reflects energy from another radiating source<br />

with a higher temperature, the apparent temperature that is calculated for the object will be<br />

higher than its true temperature. Likewise, if an object reflects energy from another radiating<br />

source with a lower temperature, the apparent temperature that is calculated for the object will be<br />

lower than its true temperature.<br />

Given two objects with the same true<br />

temperature but different emissivity, a higher<br />

apparent temperature will be calculated for<br />

the object with higher emissivity (see figure<br />

below). Given two objects with the same<br />

emissivity but different true temperature, a<br />

higher apparent temperature will be<br />

calculated for the object with higher true<br />

temperature. The apparent temperature of<br />

an object may be substantially different from<br />

its true temperature. Only when the<br />

emissivity of objects is known can thermal<br />

imagers compensate for emissivity and<br />

calculate true temperature.<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htm


Emissivity Makes a Temperature Difference: Blackbody Calibrator<br />

■<br />

https://www.youtube.com/embed/JElKE-ADXr8<br />

Charlie Chong/ Fion Zhang<br />

https://www.youtube.com/watch?v=JElKE-ADXr8


What are Emissivity and Reflectivity and How do They Affect <strong>Infrared</strong><br />

Temperature <strong>Reading</strong><br />

■<br />

https://www.youtube.com/embed/W5x7fNZTwXE<br />

Charlie Chong/ Fion Zhang<br />

https://www.youtube.com/watch?v=W5x7fNZTwXE


Emissivity Examples<br />

A material’s emissivity can vary at different wavelengths (selective emitter). Most<br />

materials, however, have relatively uniform emissivity throughout the wavelength<br />

range in which thermal imagers operate (grey body). For example, the emissivity of<br />

most plastics, ceramics, and metals does not vary significantly throughout the<br />

wavelength range of 7 to 14µm. Different materials can have widely different<br />

emissivity values within the range of 0 to 1.00 (see table below). Many common<br />

materials including plastics, ceramics, water, and organic materials have high<br />

emissivity. Uncoated metals may have very low emissivity. Polished stainless steel,<br />

for example, has an emissivity of approximately 0.1 and therefore emits only one tenth<br />

the amount of energy of a blackbody at the same temperature.<br />

Material Emissivity<br />

Human Skin 0.98 ; Water 0.95<br />

Aluminum (Polished) 0.10 ; Aluminum (Anodized) 0.65 ; Plastic 0.93 ; Ceramic 0.94<br />

Glass 0.87 ; Rubber 0.90 ; Cloth 0.95<br />

Note: The emissivity date in the table above are approximate values. The emissivity<br />

of a particular material depends on its specific chemical makeup and surface<br />

characteristics. Smooth, shiny surfaces, for example, tend to have higher reflectivity<br />

and thus, low emissivity.<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htm


Methods of Increasing Emissivity<br />

The surface characteristics of a material determine its emissivity. To increase<br />

a material’s emissivity, it is necessary to increase the emissivity of its surface.<br />

Following are several methods of altering a material's surface to increase its<br />

emissivity. For a given material, one or more of these approaches may be<br />

effective. Choose the method that is easiest to apply/remove and that has<br />

minimum affect on the temperature of the material. Apply coatings,<br />

treatments, liquids, tapes, or powders as thin as possible to prevent altering<br />

the thermal behavior of the original material. Example are:<br />

• Apply a thin layer of tape<br />

• Apply a thin layer of paint, lacquer, or other high emissivity coating<br />

• Apply a thin coating of baby powder or foot powder<br />

• Apply a thin layer of oil, water, or other high emissivity liquid<br />

• Apply a surface treatment such as anodizing<br />

• Roughen the surface (may require substantial roughening)<br />

Charlie Chong/ Fion Zhang<br />

http://www.optotherm.com/emiss-physics.htm


End Of <strong>Reading</strong><br />

Charlie Chong/ Fion Zhang


Good Luck<br />

Charlie Chong/ Fion Zhang


Good Luck<br />

Charlie Chong/ Fion Zhang


Charlie https://www.yumpu.com/en/browse/user/charliechong<br />

Chong/ Fion Zhang

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!