16.06.2015 Views

Understanding Infrared Thermography Reading 7 Part 2 of 2.pdf

Understanding Infrared Thermography Reading 7 Part 2 of 2.pdf

Understanding Infrared Thermography Reading 7 Part 2 of 2.pdf

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> Thermal Testing<br />

<strong>Reading</strong> VII <strong>Part</strong> 1 <strong>of</strong> 2<br />

My ASNT Level III,<br />

Pre-Exam Preparatory<br />

Self Study Notes<br />

12 June 2015<br />

Charlie Chong/ Fion Zhang


6. Basic Elements Of An In-house Program<br />

The creation <strong>of</strong> an in-house program to utilize infrared thermography would<br />

be customized to each facility’s methods <strong>of</strong> conducting operations. The basic<br />

elements <strong>of</strong> each program, however, would probably be much the same. This<br />

section outlines a generic approach to developing and implementing a<br />

comprehensive infrared thermography program. A discussion <strong>of</strong> the basic<br />

elements is followed by a sample program.<br />

Charlie Chong/ Fion Zhang


6.1 Basic Elements<br />

An in-house program can be developed by many different approaches. A<br />

program that is limited to the use <strong>of</strong> only qualitative thermal imaging<br />

instruments (as compared to radiometric/quantitative) is likely to be less<br />

comprehensive. Assuming that a program was created to make full use <strong>of</strong> a<br />

radiometric/quantitative imager and image processing s<strong>of</strong>tware, the following<br />

topics would need to be addressed:<br />

• Introduction<br />

• Definitions<br />

•Scope<br />

• Responsibilities<br />

• Precautions<br />

• Prerequisites<br />

• Conduct <strong>of</strong> the Survey<br />

• Acceptance criteria<br />

• Reporting requirements<br />

• Qualification <strong>of</strong> personnel<br />

• Scheduling<br />

• Equipment matrix<br />

• References<br />

Charlie Chong/ Fion Zhang


6.1.1 Introduction<br />

This section provides a discussion <strong>of</strong> the purpose and goal <strong>of</strong> the IR survey.<br />

6.1.2 Definitions<br />

In order to put the program in the proper context, the definitions should be at<br />

the front. This will allow the reader or reviewer to have an easy reference for<br />

the terminology that follows.<br />

6.1.3 Scope<br />

The scope <strong>of</strong> the program should be very specific as to what is covered and<br />

what is not. The applications for infrared thermography are very broad.<br />

Inspections <strong>of</strong> ro<strong>of</strong>s and buildings should not be addressed in a document<br />

that has inspections <strong>of</strong> safety-related equipment as its main purpose. An<br />

addendum to the main procedure should be used to avoid confusion.<br />

Charlie Chong/ Fion Zhang


6.1.4 Responsibilities<br />

This section should clearly delineate who is responsible for the various<br />

aspects <strong>of</strong> the program from administration through corrective action. The<br />

main areas <strong>of</strong> responsibility are administration, inspection (<strong>Infrared</strong><br />

Thermographer), and corrective action. Most <strong>of</strong> the difficulty in applying this<br />

technology is in image interpretation and diagnosis. It might be necessary to<br />

use others in this effort and, if so, their role should be specifically identified.<br />

6.1.5 Precautions<br />

Many <strong>of</strong> the infrared inspections necessitate that panels be removed from<br />

energized electrical equipment. Precautions as to electrical and personnel<br />

safety should be included.<br />

Charlie Chong/ Fion Zhang


IR Viewing Window<br />

Charlie Chong/ Fion Zhang<br />

http://www.testequipmentdepot.com/fluke/ir-windows/075-<br />

clkt.htm?utm_source=bing&utm_medium=cpc&utm_campaign=Bing%20Product%20Ad<br />

s&utm_term=%7BQueryString%7D


IR Viewing Window – Opaque Polymer Grill<br />

Charlie Chong/ Fion Zhang<br />

http://irviewingwindows.com/


6.1.6 Prerequisites<br />

All <strong>of</strong> the prerequisites for conducting the survey should be identified here.<br />

This should include the qualification <strong>of</strong> personnel, calibration <strong>of</strong> equipment,<br />

approvals needed from Operations and/or Management, and the required<br />

resources (equipment and personnel).<br />

6.1.7 Conduct <strong>of</strong> the Survey<br />

This section could reference or include specific procedures for inspections.<br />

Specific techniques and a suggested sequence <strong>of</strong> inspections could also be<br />

included.<br />

Charlie Chong/ Fion Zhang


6.1.8 Acceptance Criteria<br />

All survey results should be compared to either a baseline thermogram or<br />

other industry accepted standards. Problems or anomalies should then be<br />

reviewed for determination <strong>of</strong> which corrective action, if any, should be<br />

undertaken. The following acceptance criteria provide a generic example but<br />

would need adaptation for component-specific use.<br />

An alternative to the above classification is that used in Military Standard MIL-<br />

STD-2194 (1988). The MIL Standard uses four categories as follows:<br />

Charlie Chong/ Fion Zhang


The main difference between the two methods <strong>of</strong> problem classification is that<br />

the MIL Standard references temperature rise above ambient and the guide<br />

classification relates to a temperature rise above a reference value. That<br />

reference value could be ambient or, in the case <strong>of</strong> three-phase electrical<br />

circuits, a temperature rise above an adjacent phase. Each facility should<br />

adopt criteria that provide a balance between maintenance requirements and<br />

operational considerations.<br />

Charlie Chong/ Fion Zhang


6.1.9 Reporting Criteria<br />

A rigid process should be established when reporting the results <strong>of</strong> infrared<br />

inspections. This rigidity is necessary due to the ease <strong>of</strong> misinterpretation <strong>of</strong><br />

the thermograms by untrained personnel. A typical quarterly survey <strong>of</strong><br />

electrical equipment might result in 25 to 50 problems in 200 pieces <strong>of</strong><br />

inspected equipment. The vast majority <strong>of</strong> these problems might be minor in<br />

nature and require corrective action on a low priority. The process that works<br />

best, based on industry responses, is one that keeps the report distribution<br />

and decision-making in the hands <strong>of</strong> the right people (operations,<br />

maintenance, and/or program managers). The format for the report should<br />

also be consistent.<br />

Charlie Chong/ Fion Zhang


At a minimum, it should include the following:<br />

•Time/date<br />

• Equipment identification<br />

• Location<br />

• Specific problem<br />

• Corrective action recommended<br />

• Problem action criteria<br />

• Visible photograph<br />

• <strong>Infrared</strong> photograph<br />

• Inspector’s name and signature<br />

Charlie Chong/ Fion Zhang


6.1.10 Qualification <strong>of</strong> Personnel<br />

Personnel responsible for conducting the surveys and interpreting the results<br />

should be trained in the use <strong>of</strong> the equipment and certified by their employer.<br />

The training and certification criteria, established by the American Society for<br />

Nondestructive Testing (ASNT), should be adapted and incorporated into the<br />

program. These criteria are outlined in their document SNT-TC-1A and will be<br />

discussed in more detail in Section 7.<br />

6.1.11 Scheduling<br />

The documentation requirements and listing <strong>of</strong> equipment to be evaluated<br />

during the survey should be established in advance so that trends in<br />

equipment operation can be translated easily into predictions <strong>of</strong> future results.<br />

This is the key to predictive maintenance. The program must also be flexible<br />

enough to accommodate emergency inspections and inspections during<br />

unplanned outages. Typically, the administrator <strong>of</strong> the IR program provides<br />

this interface.<br />

Charlie Chong/ Fion Zhang


6.1.12 Equipment Matrix<br />

The equipment to be surveyed, the selection criteria, and the locations and<br />

frequency <strong>of</strong> inspection should be compiled in a matrix. Typically, the<br />

electrical equipment is grouped together, as are the other major component<br />

groups. An alternate approach would be to list the equipment in a route <strong>of</strong><br />

survey-format, which might save time for the infrared thermographer.<br />

6.1.13 References<br />

References to any helpful information should be provided. These typically<br />

include training materials, textbooks on the subject, and equipment operation<br />

manuals.<br />

Charlie Chong/ Fion Zhang


EPRI Licensed Material<br />

Basic Elements <strong>of</strong> an In-House Program<br />

6.2 Sample Program<br />

This section incorporates the above recommendations and could serve as the basis for a program<br />

using infrared thermography as part <strong>of</strong> a predictive maintenance program.<br />

1.0 INTRODUCTION<br />

1.1 This program is for the administration and conduct <strong>of</strong> an infrared inspection program <strong>of</strong><br />

electrical and mechanical equipment. The purpose <strong>of</strong> this program is to identify<br />

equipment that requires maintenance and to improve its reliability through the use <strong>of</strong><br />

infrared thermography (IR).<br />

1.2 This document contains the recommended scope, frequency, and corrective action criteria<br />

for routine and unscheduled infrared surveys.<br />

1.3 Requests for changes to this program and questions relative to it shall be directed to the<br />

administrator <strong>of</strong> the IR program.<br />

2.0 DEFINITIONS<br />

2.1 <strong>Infrared</strong> – Electromagnetic radiation having wavelengths that are greater than those <strong>of</strong><br />

visible light, but shorter than microwaves. As it applies to IR thermography, the<br />

wavelengths are between 3 to 15 micrometers.<br />

2.2 <strong>Infrared</strong> Survey – A comprehensive examination <strong>of</strong> components and equipment with an<br />

infrared imaging system.<br />

2.3 Emissivity – The ratio <strong>of</strong> radiance from a surface to the radiance at the same wavelength<br />

from a perfect blackbody at the same temperature. Functionally, this is the radiation<br />

efficiency <strong>of</strong> a surface in the infrared spectrum.<br />

2.4 Radiosity – Thermal energy <strong>of</strong> a surface as seen by the infrared detector.<br />

2.5 Thermogram – A recorded, displayed, or hard-copy image <strong>of</strong> the output <strong>of</strong> an infrared<br />

imaging system.<br />

2.6 Isotherm – A thermal contour on a thermogram where all <strong>of</strong> the spots along it are at the<br />

same apparent temperature.<br />

2.7 <strong>Infrared</strong> thermographer – An individual who is trained and qualified to operate infrared<br />

imaging equipment and to interpret the images.<br />

3.0 SCOPE<br />

3.1 The requirements <strong>of</strong> this procedure shall apply to all safety-related components. It shall<br />

also be applicable to non-safety-related equipment where financial benefit might be<br />

achieved by monitoring (that is, increased plant availability, decreased maintenance<br />

costs, and so on).<br />

3.2 This procedure includes guidelines for the following:<br />

• Component selection<br />

6-5


EPRI Licensed Material<br />

Basic Elements <strong>of</strong> an In-House Program<br />

• Interval selection<br />

• Determining component acceptability<br />

4.0 RESPONSIBILITIES<br />

4.1 Administrator <strong>of</strong> IR – It is the administrator’s responsibility to oversee the program. This<br />

includes making changes to the procedure. All surveys, whether they are scheduled or<br />

conducted on an emergency basis, shall be approved by the administrator or his/her<br />

designee. The administrator shall be responsible for budgeting, planning, and interfacing<br />

with outside organizations.<br />

4.2 <strong>Infrared</strong> Thermographer – The infrared thermographer is the only person trained and<br />

qualified to operate the infrared imaging equipment. He/she is responsible for conducting<br />

the surveys, interpreting the images, writing the reports, and acting as a technical<br />

resource to other plant departments. The infrared thermographer is responsible for the<br />

maintenance and calibration <strong>of</strong> the infrared imaging equipment.<br />

4.3 Cognizant Engineer – At the request <strong>of</strong> the infrared thermographer, a discipline-cognizant<br />

engineer will provide assistance in diagnosing a problem. The cognizant engineer will<br />

also suggest corrective action and provide coordination with other plant disciplines.<br />

4.4 Root Cause – Determination <strong>of</strong> root cause and the subsequent applicable action level<br />

shall be the responsibility <strong>of</strong> plant management. When necessary, the infrared<br />

thermographer shall request assistance from a cognizant systems or maintenance engineer<br />

in determining the root cause or the recommended corrective action.<br />

5.0 PRECAUTIONS<br />

5.1 Many <strong>of</strong> the components that are being inspected represent potential plant trip hazards;<br />

exercise extreme care.<br />

5.2 All safe work practices as outlined in the plant safety manual, shall be followed. These<br />

practices include exhibiting caution near energized electrical equipment, rotating<br />

equipment, and hot pipes. All surveys shall be conducted from a safe stable location.<br />

5.3 <strong>Infrared</strong> surveys within the Radiological Controls Area shall be conducted within the<br />

guidelines <strong>of</strong> the Health Physics Department. In areas <strong>of</strong> potential contamination, the<br />

infrared thermographer shall be responsible for covering the equipment with plastic as<br />

directed by Health Physics.<br />

5.4 When practical, surveys in areas <strong>of</strong> airborne contamination should be avoided. When this<br />

is not possible, a thin piece <strong>of</strong> polyethylene or plastic can be placed over the lens. If this<br />

is done, the transmittance <strong>of</strong> the covering must be taken into account.<br />

6.0 PREREQUISITES<br />

6.1 Personnel – The infrared thermographer and one craft person constitute the minimum<br />

personnel necessary to conduct a survey when the operating or opening <strong>of</strong> equipment is<br />

necessary.<br />

6-6


EPRI Licensed Material<br />

Basic Elements <strong>of</strong> an In-House Program<br />

6.2 Approvals – The required approvals to conduct a survey shall be coordinated with the IR<br />

administrator. The control room should be notified both prior to the start <strong>of</strong> the survey<br />

and at its end. If requested, the infrared thermographer will inform the control room prior<br />

to opening equipment that presents a possible plant trip hazard.<br />

6.3 Emergencies – In cases where requests for surveys are done on an emergency basis, the<br />

infrared thermographer shall fulfill the duties <strong>of</strong> the IR administrator and provide the<br />

necessary coordination.<br />

7.0 CONDUCT OF THE SURVEY<br />

7.1 The equipment survey matrix shall identify the equipment to be surveyed and the<br />

frequency <strong>of</strong> the survey.<br />

7.2 The sequence <strong>of</strong> the survey is not important unless specifically stated in the procedure or<br />

requested by either Maintenance or Operations. All equipment on the matrix must be<br />

surveyed unless it is not in operation or conditions dictate otherwise. The infrared<br />

thermographer shall note any exceptions in the inspection report.<br />

7.3 Standard practice is to videotape all surveys and to include an audio track for verbal<br />

identification and discussion.<br />

7.4 The thermal images must be <strong>of</strong> sufficient resolution to identify the components and any<br />

problem areas.<br />

7.5 When problems are identified, the thermographer shall reposition the imager and obtain<br />

more than one view. This is done to eliminate the possibility <strong>of</strong> apparent problems being<br />

caused by reflections from hot objects. The hard-copy images should be obtained from<br />

the position that provides the best image.<br />

7.6 All problems are to be photographed in the visible as well as in the infrared. This is to<br />

allow proper and easy identification <strong>of</strong> the problem areas, which will facilitate<br />

maintenance activities.<br />

7.7 The problems shall be customarily reported as a temperature rise. This rise can be<br />

calculated from ambient, thermal baseline data, or made by comparison in the cases<br />

where similar equipment exists.<br />

7.8 When absolute temperatures are requested or required, the infrared thermographer shall<br />

determine and use the target's effective emissivity to assure accuracy. A standard table <strong>of</strong><br />

effective emissivities will be developed by measurement and will be maintained by the<br />

infrared thermographer.<br />

7.9 Important information relating to test conditions, such as load, flow, and pressure shall be<br />

noted by the thermographer if it is available. This information will be used in component<br />

trend analysis.<br />

7.10 The components shall be inspected with the imager aimed along a line normal<br />

(perpendicular) to the target surface whenever possible, to minimize the potential for<br />

errors due to reflections.<br />

6-7


EPRI Licensed Material<br />

Basic Elements <strong>of</strong> an In-House Program<br />

7.11 During the infrared inspection, the components must also be inspected visually and any<br />

discolorations, questionable noise, or smell should be reported.<br />

7.12 In cases where precise measurements must be obtained, the instrument background<br />

radiation effects must be taken into account. Instrument background temperature can be<br />

determined by placing a good diffuse reflector (such as a piece <strong>of</strong> aluminum foil that has<br />

been crumpled and re-flattened) in ambient air and measuring its apparent temperature<br />

with the imager’s emissivity set to 1.0.<br />

7.13 Where external optics, such as telescopic and wide-angle lenses are used, the<br />

transmittance <strong>of</strong> the optics must be taken into account. The information that corrects the<br />

effects <strong>of</strong> these devices is supplied by the manufacturer and is entered directly into the<br />

imager s<strong>of</strong>tware.<br />

7.14 When measurements are being made on targets, the size <strong>of</strong> the target and the distance<br />

must be known. The IFOVmeas (Instantaneous Field <strong>of</strong> View for measurement) <strong>of</strong> the<br />

instrument must fit comfortably within the required target spot at the measurement<br />

distance. If these criteria are not satisfied, the instrument must be moved closer to the<br />

target and/or a higher magnification lens must be used. (See section 3.3.4 for a more<br />

detailed discussion <strong>of</strong> this subject).<br />

7.15 The survey should be done with the imager scanned at a speed that does not cause<br />

blurring <strong>of</strong> the image so that acceptable thermograms can be obtained from the videotape<br />

on playback.<br />

7.16 If requested or desired, a second (backup) measure <strong>of</strong> temperature can be obtained<br />

through the use <strong>of</strong> contact thermocouples or spot radiometers. (Care should be used in<br />

evaluating the results <strong>of</strong> measurements that are not calibrated.)<br />

7.17 In general, equipment shall be surveyed when in a normal operational state. In cases<br />

where equipment is not energized or running normally, the thermographer shall note it in<br />

the IR inspection report.<br />

7.18 Equipment such as batteries shall be surveyed during both normal operation and during<br />

discharge tests.<br />

7.19 Requests for equipment operation for the sole purpose <strong>of</strong> an infrared inspection shall be<br />

coordinated with operations by the IR administrator. In most cases, this should be<br />

avoided.<br />

7.20 All infrared inspections, whether done by on-site personnel or outside contractors, will be<br />

performed under the guidance and procedures listed in this program. Special tests outside<br />

<strong>of</strong> the normal inspection shall be reviewed and approved in advance by the IR<br />

administrator.<br />

8.0 ACCEPTANCE CRITERIA<br />

8.1 Subsequent to an initial thermal baseline, the following action levels are to be used to<br />

classify each problem:<br />

6-8


EPRI Licensed Material<br />

Basic Elements <strong>of</strong> an In-House Program<br />

Advisory (Level 1)<br />

Intermediate (Level 2)<br />

Serious (Level 3)<br />

Critical (Level 4)<br />

1°F to 15°F rise<br />

16°F to 50°F rise<br />

51°F to 100°F rise<br />

in excess <strong>of</strong> 100°F rise<br />

8.2 When indications on components fall into levels 2, 3, 4, section 9 <strong>of</strong> the program shall be<br />

followed for reporting.<br />

8.3 To determine acceptability <strong>of</strong> the inspection, the results and final report shall be<br />

compared against the criteria set forth in this program.<br />

9.0 REPORTING REQUIREMENTS<br />

9.1 Every scheduled and unscheduled infrared inspection shall be documented and reported<br />

in accordance with the requirements <strong>of</strong> this section (see Figure 6-1).<br />

9.2 At a minimum, the report shall contain the following:<br />

• Summary <strong>of</strong> inspection and findings<br />

• Equipment list<br />

• Data sheets with IR and visible photographs <strong>of</strong> anomalies<br />

• Root cause analysis and corrective action<br />

• Comments<br />

9.3 The report shall be issued to the IR administrator within five working days <strong>of</strong> the<br />

completion <strong>of</strong> the survey.<br />

9.4 A verbal report shall always be given to the on-site IR administrator upon completion <strong>of</strong><br />

the survey.<br />

9.5 The reporting <strong>of</strong> problems that fall within the four acceptance action levels are as<br />

follows:<br />

Advisory (Level 1)<br />

Normal cycle <strong>of</strong> corrective maintenance.<br />

Intermediate (Level 2) High priority during an unscheduled shutdown.<br />

Serious (Level 3)<br />

Critical (Level 4)<br />

Alert Operations—potential failure. Correct ASAP.<br />

Alert Operations, Management. Remove from service ASAP.<br />

9.6 Items classified as serious are to be immediately reported to the IR administrator who<br />

will advise Maintenance and Operations.<br />

9.7 Items classified as critical are to be immediately reported to Operations, Maintenance,<br />

and the IR administrator.<br />

6-9


EPRI Licensed Material<br />

Basic Elements <strong>of</strong> an In-House Program<br />

10.0 QUALIFICATION OF PERSONNEL<br />

10.1 The infrared thermographer shall be qualified by examination and certified by the plant to<br />

conduct the survey.<br />

10.2 The qualifying examination and training shall meet the guidelines <strong>of</strong> ASNT SNT-TC-1A<br />

(current edition).<br />

10.3 In addition to the ASNT qualifications, the thermographer shall be knowledgeable in the<br />

following areas:<br />

• Equipment-specific operation<br />

• <strong>Infrared</strong> theory<br />

• Heat transfer modes<br />

• Safety practices<br />

10.4 Certification <strong>of</strong> the thermographer shall be made through a written and a practical<br />

examination.<br />

10.5 The plant Training Department shall administer the initial and re-qualification training.<br />

11.0 SCHEDULING<br />

11.1 The IR administrator is responsible for scheduling all routine infrared inspections.<br />

11.2 The Equipment Matrix (Program, section 12.0) lists the frequency <strong>of</strong> inspection for each<br />

component.<br />

11.3 Inspections on an emergency basis or for a special test shall be scheduled and coordinated<br />

by the IR administrator.<br />

12.0 EQUIPMENT MATRIX<br />

12.1 Component Selection Criteria<br />

12.1.1 The components that are to be included in the thermographic analysis program should be<br />

selected based on the perceived or documented benefit <strong>of</strong> thermography on the type <strong>of</strong><br />

equipment and the following criteria categories:<br />

A. Critical: Critical equipment shall be defined as:<br />

• Equipment whose function is necessary and must be available at all times.<br />

• Equipment upon which thermography has been used to deviate from a specific<br />

vendor-recommended preventive maintenance activity.<br />

• Equipment necessary to maintain full-power generating capabilities (that is, nonredundant).<br />

B. Vital: Vital equipment shall be defined as those components whose function is<br />

necessary but that, through redundant design, do not have to be available at all times.<br />

6-10


EPRI Licensed Material<br />

Basic Elements <strong>of</strong> an In-House Program<br />

C. Vendor Recommended: Vendor-recommended equipment whose manufacturer or<br />

vendor recommends the periodic monitoring <strong>of</strong> the equipment with infrared<br />

thermography.<br />

D. Non-Vital: Non-Vital equipment shall be defined as:<br />

• Equipment whose replacement cost versus periodic monitoring cost does not differ<br />

greatly and does not fall into category A or B above.<br />

• Components that are used very infrequently and do not fall into category A or B.<br />

12.1.2 The IR administrator shall maintain a listing <strong>of</strong> all <strong>of</strong> the components in the<br />

thermographic analysis program, the category to which they belong, and their monitoring<br />

interval.<br />

12.1.3 Equipment in category D that has a failure history relating to thermography might be<br />

included in the program in order to determine root cause, or to prevent failure recurrences<br />

or significant inconveniences. Otherwise, equipment in category D should be omitted<br />

from the program.<br />

12.1.4 The above recommended component selection criteria should be applied predominantly<br />

to electrical equipment such as:<br />

• Motor control centers<br />

• Load centers<br />

• Transformers<br />

• Switchgear<br />

• Battery chargers<br />

• Switchyard equipment<br />

• Large motor termination<br />

12.1.5 The above criteria can also be applied to:<br />

• Pumps/motors<br />

• Steam traps<br />

• Valves<br />

12.2 Performance Intervals<br />

12.2.1 The selection <strong>of</strong> performance intervals should be based upon several factors, such as:<br />

• The impact <strong>of</strong> the component on plant operation and personnel safety if an<br />

unexpected failure were to occur.<br />

• The speed at which a component fault manifests itself into a stage <strong>of</strong> degradation,<br />

which affects the component’s operability.<br />

• Vendor/manufacturers’ recommendations.<br />

• The category <strong>of</strong> the component as stated in section 12.1.1 <strong>of</strong> the program.<br />

6-11


EPRI Licensed Material<br />

Basic Elements <strong>of</strong> an In-House Program<br />

12.2.2 When considering the vendor’s recommended frequency for thermography, the<br />

application <strong>of</strong> the equipment should be taken into consideration (that is, the run time<br />

experienced by the equipment in this installation versus what the vendor expects for<br />

typical run times). Also, if the component falls into categories A or B <strong>of</strong> 12.1.1, then the<br />

most limiting interval (between the vendor-recommended interval and the recommended<br />

interval in section 12.2.3 <strong>of</strong> the program) shall be used for the monitoring <strong>of</strong> the<br />

equipment.<br />

12.2.3 The following recommended intervals for the given categories should be used:<br />

A. Critical Equipment<br />

• Monitor quarterly for those components that are operated continuously or are optested<br />

at least quarterly.<br />

• Monitor semi-annually for those components that are operated continuously or are<br />

run-tested at least semi-annually.<br />

• At start-up, monitor when the component is placed on-line, is at a stabilized<br />

temperature, and has not been monitored for at least one monitoring interval.<br />

• Equipment less than 240 V does not require periodic monitoring.<br />

B. Vital Equipment<br />

• Monitor equipment greater than 480 V quarterly.<br />

• Monitor equipment greater than 240 V but less than 480 V semi-annually.<br />

• Equipment less than 240 V does not require periodic monitoring.<br />

12.2.4 Changes to monitoring intervals should be reviewed carefully prior to making changes in<br />

order to ensure that maximum component availability and program efficiency is<br />

provided.<br />

12.2.5 At a minimum, documentation for interval changes shall be maintained, by the IR<br />

administrator.<br />

12.2.6 Components need not be operated for the sole purpose <strong>of</strong> collecting thermography data.<br />

13.0 SUGGESTED PROGRAM REFERENCES<br />

13.1 <strong>Infrared</strong> <strong>Thermography</strong> Guide (Revision 3), (formerly NP-6973)<br />

13.2 Plant Administrative Procedures Manual<br />

13.3 Plant Safety Manual<br />

13.4 Plant Training Manual<br />

13.5 Plant Quality Assurance Procedures Manual<br />

13.6 Plant Systems Training Manual<br />

6-12


EPRI Licensed Material<br />

Basic Elements <strong>of</strong> an In-House Program<br />

13.7 <strong>Infrared</strong> Imager Instruction Manual<br />

13.8 Plant Predictive Maintenance, INPO Good Practice 89-009.<br />

13.9 Wolfe, W. L. and Zissis, G.J., The <strong>Infrared</strong> Handbook. Environmental Research Institute<br />

<strong>of</strong> Michigan (1996).<br />

13.10 Mil-Std-2194, <strong>Infrared</strong> Thermal Imaging Survey Procedure Electrical Equipment.<br />

13.11 American Society for Nondestructive Testing Standard Practice SNT-TC-1A,<br />

Qualifications Guidelines.<br />

13.12 American Society for Nondestructive Testing <strong>Infrared</strong> and Thermal Testing Handbook,<br />

2001.<br />

13.13 American Society for Nondestructive Testing Level III Study Guide: <strong>Infrared</strong> and<br />

Thermal Testing Method, 2001.<br />

6-13


EPRI Licensed Material<br />

Basic Elements <strong>of</strong> an In-House Program<br />

Figure 6-1<br />

<strong>Infrared</strong> Survey Results<br />

6-14


7. TRAINING AND CERTIFICATION<br />

This section deals solely with the efforts <strong>of</strong> the American Society <strong>of</strong><br />

Nondestructive Testing (ASNT) in the training and certification <strong>of</strong> infrared<br />

thermographers. The purpose is to provide guidelines for training individuals<br />

who will be able to deliver the best level <strong>of</strong> service possible. It is important to<br />

understand that certification via the ASNT Certification Program, does not<br />

imply authorization or licensing <strong>of</strong> the certificate holder to perform infrared<br />

thermography tasks. It is solely the employer's responsibility to review the<br />

individual’s qualification records for completeness and to authorize individuals<br />

to perform infrared thermography tasks.<br />

Charlie Chong/ Fion Zhang


7.1 Background<br />

Commercially available infrared imagers are quite easy to both use and<br />

misuse. Many small, independent contractors, from electricians to engineers,<br />

provide a wide range <strong>of</strong> services to many different industries. In the absence<br />

<strong>of</strong> formal training, most <strong>of</strong> these people have learned on the job while working<br />

with more experienced individuals. At the request <strong>of</strong> many ASNT members, a<br />

committee was formed in the fall <strong>of</strong> 1989 to propose modifying ASNT<br />

Recommended Practice No. SNT-TC-1A, the qualification guideline for<br />

nondestructive testing, to accept and recognize infrared thermography as a<br />

valid nondestructive examination method. At this writing, all <strong>of</strong> the training,<br />

qualification, and certification guidelines are in place and SNT-TC-1A has<br />

been updated (1996) to include the T/IR (Thermal <strong>Infrared</strong>) method. Two<br />

additional ASNT publications were released in 2001 to support training and<br />

certification:<br />

• ASNT <strong>Infrared</strong> and Thermal Testing Handbook, 2001<br />

• ASNT Level III Study Guide: <strong>Infrared</strong> and Thermal Testing Method, 2001<br />

Charlie Chong/ Fion Zhang


Recommended training and certification guidelines for infrared<br />

thermographers are summarized in the ASNT <strong>Infrared</strong> and Thermal Testing<br />

Handbook on pages 15 -18, and are explained in detail in SNT-TC-1A.<br />

The ASNT training program is intended to supplement equipment-specific<br />

training that might be <strong>of</strong>fered by the manufacturers. Certification is the<br />

responsibility <strong>of</strong> the individual employer. SNT-TC-1A states the following<br />

in this regard:<br />

“Written Practice. The employer shall establish a written practice for the<br />

control and administration <strong>of</strong> nondestructive personnel training, examination<br />

and certification. The employer’s written practice should describe the<br />

responsibility <strong>of</strong> each level <strong>of</strong> certification for determining the acceptability <strong>of</strong><br />

materials and components in accordance with applicable codes, standards,<br />

specifications and procedures.”<br />

Charlie Chong/ Fion Zhang


7.2 Levels <strong>of</strong> Qualification<br />

The recommended Levels <strong>of</strong> Qualification for infrared thermographers follow<br />

those <strong>of</strong> traditional NDE methods. These levels are as follows:<br />

■ Level I<br />

A Level I infrared thermographer shall be qualified to perform specific IR<br />

inspections in accordance with detailed written instructions and to record the<br />

results; the Level 1 infrared thermographer shall perform inspections under<br />

the cognizance <strong>of</strong> a Level II or Level III. The Level I shall not independently<br />

perform nor evaluate inspection results for acceptance or rejection when such<br />

inspection results are for the purpose <strong>of</strong> verifying compliance to code or<br />

regulatory requirements. (if the result is not for the purpose <strong>of</strong> verifying<br />

compliance to code or regulatory requirements; then the Level I could<br />

independently perform and evaluate inspection result?)<br />

Charlie Chong/ Fion Zhang


■ Level II<br />

A Level II infrared thermographer shall be qualified to set up and calibrate<br />

equipment, conduct inspections, and to interpret inspection results in<br />

accordance with procedure requirements. The individual shall be familiar with<br />

the limitations and scope <strong>of</strong> the method employed and shall have the ability to<br />

apply techniques over a broad range <strong>of</strong> applications within the limits <strong>of</strong> their<br />

certification. The Level II shall be able to organize and report inspection<br />

results. A Level II must have the ability to correctly identify components and<br />

parts <strong>of</strong> components within the scope <strong>of</strong> the IR inspection.<br />

■ Level III<br />

A Level III infrared thermographer is capable <strong>of</strong> designating a particular<br />

inspection technique, establishing techniques and procedures, and<br />

interpreting results. The individual shall have sufficient practical background<br />

in his/her area <strong>of</strong> expertise to develop innovative techniques and to assist in<br />

establishing acceptance criteria where none are otherwise available. The<br />

individual shall have general familiarity with other nondestructive evaluation<br />

(NDE) methods and inspection technologies. The Level III individual shall be<br />

qualified to train and examine Level I and Level II personnel for qualification<br />

and certification as an infrared thermographer.<br />

Charlie Chong/ Fion Zhang


7.3 Training Requirements<br />

The training requirements for each level <strong>of</strong> the infrared thermographer<br />

qualification parallel those for the other traditional NDE methods in that onthe-job<br />

training, educational background, and classroom work all count<br />

toward qualification. There are qualification examinations and annual requalification<br />

requirements at all levels. It is up to the utilities’ training<br />

organization and individual employers to implement the appropriate<br />

recommendations <strong>of</strong> the training program set forth in SNT-TC-1A.<br />

Charlie Chong/ Fion Zhang


The experience and education recommendations for the three levels are:<br />

• Level I A high school diploma (or equivalent) or 6 months <strong>of</strong> experience<br />

• Level II A two-year college or technical degree or 18 months <strong>of</strong> experience<br />

• Level III A four-year technical degree from a college or university or 5<br />

years <strong>of</strong> experience<br />

The required classroom training is as follows:<br />

• Level I 40 hours <strong>of</strong> instruction, 50-question written examination, classroom<br />

experiment<br />

• Level II 40 hours <strong>of</strong> instruction, 75-question written examination,<br />

classroom experiment<br />

• Level III 40 hours <strong>of</strong> instruction, 75-question written examination,<br />

procedure preparation for classroom experiment<br />

Charlie Chong/ Fion Zhang


The classroom training is based on the body <strong>of</strong> knowledge reviewed, adopted,<br />

and updated by ASNT, summarized in ASNT Recommended Practice No.<br />

SNT-TC-1A, and reviewed in ASNT Level III Study Guide: <strong>Infrared</strong> and<br />

Thermal Testing Method, 2001. The depth that is covered by these areas<br />

corresponds to the level <strong>of</strong> the training. This translates into more extensive<br />

training at Level III than Level I, even though the classroom hours are the<br />

same. The four areas for training and associated practical aspects are listed<br />

below. At the conclusion <strong>of</strong> training, the trainee will:<br />

A. Radiosity or Target Exitance<br />

• Understand the concepts <strong>of</strong> radiosity and associated parameters.<br />

• Be able to measure emissivity, reflectance, transmittance, background<br />

temperature, foreground temperature, and target temperature.<br />

• Be cognizant <strong>of</strong> potential errors in the measurement <strong>of</strong> the above<br />

parameters, caused by variation across the target surface.<br />

Charlie Chong/ Fion Zhang


B. Spatial Resolution<br />

• the concept <strong>of</strong> spatial resolution. • Understand the difference between<br />

image resolution and measurement resolution.<br />

• Understand the effect on measurement <strong>of</strong> the distance between the<br />

instrument and the target.<br />

• Be able to calculate measurement spot size.<br />

• Be able to exploit equipment- pecific aids to determine measurement<br />

adequacy.<br />

Charlie Chong/ Fion Zhang


C. Heat Transfer<br />

• Understand the fundamental concepts <strong>of</strong> heat transfer including<br />

conduction, convection, and radiation.<br />

• Understand the difference between steady state and transient heat flow<br />

and application dependence.<br />

• Understand the effect <strong>of</strong> the environmental conditions <strong>of</strong> sky temperature,<br />

view factor, wind velocity, and surface orientation.<br />

• Understand the potential problems if evaporation or condensation occur at<br />

the target surface.<br />

Charlie Chong/ Fion Zhang


D. Equipment Operation<br />

• Be able to set up and operate the necessary equipment.<br />

• Understand dynamic range and its implication in image acquisition.<br />

• Demonstrate good data acquisition practices.<br />

• Demonstrate the use <strong>of</strong> accessories.<br />

• Understand how to compensate for external optics.<br />

• Understand the implications <strong>of</strong> system spectral response.<br />

The written examination is derived from a pool <strong>of</strong> 200-300 questions that are<br />

reviewed and approved by the ASNT T/IR committee members. During<br />

training, practical exams are conducted through classroom experiments and<br />

are focused on one particular concept, such as transient thermal heat transfer.<br />

The actual practical exam is determined by the trainer and is conducted<br />

within the guidelines for each particular level. <strong>Infrared</strong> thermography was<br />

adopted as a nondestructive inspection method in the fall <strong>of</strong> 1991.<br />

Charlie Chong/ Fion Zhang


7.4 Predictive Maintenance (PdM) Level III Certification<br />

Program<br />

Recognizing that there are areas <strong>of</strong> specialization within the infrared<br />

thermography discipline, the ASNT T/IR committee has promoted the<br />

development <strong>of</strong> specialty certification. The Predictive Maintenance Level III<br />

Certification Program has been developed by ASNT in response to this effort.<br />

Developed to meet the needs <strong>of</strong> the predictive maintenance sector <strong>of</strong> the<br />

industry, this program incorporates the vibration analysis (VA) and<br />

infrared/thermal (IR) test methods. A PdM-specific body <strong>of</strong> knowledge,<br />

including knowledge <strong>of</strong> the Recommended Practice No. SNT-TC-1A and the<br />

ANSI/ASNT CP-189 standard, is used for the two-hour PdM basic<br />

examination. The VA and IR method tests are the same as those used in the<br />

ASNT NDT Level III program. A separate and distinct PdM Level III certificate<br />

is issued for this certification.<br />

Charlie Chong/ Fion Zhang


The PdM basic examination is more specific than the ASNT NDT Level III<br />

basic examination, and thus, PdM certificate holders wishing to gain<br />

traditional NDT Level III certification will still be required to sit for the ASNT<br />

NDT Level III basic examination, as well as taking an ASNT NDT Level III<br />

method test.<br />

Certification via the ASNT PdM Level III Certification Program, as with the<br />

ASNT NDT Level III program, does not imply authorization or licensing <strong>of</strong> the<br />

PdM certificate holder to perform PdM tasks. It is solely the employer’s<br />

responsibility to review the individual’s qualification records for completeness<br />

and to authorize individuals to perform PdM.<br />

Charlie Chong/ Fion Zhang


The Expert!


Appendix-A<br />

The Science Of <strong>Thermography</strong> (Practical<br />

Application Of Thermographic And Thermal<br />

Sensing Equipment)<br />

Charlie Chong/ Fion Zhang


A.1 Introduction<br />

This appendix is presented as a reference guide to provide the practical<br />

thermographer with an understanding <strong>of</strong> the science behind the<br />

measurements. It is intended as an aid in performing and understanding<br />

non-contact thermal and thermographic measurements using infrared sensing<br />

equipment. The deployment and operation <strong>of</strong> infrared sensing instruments<br />

was, at one time, cumbersome and difficult. Thermographers were <strong>of</strong>ten<br />

required to perform on-the-spot calculations in order to reduce their<br />

measurement data and determine actual temperature values; this is no longer<br />

so. Modern instruments are light in weight, portable, and rugged.<br />

Menu-driven on-board s<strong>of</strong>tware now makes it relatively simple to operate<br />

equipment and to gather data directly in terms <strong>of</strong> target temperature. Because<br />

<strong>of</strong> this very ease <strong>of</strong> operation, it is also relatively simple to misinterpret the<br />

results so easily and quickly obtained.<br />

Charlie Chong/ Fion Zhang


Erroneous conclusions can have an extremely negative effect on the<br />

measurements program and on the credibility <strong>of</strong> the thermographer.<br />

A solid understanding <strong>of</strong> the basis on which thermographic measurements<br />

are made will go a long way toward minimizing operator error and ensuring<br />

the success <strong>of</strong> the thermographic program.<br />

The subject matter in this appendix begins with a discussion <strong>of</strong> heat transfer<br />

and how radiative heat transfer is the basis for infrared thermography. The<br />

basic physics <strong>of</strong> infrared radiation and how it applies to instrument<br />

performance is explained. Finally, the performance parameters <strong>of</strong> infrared<br />

point-sensing and imaging instruments are discussed, including how to select,<br />

calibrate, and evaluate the performance <strong>of</strong> the instrument that is best suited<br />

to your application.<br />

Charlie Chong/ Fion Zhang


A.2 Heat Transfer and Radiation Exchange Basics for<br />

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

This section is to provide the reader with an understanding <strong>of</strong> how heat<br />

transfer phenomena affect non-contact infrared thermal sensing and<br />

thermographic measurements. <strong>Infrared</strong> thermography depends on measuring<br />

the distribution <strong>of</strong> radiant thermal energy (heat) emitted from a target surface,<br />

thus, the thermographer requires an understanding <strong>of</strong> heat, temperature, and<br />

the various types <strong>of</strong> heat transfer as an essential prerequisite in preparing to<br />

undertake a program <strong>of</strong> IR thermography.<br />

Charlie Chong/ Fion Zhang


A.2.1 Heat and Temperature<br />

What is <strong>of</strong>ten referred to as a heat source (like an oil furnace or an electric<br />

heater) is really one form or another <strong>of</strong> energy conversion; the energy stored<br />

in one object is converted to heat and flows to another object.<br />

Heat can be defined as thermal energy in transition. It flows from one place or<br />

object to another as a result <strong>of</strong> temperature difference, and the flow <strong>of</strong> heat<br />

changes the energy levels in the objects.<br />

Temperature is a property <strong>of</strong> matter and not a complete (that means it need<br />

other input to completely quantify the internal energy) measurement <strong>of</strong><br />

internal energy. It defines the direction <strong>of</strong> heat when another temperature is<br />

known. Heat always flows from the object that is at the higher temperature to<br />

the object that is at the lower temperature. As a result <strong>of</strong> heat transfer, hotter<br />

objects tend to become cooler and cooler objects become hotter, approaching<br />

thermal equilibrium. To maintain a steady-state condition, energy needs to be<br />

continuously supplied to the hotter object by some means <strong>of</strong> energy<br />

conversion so that the temperature and, hence, the heat flow remains<br />

constant.<br />

Charlie Chong/ Fion Zhang


A.2.2 Converting Temperature Units<br />

Temperature is expressed in either absolute or relative terms. There are two<br />

absolute scales called degree Rankine (English system) and Kelvin (metric<br />

system). There are two corresponding relative scales called Fahrenheit<br />

(English system) and Celsius or Centigrade (metric system). Absolute zero is<br />

the temperature at which no molecular action takes place. This is expressed<br />

as zero Kelvins or zero Rankines (0 K or 0 °R). Relative temperature is<br />

expressed as degrees Celsius or degrees Fahrenheit (°C or °F). The<br />

numerical relations among the four scales are as follows:<br />

T Celsius = 5/9 (T Fahrenheit - 32 )<br />

T Fahrenheit = 9/5 T Celsius + 32<br />

T Rankine = T Fahrenheit + 459.7<br />

T Kelvin = T Celsius + 273.16<br />

Absolute zero is equal to -273.1°C and is also equal to -459.7°F.<br />

Charlie Chong/ Fion Zhang


To convert changes in temperature or delta T between the English and Metric<br />

systems, the simple 9/5 (1.8 to 1) relationship is used:<br />

ΔT Fahrenheit (or Rankine) = 1.8 ΔT Celsius (or Kelvin)<br />

Table A-1 is a conversion table to allow for the rapid conversion <strong>of</strong><br />

temperature between Fahrenheit and Celsius values. Instructions for the use<br />

<strong>of</strong> the table are shown at the top.<br />

(ΔT ≡ temperature interval)<br />

Charlie Chong/ Fion Zhang


Table A-1 Temperature Conversion Chart Instructions for Use:<br />

1. Start in the Temp. column and find the temperature that you wish to convert.<br />

2. If the temperature to be converted is in °C, scan to the right column for the °F equivalent.<br />

3. If the temperature to be converted is in °F, scan to the left column for the °C equivalent.<br />

Charlie Chong/ Fion Zhang


A.2.3 The Three Modes <strong>of</strong> Heat Transfer<br />

There are three modes <strong>of</strong> heat transfer: (1) conduction, (2) convection, and (3)<br />

radiation (and nothing else) . All heat transfer processes occur by one or<br />

more <strong>of</strong> these three modes. <strong>Infrared</strong> thermography is based on the<br />

measurement <strong>of</strong> radiative heat flow radiation (and nothing else) and is,<br />

therefore, most closely related to the radiation mode <strong>of</strong> heat transfer.<br />

A.2.4 Conduction<br />

Conduction is the transfer <strong>of</strong> heat in stationary media. It is the only mode <strong>of</strong><br />

heat flow in solids, but can also take place in liquids and gases. It occurs as<br />

the result <strong>of</strong> molecular collisions (in liquids) (fluid, both liquid and gas) and<br />

atomic vibrations (in solids), whereby energy is moved one molecule at a time,<br />

from higher temperature sites to lower temperature sites. Figure A-1 is an<br />

illustration <strong>of</strong> conductive heat flow. The Fourier conduction law expresses the<br />

conductive heat flow through the slab shown in Figure A-1.<br />

Charlie Chong/ Fion Zhang


Figure A-1 Conductive Heat Flow<br />

Charlie Chong/ Fion Zhang


The Fourier Conduction Law:<br />

Q/A<br />

Q<br />

= K (T 1 -T 2 ) / L<br />

= K∙ΔT∙A / L<br />

Where:<br />

Q/A<br />

L<br />

T 1<br />

T 2<br />

K<br />

= the rate <strong>of</strong> heat transfer through the slab per unit area<br />

perpendicular to the flow<br />

= the thickness <strong>of</strong> the slab<br />

= the higher temperature (at the left)<br />

= the lower temperature (at the right)<br />

= the thermal conductivity <strong>of</strong> the slab material<br />

Charlie Chong/ Fion Zhang


Thermal conductivity is analogous to electrical conductivity and is inversely<br />

proportional to thermal resistance, as shown in the lower portion <strong>of</strong> Figure A-1.<br />

The temperatures, T 1 and T 2 , are analogous to voltages V1 and V2, and the<br />

heat flow, Q/A, is analogous to electrical current, I, so that: if:<br />

R electrical = V1 - V2/ I<br />

then:<br />

R thermal<br />

= T1 - T2 / Q /A = L/K<br />

Heat flow is usually expressed in English units. K is expressed in<br />

BTU/hr∙ft²∙°F and thermal resistance (1/K) would then be expressed in<br />

°F∙hr∙ft²/BTU.<br />

Charlie Chong/ Fion Zhang


A.2.5 Convection<br />

Convective heat transfer takes place in a moving medium and is almost<br />

always associated with transfer between a solid and a moving fluid (such as<br />

air). Forced convection takes place when an external driving force, such as<br />

wind or an air pump, moves the fluid. Free convection takes place when the<br />

temperature difference necessary for heat transfer produces density changes<br />

in the fluid and the warmer fluid rises as a result <strong>of</strong> increased buoyancy. In<br />

convective heat flow, heat transfer takes effect by means <strong>of</strong> two mechanisms,<br />

(1) the direct conduction through the fluid and (2) the motion <strong>of</strong> the fluid itself.<br />

Figure A-2 illustrates convective heat transfer between a flat plate and a<br />

moving fluid. The presence <strong>of</strong> the plate causes the velocity <strong>of</strong> the fluid to<br />

decrease to zero at the surface and influences its velocity throughout the<br />

thickness <strong>of</strong> a boundary layer. The thickness <strong>of</strong> the boundary layer depends<br />

on the free velocity, V∞, <strong>of</strong> the fluid. It is greater for free convection and<br />

smaller for forced convection. The rate <strong>of</strong> heat flow depends on the thickness<br />

<strong>of</strong> the convection layer, as well as the temperature difference between Ts and<br />

T∞ (Ts is the surface temperature, T∞ is the free field fluid temperature<br />

outside <strong>of</strong> the boundary layer.)<br />

Charlie Chong/ Fion Zhang


Newton’s cooling law defines the convective heat transfer coefficient:<br />

(h is expressed in BTU/hr-ft²-°F)<br />

rearranged:<br />

= ΔT∙h<br />

where:<br />

Rc = 1/h and is the resistance to convective heat flow<br />

Rc is also analogous to electrical resistance and is easier to use when<br />

determining combined conductive and convective heat transfer.<br />

Charlie Chong/ Fion Zhang


Figure A-2 Convective Heat Flow<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang


A.2.6 Radiation<br />

Radiative heat transfer is unlike the other two modes in several respects:<br />

1. It can take place in a vacuum.<br />

2. It occurs by electromagnetic emission and absorption.<br />

3. It occurs at the speed <strong>of</strong> light.<br />

4. The energy transferred is proportional to the fourth power <strong>of</strong> the<br />

temperature difference between the objects (ΔT 4 or T 4 ?) .<br />

The electromagnetic spectrum is illustrated in Figure A-3. Radiative heat<br />

transfer takes place in the infrared portion <strong>of</strong> the spectrum, between 0.75 µm<br />

and about 100 µm (0.1mm) , although most practical measurements can be<br />

made out to 20 µm. (µ or µm stands for micrometers or microns. A micron is<br />

one-millionth <strong>of</strong> a meter and is the measurement unit for radiant energy<br />

wavelength.) (radiative heat only take place at the aforementioned portion <strong>of</strong><br />

spectrum?)<br />

Charlie Chong/ Fion Zhang


Figure A-3 <strong>Infrared</strong> in the Electromagnetic Spectrum<br />

Charlie Chong/ Fion Zhang


A.2.7 Radiation Exchange at the Target Surface<br />

The measurement <strong>of</strong> thermal infrared radiation is the basis for non-contact<br />

temperature measurement and thermal imaging (or thermography). The<br />

process <strong>of</strong> thermal infrared radiation leaving a surface is called exitance or<br />

radiosity. It can be emitted from the surface, reflected <strong>of</strong>f <strong>of</strong> the surface, or<br />

transmitted through the surface. This is illustrated in Figure A-4. The total<br />

radiosity is equal to the sum <strong>of</strong> the emitted component (E), the reflected<br />

component (R), and the transmitted component (T). The surface temperature<br />

is related to E, the emitted component only.<br />

Charlie Chong/ Fion Zhang


Thermal infrared radiation impinging on a surface can be absorbed, reflected,<br />

or transmitted as illustrated in Figure A-5. Kirchh<strong>of</strong>f’s law states that the sum<br />

<strong>of</strong> the three components is always equal to the received radiation (the<br />

percentage sum <strong>of</strong> the three components equals unity):<br />

A (absorptivity) + R (reflectivity) + T (transmissivity) = 1<br />

(ε + ρ + τ = 1)<br />

When making practical measurements, the specularity or diffusivity <strong>of</strong> a target<br />

surface is taken into effect by accounting for the emissivity <strong>of</strong> the surface.<br />

Emissivity is discussed as part <strong>of</strong> the detailed discussion <strong>of</strong> the<br />

characteristics <strong>of</strong> infrared thermal radiation in section A.3.<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang


Figure A-4 Radiative Heat Flow<br />

Charlie Chong/ Fion Zhang


Figure A-4 Radiative Heat Flow<br />

W ε = σεT e<br />

4<br />

W ρ = σρT r<br />

4<br />

W τ = στT t<br />

4<br />

Charlie Chong/ Fion Zhang


Figure A-5 Radiation Exchange at the Target Surface<br />

Charlie Chong/ Fion Zhang


A.2.8 Specular and Diffuse Surfaces<br />

It should be noted that the roughness or structure <strong>of</strong> a surface will determine<br />

the type and direction <strong>of</strong> reflection <strong>of</strong> incident radiation. A smooth surface will<br />

reflect incident energy at an angle complementary to the angle <strong>of</strong> incidence.<br />

This is called a specular reflector. A rough or structured surface will scatter or<br />

disperse some <strong>of</strong> the incident radiation; this is a diffuse reflector.<br />

No perfectly specular or perfectly diffuse surface can exist in nature. All real<br />

surfaces have some diffusivity and some specularity.<br />

Charlie Chong/ Fion Zhang


Specular or Diffuse Surfaces<br />

Charlie Chong/ Fion Zhang


Specular or Diffuse Surfaces<br />

Diffuse<br />

Reflector<br />

Specular<br />

Reflector?<br />

Charlie Chong/ Fion Zhang<br />

http://www.hunantv.com/v/3/56616/f/750962.html?f=lb#


Specular and Diffuse Surfaces<br />

Diffuse<br />

Reflector<br />

Specular<br />

Reflector?<br />

Charlie Chong/ Fion Zhang


Specular and Diffuse Surfaces<br />

Confused<br />

Specular<br />

Reflector?<br />

Charlie Chong/ Fion Zhang


Specular or Diffuse Surfaces<br />

Charlie Chong/ Fion Zhang


Specular reflection is the mirror-like reflection <strong>of</strong> light (or <strong>of</strong> other kinds <strong>of</strong><br />

wave) from a surface, in which light from a single incoming direction (a ray) is<br />

reflected into a single outgoing direction. Such behavior is described by the<br />

law <strong>of</strong> reflection, which states that the direction <strong>of</strong> incoming light (the incident<br />

ray), and the direction <strong>of</strong> outgoing light reflected (the reflected ray) make the<br />

same angle with respect to the surface normal, thus the angle <strong>of</strong> incidence<br />

equals the angle <strong>of</strong> reflection θ 2 = θ 1 in the figure), and that the incident,<br />

normal, and reflected directions are coplanar.<br />

Charlie Chong/ Fion Zhang


Reflections <strong>of</strong>f Specular and Diffuse Surfaces<br />

Charlie Chong/ Fion Zhang


Reflections <strong>of</strong>f Specular and Diffuse Surfaces<br />

Charlie Chong/ Fion Zhang


A.2.9 Transient Heat Exchange The discussions <strong>of</strong> the three types <strong>of</strong> heat<br />

exchange in sections A.2.4, A.2.5, and A.2.6 deal with steady-state heat<br />

exchange for reasons <strong>of</strong> simplicity and easier understanding. Two fixed<br />

temperatures are assumed to exist at the two points between which the heat<br />

flows. In many applications, however, temperatures are in transition, so that<br />

the values shown for energy radiated from a target surface are the<br />

instantaneous values from the moment that measurements are made. There<br />

are numerous instances where existing transient thermal conditions are<br />

exploited in order to use thermography to reveal material or structural<br />

characteristics in test articles.<br />

Charlie Chong/ Fion Zhang


The thermogram <strong>of</strong> the outside surface <strong>of</strong> an insulated vessel carrying heated<br />

liquid, for example, should be relatively isothermal and somewhat warmer<br />

than the ambient air. Insulation voids or defects will cause warm anomalies to<br />

appear on the thermogram, allowing the thermographer to pinpoint areas <strong>of</strong><br />

defective or damaged insulation. Here a passive approach can be taken<br />

because the transient heat flow (or it is a steady state heat flow?) from the<br />

liquid through the insulation to the outside air produces the desired<br />

characteristic thermal pattern on the product surface. Similarly, water<br />

saturated areas on flat ro<strong>of</strong>s will retain solar heat well into the night; long after<br />

the dry sections have radiated their stored heat to the cold night sky, the<br />

saturated sections will continue to radiate and exhibit distinct anomalies to the<br />

thermographer. When there is no heat flow through the material or the test<br />

article to be evaluated, an active, or thermal injection, approach is used to<br />

generate a transient heat flow.<br />

Comment: In general steady state heat flow always lead to thermal<br />

equilibrium, for IRT, transient heat flows are exploited to reveal abnormalities.<br />

Charlie Chong/ Fion Zhang


This approach requires the generation <strong>of</strong> a controlled flow <strong>of</strong> thermal energy<br />

across the laminar structure <strong>of</strong> the sample material under test, thermography<br />

monitoring <strong>of</strong> one <strong>of</strong> the surfaces (or sometimes both) <strong>of</strong> the sample, and a<br />

search for anomalies in the thermal patterns that will indicate a defect in<br />

accordance with established accept-reject criteria. This approach has been<br />

used extensively and successfully by the aerospace community in the<br />

evaluation <strong>of</strong> composite structures for impurities, flaws, voids, disbonds,<br />

delaminations, and variations in structural integrity. Most recently, time-based<br />

heat injection methods have been applied successfully to measure the depth<br />

<strong>of</strong> voids, as well as their location. This is effective because thinner sections <strong>of</strong><br />

a given material will heat more rapidly than thicker sections.<br />

Charlie Chong/ Fion Zhang


Steady-state conduction<br />

Steady state conduction is the form <strong>of</strong> conduction that happens when the temperature<br />

differences ΔT driving the conduction are constant, so that (after an equilibration time), the<br />

spatial distribution <strong>of</strong> temperatures (temperature field) in the conducting object does not change<br />

any further. Thus, all partial derivatives <strong>of</strong> temperature with respect to space may either be zero<br />

or have nonzero values, but all derivatives <strong>of</strong> temperature at any point with respect to time are<br />

uniformly zero. In steady state conduction, the amount <strong>of</strong> heat entering any region <strong>of</strong> an object is<br />

equal to amount <strong>of</strong> heat coming out (if this were not so, the temperature would be rising or falling,<br />

as thermal energy was tapped or trapped in a region).<br />

For example, a bar may be cold at one end and hot at the other, but after a state <strong>of</strong> steady state<br />

conduction is reached, the spatial gradient <strong>of</strong> temperatures along the bar does not change any<br />

further, as time proceeds. Instead, the temperature at any given section <strong>of</strong> the rod remains<br />

constant, and this temperature varies linearly in space, along the direction <strong>of</strong> heat transfer.<br />

In steady state conduction, all the laws <strong>of</strong> direct current electrical conduction can be applied to<br />

"heat currents". In such cases, it is possible to take "thermal resistances" as the analog to<br />

electrical resistances. In such cases, temperature plays the role <strong>of</strong> voltage, and heat transferred<br />

per unit time (heat power) is the analog <strong>of</strong> electrical current. Steady state systems can be<br />

modelled by networks <strong>of</strong> such thermal resistances in series and in parallel, in exact analogy to<br />

electrical networks <strong>of</strong> resistors. See purely resistive thermal circuits for an example <strong>of</strong> such a<br />

network.<br />

Charlie Chong/ Fion Zhang<br />

https://en.wikipedia.org/wiki/Thermal_conduction


Transient conduction<br />

In general, during any period in which temperatures change in time at any place within an object,<br />

the mode <strong>of</strong> thermal energy flow is termed transient conduction. Another term is "non steadystate"<br />

conduction, referring to time-dependence <strong>of</strong> temperature fields in an object. Non-steadystate<br />

situations appear after an imposed change in temperature at a boundary <strong>of</strong> an object. They<br />

may also occur with temperature changes inside an object, as a result <strong>of</strong> a new source or sink <strong>of</strong><br />

heat suddenly introduced within an object, causing temperatures near the source or sink to<br />

change in time.<br />

When a new perturbation <strong>of</strong> temperature <strong>of</strong> this type happens, temperatures within the system<br />

change in time toward a new equilibrium with the new conditions, provided that these do not<br />

change. After equilibrium, heat flow into the system once again equals the heat flow out, and<br />

temperatures at each point inside the system no longer change. Once this happens, transient<br />

conduction is ended, although steady-state conduction may continue if heat flow continues. If<br />

changes in external temperatures or internal heat generation changes are too rapid for<br />

equilibrium <strong>of</strong> temperatures in space to take place, then the system never reaches a state <strong>of</strong><br />

unchanging temperature distribution in time, and the system remains in a transient state.<br />

An example <strong>of</strong> a new source <strong>of</strong> heat "turning on" within an object, causing transient conduction,<br />

is an engine starting in an automobile. In this case the transient thermal conduction phase for the<br />

entire machine is over, and the steady state phase appears, as soon as the engine reaches<br />

steady-state operating temperature. In this state <strong>of</strong> steady-state equilibrium, temperatures vary<br />

greatly from the engine cylinders to other parts <strong>of</strong> the automobile, but at no point in space within<br />

the automobile does temperature increase or decrease. After establishing this state, the transient<br />

conduction phase <strong>of</strong> heat transfer is over.<br />

Charlie Chong/ Fion Zhang<br />

https://en.wikipedia.org/wiki/Thermal_conduction


A.3 The Basic Physics <strong>of</strong> <strong>Infrared</strong> Radiation and Sensing<br />

All targets radiate energy in the infrared spectrum. The hotter the target, the<br />

more energy is radiated (∝T 4 ). Very hot targets radiate in the visible as well,<br />

and our eyes can see this because they are sensitive to light. The sun for<br />

example, at about 6000 K, appears to glow white-hot; a tungsten filament, at<br />

about 3000 K, has a yellowish glow, and an electric stove element, at 800 K,<br />

glows red. As the stove element cools, it loses its visible glow but it continues<br />

to radiate. We can feel it with a hand placed near the surface but we can’t see<br />

the glow because the energy has shifted from red to infrared. <strong>Infrared</strong><br />

detectors can sense infrared radiant energy and produce useful electrical<br />

signals proportional to the temperature <strong>of</strong> target surfaces. Instruments that<br />

use infrared detectors and optics to gather and focus energy from the targets<br />

onto these detectors are capable <strong>of</strong> measuring target surface temperatures<br />

with sensitivities better than 0.1°C, and with response times as fast as<br />

microseconds. Instruments that combine this measurement capability with<br />

capabilities for scanning the target surface are called infrared thermal imagers.<br />

Charlie Chong/ Fion Zhang


They can produce thermal maps or thermograms where the brightness<br />

intensity or color hue <strong>of</strong> any spot on the map is representative <strong>of</strong> the<br />

temperature <strong>of</strong> the surface at that point. In most cases, thermal imagers can<br />

be considered as extensions <strong>of</strong> radiation thermometers or as a radiation<br />

thermometer with scanning capability. The performance parameters <strong>of</strong><br />

thermal imagers are extensions <strong>of</strong> the performance parameters <strong>of</strong> radiation<br />

thermometers.<br />

Charlie Chong/ Fion Zhang


A.3.1 Some Historical Background<br />

The color <strong>of</strong> a glowing metal is a fair indication <strong>of</strong> its temperature (the higher<br />

the temperature, the whiter the color). The ancient sword-maker and<br />

blacksmith knew from the color <strong>of</strong> a heated part when it was time to quench<br />

and temper. This technique is still in use today; precision optical matching<br />

pyrometers are used to match the brightness in color <strong>of</strong> a product with that <strong>of</strong><br />

a glowing filament. The brightness <strong>of</strong> the filament is controlled by adjusting a<br />

knob that is calibrated in temperature. The next logical step is to substitute a<br />

photomultiplier for the operator’s eye and, thus, calibrate the measurement.<br />

Finally, a differential measurement is made between what the brightness <strong>of</strong><br />

the product is and what it should be (the set point), and the differential signal<br />

is injected into the process and used to drive the product temperature to the<br />

set point. With the advent <strong>of</strong> modern infrared detectors, the precision<br />

measurement <strong>of</strong> thermal energy radiating from surfaces that do not glow<br />

became possible. Measurements <strong>of</strong> cool surfaces, well below 0°C, are<br />

accomplished routinely with even the least expensive <strong>of</strong> infrared sensors.<br />

Charlie Chong/ Fion Zhang


A.3.2 Non-Contact Thermal Measurements<br />

<strong>Infrared</strong> non-contact thermal sensing instruments are classified as infrared<br />

radiation thermometers by the American Society <strong>of</strong> Testing and Materials<br />

(ASTM), even though they don’t always read out in temperatures. The laws <strong>of</strong><br />

physics allow for the conversion <strong>of</strong> infrared radiation measurements to<br />

temperature measurements. This is done by first measuring the self-emitted<br />

radiation in the infrared portion <strong>of</strong> the electromagnetic spectrum <strong>of</strong> target<br />

surfaces, and then converting these measurements to electrical signals. In<br />

making these measurements, three sets <strong>of</strong> characteristics need to be<br />

considered:<br />

• The target surface<br />

• The transmitting medium between the target and the instrument<br />

• The measuring instrument<br />

Charlie Chong/ Fion Zhang


A.3.3 The Target Surface<br />

The chart <strong>of</strong> the electromagnetic spectrum (Figure A-3) indicates that the<br />

infrared portion <strong>of</strong> the spectrum lies adjacent to the visible. Every target<br />

surface above absolute zero (0 Kelvins or -273° Centigrade) radiates energy<br />

in the infrared. The hotter the target, the more radiant energy is emitted.<br />

When targets are hot enough, they radiate or glow in the visible part <strong>of</strong> the<br />

spectrum as well ( and beyond that, again becoming invisible again, example<br />

UV & ɣ ray) . As they cool, the eye becomes no longer able to see the emitted<br />

radiation and the targets appear to not glow at all. <strong>Infrared</strong> sensors are<br />

employed here to measure the radiation in the infrared, which can be related<br />

to target surface temperature. The visible spectrum extends from energy<br />

wavelengths <strong>of</strong> 0.4 µm for violet light to about 0.75 µm for red light. (µ or µm<br />

stands for micrometers or microns. A micron is one-millionth <strong>of</strong> a meter and is<br />

the measurement unit for radiant energy wavelength.) For practical purposes<br />

<strong>of</strong> temperature measurement, the infrared spectrum extends from 0.75 µm to<br />

about 20 µm.<br />

Charlie Chong/ Fion Zhang


The visible spectrum extends from<br />

energy wavelengths <strong>of</strong><br />

0.4 µm for violet light to about 0.75<br />

µm for red light. For practical purposes <strong>of</strong><br />

temperature measurement, the infrared spectrum<br />

extends from 0.75 µm to about 20 µm.<br />

for my ASNT Exam<br />

Charlie Chong/ Fion Zhang


Figure A-6 shows the distribution <strong>of</strong> emitted energy over the electromagnetic<br />

spectrum <strong>of</strong> targets at various temperatures. The sun, at 6000 K, appears<br />

white hot because its emitted energy is centered over the visible spectrum<br />

with a peak at 0.5 µm. Other targets, such as a tungsten filament at 3000 K, a<br />

red-hot surface at 800 K, and the ambient earth at 300 K (about 30°C), are<br />

also shown in this illustration. It becomes apparent that, as surfaces cool, not<br />

only do they emit less energy, but the wavelength distribution shifts to longer<br />

infrared wavelengths. Even though the eye becomes no longer capable <strong>of</strong><br />

sensing this energy, infrared sensors can detect these invisible longer<br />

wavelengths. They enable us to measure the self-emitted radiant energy from<br />

even very cold targets and, thereby, determine the temperatures <strong>of</strong> target<br />

surfaces remotely and without contact.<br />

Keypoints:<br />

The visible spectrum extends from energy wavelengths <strong>of</strong> 0.4 µm for violet<br />

light to about 0.75 µm for red light.<br />

For practical purposes <strong>of</strong> temperature measurement, the infrared spectrum<br />

extends from 0.75 µm to about 20 µm.<br />

Charlie Chong/ Fion Zhang


Figure A-6<br />

Blackbody Curves<br />

at Various<br />

Temperatures<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang


λ m = b/T = (2897/T μm)<br />

Charlie Chong/ Fion Zhang<br />

http://www.nasa.gov/centers/goddard/news/topstory/2004/0107filament.html


Two physical laws define the radiant behavior illustrated in Figure A-6:<br />

The Stephan-Boltzmann Law (1):<br />

W = εδT 4<br />

and Wien’s Displacement Law (2):<br />

λ m = b/T = (2897/T μm)<br />

Where:<br />

W = Radiant flux emitted per unit are a (watts/cm²)<br />

ε = Emissivity (unity for a blackbody target)<br />

δ = Stephan-Boltzmann constant = 5.673 x10 -12 watts cm -2<br />

T = Absolute temperature <strong>of</strong> target (K)<br />

λ m = Wavelength <strong>of</strong> maximum radiation (µm)<br />

b = Wien’s displacement constant = 2897 (µm∙K)<br />

Charlie Chong/ Fion Zhang


According to (1), the radiant energy emitted from the target surface (W)<br />

equals two constants multiplied by the fourth power <strong>of</strong> the absolute<br />

temperature (T 4 ) <strong>of</strong> the target. The instrument measures W and calculates T.<br />

One <strong>of</strong> the two constants, δ, is a fixed number.<br />

Emissivity (ξ) is the other constant and is a surface characteristic that is only<br />

constant for a given material over a given range <strong>of</strong> temperatures.<br />

For point measurements, one can usually estimate the emissivity setting<br />

needed to dial into the instrument from available tables and charts. One can<br />

also learn, experimentally, the proper setting needed to make the instrument<br />

produce the correct temperature reading by using samples <strong>of</strong> the actual target<br />

material. This more practical setting value is called effective emissivity (e*).<br />

Charlie Chong/ Fion Zhang


According to (2), the wavelength at which a target radiates its peak energy is<br />

defined as simply a constant (b = 2897≈ 3000) divided by the target<br />

temperature (T) in Kelvins. For the 300 K ambient earth, for example, the<br />

peak wavelength would be (λ max = 2897/300) or ≈ 10 µm. This quick<br />

calculation is important in selecting the proper instrument for a measurement<br />

task, as will be discussed in section A.4.<br />

Target surfaces can be classified in three categories: (1) black bodies, (2)<br />

gray bodies, and (3) non-gray bodies.<br />

The targets shown in Figure A-6 are all blackbody radiators (or black bodies).<br />

A blackbody radiator is a theoretical surface having unity emissivity at all<br />

wavelengths and absorbing all <strong>of</strong> the energy available at its surface. This<br />

would be an ideal target to measure because the temperature calculation<br />

within the instrument would be simply mechanized and always constant.<br />

Fortunately, although blackbody radiators do not exist in practice, the<br />

surfaces <strong>of</strong> most solids are gray bodies, that is, surfaces whose emissivities<br />

are high and fairly constant with wavelength.<br />

Charlie Chong/ Fion Zhang


Figure A-7 shows the comparative spectral distribution <strong>of</strong> energy emitted by a<br />

blackbody, a gray body, and a non-gray body (also called a spectral body), all<br />

at the same temperature. For gray body measurements, a simple emissivity<br />

correction can usually be dialed in when absolute measurements are required.<br />

For non-gray bodies, the solutions are more difficult. To understand the<br />

reason for this, it is necessary to see what an instrument sees when it is<br />

aimed at a non-gray target surface.<br />

Keywords:<br />

non-gray body (also called a spectral body)<br />

Charlie Chong/ Fion Zhang


Figure A-7 Spectral Distribution <strong>of</strong> a Blackbody, a Gray Body, and a Non-<br />

Gray Body<br />

Charlie Chong/ Fion Zhang


Figure A-7 Spectral Distribution <strong>of</strong> a Blackbody, a Gray Body, and a Non-<br />

Gray Body<br />

Charlie Chong/ Fion Zhang


Figure A-8 shows that the instrument sees three components <strong>of</strong> energy: first,<br />

emitted energy (ε); second, reflected energy from the environment (ρ); and<br />

third, energy transmitted through the target from sources behind the target (τ).<br />

The percentage sum <strong>of</strong> these components is always unity (1). The instrument<br />

sees only ε, the emitted energy, when aimed at a blackbody target because a<br />

blackbody reflects and transmits nothing. For a gray body, the instrument<br />

sees ε and ρ, the emitted and reflected energy. The instrument sees all three<br />

components when aimed at a nongray body because a non-gray body is<br />

partially transparent.<br />

Keywords:<br />

because a non-gray body is partially transparent.(?)<br />

Charlie Chong/ Fion Zhang


Figure A-8 Components <strong>of</strong> Energy Reaching the Measuring Instrument<br />

Charlie Chong/ Fion Zhang


If the emissivity <strong>of</strong> a gray body is very low, as in the case <strong>of</strong> polished metal<br />

surfaces, the reflectance becomes high (reflectance = 1 - emissivity) and can<br />

generate erroneous readings if not properly handled. Reflected energy from a<br />

specific source can generally be redirected by proper orientation <strong>of</strong> the<br />

instrument with respect to the target surface, as shown in Figure A-9. This<br />

illustrates the proper and improper orientation that is necessary to avoid<br />

reflected energy from a specific source.<br />

Charlie Chong/ Fion Zhang


Figure A-9 Aiming the Instrument to Avoid Point Source Reflections<br />

Charlie Chong/ Fion Zhang


Under certain conditions, an error in temperature indication can occur as the<br />

result <strong>of</strong> a high temperature background, such as a boiler wall (behind the<br />

instrument), reflecting <strong>of</strong>f <strong>of</strong> a reflective target surface and contributing to the<br />

apparent temperature <strong>of</strong> the target. Most instrument manufacturers provide a<br />

background temperature correction to compensate for this condition. Often, in<br />

practice, the troublesome component is T, the energy transmitted through a<br />

non-gray target from sources behind the target. A discussion <strong>of</strong> solutions to<br />

this type <strong>of</strong> problem is included in section A.4.<br />

Non-Gray body – An object whose emissivity varies with wavelength over the<br />

wavelength interval <strong>of</strong> interest. A radiating object that does not have a<br />

spectral radiation distribution similar to a blackbody; also called a “colored<br />

body” or “realbody”. Glass and plastic films are examples <strong>of</strong> non-graybodies.<br />

An object can be a graybody over one wavelength interval and a non-gray<br />

body over another. http://www.infraredtraininginstitute.com/thermography-terms-definitions/<br />

Charlie Chong/ Fion Zhang


Blackbody, Graybody & Non-graybody (colored body or real body)<br />

Charlie Chong/ Fion Zhang<br />

http://www.moistureview.com/resources/infrarods-blog/page/4


EXAM score!<br />

Non-graybody<br />

(colored body or real body)<br />

for my ASNT exam<br />

for my ASNT exam<br />

Charlie Chong/ Fion Zhang


A.3.4 The Transmitting Medium<br />

The transmission characteristics <strong>of</strong> the medium in the measurement path<br />

between the target and the instrument need to be considered in making nonontact<br />

thermal measurements. No loss <strong>of</strong> energy is encountered when<br />

measuring through a vacuum. For short path lengths, a few feet for example,<br />

most gases including the atmosphere, absorb very little energy and can be<br />

ignored (except where measurements <strong>of</strong> precision temperature values are<br />

required). As the path length increases to hundreds <strong>of</strong> feet, or as the air<br />

becomes heavy with water vapor, the absorption might become a factor. It is<br />

then necessary to consider the infrared transmission characteristics <strong>of</strong> the<br />

atmosphere.<br />

Charlie Chong/ Fion Zhang


Figure A-10 illustrates the spectral transmission characteristics <strong>of</strong> 0.3 km <strong>of</strong><br />

ground level atmosphere (what is the object to detector distance in tabulating<br />

the chart? or this is not a factor as the transmittance is given as a ratio (%)<br />

with respect to transmittance in vacuum (Transmittance in vacuum=100%)).<br />

Two spectral intervals can be seen to have very high transmission. These are<br />

known as the 1.5 µm and the 8.14 µm atmospheric windows, and almost all<br />

infrared sensing and scanning instruments are designed to operate in one or<br />

the other <strong>of</strong> these windows. (unless) Usually, the difficulties encountered with<br />

transmitting media occur when the target is viewed by the instrument through<br />

another solid object such as a glass or quartz viewing port in a process.<br />

Keywords:<br />

These are known as the 1.5 µm and the 8.14 µm atmospheric windows.<br />

Charlie Chong/ Fion Zhang


Figure A-10 <strong>Infrared</strong> Transmission <strong>of</strong> 0.3 km <strong>of</strong> Sea Level Atmosphere<br />

Charlie Chong/ Fion Zhang


Figure A-10 <strong>Infrared</strong> Transmission <strong>of</strong> 0.3 km <strong>of</strong> Sea Level Atmosphere<br />

Charlie Chong/ Fion Zhang


Figure A-11 shows transmission curves for various samples <strong>of</strong> glass and<br />

quartz. Upon seeing these, our first impression is that glass is opaque at 10<br />

µm where ambient (30°C) surfaces radiate their peak energy. This impression<br />

is correct and, although in theory, infrared measurements can be made <strong>of</strong><br />

30°C targets through glass, it is hardly practical. The first approach to the<br />

problem is to attempt to eliminate the glass, or at least a portion <strong>of</strong> it, through<br />

which the instrument can be aimed at the target. If, for reasons <strong>of</strong> hazard,<br />

vacuum, or product safety, a window must be present; a material that<br />

transmits in the longer wavelengths might be substituted.<br />

Charlie Chong/ Fion Zhang


Figure A-11 <strong>Infrared</strong> Spectral Transmission <strong>of</strong> Glass<br />

Charlie Chong/ Fion Zhang


Figure A-11 <strong>Infrared</strong> Spectral Transmission <strong>of</strong> Glass<br />

Charlie Chong/ Fion Zhang


Figure A-11 shows transmission curves for various samples <strong>of</strong> glass and<br />

quartz. Upon seeing these, our first impression is that glass is opaque at 10<br />

µm where ambient (30°C) surfaces radiate their peak energy (?). This<br />

impression is correct and, although in theory, infrared measurements can be<br />

made <strong>of</strong> 30°C targets through glass, it is hardly practical. The first approach<br />

to the problem is to attempt to eliminate the glass, or at least a portion <strong>of</strong> it,<br />

through which the instrument can be aimed at the target. If, for reasons <strong>of</strong><br />

hazard, vacuum, or product safety, a window must be present; a material that<br />

transmits in the longer wavelengths might be substituted.<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang<br />

http://www.technicalglass.com/fused_quartz_transmission.html


EXAM score!<br />

Glass is opaque<br />

λ > 5µm at 30ºC?<br />

for my ASNT exam<br />

Charlie Chong/ Fion Zhang


Figure A-12 shows the spectral transmission characteristics <strong>of</strong> several <strong>of</strong><br />

these materials, many <strong>of</strong> which transmit energy past 10 µm. These materials<br />

are <strong>of</strong>ten used as lenses and optical elements in low-temperature infrared<br />

sensors. Of course, as targets become hotter and the emitted energy shifts to<br />

the shorter wavelengths, glass and quartz windows pose less <strong>of</strong> a problem<br />

and are even used as elements and lenses in high-temperature sensing<br />

instruments.<br />

Charlie Chong/ Fion Zhang


Figure A-12 Characteristics <strong>of</strong> IR Transmitting Materials<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang


The characteristics <strong>of</strong> the window material will always have some effect on<br />

the temperature measurement, but the attenuation can always be corrected<br />

by pre-calibrating the instrument with a sample window placed between the<br />

instrument and a target <strong>of</strong> known temperature. In closing the discussion <strong>of</strong> the<br />

transmitting medium, it is important to note that infrared sensors can only<br />

work when all <strong>of</strong> the following spectral ranges coincide or overlap:<br />

1. The spectral range over which the target emits<br />

2. The spectral range over which the medium transmits<br />

3. The spectral range over which the instrument operates<br />

3<br />

2<br />

1<br />

Charlie Chong/ Fion Zhang


IR Lenses – Sapphire Lens<br />

Charlie Chong/ Fion Zhang<br />

http://www.ecvv.com/product/3411419.html


IR Lenses – LWIR Len<br />

Charlie Chong/ Fion Zhang<br />

http://eom.umicore.com/en/infrared-optics/product-range/25-mm-f-1.2/


IR Lenses – Fresnel Len<br />

Charlie Chong/ Fion Zhang<br />

http://www.glolab.com/pirparts/pirparts.html


IR Lenses – Fresnel Len<br />

Charlie Chong/ Fion Zhang<br />

http://www.glolab.com/pirparts/pirparts.html


A.3.5 The Measuring Instrument<br />

Figure A-13 shows the necessary components <strong>of</strong> an infrared radiation<br />

thermometer. Collecting optics (an infrared lens, for example) is necessary in<br />

order to focus the energy emitted by the target onto the sensitive surface <strong>of</strong><br />

an infrared detector, which, in turn, converts this energy into an electrical<br />

signal.<br />

Charlie Chong/ Fion Zhang


Figure A-13 Components <strong>of</strong> an <strong>Infrared</strong> Radiation Thermometer<br />

Thermal or photon<br />

detector, single<br />

element or FPA.<br />

Charlie Chong/ Fion Zhang


When an infrared radiation thermometer (point-sensing instrument) is aimed<br />

at a target, it collects energy within a collecting beam, the shape <strong>of</strong> which is<br />

determined by the configuration <strong>of</strong> the optics and the detector.<br />

The cross- section <strong>of</strong> this collecting beam is called the field <strong>of</strong> view <strong>of</strong> the<br />

instrument, and it determines the size <strong>of</strong> the area (spot size) on the target<br />

surface that is measured by the instrument.<br />

On thermal imaging instruments, this is called the instantaneous field <strong>of</strong> view<br />

(IFOV) and becomes one picture element on the thermogram.<br />

Comment:<br />

for single element detector; FOV = IFOV<br />

for FPA multi element detector; IFOV (D) = θ(rad) x d<br />

Where, d= focal to object distance<br />

Charlie Chong/ Fion Zhang


<strong>Infrared</strong> optics are available in two general configurations, refractive and<br />

reflective;<br />

■<br />

Refractive optics (lenses), which are at least partly transparent to the<br />

wavelengths <strong>of</strong> interest, are used most <strong>of</strong>ten for high- temperature<br />

applications where their throughput losses can be ignored.<br />

■<br />

Reflective optics (mirrors), which are more efficient but somewhat<br />

complicate the optical path, are used more <strong>of</strong>ten for low-temperature<br />

applications, where the energy levels cannot warrant throughput energy<br />

losses.<br />

An infrared interference filter is <strong>of</strong>ten placed in front <strong>of</strong> the detector to limit the<br />

spectral region or band <strong>of</strong> the energy reaching the detector. The reasons for<br />

spectral selectivity will be discussed later in this section.<br />

Charlie Chong/ Fion Zhang


The processing electronics unit amplifies and conditions the signal from the<br />

infrared detector and introduces corrections for such factors as (1) detector<br />

ambient temperature drift and (2) target surface emissivity. Generally, a meter<br />

indicates the target temperature and an analog output is provided. The analog<br />

signal is used to record, display, alarm, control, correct, or any combination <strong>of</strong><br />

these.<br />

Figure A-14 illustrates the configuration <strong>of</strong> a typical instrument employing all<br />

<strong>of</strong> the elements outlined. The germanium lens collects the energy from a spot<br />

on the target surface and focuses it on the surface <strong>of</strong> the radiation thermopile<br />

detector. The 8.14 µm filter limits the spectral band <strong>of</strong> the energy reaching the<br />

detector so that it falls within the atmospheric window. The detector generates<br />

a dc emf proportional to the energy emitted by the target surface. The autozero<br />

amplifier senses ambient temperature changes and prevents ambient<br />

drift errors. The output electronics unit conditions the signal and computes the<br />

target surface temperature based on a manual emissivity setting. The analog<br />

output terminals accept a 15 - 30 VDC loop supply and generate a 4 - 20<br />

milliampere signal, proportional to target surface temperature.<br />

Charlie Chong/ Fion Zhang


All infrared detector-transducers exhibit some electrical change in response<br />

to the radiant energy impinging on their sensitive surfaces. Depending on the<br />

type <strong>of</strong> detector this can be (1) an impedance change, (2) a capacitance<br />

change, (3) the generation <strong>of</strong> an emf (voltage), or (4) the release <strong>of</strong> photons.<br />

Detectors are available with response times as fast as nanoseconds or as<br />

slow as fractions <strong>of</strong> seconds. Depending on the requirement, either a<br />

broadband detector or a spectrally limited detector can be selected.<br />

Keywords:<br />

Depending on the type <strong>of</strong> detector this can be<br />

(1) an impedance change, (Z) (thermal detector?)<br />

(2) a capacitance change, (C) (thermal detector?)<br />

(3) the generation <strong>of</strong> an emf (voltage), (Emf) (thermal detector?)<br />

(4) the release <strong>of</strong> photons. (E=hѵ) (photon detector?)<br />

Charlie Chong/ Fion Zhang


Sequences <strong>of</strong> Events:<br />

1. The germanium lens collects the energy from a spot on the target surface<br />

2. focuses it on the surface <strong>of</strong> the radiation thermopile detector.<br />

3. The 8.14 µm filter (pass) limits the spectral band <strong>of</strong> the energy reaching<br />

the detector so that it falls within the atmospheric window.<br />

4. The detector generates a dc emf proportional to the energy emitted by the<br />

target surface. (thermal detector)<br />

5. The auto-zero amplifier senses ambient temperature changes and<br />

prevents ambient drift errors. (electronic)<br />

6. The output electronics unit conditions the signal and computes the target<br />

surface temperature based on a manual emissivity setting. (W = εσT 4 )<br />

7. The analog output terminals accept a 15 - 30 Volt, DC loop supply and<br />

generate a 4 - 20 milliampere signal, proportional to target surface<br />

temperature.<br />

Charlie Chong/ Fion Zhang


Figure A-14 Typical <strong>Infrared</strong> Radiation Thermometer Schematic<br />

Charlie Chong/ Fion Zhang


Germanium Len<br />

Charlie Chong/ Fion Zhang


Germanium Len<br />

Charlie Chong/ Fion Zhang


Germanium Len<br />

Charlie Chong/ Fion Zhang


Thermopile Detector<br />

Charlie Chong/ Fion Zhang


Thermopile Detector<br />

Charlie Chong/ Fion Zhang


Thermopile Detector<br />

Charlie Chong/ Fion Zhang<br />

http://wanda.fiu.edu/teaching/courses/Modern_lab_manual/stefan_boltzmann.html


Thermopile Detector<br />

Charlie Chong/ Fion Zhang<br />

https://www.adafruit.com/products/2023


Thermopile Detector<br />

Charlie Chong/ Fion Zhang


Thermopile Detector<br />

Charlie Chong/ Fion Zhang<br />

https://www.adafruit.com/products/2023


Thermopile Detector<br />

Charlie Chong/ Fion Zhang<br />

https://www.adafruit.com/products/2023


Thermopile Detector<br />

The Grid-EYE 64-thermopile infrared array sensor from Panasonic adds state-<strong>of</strong>-the-art sensing technology to Avnet Abacus'<br />

passives portfolio. Based on Panasonic’s advanced MEMS technology, the 8x8 grid format infrared array sensor combines a builtin<br />

thermistor and an integrated circuit for temperature sensing in a small SMT package measuring only 11.6x4.3x8.0mm. Grid-<br />

EYE enables contactless temperature detection over the entire specified area. It can use passive infrared detection to determine<br />

temperature differentiation allowing it to detect multiple objects simultaneously. It is able to measure actual temperature and<br />

temperature gradients, providing thermal images and identifying the direction <strong>of</strong> movement <strong>of</strong> people or objects. The device’s 64<br />

pixel range yields accurate temperature sensing, within the range <strong>of</strong> -20°C to 100°C, over a viewing angle <strong>of</strong> 60° provided by a<br />

silicon lens. It uses an external I²C communication interface, enabling temperature measurement at speeds <strong>of</strong> 1 or 10 frames/s.<br />

An interrupt function is also available. The operating voltage <strong>of</strong> the device is 3.3 or 5.0V.<br />

Charlie Chong/ Fion Zhang<br />

http://www.electronics-eetimes.com/en/64-thermopile-infrared-array-sensor-available-fromavnet-abacus.html?cmp_id=7&news_id=222915463


Thermopile Detector<br />

Charlie Chong/ Fion Zhang<br />

http://www.electronics-eetimes.com/en/64-thermopile-infrared-array-sensor-available-fromavnet-abacus.html?cmp_id=7&news_id=222915463


Thermopile Detector - DR46 Thermopile Detector<br />

Features- A two-channel or a one-channel compensated thin-film thermopile in a TO-8 package. Each active area is 4mm x<br />

0.6mm. Offers high output with excellent signal-to-noise ratio. An internal aperture minimizes channel-to-channel crosstalk<br />

increasing sensitivity. Applications: Gas analysis, non-contact temperature measurement, fire detection / suppression.<br />

Charlie Chong/ Fion Zhang<br />

http://www.dexterresearch.com/?module=Page&sID=dr46


The IR Detectors<br />

<strong>Infrared</strong> detectors fall into two broad categories:<br />

■<br />

■<br />

thermal detectors, which have broad, uniform spectral responses,<br />

somewhat lower sensitivities, and slower response times (on the order <strong>of</strong><br />

milliseconds), and<br />

photodetectors, (or photon detectors), which have limited spectral<br />

responses, higher peak sensitivities, and faster response times (on the<br />

order <strong>of</strong> microseconds).<br />

Thermal detectors will generally operate at or near room temperature, while<br />

photodetectors are generally cooled to optimize performance. The mercury-<br />

Cadmium-telluride (HgCdTe) detector, for example, is a photodetector cooled<br />

to 77 K for 8.14 µm operation and to 195 K for 3.5 µm operation. Because <strong>of</strong><br />

its fast response, this detector is used extensively in high-speed scanning<br />

and imaging applications.<br />

Charlie Chong/ Fion Zhang


The radiation thermopile, on the other hand, is a broadband thermal detector<br />

operating uncooled. It is used extensively for spot measurements <strong>of</strong> cool<br />

targets. It generates a dc emf proportional to the radiant energy reaching its<br />

surface and is ideal for use in portable, battery powered instruments. Figure<br />

A-15 illustrates the spectral responses <strong>of</strong> various infrared detectors.<br />

Charlie Chong/ Fion Zhang


Figure A-15 Spectral Sensitivity <strong>of</strong> Various <strong>Infrared</strong> Detectors<br />

Charlie Chong/ Fion Zhang


Thermal Detectors & Photon Detectors<br />

Photon<br />

Detector<br />

Thermal<br />

Detector<br />

Charlie Chong/ Fion Zhang


The Mercury- Cadmium-telluride (Hgcdte) Detector,<br />

Charlie Chong/ Fion Zhang


Discussion<br />

Subject: Why there are many curves for HgCdTe.<br />

Charlie Chong/ Fion Zhang<br />

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


The Mercury- Cadmium-Telluride (HgCdTe) Detector – FPA<br />

WISE Mercury Cadmium Telluride Focal Plane Mount Assembly (HgCdTe FPMA). This picture<br />

shows one <strong>of</strong> the four WISE detectors. The sensitive area shows as green and contains 1 million<br />

pixel elements.<br />

Charlie Chong/ Fion Zhang<br />

http://wise.ssl.berkeley.edu/gallery_detector.html


The Mercury- Cadmium-Telluride (HgCdTe) Detector – FPA<br />

Charlie Chong/ Fion Zhang<br />

http://spie.org/x91246.xml


The Mercury- Cadmium-Telluride (HgCdTe) Detector – FPA<br />

October 24, 2011 - All Eyes on Oldest Recorded Supernova<br />

This image combines data from four different space telescopes to create a multi-wavelength view <strong>of</strong> all that remains <strong>of</strong> the oldest documented example <strong>of</strong> a<br />

supernova, called RCW 86. The Chinese witnessed the event in 185 A.D., documenting a mysterious "guest star" that remained in the sky for eight months.<br />

X-ray images from the European Space Agency's XMM-Newton Observatory and NASA's Chandra X-ray Observatory are combined to form the blue and<br />

green colors in the image. The X-rays show the interstellar gas that has been heated to millions <strong>of</strong> degrees by the passage <strong>of</strong> the shock wave from the<br />

supernova.<br />

Charlie Chong/ Fion Zhang<br />

http://wise.ssl.berkeley.edu/gallery_detector.html


Discussion<br />

Subject: Why it wasn’t pixel-like correspond to the spatial resolution <strong>of</strong> 10 6 ?<br />

Charlie Chong/ Fion Zhang<br />

http://wise.ssl.berkeley.edu/gallery_detector.html


The Mercury- Cadmium-Telluride (HgCdTe) Detector – FPA<br />

Sept 29, 2011 - Portrait <strong>of</strong> Two Asteroids in Different Light - This animation illustrates the benefits <strong>of</strong> observing asteroids in infrared light. It begins by<br />

showing two artistic interpretations <strong>of</strong> asteroids up close. They are about the same size but the one on the right is darker. The animation zooms away to<br />

show how a visible-light telescope would see these two space rocks, located at the same distance millions <strong>of</strong> miles away from Earth, against a background<br />

<strong>of</strong> more distant stars. The one on the left would be much easier to see because it reflects more visible light from the sun. The animation then transitions to<br />

an infrared view <strong>of</strong> the same two objects. Both asteroids are equally as bright because the telescope is picking up infrared light coming from the bodies<br />

themselves, as a result <strong>of</strong> being heated by the sun. The measurements are not strongly affected by how light or dark an asteroid is, a property called albedo.<br />

Instead, the brightness is more directly related to an asteroid's size. Therefore, infrared telescopes like WISE are better at both finding the small, dark<br />

asteroids and determining asteroid sizes.<br />

Charlie Chong/ Fion Zhang<br />

http://wise.ssl.berkeley.edu/gallery_detector.html


Sept 29, 2011 - Portrait <strong>of</strong> Two Asteroids in Different Light - This animation<br />

illustrates the benefits <strong>of</strong> observing asteroids in infrared light.<br />

■<br />

http://wise.ssl.berkeley.edu/video/quicktime/V2-TwoAsteroids-HD.mov<br />

Charlie Chong/ Fion Zhang<br />

http://wise.ssl.berkeley.edu/gallery_detector.html


Point-sensing instruments for measuring very hot targets, usually operate in<br />

shorter wavelengths (0.9 - 1.1 µm, for example), and instruments for<br />

measuring cooler targets usually operate in longer wavelengths (3.5 µm or<br />

8.14 µm, for example). Most infrared thermal imagers operate in either the 3.5<br />

µm or 8.14 µm spectral region.<br />

The spectral Selectivity:<br />

■ Very hot - 0.9 - 1.1 µm<br />

■ Hot - 3.5 µm<br />

■ Cool - 8.14 µm<br />

Charlie Chong/ Fion Zhang


A.3.6 Introduction to Thermal Scanning and Imaging Instruments<br />

When problems in temperature monitoring and control cannot be solved by<br />

the measurement <strong>of</strong> one or several discrete points on a target surface, it<br />

becomes necessary to spatially scan (that is, to move the collecting beam<br />

(field <strong>of</strong> view) <strong>of</strong> the instrument relative to the target). This can be<br />

accomplished by:<br />

(1) inserting a movable optical element into the collecting beam, or<br />

(2) by employing a multi-detector array or mosaic, and scanning the array<br />

electronically. (line scanner & FPA)<br />

A brief overview <strong>of</strong> scanning and imaging instruments follows. A more<br />

detailed overview can be found in section 2.<br />

Charlie Chong/ Fion Zhang


A.3.6.1 Line Scanning<br />

The purpose <strong>of</strong> spatial scanning is to derive information concerning the<br />

distribution <strong>of</strong> radiant energy over a target scene. Quite <strong>of</strong>ten, a single straight<br />

line scanned on the target is all that is necessary to locate a critical thermal<br />

anomaly. The instantaneous position <strong>of</strong> the scanning element (or the position<br />

<strong>of</strong> the element in the linear array) is controlled or sensed, so that the<br />

radiometric output signal can be accompanied by a position signal output and<br />

be displayed on a chart recorder, an oscilloscope, or some other recording<br />

device.<br />

A typical high-speed commercial line scanner develops a high-resolution<br />

thermal map by scanning normal to the motion <strong>of</strong> a moving target, such as a<br />

paper web or a strip steel process. The resulting output is a thermal strip map<br />

<strong>of</strong> the process as it moves normal to the scan line (as illustrated in Figure A-<br />

16). The output signal information is in real-time computer compatible format<br />

and can be used to monitor, control or predict the behavior <strong>of</strong> the target.<br />

Charlie Chong/ Fion Zhang


Figure A-16 Scanning Configuration <strong>of</strong> an <strong>Infrared</strong> Line Scanner<br />

The line scanner<br />

could be a single<br />

element or linear<br />

array detector.<br />

Charlie Chong/ Fion Zhang


A.3.6.2 Two-Dimensional Scanning<br />

The purpose <strong>of</strong> spatial scanning is to derive information concerning the<br />

distribution <strong>of</strong> infrared radiant energy over a target scene. Scanning can be<br />

accomplished either opto-mechanically or electronically.<br />

Opto-mechanical scanning can be done by moving the target with the<br />

instrument fixed, or by moving (translating or panning) the instrument, but it is<br />

more practically accomplished by inserting movable optical elements into the<br />

collected beam. Although an almost infinite variety <strong>of</strong> scanning patterns can<br />

be generated using two moving elements, the most common pattern is<br />

rectilinear, and this is most <strong>of</strong>ten accomplished by two elements, each<br />

scanning a line normal to the other. A typical rectilinear scanner employs two<br />

rotating prisms behind the primary lens system (refractive scanning). An<br />

alternate configuration uses two oscillating mirrors behind the primary lens<br />

(reflective scanning). This is also commonly used in commercial scanners, as<br />

are combinations <strong>of</strong> reflective and refractive scanning elements.<br />

Charlie Chong/ Fion Zhang


Now, electronically scanned thermal imaging is accomplished by means <strong>of</strong> an<br />

infrared focal plane array (IRFPA), whereby a two-dimensional staring array<br />

<strong>of</strong> detectors collects radiant energy from the target and is digitally scanned to<br />

produce the thermogram. In the case <strong>of</strong> the line scanner (Figure A-16), the<br />

opto-mechanical scanning approach is gradually being superceded by<br />

replacement <strong>of</strong> the single-element detector with an electronically scanned<br />

linear focal plane array (a line <strong>of</strong> detectors), thus eliminating the scanning<br />

mechanism entirely. At the time <strong>of</strong> this writing, focal plane array imagers have<br />

all but completely replaced optomechanically scanned imagers in<br />

manufacturers’ inventory and product literature. Because many optomechanically<br />

scanned line scanners and imagers are still in use throughout<br />

the predictive maintenance community, the following discussion is included in<br />

this appendix.<br />

Charlie Chong/ Fion Zhang


Opto-mechanical Scanner<br />

A typical commercial rectilinear opto-mechanical scanner is shown<br />

schematically in Figure A-17. It employs two oscillating mirrors (reflective<br />

scanning) behind the primary lens and is commonly used in commercially<br />

available scanners. This approach has the advantage <strong>of</strong> a broad spectral<br />

response limited only by the spectral characteristics <strong>of</strong> the detector and the<br />

primary lens system. The main disadvantage is that the elements and their<br />

associated drive mechanisms must be arranged so that there is no optical or<br />

mechanical interference. This makes compact design more difficult. An<br />

alternate approach to scanning employs two rotating prisms behind the<br />

primary lens system. This instrument, using refractive scanning elements, has<br />

the advantage <strong>of</strong> compact design, because all <strong>of</strong> the scanning elements can<br />

be arranged in a line. It has the disadvantage <strong>of</strong> spectral limitation in that<br />

each element must transmit the entire portion <strong>of</strong> the infrared spectrum for<br />

which the instrument was designed. Some energy is absorbed by each<br />

refractive element, reducing the throughput somewhat, and the rather high<br />

cost <strong>of</strong> infrared transmitting materials add to the instrument cost. It should be<br />

pointed out that opto-mechanical scanners can employ refractive or reflective<br />

scanning elements or even combinations <strong>of</strong> both elements.<br />

Charlie Chong/ Fion Zhang


Figure A-17 Schematic <strong>of</strong> a Typical Opto-Mechanically Scanned Imager<br />

Charlie Chong/ Fion Zhang


Electronic scanning<br />

Electronic scanning involves no mechanical scanning elements.the surface is<br />

scanned electronically. The earliest type <strong>of</strong> electronically scanned thermal<br />

imager is the pyrovidicon.<br />

Pyrovidicon thermal imagers<br />

Pyrovidicon thermal imagers (pyroelectric vidicons) or thermal video<br />

systems are devices in which charge proportional to target temperature is<br />

collected on a single pyroelectric detector surface within an electronic<br />

picture tube, and scanning is accomplished by an electronic scanning<br />

beam. The pyrovidicon is a video camera tube that operates in the<br />

infrared (2.14 µm) region instead <strong>of</strong> in the visible spectrum. Electronically<br />

scanned thermal imaging systems based on pyrovidicons and operating<br />

in the 8.14 µm atmospheric window are in common use today. They<br />

provide qualitative thermal images and are classified as thermal viewers.<br />

Charlie Chong/ Fion Zhang


Focal plane array (FPA) imagers<br />

Focal plane array (FPA) imagers have, over the last decade, become the<br />

imagers <strong>of</strong> choice over opto-mechanically scanned imagers, replacing<br />

them in virtually all commercial applications. Manufacturers <strong>of</strong> FPA<br />

imagers <strong>of</strong>fer a wide choice <strong>of</strong> both cooled and uncooled detector arrays,<br />

with a wide selection <strong>of</strong> spectral ranges for both measuring (quantitative)<br />

and non-measuring (qualitative) applications. A more detailed discussion<br />

<strong>of</strong> focal plane array imagers can be found in Section 2.<br />

Published performance characteristics <strong>of</strong> currently available infrared<br />

commercial thermal imaging systems, including detailed discussions <strong>of</strong><br />

diagnostic s<strong>of</strong>tware and image recording methods, can also be found in<br />

Section 2, Table 2-1.<br />

Figure A-18 is a schematic <strong>of</strong> a typical focal plane array based thermal<br />

imager.<br />

Charlie Chong/ Fion Zhang


Figure A-18 Schematic <strong>of</strong> a Typical (Staring) FPA-Based Thermal<br />

Imager<br />

Charlie Chong/ Fion Zhang


Staring Array<br />

A staring array, staring-plane array, focal-plane array (FPA), or focal-plane is<br />

an image sensing device consisting <strong>of</strong> an array (typically rectangular) <strong>of</strong> lightsensing<br />

pixels at the focal plane <strong>of</strong> a lens. FPAs are used most commonly for<br />

imaging purposes (e.g. taking pictures or video imagery), but can also be<br />

used for non-imaging purposes such as spectrometry, LIDAR, and wave-front<br />

sensing.<br />

In radio astronomy the term "FPA" refers to an array at the focus <strong>of</strong> a radiotelescope<br />

(see full article on Focal Plane Arrays). At optical and infrared<br />

wavelengths it can refer to a variety <strong>of</strong> imaging device types, but in common<br />

usage it refers to two-dimensional devices that are sensitive in the infrared<br />

spectrum. Devices sensitive in other spectra are usually referred to by other<br />

terms, such as CCD (charge-coupled device) and CMOS image sensor in the<br />

visible spectrum. FPAs operate by detecting photons at particular<br />

wavelengths and then generating an electrical charge, voltage, or resistance<br />

in relation to the number <strong>of</strong> photons detected at each pixel. This charge,<br />

voltage, or resistance is then measured, digitized, and used to construct an<br />

image <strong>of</strong> the object, scene, or phenomenon that emitted the photons.<br />

Charlie Chong/ Fion Zhang<br />

http://military.wikia.com/wiki/Staring_arrayc


Applications for infrared FPAs include missile or related weapons guidance<br />

sensors, infrared astronomy, manufacturing inspection, thermal imaging for<br />

firefighting, medical imaging, and infrared phenomenology (such as observing<br />

combustion, weapon impact, rocket motor ignition and other events that are<br />

interesting in the infrared spectrum).<br />

Comparison To Scanning Array<br />

Staring arrays are distinct from scanning array and TDI (time-domain<br />

integration) imagers in that they image the desired field <strong>of</strong> view without<br />

scanning. Scanning arrays are constructed from linear arrays (or very narrow<br />

2-D arrays) that are rastered across the desired field <strong>of</strong> view using a rotating<br />

or oscillating mirror to construct a 2-D image over time. A TDI imager<br />

operates in similar fashion to a scanning array except that it images<br />

perpendicularly to the motion <strong>of</strong> the camera. A staring array is analogous to<br />

the film in a typical camera; it directly captures a 2-D image projected by the<br />

lens at the image plane.<br />

Charlie Chong/ Fion Zhang<br />

http://military.wikia.com/wiki/Staring_arrayc


A scanning array is analogous to piecing together a 2D image with photos<br />

taken through a narrow slit. A TDI imager is analogous to looking through a<br />

vertical slit out the side window <strong>of</strong> a moving car, and building a long,<br />

continuous image as the car passes the landscape.<br />

Scanning arrays were developed and used because <strong>of</strong> historical difficulties in<br />

fabricating 2-D arrays <strong>of</strong> sufficient size and quality for direct 2-D imaging.<br />

Modern FPAs are available with up to 2048 x 2048 pixels, and larger sizes<br />

are in development by multiple manufacturers. 320 x 256 and 640 x 480<br />

arrays are available and affordable even for non-military, non-scientific<br />

applications.<br />

Charlie Chong/ Fion Zhang<br />

http://military.wikia.com/wiki/Staring_arrayc


Staring<br />

Charlie Chong/ Fion Zhang


A.4 Performance Parameters <strong>of</strong> Thermal-Sensing<br />

Instruments<br />

To select an instrument suitable to a particular application, the thermographer<br />

needs to understand how to determine and specify its required performance.<br />

This section provides information regarding the performance parameters <strong>of</strong> (1)<br />

point-sensing instruments and (2) scanning & imaging instruments.<br />

Charlie Chong/ Fion Zhang


A.4.1 Point-Sensing Instruments<br />

For point-sensing instruments (infrared radiation thermometers), the following<br />

performance parameters should be considered:<br />

• Temperature range: The high and low limits over which the target<br />

emperature can vary<br />

• Absolute accuracy: As related to the National Institute <strong>of</strong> Standards and<br />

Technology (NIST) standard<br />

• Repeatability: How faithfully a reading is repeated for the same target<br />

• Temperature sensitivity: The smallest target temperature change that the<br />

instrument needs to detect<br />

• Speed <strong>of</strong> response: How fast the instrument responds to a temperature<br />

change at the target surface<br />

• Target spot size and working distance: The size <strong>of</strong> the spot on the target to<br />

be measured, and its distance from the instrument (FOV/IFOV)<br />

• Output requirements: How the output signal is to be used<br />

• Spectral range: The portion <strong>of</strong> the infrared spectrum over which the<br />

instrument will operate<br />

• Sensor environment: The ambient conditions under which the instrument<br />

will operate<br />

Charlie Chong/ Fion Zhang


Temperature range and absolute accuracy will always be interrelated; for<br />

example, the instrument might be expected to measure a range <strong>of</strong><br />

temperatures from 0 to 200°C with an absolute accuracy ± 2°C over the<br />

entire range. This could alternately be specified as ± 1% absolute accuracy<br />

over full scale. On the other hand, we might require the best accuracy at<br />

some specific temperature, say 100°C. In this case, the manufacturer should<br />

be so informed. The instrument can then be calibrated to exactly match the<br />

manufacturer’s laboratory calibration standard at that temperature.<br />

It is difficult for a manufacturer to comply with a tight specification for absolute<br />

accuracy because absolute accuracy is based on traceability to the National<br />

Institute <strong>of</strong> Standards and Technology (NIST) standard. An absolute accuracy<br />

<strong>of</strong> ±0.5°C ± 1% <strong>of</strong> full scale is about as tight as can be reasonably specified.<br />

Repeatability, on the other hand, can be more easily assured by the<br />

manufacturer, and is usually more important to the user.<br />

Charlie Chong/ Fion Zhang


Temperature sensitivity is also called thermal resolution (≠ spatial resolution)<br />

or noise equivalent temperature difference. It is the smallest temperature<br />

change at the target surface that must be clearly sensed at the output <strong>of</strong> the<br />

instrument. This is almost always closely associated with the cost <strong>of</strong> the<br />

instrument, so unnecessarily fine temperature sensitivity should not be<br />

specified.<br />

An important rule to remember is that, for any given instrument, target<br />

sensitivity will improve for hotter targets where there is more energy available<br />

for the instrument to measure. We should specify temperature sensitivity,<br />

therefore, at a particular target temperature, and this should be near the low<br />

end <strong>of</strong> the range <strong>of</strong> interest. We might, for example, specify temperature<br />

sensitivity to be 0.25°C at a target temperature <strong>of</strong> 25°C, and be confident that<br />

the sensitivity <strong>of</strong> the instrument will be at least that for targets hotter than<br />

25°C.<br />

Keywords<br />

Temperature sensitivity is also called thermal resolution or noise equivalent<br />

temperature difference (NETD).<br />

Charlie Chong/ Fion Zhang


EXAM score!<br />

Temperature sensitivity is also<br />

called thermal resolution or<br />

noise equivalent temperature<br />

difference (NETD).<br />

for my ASNT exam<br />

Charlie Chong/ Fion Zhang


EXAM score!<br />

thermal resolution<br />

(≠ spatial resolution)<br />

for my ASNT exam<br />

Charlie Chong/ Fion Zhang


NETD - Noise Equivalent Temperature Difference<br />

Noise Equivalent Temperature Difference is used to measure the<br />

performance <strong>of</strong> a infrared cameras ability discern the minimum level <strong>of</strong><br />

thermal sensitivity and is very similar to the MRTD with the exception that the<br />

test is based on the output <strong>of</strong> the detector only, without taking into<br />

consideration the performance <strong>of</strong> the infrared cameras image as it would be<br />

displayed to a thermographer. The results are usually expressed as the<br />

NETD. A common specification for an IR cameras NETD is 0.02 deg. C at 30<br />

deg. C.<br />

MRTD - Minimum Resolvable Temperature Difference<br />

Minimum Resolvable Temperature Difference is a test developed by the<br />

Department <strong>of</strong> Defense (ASTM Standard E1213) and used to measure the<br />

performance <strong>of</strong> a infrared cameras ability discern the minimum level <strong>of</strong><br />

thermal sensitivity that a operator <strong>of</strong> the camera can see. The test involves<br />

selecting the smallest test pattern (4 bars with a 7:1 length to width aspect<br />

ratio) that can be clearly distinguished by the operator as viewed on a display.<br />

Charlie Chong/ Fion Zhang<br />

http://www.prothermographer.com/training/IRBasics/qualitative_thermography/mrtd<br />

_minimum_resolvable_temperature_difference.htm


NETD - Noise Equivalent Temperature Difference<br />

NETD is used to measure the performance <strong>of</strong> a infrared cameras ability<br />

discern the minimum level <strong>of</strong> thermal sensitivity and is very similar to the<br />

MRTD with the exception that the test is based on the output <strong>of</strong> the detector<br />

only, without taking into consideration the performance <strong>of</strong> the infrared<br />

cameras image as it would be displayed to a thermographer. The results are<br />

usually expressed as the NETD. A common specification for an IR cameras<br />

NETD is 0.02 deg. C at 30 deg. C.<br />

MRTD - Minimum Resolvable Temperature Difference<br />

METD is a test developed by the Department <strong>of</strong> Defense (ASTM Standard<br />

E1213) and used to measure the performance <strong>of</strong> a infrared cameras ability<br />

discern the minimum level <strong>of</strong> thermal sensitivity that a operator <strong>of</strong> the camera<br />

can see. The test involves selecting the smallest test pattern (4 bars with a<br />

7:1 length to width aspect ratio) that can be clearly distinguished by the<br />

operator as viewed on a display.<br />

Charlie Chong/ Fion Zhang<br />

http://www.prothermographer.com/training/IRBasics/qualitative_thermography/mrtd<br />

_minimum_resolvable_temperature_difference.htm


Speed <strong>of</strong> response is generally defined as the time it takes the instrument<br />

output to respond to 95% <strong>of</strong> a step change at the target surface.<br />

Figure A-19 shows this graphically. Note that the sensor time constant is<br />

defined by convention to be the time required to reach 63% <strong>of</strong> a step change<br />

at the target surface. Instrument speed <strong>of</strong> response is about 5 time constants,<br />

and is generally limited by the detector used. As previously discussed, this<br />

limit is on the order <strong>of</strong> microseconds for photodetectors and milliseconds for<br />

thermal detectors. There is, however, a trade<strong>of</strong>f between speed <strong>of</strong> response<br />

and temperature sensitivity. As in all instrumentation systems, as the speed <strong>of</strong><br />

response becomes faster (wider information bandwidth), the sensitivity<br />

becomes poorer (lower signal-to-noise ratio). We learn from this that the<br />

speed <strong>of</strong> response should not be over-specified.<br />

Keywords:<br />

■ 63%<br />

■ 95%<br />

Charlie Chong/ Fion Zhang


Figure A-19 Instrument Speed <strong>of</strong> Response and Time Constant<br />

Charlie Chong/ Fion Zhang


Target spot size (also called spatial resolution) and working distance can be<br />

specified as just that (1 cm at 1 meter, for example), or we can put it in more<br />

general terms such as field <strong>of</strong> view angle (10 milliradians, 1 degree, 2<br />

degrees) or a field <strong>of</strong> view (spot size-to -working distance) ratio (D/15, D/30,<br />

D/75). A D/15 ratio means that the instrument measures the emitted energy <strong>of</strong><br />

a spot one-fifteenth the size <strong>of</strong> the working distance (3 cm at 45 cm, for<br />

example).<br />

Figure A-20 illustrates the fields <strong>of</strong> view for several instruments and how an<br />

instrument can be selected based on the spot size and working distance<br />

required. An examination <strong>of</strong> the collecting beams <strong>of</strong> the instruments shown<br />

also shows that, at very close working distances, this simple ratio does not<br />

always apply. If close-up information is not clearly provided in the product<br />

literature, the instrument manufacturer should be consulted. For quick<br />

reference, a method <strong>of</strong> approximating spot size based on manufacturerrovided<br />

information is illustrated in Appendix C, Plate 2.<br />

Charlie Chong/ Fion Zhang


Figure A-20 Fields <strong>of</strong> View <strong>of</strong> <strong>Infrared</strong> Radiation Thermometers<br />

Charlie Chong/ Fion Zhang


Figure A-20 Fields <strong>of</strong> View <strong>of</strong> <strong>Infrared</strong> Radiation Thermometers<br />

Charlie Chong/ Fion Zhang


Figure A-20 Fields <strong>of</strong> View <strong>of</strong> <strong>Infrared</strong> Radiation Thermometers<br />

Charlie Chong/ Fion Zhang


Figure A-20 Fields <strong>of</strong> View <strong>of</strong> <strong>Infrared</strong> Radiation Thermometers<br />

An examination <strong>of</strong> the collecting<br />

beams <strong>of</strong> the instruments shown also<br />

shows that, at very close working<br />

distances, this simple ratio does not<br />

always apply. If close-up information<br />

is not clearly provided in the product<br />

literature, the instrument<br />

manufacturer should be consulted.<br />

Charlie Chong/ Fion Zhang


The output requirements are totally dependent on the user’s needs. If a<br />

readout indicator is required, a wide selection is usually <strong>of</strong>fered. An analog<br />

output suitable for recording, monitoring, and control is commonly provided. In<br />

addition, most manufacturers <strong>of</strong>fer a broad selection <strong>of</strong> output functions<br />

including digital (BCD coded) outputs, high, low, and proportional set-points,<br />

signal peak or valley sensors, sample and hold circuits, and even closed-loop<br />

controls for specific applications. Many currently available instruments, even<br />

portable hand-held units, include microprocessors that provide many <strong>of</strong> the<br />

above functions on standard models.<br />

Charlie Chong/ Fion Zhang


As previously noted, the operating spectral range <strong>of</strong> the instrument is <strong>of</strong>ten<br />

critical to its performance. For cooler targets, up to about 500°C, most<br />

manufacturers <strong>of</strong>fer instruments operating in the 8.14 µm atmospheric<br />

window. For hotter targets, shorter operating wavelengths are selected,<br />

usually shorter than 3 µm.<br />

One reason for choosing shorter wavelengths is that this enables<br />

manufacturers to use commonly available and less expensive quartz and<br />

glass optics, which have the added benefit <strong>of</strong> being visibly transparent for<br />

more convenient aiming and sighting. Another reason is that estimating<br />

effective emissivity incorrectly will result in smaller temperature errors when<br />

measurements are made at shorter wavelengths. A good general rule to<br />

follow, particularly when dealing with targets <strong>of</strong> low or uncertain effective<br />

emissivities, is to work at the shortest wavelengths possible without<br />

compromising sensitivity or risking susceptibility to reflections from visible<br />

energy sources.<br />

Charlie Chong/ Fion Zhang


Spectrally selective instruments employ interference filters to allow only a<br />

very specific broad or narrow band <strong>of</strong> wavelengths to reach the detector. (A<br />

combination <strong>of</strong> a spectrally selective detector and a filter can also be used.)<br />

This can make the instrument highly selective to a specific material whose<br />

temperature is to be measured in the presence <strong>of</strong> an intervening medium or<br />

an interfering background. For example, for measuring the temperature <strong>of</strong><br />

objects from 200°C to 1000°C inside a heating chamber with a glass port, or<br />

inside a glass bell jar, an instrument operating in the 1.5 to 2.5 µm band will<br />

see through the glass and make the measurement easily. A very important<br />

generic example <strong>of</strong> the need for spectral selectivity is in the measurement <strong>of</strong><br />

plastics in the process <strong>of</strong> being formed into films and other configurations.<br />

Keywords:<br />

interference filters<br />

Charlie Chong/ Fion Zhang


Thin films <strong>of</strong> many plastics are virtually transparent to most infrared<br />

wavelengths but do emit at certain wavelengths. Polyethylene, polypropylene,<br />

and other related materials, for example, have a very strong, though narrow,<br />

absorption band at 3.45 µm. Polyethylene film is formed at about 200°C in the<br />

presence <strong>of</strong> heaters that are at about 700°C.<br />

Figure A-21 shows the transmission spectra <strong>of</strong> 1.5- mil thick polyethylene film<br />

and the narrow absorption band at 3.45 µm. The instrument selected for<br />

measuring the surface <strong>of</strong> the film has a broadband thermal detector and a<br />

3.45 µm spike band pass filter. The filter makes the instrument blind to all<br />

energy outside <strong>of</strong> 3.45 µm, and enables it to measure the temperature <strong>of</strong> the<br />

surface <strong>of</strong> the plastic film without seeing through the film to the heaters.<br />

Charlie Chong/ Fion Zhang


Figure A-21 Spectral Filtering for Polyethylene Temperature<br />

Measurement<br />

Charlie Chong/ Fion Zhang


The object is opaque to 3.45 µm radiation, by using 3.45 µm pass filter, only<br />

the object’s 3.45 µm is monitored, all other bandwidth from the object or<br />

transmitted from the process hot roller are filtered <strong>of</strong>f.<br />

3.45 µm pass filter<br />

Charlie Chong/ Fion Zhang


Figure A-22 shows a similar solution for 0.5-mil thick polyester (Mylar) film<br />

under about the same temperature conditions. Here, the strong polyester<br />

absorption band, from 7.7 to 8.2 µm, dictates the use <strong>of</strong> a 7.9 µm spike filter<br />

placed in front <strong>of</strong> the same broadband detector.<br />

Charlie Chong/ Fion Zhang


Figure A-22 Spectral Filtering for Polyester Temperature Measurement<br />

Charlie Chong/ Fion Zhang


A.4.2 Scanners and Imagers.Qualitative and Quantitative<br />

The parameters used for assessing the performance <strong>of</strong> infrared thermal<br />

imaging scanners are complex and the methods used for testing performance<br />

have generated some controversy among manufacturers and users <strong>of</strong> these<br />

instruments. A thermal image is made up <strong>of</strong> a great number <strong>of</strong> discrete point<br />

measurements, however, many <strong>of</strong> the performance parameters <strong>of</strong> infrared<br />

thermal imagers are the same as those <strong>of</strong> radiation thermometers (pointsensing<br />

infrared radiometers that read out in temperature). Others derive from,<br />

or are extensions <strong>of</strong>, radiation thermometer performance parameters.<br />

Qualitative (non-measuring) thermal imagers, also called thermal viewers,<br />

differ from quantitative (measuring) thermal imagers, also called imaging<br />

radiometers, in that thermal viewers do not provide temperature or thermal<br />

energy measurements.<br />

It should be noted, therefore, that for users requiring qualitative rather than<br />

quantitative thermal images, many <strong>of</strong> the parameters discussed herein are <strong>of</strong><br />

no importance.<br />

Charlie Chong/ Fion Zhang


A.4.3 Performance Parameters <strong>of</strong> Imaging Radiometers<br />

The Environmental Research Institute, Michigan (ERIM) <strong>Infrared</strong> Handbook<br />

[13] provides an extensive table <strong>of</strong> terms and definitions (section 19.1.2) and<br />

a list <strong>of</strong> specimen specifications (section 19.4.1). The section <strong>of</strong> the<br />

Handbook covering infrared imaging systems does not, however, deal with<br />

the imager as a quantitative measurement instrument, and so the<br />

performance parameters related with temperature measurement need to be<br />

added. Some simplifications can be made, which result in some acceptable<br />

approximations. Bearing these qualifications in mind, the following definitions<br />

<strong>of</strong> the key performance parameters <strong>of</strong> infrared thermal scanners are <strong>of</strong>fered:<br />

Charlie Chong/ Fion Zhang


• Total field <strong>of</strong> view (TFOV): the image size, in terms <strong>of</strong> scanning angle.<br />

(example: TFOV = 20°V x 30°H)<br />

• Instantaneous field <strong>of</strong> view (IFOV): the angular projection <strong>of</strong> the detector<br />

element at the target plane; imaging spatial resolution. (example: IFOV= 2<br />

milliradians )<br />

• Measurement spatial resolution (IFOVmeas): the spatial resolution<br />

describing the minimum target spot size on which an accurate temperature<br />

measurement can be made. (example: IFOVmeas = 5 milliradians)<br />

• Frame repetition rate: The number <strong>of</strong> times every point on the target is<br />

scanned in one second. (example: Frame rate = 30 /second)<br />

• Minimum resolvable temperature (MRT) (NETD? / MRDT?) : The smallest<br />

blackbody equivalent target temperature difference that can be observed;<br />

temperature sensitivity (example: MRT=0.1°C @ 30°C target temperature)<br />

Charlie Chong/ Fion Zhang


Minimum resolvable temperature MRT<br />

MRT and the terms relating to spatial resolution are interrelated and cannot<br />

be considered independently. (unlike the point sensing: IR thermometer)<br />

Other parameters, such as spectral ranges, target temperature ranges,<br />

accuracy and repeatability, and focusing distances, are essentially the same<br />

as those defined previously for infrared radiation thermometers, although they<br />

can be expressed differently. Dynamic range and reference level range, for<br />

example, are the terms that define the target temperature ranges for thermal<br />

imagers. While the operating spectral range <strong>of</strong> a radiation thermometer is<br />

<strong>of</strong>ten critical to its performance, the spectral range <strong>of</strong> operation <strong>of</strong> a thermal<br />

imager is not usually as critical to the user, except for a few specialized<br />

applications. Most commercial thermal imagers operate in either the 2.5 µm<br />

or the 8.12 µm atmospheric window, depending on the manufacturer’s choice<br />

<strong>of</strong> detector. Filter wheels or slides are usually available to enable users to<br />

insert special interference filters and perform spectrally selective<br />

measurements when necessary.<br />

Charlie Chong/ Fion Zhang


Despite some manufacturers’ claims to the contrary, there is usually little<br />

difference in overall performance between an imager operating in the 2.5 µm<br />

band and an imager operating in the 8.12 µm band, all other parameters<br />

being equal.<br />

For a specific application, however, there might be a clear choice. One<br />

example <strong>of</strong> this would be selecting an imager operating in the 2 .5 µm band to<br />

observe a target through a quartz window. There would be no alternative<br />

because quartz is virtually opaque in the 8.12 µm region. Another example<br />

would be selecting an imager operating in the 8.12 µm band to observe a cool<br />

target through a long atmospheric path. The choice would be obvious<br />

because long-path atmospheric absorption is substantially greater in the 2.5<br />

µm window than in the 8.12 µm window.<br />

Charlie Chong/ Fion Zhang


For qualitative (non measuring) thermal viewers, parameters relating to<br />

temperature range are only applicable in the broadest sense. Absolute<br />

accuracy and stability parameters are not applicable. MRT is applicable only<br />

as an approximation because stability cannot be assured. IFOV meas is not<br />

applicable.<br />

Secondary features, such as field uniformity and spatial distortion, are design<br />

parameters and are assumed to be handled by responsible manufacturers. A<br />

discussion <strong>of</strong> the significant performance parameters (figures <strong>of</strong> merit) follows.<br />

Charlie Chong/ Fion Zhang


A.4.3.1 Temperature Sensitivity, Minimum Resolvable Temperature<br />

Difference<br />

(MRTD) or Minimum Resolvable Temperature (MRT) Temperature sensitivity,<br />

also called thermal resolution or noise equivalent temperature difference<br />

(NETD) for a radiation thermometer, is the smallest temperature change at<br />

the target surface and can be clearly sensed at the output <strong>of</strong> the instrument.<br />

For an imaging system, the MRT or MRTD defines temperature sensitivity but<br />

also implies spatial resolution (IFOV). MRTD is expressed as a function <strong>of</strong><br />

angular spatial frequency. Testing for MRTD is usually accomplished by<br />

means <strong>of</strong> a subjective procedure developed by the Department <strong>of</strong> Defense<br />

community.<br />

Keywords:<br />

■ the MRT or MRTD defines temperature sensitivity but also implies spatial<br />

resolution (IFOV) (for 2D thermography; both thermal viewer & thermal<br />

radiometric imaging) .<br />

■ MRTD is expressed as a function <strong>of</strong> angular spatial frequency. (?)<br />

Charlie Chong/ Fion Zhang


This involves selecting the smallest (highest frequency) standard periodic test<br />

pattern (four bars, 7:1 length-to-width aspect ratio) that can be distinguished<br />

as a 4 bar contrast target by the observer, and recording the smallest<br />

detectable element-to-element temperature difference between two<br />

blackbody elements on this pattern. Unlimited viewing time and optimization<br />

<strong>of</strong> controls is allowed and the target is oriented with the bars normal to the<br />

horizontal scan line.<br />

Figure A-23 illustrates the setup using an ambient pattern and a heated<br />

background. The MRTD curve shown is a function <strong>of</strong> spatial frequency<br />

(cycles/mRad). Additional points on the curve are achieved by changing the<br />

pattern size or the distance to the scanner.<br />

Charlie Chong/ Fion Zhang


Figure A-23 Test Setup for MRTD Measurement, MRTD Curve<br />

heated background<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang


A.4.3.2 Spot Size (FOV) , Instantaneous Field <strong>of</strong> View (IFOV), Imaging<br />

Spatial Resolution (?) , Measurement Spatial Resolution (IFOVmeas)<br />

For thermal imagers, the instantaneous field <strong>of</strong> view (IFOV) expresses spatial<br />

resolution for imaging purposes but not for measurement purposes.<br />

Measurement instantaneous field <strong>of</strong> view (IFOVmeas) expresses spatial<br />

resolution for measurement purposes.<br />

The modulation transfer function (MTF) is a measure <strong>of</strong> imaging spatial<br />

resolution. Modulation is a measure <strong>of</strong> radiance contrast and is expressed:<br />

Modulation = (L max -L min ) / (L max + L min )<br />

L = luminosity?<br />

Modulation transfer is the ratio <strong>of</strong> the modulation in the observed image to<br />

that in the actual object.<br />

Charlie Chong/ Fion Zhang


For any system, MTF will vary with scan angle and background, and will <strong>of</strong>ten<br />

be different when measured along the horizontal than it is when measured<br />

along the vertical.<br />

For this reason, a methodology was established and accepted by<br />

manufacturers and users alike to measure the MTF <strong>of</strong> an imager and, thereby,<br />

to verify the spatial resolution for imaging (night vision) purposes. A sample<br />

procedure follows for a system where IFOV is specified at 2.0 milliradians.<br />

This is shown in Figure A-24 and uses the same setup as illustrated in Figure<br />

A-23:<br />

Charlie Chong/ Fion Zhang


■<br />

A standard 4 bar (slit) resolution target (7:1 aspect ratio) with a 2-mm slit<br />

width is placed in front <strong>of</strong> a heated blackbody reference surface at a<br />

distance <strong>of</strong> 1 meter from the primary optic <strong>of</strong> the instrument. The ratio <strong>of</strong><br />

the 2-mm slit width to the 1-meter working distance is 2 milliradians). The<br />

target is centered in the scanned field (oriented so that the horizontal axis<br />

is normal to the slit), and a single line scan output signal is monitored.<br />

The analog signal value <strong>of</strong> the 4 peaks (Vmax), as the slits are scanned,<br />

and the analog signal value <strong>of</strong> the 3 valleys (Vmin), are recorded using<br />

the bar target surface ambient temperature as a base reference. The<br />

MTF is expressed as a ratio equal to (Vmax -Vmin) / (Vmax + Vmin). If<br />

this ratio is at least 0.35, the 2 milliradian IFOV is verified.<br />

There are some disagreements among users and manufacturers regarding<br />

the acceptable minimum value <strong>of</strong> MTF to verify imaging spatial resolution,<br />

with values varying between 0.35 and 0.5, depending on the manufacturer<br />

and the purpose <strong>of</strong> the instrument. For most users, a tested value <strong>of</strong> MTF,<br />

equal to or greater than 0.35 for a slit width representing a specified spatial<br />

resolution is generally considered sufficient to demonstrate that spatial<br />

resolution for imaging purposes.<br />

Charlie Chong/ Fion Zhang


Figure A-24 Modulation Transfer Function, Imager Spatial Resolution<br />

Charlie Chong/ Fion Zhang


Both MRTD and MTF are functions <strong>of</strong> spatial frequency for any given system.<br />

This is illustrated in Figure A-25, reprinted from J.M. Lloyd, Thermal Imaging<br />

Systems [14], for a typical system rated by the manufacturer to be 1<br />

milliradian. The cut-<strong>of</strong>f frequency is where the IFOV equals 1 cycle (one bar<br />

and one slit) so that the intersection <strong>of</strong> the two curves at the half-cut-<strong>of</strong>f<br />

frequency represents the actual performance <strong>of</strong> the system for an MRTD <strong>of</strong><br />

1°C. MTF is seen to be about 0.22 for this system.<br />

Charlie Chong/ Fion Zhang


Figure A-25 MRTD and MTF for a System Rated at 1.0 Milliradian<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang


For measurement purposes, <strong>of</strong> course the slit width should, ideally, be<br />

increased until the modulation reaches unity. For this reason the MTF method<br />

was found to be unsatisfactory for commercial thermal imagers where<br />

quantitative temperature measurement and control are <strong>of</strong>ten necessary.<br />

Another procedure, called the Slit Response Function (SRF), was developed<br />

for this purpose and is generally accepted for measuring IFOVmeas. In this<br />

method, illustrated in Figure A-26, a single variable slit is placed in front <strong>of</strong> a<br />

blackbody source and the slit width is varied until the resultant single-line- can<br />

signal approaches the signal <strong>of</strong> the blackbody reference. Because there are<br />

other errors in the optics, the 100% level <strong>of</strong> SRF is approached rather slowly,<br />

as shown in the curve <strong>of</strong> Figure A-26. The slit width at which the SRF<br />

reaches 0.9, divided by the distance to the slit (W/d), is usually accepted<br />

as the IFOVmeas <strong>of</strong> the instrument under test. Figures A-23 and A-26 are<br />

adapted from the Ohman paper, .Measurement Versus Imaging in<br />

<strong>Thermography</strong>. [15], which provides a detailed description <strong>of</strong> the Slit<br />

Response Method, setup diagrams, and a discussion <strong>of</strong> imaging and<br />

measurement spatial resolution figures <strong>of</strong> merit. The step-by-step procedure<br />

for measuring SRF is described in detail in Appendix C, Plate 6.<br />

Charlie Chong/ Fion Zhang


IFOVmeas: The slit width at which the SRF reaches 0.9, divided by the<br />

distance to the slit (W/d), is usually accepted as the IFOVmeas <strong>of</strong> the<br />

instrument under test.<br />

for IFOV or IFOV geometric :<br />

D = σ∙d, σ (IFOV) = D / d<br />

σ<br />

D<br />

d<br />

for IFOV meas = D meas /d<br />

σ<br />

D meas<br />

d<br />

Charlie Chong/ Fion Zhang


Figure A-26 Setup and Curves for Slit Response Function Test<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang


Note: Because FPA imagers have all but replaced opto-mechanically<br />

scanned imagers, many experienced thermographers suggest that the SRF<br />

measurement procedure be performed in both the horizontal and vertical<br />

scan-line direction. The larger <strong>of</strong> the two results is then accepted as the<br />

IFOV meas <strong>of</strong> the imager under test.<br />

Charlie Chong/ Fion Zhang


FPA<br />

Charlie Chong/ Fion Zhang<br />

http://spie.org/x34358.xml


FPA<br />

Charlie Chong/ Fion Zhang


FPA<br />

Charlie Chong/ Fion Zhang


FPA<br />

Charlie Chong/ Fion Zhang


A.4.3.3 Speed <strong>of</strong> Response and Frame Repetition Rate<br />

Speed <strong>of</strong> response <strong>of</strong> a radiation thermometer is generally defined as the time<br />

it takes the instrument output to respond to 95% <strong>of</strong> a step change at the<br />

target surface (about 5 time constants). This parameter is not applicable to<br />

thermal imagers. Frame repetition rate is the measure <strong>of</strong> the data update <strong>of</strong> a<br />

thermal imager. This is not the same as field repetition rate. (Manufacturers<br />

might use fast field rates with not all <strong>of</strong> the picture elements included in any<br />

one scan, and then interlace the fields so that it takes multiple fields to<br />

complete a full frame. This might produce a more flicker-free image and be<br />

more pleasing to the eye than scanning full data frames at a slower rate.<br />

Frame repetition rate is the number <strong>of</strong> times per second every picture element<br />

is scanned.<br />

Charlie Chong/ Fion Zhang


Figure A-19 Instrument Speed <strong>of</strong> Response and Time Constant<br />

Charlie Chong/ Fion Zhang


A.4.4 Thermal Imaging S<strong>of</strong>tware<br />

In order to optimize the effectiveness <strong>of</strong> thermography measurement<br />

programs, the thermographer needs a basic understanding <strong>of</strong> thermal image<br />

processing techniques. The following is a broad discussion <strong>of</strong> thermal image<br />

processing and diagnostics. A detailed description <strong>of</strong> thermal imaging and<br />

diagnostic s<strong>of</strong>tware currently available from manufacturers is provided in<br />

section 2.<br />

Thermal imaging s<strong>of</strong>tware can be categorized into the following groupings:<br />

• Quantitative thermal measurements <strong>of</strong> targets<br />

• Detailed processing and image diagnostics<br />

• Image recording, storage, and recovery<br />

• Image comparison<br />

• Archiving and database*<br />

*Although data and image database development is not an exclusive<br />

characteristic <strong>of</strong> thermal imaging s<strong>of</strong>tware, it should be considered an<br />

important part <strong>of</strong> the thermographer’s tool kit.<br />

Charlie Chong/ Fion Zhang


With the introduction <strong>of</strong> computer-assisted thermal image storage and<br />

processing, thermography has become a far more exact science, and the<br />

ability to perform image analysis and trend analysis has greatly expanded its<br />

reach. Innovative s<strong>of</strong>tware has been tailored specifically for detailed image<br />

and thermal data analysis, and has been rapidly updated and expanded.<br />

Most s<strong>of</strong>tware packages for thermography image analysis and diagnostics<br />

<strong>of</strong>fer a number <strong>of</strong> standard features. These include spot temperature readout,<br />

multiple X and Y analog traces, monochrome and multiple-color scale<br />

selection, image shift, rotation and magnification, area analysis with<br />

histogram display, image averaging and filtering, and permanent disk storage<br />

and retrieval. Some <strong>of</strong> these capabilities are <strong>of</strong>fered as part <strong>of</strong> the basic<br />

instrument and some are found in a diagnostics package <strong>of</strong>fered separately.<br />

Charlie Chong/ Fion Zhang


The newest field-portable instruments allow the thermographer to store<br />

images to disc (or data card) during field measurements, and perform detailed<br />

image analysis upon return to home base (see Section 2 for details). The<br />

ability to perform differential thermography is a most powerful feature <strong>of</strong><br />

thermographic s<strong>of</strong>tware routines. This is the capability for archiving thermal<br />

images <strong>of</strong> acceptable operating components, and assemblies and<br />

mechanisms, and using these stored images as models for comparison to<br />

subsequently inspected items. Subtractive routines produce differential<br />

images, illustrating the deviation <strong>of</strong> each pixel (picture element) from its<br />

corresponding model.<br />

Another powerful routine that was recently introduced is an emissivity<br />

determination and correction program, which produces true surfacetemperature<br />

thermograms <strong>of</strong> microelectronic devices and other very small<br />

targets.<br />

Keywords:<br />

Subtractive routines<br />

Charlie Chong/ Fion Zhang


To perform this function, the unpowered device is heated sequentially to two<br />

known low-level temperatures, and the stored thermal images are used to<br />

allow the computer to calculate emissivity <strong>of</strong> each pixel. The device is then<br />

powered and the image produced is corrected, point by point, for the<br />

emissivities previously computed. There is great interest in applying this<br />

spatial emissivity correction to larger targets such as circuit cards. The<br />

difficulty in developing a reliable emissivity matrix lies in achieving tight<br />

control over the temperature and temperature uniformity while heating a<br />

target <strong>of</strong> this size.<br />

For the pr<strong>of</strong>essional thermographer, the maintenance <strong>of</strong> an historical<br />

database is most critical, and thermography s<strong>of</strong>tware allows this to be done<br />

systematically. The historical data included with stored images (time, date,<br />

location, ambient conditions, distance to target, estimated effective emissivity,<br />

scanner serial number, and additional stored comments) serve as important<br />

inputs and subsequent backup for the written report. New s<strong>of</strong>tware to aid the<br />

thermographer in the efficient and rapid preparation <strong>of</strong> pr<strong>of</strong>essional looking<br />

reports is also available from most manufacturers <strong>of</strong> thermal imagers (see<br />

Section 2).<br />

Charlie Chong/ Fion Zhang


Appendix B<br />

Measuring Emissivity, Reflectance & Transmittance<br />

Charlie Chong/ Fion Zhang


B.1 Introduction<br />

An infrared radiometer measures the sum <strong>of</strong> the emitted (We), reflected (Wr),<br />

and transmitted (Wt) energies coming from the target <strong>of</strong> interest. Figure B-1<br />

(repeated from Appendix A, Figure A-8) demonstrates this graphically. The<br />

sum <strong>of</strong> We + Wr + Wt is called Exitance or Radiosity. To determine the<br />

temperature <strong>of</strong> the target, the emitted energy must first be subtracted from the<br />

reflected and transmitted energies. This value must then be corrected to<br />

account for the emissivity <strong>of</strong> the target and to obtain a blackbody equivalent<br />

value. The blackbody equivalent value is then converted to temperature by<br />

referencing a calibration curve. All <strong>of</strong> the techniques discussed below for<br />

measuring emissivity, reflectance, and transmittance assume that the user<br />

has a thermal imager. Also note that the values for emissivity, reflectance,<br />

and transmittance are valid only for the spectral range <strong>of</strong> that instrument.<br />

Charlie Chong/ Fion Zhang


Figure B-1 Target Radiosity<br />

Charlie Chong/ Fion Zhang


B.2 Measuring Emissivity<br />

There are several common techniques for the measurement <strong>of</strong> emissivity<br />

using a single band radiometer, two <strong>of</strong> which are illustrated below.<br />

■ The first technique, known as the reference emitter technique, is<br />

accomplished by direct comparison with a known emitter at the same<br />

temperature.<br />

■ The second technique, known as the reflective emissivity technique, is<br />

accomplished by calculating emissivity indirectly using measured values <strong>of</strong><br />

reflectance (and transmittance if applicable).<br />

Charlie Chong/ Fion Zhang


The reference emitter technique works well when the target is at a different<br />

temperature than the background, such as in the case <strong>of</strong> a steam inlet valve<br />

whose body is at system operating temperature, while the applied emissivity<br />

reference is at the same temperature as the target.<br />

The reflective emissivity technique works well for smooth surfaces such as an<br />

electrical connection. The reflective emissivity technique is independent <strong>of</strong><br />

target temperature, although the temperature <strong>of</strong> the target must remain<br />

constant throughout the measurement.<br />

A third, field-type method for estimating the effective emissivity <strong>of</strong> a specific<br />

target under specific conditions, is described in Section 3.3.3 and is illustrated<br />

in Appendix C, Plate 5.<br />

Charlie Chong/ Fion Zhang


B.2.1 Reference Emitter Technique<br />

The reference emitter technique assumes both that the transmittance through<br />

the target is zero, and that a constant temperature difference between the<br />

target and the background is maintained. Ideally, this temperature difference,<br />

either hotter or colder, should be in the range <strong>of</strong> at least 15°F to 25°F. If the<br />

target is colder than the background, it should be above the dew point so that<br />

condensation on the surface <strong>of</strong> the target cannot occur.<br />

The reference emitter technique will only work if a reference emitter is applied<br />

to the surface <strong>of</strong> the target.<br />

Good reference emitters (E) are foot-powder, dye check developer, or black<br />

electrician’s tape, as previously discussed in sections 4.1.2 through 4.1.4.<br />

The procedure for determining the effective emissivity <strong>of</strong> a target using the<br />

reference emitter is as follows (refer to Figure B-2):<br />

Charlie Chong/ Fion Zhang


1. Apply the reference emitter (E) to a portion <strong>of</strong> the target (an area <strong>of</strong> at least<br />

one square inch is normally adequate).<br />

2. Set the imager to measure isotherm units.<br />

3. Measure the background thermal level (B) adjacent to the target. Do this<br />

by placing a piece <strong>of</strong> cardboard to which is applied a crumpled, flattened<br />

piece <strong>of</strong> aluminum foil. Take this measurement over a large area <strong>of</strong> the foil.<br />

(An area <strong>of</strong> at least one square foot is normally adequate.)<br />

4. Measure the target thermal level (T).<br />

5. Measure the reference emitter level (R). The reference emitter must be in<br />

thermal equilibrium with the target. This thermal equilibrium condition will<br />

be apparent when the reference emitter thermal level is not changing. (In<br />

the case <strong>of</strong> dye check developer, its application cools the surface as the<br />

propellant evaporates. Wait at least 15 minutes after application unless the<br />

target is very warm.)<br />

6. Calculate the emissivity by using the equation: Emissivity=(T-B)/(R-B)<br />

7. Measure the emissivity several times. Determine the final value by taking<br />

an average <strong>of</strong> all measured emissivity values.<br />

Charlie Chong/ Fion Zhang


Figure B-2 Using the Reference Emitter Technique<br />

Charlie Chong/ Fion Zhang


B.2.2 Reflective Emissivity Technique<br />

The reflective emissivity technique involves measuring the reflectance <strong>of</strong> the<br />

target and subtracting it from 1.0 (emissivity = 1 minus target reflectance).<br />

ε= 1-ρ<br />

The procedure for determining emissivity using the reflective emissivity<br />

technique works best when dealing with highly reflected or mirrored surfaces,<br />

such as mirror insulation, and when dealing with pipes or electrical contacts.<br />

Some <strong>of</strong> these surfaces naturally have a low emissivity. In this technique, the<br />

target should not be coated with a reference emitter and must be kept at a<br />

constant temperature. Also, once a range is chosen for measuring<br />

temperature, both measurements must be made on that range. This<br />

technique is temperature independent. The emissivity, using the reflective<br />

emissivity technique, is calculated from the ratio <strong>of</strong> the thermal level<br />

differences. The procedure for determining the reflective emissivity technique<br />

follows (refer to Figure B-3). Note: The temperatures <strong>of</strong> the two sources must<br />

be constant and with a substantial spread between them (15°F to 25°F).<br />

Charlie Chong/ Fion Zhang


1. Establish that the two sources are at different temperatures and are<br />

thermally stable. This can be adequately accomplished with a hand-held<br />

contact pyrometer. The exact temperature <strong>of</strong> each surface does not need<br />

to be known, only the ΔT. The ΔT, however, is limited by the temperature<br />

range <strong>of</strong> the imager.<br />

2. Aim the imager at each source and measure the direct isotherm levels (S a<br />

and S b ).<br />

3. Reposition the imager so that the sources are reflected <strong>of</strong>f the target.<br />

Measure the reflected isotherm levels (T a and T b ). In most situations, this<br />

requires reflecting one source at a time (the exception is when they are<br />

reflected <strong>of</strong>f a large uniform surface).<br />

4. Calculate the target reflectance: Reflectance = (T a -T b ) /(S a -S b ) To ensure<br />

that the data is reliable, take the average <strong>of</strong> several <strong>of</strong> these<br />

measurements over several parts <strong>of</strong> the surface, particularly if the surface<br />

is non-uniform in appearance. The exception to this is when an imager,<br />

either directly or through s<strong>of</strong>tware, allows an area to be defined and<br />

averaged.<br />

Charlie Chong/ Fion Zhang


Figure B-3 Using the Reflective Emissivity Technique<br />

Reflectance = (T a -T b ) /(S a -S b )<br />

Charlie Chong/ Fion Zhang


B.2.3 Transmittance Measurement<br />

The transmittance <strong>of</strong> non-opaque targets is measured similar to the<br />

reflectance measurement technique. As shown in Figure B-4, two sources are<br />

again used. In this case, the target is placed directly in front <strong>of</strong> the two<br />

sources rather than reflected <strong>of</strong>f <strong>of</strong> it. To calculate transmittance, substitute<br />

the reflected levels in the equation cited previously for reflectance (Section<br />

B.2.2) with the transmitted thermal levels.<br />

Charlie Chong/ Fion Zhang


Figure B-4 Using the Transmittance Technique (Measuring<br />

Transmittance)<br />

Transmittance = (T a -T b ) /(S a -S b )<br />

Charlie Chong/ Fion Zhang


B.2.4 Generic Emissivity Values<br />

Table B-1 lists broadband, generic normal emissivity values for several<br />

common materials (repeated from Section 4, Table 4-1. These values should<br />

only be used as references until the user can compile a library <strong>of</strong> values<br />

based on actual measurements.<br />

Table B-1 Normal Emissivity Values <strong>of</strong> Common Materials<br />

Charlie Chong/ Fion Zhang


Appendix C<br />

Quick Reference Charts And Plates<br />

Charlie Chong/ Fion Zhang


Calculating Instantaneous Field Of View, Quick Calculation<br />

Charlie Chong/ Fion Zhang


MTF Determination<br />

Using An Ir Imager<br />

Charlie Chong/ Fion Zhang


Minimum Resolvable<br />

Temperature Difference<br />

(MRTD) Estimate Using An Ir<br />

Imager<br />

Charlie Chong/ Fion Zhang


Measuring And Setting Effective Emissivity Using An Imager Or A Point<br />

Sensor<br />

Charlie Chong/ Fion Zhang


MEASURING IFOV meas OF AN IMAGER<br />

USING THE SLIT RESPONSE FUNCTION (SRF)<br />

Charlie Chong/ Fion Zhang


Classification Of Faults (Guidelines) Relating To 50% Of Maximum Load<br />

Joule.s Law: P = I 2 R. Use this to proportion the temperature rise to 50% <strong>of</strong> the<br />

load.<br />

For example:<br />

At 20% <strong>of</strong> load, an 8°C rise is seen. To proportion it to 50% <strong>of</strong> load, multiply<br />

by the square <strong>of</strong> the load ratio as follows:<br />

(50/20)² = 6.25; 6.25 x 8°C = 50°C equivalent temperature rise<br />

Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang


Charlie Chong/ Fion Zhang


End Of <strong>Reading</strong> Three<br />

Charlie Chong/ Fion Zhang


<strong>Reading</strong>: Four<br />

Emissivity: Understand the difference<br />

between apparent and actual IR<br />

temperatures<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Emissivity: Understand the difference between<br />

apparent and actual IR temperatures<br />

Taking infrared temperature measurements is certainly a lot easier than it<br />

used to be. The tricky part is understanding when an infrared reading is<br />

accurate as is and when you need to account for certain properties <strong>of</strong> the<br />

materials you’re measuring, or for other things like heat transfer.<br />

The most common use <strong>of</strong> infrared temperature measurement is for the<br />

inspection <strong>of</strong> electrical power distribution equipment. Let’s look at a typical<br />

three-phase fused power disconnect (Figure 1) and the corresponding<br />

infrared image (Figure IR1) below.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Figure 1 shows a typical three-phase fused power disconnect. The<br />

corresponding infrared image, figure IR1, was taken with the emissivity<br />

setting at 1 on our thermal imager. The temperature span and color scale for<br />

the infrared image is set to 95.5 ºF referring to black, with warmer<br />

temperatures indicated progressively by blue (105 ºF), green (115 ºF), red<br />

(125 ºF) and white (133 ºF and hotter). We also measured the load in phase<br />

A, B and C (from left to right), at approximately 34 amps each.<br />

A simple analysis <strong>of</strong> the thermal image indicates that Phase A is significantly<br />

hotter than phases B and C. The fuse clip at the top <strong>of</strong> Phase A indicates<br />

133.4 F, while the end <strong>of</strong> the fuse, specifically the metal cap <strong>of</strong> the top <strong>of</strong> the<br />

fuse, appears much cooler with a temperature <strong>of</strong> 103.6 ºF and the fuse body<br />

just below the cap appears to be 121.9 ºF.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Figure 1: Fused power disconnect.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Figure IR1: Corresponding infrared image.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Can this be true? Is the metal cap only 103 ºF? No. You are seeing an<br />

example <strong>of</strong> the apparent temperature and the effect <strong>of</strong> emissivity. The fuse<br />

end cap is a highly reflective metal, in this case copper. Notice that the body<br />

<strong>of</strong> the fuse also appears hotter than the metal cap. The temperature <strong>of</strong> the<br />

cap is actually as hot as the fuse body it’s in contact with.<br />

To explain why the apparent temperature seen through a thermal imager can<br />

be significantly different than the actual temperature, let’s review our<br />

knowledge <strong>of</strong> physics.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Thermal radiation and properties <strong>of</strong> materials<br />

All objects emit infrared (thermal) radiation. The intensity <strong>of</strong> the radiation<br />

depends on the temperature and nature <strong>of</strong> the material’s surface. At lower<br />

temperatures, the majority <strong>of</strong> this thermal radiation is at longer wavelengths.<br />

As the object becomes hotter, the radiation intensity rapidly increases and the<br />

peak <strong>of</strong> the radiation shifts towards shorter wavelengths. The relationship<br />

between total radiation intensity (all wavelengths) and temperature is defined<br />

by the Stefan-Boltzmann Law: (broad band)<br />

Q = eσT 4<br />

where:<br />

Q = radiation intensity<br />

e = emissivity <strong>of</strong> material<br />

σ = Stefan-Boltzmann constant<br />

T = absolute temperature<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


At a given temperature, the maximum radiation is achieved when the object<br />

has an emissivity <strong>of</strong> 1. This is referred to as blackbody radiation, because<br />

with an emissivity <strong>of</strong> 1, the object is a perfect radiator. However in our real<br />

world, there are no true blackbodies – that is, no perfect radiators. Since real<br />

materials are less than perfect radiators, the relevant issue is “how much less<br />

than perfect are they?” Emissivity is defined as the measure <strong>of</strong> how much<br />

less than perfectly a material radiates when compared to a blackbody. But,<br />

emissivity is only one <strong>of</strong> three factors that cause an object to be less than a<br />

perfect radiator.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


The thermal nature <strong>of</strong> materials.<br />

Materials (objects in everyday life, whether they be solids, liquids or gases)<br />

are constantly affected by their surroundings. Thermally, all objects attempt to<br />

exchange energy with other objects in their natural drive toward thermal<br />

equilibrium with their surroundings. In this search for thermal equilibrium, heat<br />

is exchanged between objects via three mechanisms: conduction, convection<br />

and radiation.<br />

Conduction is defined as heat transfer between two solid bodies that are in<br />

physical contact with each other. Convection is heat transfer usually between<br />

a solid material and a liquid or gas. Conduction and convection are<br />

dependent on physical contact between materials. Radiation is a process <strong>of</strong><br />

heat transfer, characteristic <strong>of</strong> all matter (at temperatures above absolute<br />

zero). Radiation passes through a vacuum and can also pass through gasses,<br />

liquids and even solids.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


When radiative power is incident on an object, a fraction <strong>of</strong> the power will be<br />

reflected (ρ), another portion will be absorbed (α), and the final portion will be<br />

transmitted through the object (τ). The transmitted fraction is τ. All <strong>of</strong> this is<br />

described by the Total Power Law:<br />

ρ + α + τ = 1<br />

where:<br />

ρ = fraction reflected<br />

α = fraction absorbed<br />

τ = fraction transmitted<br />

The ability <strong>of</strong> an object to absorb radiation is also related to its ability to emit<br />

radiation. This is defined by Kirchh<strong>of</strong>f's Law<br />

α = ε<br />

where<br />

α = absorbance coefficient<br />

ε = emissive coefficient<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


So in plain English, when the thermal imager observes the thermal radiation<br />

from real objects, part <strong>of</strong> what the thermal imager sees is reflected from the<br />

surface <strong>of</strong> the object, part is emitted by the object, and part may be<br />

transmitted through the object. In our example <strong>of</strong> a steel part, the<br />

transmission is zero (opaque, τ = 0), but to the degree that the part is<br />

reflective, it is less emissive and therefore real objects will usually appear<br />

cooler than they actually are. Except when there is something hotter in the<br />

vicinity; since with opaque materials, the lower the emissivity, the higher the<br />

reflectivity. The result in this case is materials appear to be hotter than they<br />

actually are! Let’s examine some real objects to illustrate these effects.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Applying emissivity to real objects<br />

In the figure IR1 example, not only is the fuse end cap temperature actually<br />

much hotter than the 103.6 ºF that it appears, the hot spot above it is most<br />

assuredly hotter than the 133.4 ºF that it appears.<br />

So, how much hotter might it be? This fused power disconnect is electrically<br />

energized, so let’s conduct a simple experiment with a metal part that is not<br />

electrically energized. Note: While this experiment may not be shocking, it<br />

can still burn you.<br />

Picture a round stainless steel block sitting at ambient temperature.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Figure 2: Stainless steel block. (at ambient temperature)<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Observed with our thermal imager (with emissivity set to 1), the metal<br />

appears to vary in temperature from about 74 ºF to 87 ºF. This seems to<br />

make sense, since the block could have picked up a little heat from our hands<br />

during handling. Actually, the metal block is very uniform in temperature. The<br />

apparent hot spot is a reflection <strong>of</strong> my face on the surface <strong>of</strong> the metal. Can<br />

you see my eye glasses in the image? (Figure IR2)<br />

Figure IR2: Thermal image <strong>of</strong> stainless steel block.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Observed with our thermal imager (with emissivity set to 1), the metal<br />

appears to vary in temperature from about 74 ºF to 87 ºF. This seems to<br />

make sense, since the block could have picked up a little heat from our hands<br />

during handling. Actually, the metal block is very uniform in temperature. The<br />

apparent hot spot is a reflection <strong>of</strong> my face on the surface <strong>of</strong> the metal. Can<br />

you see my eye glasses in the image? (Figure IR2)<br />

Figure IR2: Thermal image <strong>of</strong> stainless steel block.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Can you see my eye glasses in the image? (Figure IR2)<br />

Charlie Chong/ Fion Zhang


Can you see my eye glasses in the image?<br />

Charlie Chong/ Fion Zhang<br />

https://en.wikipedia.org/wiki/Douglas_MacArthur


The block appears to vary in temperature from about 92 ºF to 110 ºF – and<br />

you can see the image <strong>of</strong> my face in the warm metal surface even more<br />

clearly than before. Using a thermocouple, we measure the surface<br />

temperature and find that it’s actually 169 ºF (see Figure 2a).<br />

Figure 2a: DMM with thermocouple, measuring surface temperature <strong>of</strong> the<br />

steel block.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


How can the thermal imager’s readings appear reasonable when the metal<br />

part is at room temperature and be so wrong (still producing a mirror image <strong>of</strong><br />

my face on the hot surface) when the part is 169 ºF?<br />

At room temperature, the block appears to be room temperature because the<br />

block is primarily reflecting the thermal radiation from everything around it.<br />

Since the ambient temperature in the room is in the 70s, the reflection from<br />

the surface <strong>of</strong> the block appears also to be similar. When the same part is<br />

heated in the oven, the part becomes much hotter than the surroundings, so<br />

the thermal imager is able to see an increase in radiant energy, albeit 尽 然<br />

much lower in apparent temperature because <strong>of</strong> the low emissivity value <strong>of</strong><br />

the surface.<br />

Let’s modify our experiment to better demonstrate what the thermal imager<br />

sees. We take another stainless steel block and paint half <strong>of</strong> it with a flat black<br />

paint (flat black paint has an emissivity <strong>of</strong> 1 or 0.98 to be a little precise) and<br />

bake it (in a slightly warmer oven) another three hours (Figure 3, Figure IR3).<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Figure 3: Steel block, left side painted black.<br />

Figure IR3: Corresponding thermal image <strong>of</strong> steel block.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


When we remove the block from the oven this time, the unpainted side<br />

appears to be 92 ºF, but the thermal imager now indicates the painted sided<br />

to be 198 ºF. We can make a very good estimation <strong>of</strong> the actual emissivity <strong>of</strong><br />

this material by observing the unpainted surface with our IR camera and<br />

adjusting the emissivity value on the thermal imager until the reading matches<br />

the temperature observed on the painted side. In this case, the emissivity is<br />

found to be approximately 0.12.<br />

Assumed<br />

ε = 0.98<br />

Adjusted<br />

ε = 0.12<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Emissivity is a cantankerous 脾 气 坏 且 抱 怨 不 休 的 variable<br />

As we’ve seen, emissivity varies by surface condition, but also by viewing<br />

angle, and even by temperature and by spectral wavelength. A table <strong>of</strong><br />

common emissivity values is published in the operating manual for your<br />

thermal imager. The table should be considered only a rough guide in<br />

estimating an emissivity value to use with any particular material. If actual<br />

temperature values are required, it is best to perform experiments as<br />

described here, to properly characterize the emissivity for the material and its<br />

application.<br />

The two most common techniques for providing a higher emissivity reference<br />

surface are the application <strong>of</strong> a flat black high emissivity paint to the surface<br />

(as discussed in the previous section), or application <strong>of</strong> common black<br />

electrical tape to the material’s surface. Both black electrical tape and flat<br />

black tape have an emissivity <strong>of</strong> approximately 0.96. Another option is to use<br />

an infrared thermometer with adjustable emissivity, and a contact probe,<br />

adjusting the emissivity until the contact probe and infrared temperature<br />

displays equilibrate.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


In this experiment we see that the difference between the apparent<br />

temperature on the unpainted side and actual temperature is an error <strong>of</strong> 106<br />

ºF. If we were to conduct a similar experiment with a high-temperature<br />

infrared sensor, and examine steel at 2,000 ºF, the error between the actual<br />

and apparent temperatures could be more than 400 ºF. Of course, neither<br />

black paint or tape could survive 2,000 ºF. It’s <strong>of</strong>ten useful to use a narrow<br />

spectral band similar to the wavelength <strong>of</strong> the object’s radiant energy.<br />

Wien’s displacement law helps us determine the peak wavelength <strong>of</strong> the<br />

object’s peak radiant energy for an object at a certain temperature.<br />

λ max = b / T<br />

where:<br />

λ max = peak wavelength <strong>of</strong> radiant energy<br />

b = 2897 μm/ °K<br />

T = temperature (Kelvin)<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Wien’s Displacement Law<br />

Charlie Chong/ Fion Zhang<br />

http://www.sun.org/encyclopedia/electromagnetic-spectrum


Charlie Chong/ Fion Zhang<br />

http://www.sun.org/encyclopedia/electromagnetic-spectrum


When you are working with high-temperature materials, you can greatly<br />

reduce the errors due to uncertainty in emissivity by selecting infrared<br />

detectors that operate at narrow wavelength bands at shorter wavelengths.<br />

The math and physics necessary to prove this is beyond the scope <strong>of</strong> this<br />

application note. However, calculations demonstrate that by choosing an<br />

infrared sensor with a wavelength band close to 1 μm (rather than the 8 μm<br />

to 14 μm spectral band used by most thermal imagers), the maximum<br />

difference between the 2,000 ºF actual and apparent temperatures would be<br />

closer to 50 degrees (without knowing the precise emissivity <strong>of</strong> the material<br />

with better certainty).<br />

(the reflected low temperature spectrum is filter out leaving the high energy<br />

narrow wavelength representing the high temperature λ max )<br />

To summarize: Temperature measurement without knowledge in this case<br />

would result in an error <strong>of</strong> more than 400 ºF. Making the same measurement<br />

with knowledge would reduce the error to 50 ºF, with no better determination<br />

<strong>of</strong> the material’s emissivity.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Using a narrow band filter<br />

to measure this<br />

temperature range<br />

The ambient reflection<br />

ρ is filter out<br />

Charlie Chong/ Fion Zhang<br />

http://www.sun.org/encyclopedia/electromagnetic-spectrum


Discussion<br />

Subject: Are the following statement true?<br />

• Is the absolute emittance (≠ emissivity, ε) <strong>of</strong> an object constant with<br />

disregard <strong>of</strong> reflectance ρ?<br />

• Is the 1= ε + ρ + τ, a weighted ratio <strong>of</strong> three contributing factor meant for<br />

IR thermographic measurement purpose and not a physical property <strong>of</strong> the<br />

object?<br />

• What ever the reflectance be (by surface conditioning, texture, shielding,<br />

by raising the T amb to T obj etc.) , the emittance from the object is always the<br />

same as long the T obj remain the same?<br />

• Could be say that will reflectance, transmittance coupled with the object<br />

emissivity, the actual power radiating from the object is higher than the<br />

black body?<br />

Charlie Chong/ Fion Zhang


Emissivity, the variable’s variable!<br />

Back to our steel block example, let’s discuss another very significant<br />

phenomena. We will take our unpainted metal block and drill three holes part<br />

way into the body. All three holes are one-eighth <strong>of</strong> an inch in diameter. The<br />

first is one-eighth-inch deep, the second is one-fourth-inch deep, and the third<br />

is three-eighths-inch deep.<br />

Figure IR4: Thermal image <strong>of</strong> steel block with three holes.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Bake the block for another three hours, then remove the block and observe it<br />

again with the camera. Interestingly, the hot block surface appears to be<br />

about 84 F, and now appears to have three hot spots.<br />

■<br />

■<br />

■<br />

The one-eighth-inch deep hole appears to be 106 ºF.<br />

The one-fourth-inch deep hole appears to be 112 ºF; and<br />

the three-eighth-inch deep hole appears to be 125 ºF.<br />

We know that the metal block is actually about 175 ºF (measured by a<br />

thermocouple) and the surface finish is uniform and has an emissivity <strong>of</strong><br />

approximately 0.12. The reason the temperature appears to be higher in the<br />

holes is that a hole in a body enhances the emissivity. The greater the<br />

depth/diameter ratio <strong>of</strong> the hole, the greater the emissivity enhancement. By<br />

adjusting the emissivity on the thermal imager to match the actual<br />

temperature at each hole, we find that the emissivity appears to be 0.25 for<br />

the one-eighth-inch deep hole. The emissivity <strong>of</strong> the one-fourth-inch deep<br />

hole appears to be 0.35 and the three-eighth-inch deep hole appears to have<br />

an emissivity <strong>of</strong> 0.45. This is an extremely important effect. Let’s look at<br />

another piece <strong>of</strong> electrical equipment to see why.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Emissivity and electrical equipment<br />

In Figures 5 and IR5, you see another power disconnect with the conductors<br />

bolted in place using Allen head bolts. The corresponding infrared image<br />

shows a hot connection on the middle phase.<br />

Figure 5: 3-phase power disconnect.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Figure IR5: Corresponding thermal image.<br />

Notice the apparent hot spot in the hot Allen socket head. The well <strong>of</strong> the bolt<br />

head appears hotter primarily because the well illustrates the blackbody effect<br />

<strong>of</strong> a hole.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


In manufacturing processes, steel or aluminum rolls are <strong>of</strong>ten used to heat or<br />

cool a material such as in paper or plastic film processing. These rolls are<br />

usually polished metal surfaces, and it’s important to understand the thermal<br />

pr<strong>of</strong>ile since the manufacturing process depends on thermal uniformity across<br />

the rolls. The temperature <strong>of</strong> these rolls can be difficult to measure with a<br />

thermal imager because they have very low emissivities. However, there are<br />

<strong>of</strong>ten points where the material passes between two rolls. The tangent point<br />

between two rolls also tends to simulate the blackbody effect, allowing for<br />

effective temperature measurement in an otherwise difficult situation.<br />

This effect is illustrated in common electrical equipment as well. Look at<br />

Figure 6.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Figure 6: Power<br />

disconnect with knife<br />

blade connectors.<br />

Figure IR6:<br />

Corresponding thermal<br />

image.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


In this case, we have another power disconnect with knife blade switches.<br />

This type <strong>of</strong> switch utilizes shiny metal blades, and the proximity <strong>of</strong> the blades<br />

with narrow gaps simulates the blackbody effect for greatly improved effective<br />

emissivity. The important message here is to develop your understanding <strong>of</strong><br />

apparent and actual temperature measurement. Actual temperature<br />

measurement requires an intimate understanding <strong>of</strong> physics, heat transfer<br />

and characteristics <strong>of</strong> materials.<br />

Aimed here<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Qualitative vs. quantitative infrared thermography<br />

Emissivity difficulties are not a barrier to effectively using infrared<br />

thermography for predictive maintenance (PdM). ASTM standards exist to<br />

guide thermographic PdM inspections. These standards describe the use <strong>of</strong><br />

thermal imagers for qualitative and quantitative infrared inspections.<br />

Quantitative infrared inspections require determining the emissivity <strong>of</strong> each<br />

component, to make accurate temperature measurements possible. This<br />

practice may not always be necessary for routine inspections, unless the<br />

exact temperature value is needed for long term tracing. Qualitative methods,<br />

in contrast, allow you to leave the emissivity at 1.0 and evaluate the<br />

equipment on a relative basis: Has it changed, or is it different? The basis for<br />

qualitative evaluation is comparing similar equipment under similar loads.<br />

Looking back at Figure 1 and IR1, you can see that there is little value to be<br />

gained in spending time estimating or debating the emissivity <strong>of</strong> the various<br />

parts in the power disconnect. The value is in understanding that Phase A is<br />

hotter than phase B and C. In addition to realizing that a phase is hotter, it is<br />

essential to measure the load <strong>of</strong> the three phases.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Figure IR1: Corresponding infrared image.<br />

there is little value to be gained in<br />

spending time estimating or debating the<br />

emissivity <strong>of</strong> the various parts in the<br />

power disconnect. The value is in<br />

understanding that Phase A is hotter than<br />

phase B and C.<br />

A B C<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Greater electrical load inherently means more heat is present<br />

W = I 2 R<br />

where<br />

W = power in watts (heat)<br />

I = current in amps<br />

R = resistance in ohms<br />

The first rule <strong>of</strong> thermography in predictive maintenance PDM is to compare<br />

comparable equipment under comparable loads. In electrical power<br />

distribution, comparable equipment is usually the easy part since each<br />

electrical phase is usually similar in materials to the phase next to it. Load is a<br />

very different matter. Figure 7 illustrates an electrician measuring the<br />

electrical load.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Figure 7: Measuring the loads on a power disconnect.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


So, just observing that there is a hot spot does not indicate a problem.<br />

Electrical components can be appropriately hot for the electrical load and<br />

conditions. If you measure the loads, you can determine if the presence <strong>of</strong> a<br />

thermal anomaly indicates a problem. Thermal imagers do not identify<br />

thermal problems – trained, knowledgeable, qualified people make educated<br />

assessments <strong>of</strong> equipment. This leads to real value in preventive<br />

maintenance and reduced frequency <strong>of</strong> equipment breakdowns.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


Summary<br />

Predictive maintenance PDM with a thermal imager can be effectively<br />

performed by utilizing qualitative analysis <strong>of</strong> equipment.<br />

Qualitative techniques allow the emissivity setting on the thermal imager to be<br />

kept at 1.0 and apparent temperatures used for comparisons between similar<br />

equipment under similar load. With basic training, most technicians can<br />

reliably perform qualitative analysis.<br />

Quantitative infrared analysis requires a deeper understanding <strong>of</strong> thermal<br />

theory and application to be truly effective. It refers to the attempt to measure<br />

actual temperatures <strong>of</strong> materials using infrared thermography. Actual<br />

temperature measurement involves more than simply adjusting for emissivity.<br />

Total incident radiance requires dealing with the effect <strong>of</strong> reflection and<br />

transmission in addition to emissivity.<br />

Today’s thermal imagers are becoming increasingly affordable and easy to<br />

use. But, what does easy mean? The practice <strong>of</strong> infrared thermography looks<br />

straight forward and simple; but it has its tricks. It is much like most<br />

endeavors in life: the more you learn, the more you discover that there is<br />

more to learn.<br />

Charlie Chong/ Fion Zhang<br />

http://reliableplant.com/Read/14134/emissivity-underst-difference-between-apparent,-actual-ir-temps


It is much like most endeavors in life: the more you learn, the more you<br />

discover that there is more to learn.<br />

Charlie Chong/ Fion Zhang


It is much like most endeavors in life: the more you learn, the more you<br />

discover that there is more to learn.<br />

Charlie Chong/ Fion Zhang


Good Luck<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!