Sie sind auf Seite 1von 5

22/8/2017 Thermogravimetric analysis - Wikipedia

Thermogravimetric analysis
From Wikipedia, the free encyclopedia

Thermogravimetric analysis or thermal gravimetric analysis


(TGA) is a method of thermal analysis in which changes in physical Thermogravimetric analysis
and chemical properties of materials are measured as a function of Acronym TGA
increasing temperature (with constant heating rate), or as a function Classification
of time (with constant temperature and/or constant mass loss).[1] Thermal analysis
TGA can provide information about physical phenomena, such as
second-order phase transitions, including vaporization, sublimation,
absorption and desorption. Likewise, TGA can provide information
about chemical phenomena including chemisorptions, desolvation
(especially dehydration), decomposition, and solid-gas reactions
(e.g., oxidation or reduction).[1]

TGA is commonly used to determine selected characteristics of


materials that exhibit either mass loss or gain due to decomposition,
oxidation, or loss of volatiles (such as moisture). Common
applications of TGA are (1) materials characterization through
analysis of characteristic decomposition patterns, (2) studies of
degradation mechanisms and reaction kinetics, (3) determination of
organic content in a sample, and (4) determination of inorganic (e.g.
ash) content in a sample, which may be useful for corroborating
predicted material structures or simply used as a chemical analysis. A typical TGA system
It is an especially useful technique for the study of polymeric
materials, including thermoplastics, thermosets, elastomers, Other techniques
composites, plastic films, fibers, coatings and paints. Discussion of Related Isothermal microcalorimetry
the TGA apparatus, methods, and trace analysis will be elaborated Differential scanning calorimetry
upon below. Thermal stability, oxidation, and combustion, all of Dynamic mechanical analysis
which are possible interpretations of TGA traces, will also be Thermomechanical analysis
discussed.
Differential thermal analysis
Dielectric thermal analysis

Contents
1 Instrumental apparatus
2 Methods
3 Trace analysis
4 Ceramic yield
5 Thermal stability
6 Oxidation processes
7 Combustion
8 References

Instrumental apparatus
Thermogravimetric analysis (TGA) relies on a high degree of precision in three measurements: mass change,
temperature, and temperature change. Therefore, the basic instrumental requirements for TGA are a precision balance
with a pan loaded with the sample, and a programmable furnace. The furnace can be programmed either for a constant
heating rate, or for heating to acquire a constant mass loss with time.

Though a constant heating rate is more common, a constant mass loss rate can illuminate specific reaction kinetics. For
example, the kinetic parameters of the carbonization of polyvinyl butyral were found using a constant mass loss rate of
0.2 wt %/min.[2] Regardless of the furnace programming, the sample is placed in a small, electrically heated furnace
equipped with a thermocouple to monitor accurate measurements of the temperature by comparing its voltage output
with that of the voltage-versus-temperature table stored in the computers memory.[3] A reference sample may be
placed on another balance in a separate chamber. [4] The atmosphere in the sample chamber may be purged with an
https://en.wikipedia.org/wiki/Thermogravimetric_analysis 1/5
22/8/2017 Thermogravimetric analysis - Wikipedia

placed on another balance in a separate chamber.[4]


The atmosphere in the sample chamber may be purged with an
inert gas to prevent oxidation or other undesired reactions. A different process using a quartz crystal microbalance has
been devised for measuring smaller samples on the order of a microgram (versus milligram with conventional TGA).

Methods
The TGA instrument continuously weighs a sample as it is heated to temperatures of up to 2000 C for coupling with
FTIR and Mass spectrometry gas analysis. As the temperature increases, various components of the sample are
decomposed and the weight percentage of each resulting mass change can be measured. Results are plotted with
temperature on the X-axis and mass loss on the Y-axis. The data can be adjusted using curve smoothing and first
derivatives are often also plotted to determine points of inflection for more in-depth interpretations.

Trace analysis
If the identity of the product after heating is known, then the ceramic yield can be found from analysis of the ash
content (see discussion below). By taking the weight of the known product and dividing it by the initial mass of the
starting material, the mass percentage of all inclusions can be found. Knowing the mass of the starting material and the
total mass of inclusions, such as ligands, structural defects, or side-products of reaction, which are liberated upon
heating, the stoichiometric ratio can be used to calculate the percent mass of the substance in a sample. The results
from thermogravimetric analysis may be presented by (1) mass versus temperature (or time) curve, referred to as the
thermogravimetric curve, or (2) rate of mass loss versus temperature curve, referred to as the differential
thermogravimetric curve. Though this is by no means an exhaustive list, simple thermogravimetric curves may contain
the following features:

A horizontal portion, or plateau that indicates constant sample weight


A curved portion; the steepness of the curve indicates the rate of mass loss
An inflection (at which is a minimum, but not zero)

Certain features in the TGA curve that are not readily seen can be more clearly discerned in the first derivative TGA
curve. For example, any change in the rate of weight loss can immediately be seen in the first derivative TGA curve as
a trough, or as a shoulder or tail to the peak, indicating two consecutive or overlapping reactions. Differential TGA
curves also can show considerable similarity to differential thermal analysis (DTA) curves, which can permit easy
comparisons to be made.[1]

Ceramic yield
Ceramic yield is defined as the mass percent of starting material found in the end product. From this, stoichiometry can
then be used to calculate the percent mass of the substance in the sample.

Metal aluminates (MAl2O4) are an important type of mixed-cation oxide ceramics that have many applications.[5] The
metal aluminate CaAl2O4 is used in the cement industry as a hydraulic material.[5] Its precursor is
CaAl2C18H37O9N3.[5] The formation of CaAl2O4 occurs during the thermogravimetric analysis. This is how the
theoretical ceramic yield is calculated for this example:

1. Calculate molecular weight of CaAl2O4:


2. Calculate molecular weight of CaAl2C18H37O9N3:

3. Calculate the percentage that CaAl2O4 is of CaAl2C18H37O9N3:

Therefore, the theoretical ceramic yield for the thermogravimetric analysis of CaAl2C18H37O9N3 is 29.6%. This
correlates well with the experimentally determined ceramic yield of 28.9%.

https://en.wikipedia.org/wiki/Thermogravimetric_analysis 2/5
22/8/2017 Thermogravimetric analysis - Wikipedia

As another example of calculating theoretical ceramic yield, take the TGA of calcium oxalate monohydrate. Using the
same process detailed above, the theoretical ceramic yield can be calculated: the formula weight of calcium oxalate
monohydrate is 146 g/mol. The final ceramic product is CaO, with a formula weight of 56 g/mol. The theoretical
ceramic yield is therefore 38.4%. The actual yield from the TGA was found to be 39.75%. Some reasons for
discrepancies between the theoretical and actual yields are trapped CO2 and the formation of metal carbides.

In the TGA trace of calcium oxalate monohydrate, the first mass loss corresponds to loss of water of hydration. The
second mass loss corresponds to decomposition of dehydrated calcium oxalate to calcium carbonate and carbon
monoxide and carbon dioxide. The last mass loss is due to the decomposition of calcium carbonate to calcium oxide
and carbon dioxide.

The differences between thermograms can be seen in the example of four different chloro-polymers: (a) polyvinyl
chloride, (b) chlorinated polyvinyl chloride, (c) chlorinated rubber, and (d) polyvinylidene chloride.[6] There are two
stages of degradation in these four polymers. The first stage is the loss of hydrogen chloride, and is complete around
250 C. This first step occurs at lower temperatures for the polymers containing more chlorine (chlorinated polyvinyl
chloride, chlorinated rubber, and polyvinylidene chloride), implying that these chloride groupings are less stable than
in polyvinyl chloride.[6]

The second stage is the carbonization of the polymer, and takes place between 250 C and 500 C. This is seen by the
large loss of mass between 250 C and 500 C. Tar and simple gases, such as hydrogen and methane, are evolved and
the carbon that remains loses very little mass between 500 C and 900 C. In this second stage, the higher the chlorine
content of the polymer, the lower the yield of tar. This is because chlorine is able to remove hydrogen, which would
otherwise be used in the compounds that form tar.[6]

Thermal stability
TGA can be used to evaluate the thermal stability of a material. In a desired temperature range, if a species is thermally
stable, there will be no observed mass change. Negligible mass loss corresponds to little or no slope in the TGA trace.
TGA also gives the upper use temperature of a material. Beyond this temperature the material will begin to degrade.

TGA has a wide variety of applications, including analysis of ceramics and thermally stable polymers. Ceramics
usually melt before they decompose as they are thermally stable over a large temperature range, thus TGA is mainly
used to investigate the thermal stability of polymers. Most polymers melt or degrade before 200 C. However, there is
a class of thermally stable polymers that are able to withstand temperatures of at least 300 C in air and 500 C in inert
gases without structural changes or strength loss, which can be analyzed by TGA.[7] For example, the polyimide
Kapton loses less than 10% mass when held in 400 C air for 100 hours.[7]

High performance fibers can be compared using TGA as an evaluation of thermal stability. From the TGA,
polyoxazole (PBO) has the highest thermal stability of the four fibers as it is stable up to ca. 500 C. Ultra-high-
molecular-weight polyethylene (UHMW-PE) has the lowest thermal stability, as it begins to degrade around 200 C.
Often the onset of mass loss is seen more prominently in the first derivative of the mass loss curve. High performance
fibers used in bulletproof vests must remain strong enough mechanically so as to protect the user from incoming
projectiles. The thermal and photochemical degradation of the fibers causes the mechanical properties of the vests to
decrease, effectively rendering the armor useless. Thus, thermal stability is a key property when designing these
vests.[8]

Three ways a material can lose mass during heating are through chemical reactions, the release of adsorbed species,
and decomposition. All of these indicate that the material is no longer thermally stable. Out of the four fibers shown in
the previous example, only Terlon shows loss of adsorbed species, most likely water, as the mass loss occurs after
100 C. Because the TGA is performed in air, oxygen reacts with the organic fibers which eventually degrade
completely, evidenced by the 100% mass loss. It is important to link thermal stability to the gas in which the TGA is
performed. PBO, which completely decomposes when heated in air, retains ~60% mass when heated in N2.[9] Thus,
PBO is thermally stable in nitrogen up to 630 C, whereas in air, PBO has almost completely decomposed at that
temperature.

Oxidation processes
Oxidative mass losses are the most common observable losses in TGA.[10]
https://en.wikipedia.org/wiki/Thermogravimetric_analysis 3/5
22/8/2017 Thermogravimetric analysis - Wikipedia

Studying the resistance to oxidation in copper alloys is very important. For example, NASA (National Aeronautics and
Space Administration) is conducting research on advanced copper alloys for their possible use in combustion engines.
However, oxidative degradation can occur in these alloys as copper oxides form in atmospheres that are rich in oxygen.
Resistance to oxidation is very important because NASA wants to be able to reuse shuttle materials. TGA can be used
to study the static oxidation of materials such as these for practical use.

Some researchers have been studying ways in which to protect certain oligomers or polymers from oxidation
processes. One example is inserting an oligomer into a multiblock copolymer.[11] An example is the TGA traces of
both the oligomer and the oligomer/multiblock copolymer in N2 and in air.[11] When the TGAs were run under a
nitrogen atmosphere, there is no oxidation of the substrate. When the TGA of the oligomer was run under air, an
oxidation process can be seen between 200 C-350 C. This process is not seen for the oligomer/multiblock copolymer.
The authors of this paper explained this disappearance by suggesting that the oxidative process involved hydroxyl end
groups on the oligomer. The encasing of the oligomer by the multiblock copolymer prevented this from happening.[11]

Combustion
Combustion during TG analysis is identifiable by distinct traces made in the TGA thermograms produced. One
interesting example occurs with samples of as-produced unpurified carbon nanotubes that have a large amount of metal
catalyst present. Due to combustion, a TGA trace can deviate from the normal form of a well-behaved function. This
phenomenon arises from a rapid temperature change. When the weight and temperature are plotted versus time, a
dramatic slope change in the first derivative plot is concurrent with the mass loss of the sample and the sudden increase
in temperature seen by the thermocouple. The mass loss could be the result of particles of smoke released from burning
caused by inconsistencies in the material itself, beyond the oxidation of carbon due to poorly controlled weight loss.

References

1. Coats, A. W.; Redfern, J. P. (1963). "Thermogravimetric Analysis: A Review". Analyst. 88 (1053): 906924.
Bibcode:1963Ana....88..906C (http://adsabs.harvard.edu/abs/1963Ana....88..906C). doi:10.1039/AN9638800906
(https://doi.org/10.1039%2FAN9638800906).
2. Tikhonov, N. A.; Arkhangelsky, I. V.; Belyaev, S. S.; Matveev, A. T. (2009). "Carbonization of polymeric
nonwoven materials". Thermochimica Acta. 486: 6670. doi:10.1016/j.tca.2008.12.020 (https://doi.org/10.101
6%2Fj.tca.2008.12.020).
3. "Thermogravimetric Analysis" (http://web.archive.org/web/20120610045047/http://vedyadhara.ignou.ac.in:80/w
iki/images/1/1c/UNIT_10_THERMOGRAVIMETRIC_ANALYSIS.pdf) (PDF). Archived from the original (htt
p://vedyadhara.ignou.ac.in/wiki/images/1/1c/UNIT_10_THERMOGRAVIMETRIC_ANALYSIS.pdf) (PDF) on
2012-06-10. Retrieved 2017-07-15.
4. "Apollo thermogravimetric analyzer TGA" (https://www.impautomation.com/Analytical-instruments/Thermogra
vimetric-analysers/Apollo-thermogravimetric-analyzer.html). www.impautomation.com. Retrieved 2017-07-15.
5. Narayanan, R.; Laine, R. M. (1997). "Synthesis and Characterization of Precursors for Group II Metal
Aluminates". Appl. Organomet. Chem. 11: 919927. doi:10.1002/(SICI)1099-
0739(199710/11)11:10/11<919::AID-AOC666>3.0.CO;2-Z (https://doi.org/10.1002%2F%28SICI%291099-073
9%28199710%2F11%2911%3A10%2F11%3C919%3A%3AAID-AOC666%3E3.0.CO%3B2-Z).
6. Gilbert, J. B.; Kipling, J. J.; McEnaney, B.; Sherwood, J. N. (1962). "Carbonization of Polymers I -
Thermogravimetric Analysis". Polymer. 3: 110.
7. Marvel, C. S. (1972). "Synthesis of Thermally Stable Polymers". Ft. Belvoir: Defense Technical Information
Center.
8. Liu, X.; Yu, W. (2006). "Evaluating the Thermal Stability of High Performance Fibers by TGA". Journal of
Applied Polymer Science. 99: 937944. doi:10.1002/app.22305 (https://doi.org/10.1002%2Fapp.22305).
9. Tao, Z.; Jin, J.; Yang, S.; Hu, D.; Li, G.; Jiang, J. (2009). "Synthesis and Characterization of Fluorinated PBO
with High Thermal Stability and Low Dielectric Constant". Journal of Macromolecular Science, Part B. 48:
11141124. doi:10.1080/00222340903041244 (https://doi.org/10.1080%2F00222340903041244).
10. Voitovich, V. B.; Lavrenko, V. A.; Voitovich, R. F.; Golovko, E. I. (1994). "The Effect of Purity on High-
Temperature Oxidation of Zirconium". Oxidation of Metals. 42: 223237. doi:10.1007/BF01052024 (https://doi.
org/10.1007%2FBF01052024).
11. D'Antone, S.; Bignotti, F.; Sartore, L.; DAmore, A.; Spagnoli, G.; Penco, M. (2001). "Thermogravimetric
investigation of two classes of block copolymers based on poly(lactic-glycolic acid) and poly(-caprolactone) or
poly(ethylene glycol)". Polymer Degradation and Stability. 74: 119124. doi:10.1016/S0141-3910(01)00110-0
(https://doi.org/10.1016%2FS0141-3910%2801%2900110-0).
https://en.wikipedia.org/wiki/Thermogravimetric_analysis 4/5
22/8/2017 Thermogravimetric analysis - Wikipedia

Retrieved from "https://en.wikipedia.org/w/index.php?title=Thermogravimetric_analysis&oldid=794420534"

This page was last edited on 7 August 2017, at 22:42.


Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By
using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia is a registered trademark of the
Wikimedia Foundation, Inc., a non-profit organization.

https://en.wikipedia.org/wiki/Thermogravimetric_analysis 5/5

Das könnte Ihnen auch gefallen