Sunteți pe pagina 1din 9

TIP FROM TECHNOLOGY

UV Effect on Polyethylene

UV light characteristics
All polyethylene (PE) is susceptible to degradation upon long-term exposure to sunlight. This degradation is brought about by
physical changes, which occur in the polyethylene as a result of exposure to the ultraviolet (UV) portion of sunlight. Figure 1
shows that UV light contains shorter wavelengths than visible light. The shorter the wavelength, the more energy it contains
and thus, the more damage it does. This is explained in the next section in more detail. Fluorescent lighting also contains a
band of UV light, but only at an intensity of around 15% of normal sunlight.

Figure 1: UV light and its position in the Electromagnetic spectrum

Visible

X-rays ultra- heat


gamma rays violet infrared wireless

extreme micro- radio


ultraviolet waves broadcast

60 50 40 30 20 10 0 octaves
22 20 18 16 14 12 10 8 6
10 10 10 10 10 10 10 10 10 vib/s

10-14 10-12 10-10 10-8 10-6 10-4 10-2 1 102 104 m


2 3 4 5 6 7 8
Wavelengths 1 10 10 10 10 10 10 10 10 Angstroms

Why does PE change when exposed to UV light?


Table 1 shows why UV light alters the physical characteristics of polyethylene (PE). It does this by breaking the carbon and
hydrogen bonds, creating free radicals, which, in turn, break the PE into shorter molecules and thus, a more brittle polymer.
Effectively, UV light creates a higher melt index polyethylene, especially on the exposed surface area. This shows up as a
reduction in break elongation and impact properties, typical of higher melt index PE. The subsequent attachment of oxygen
to these broken sites leads to further accelerated degradation and the formation of oxidized species such as carbonyl and car-
boxyl structures, which are often used as analytical indicators of UV degradation. Table 1 shows the importance of choosing
the correct wavelengths for exposure testing that realistically simulates the weathering effects on PE.

Table 1: Why UV light changes PE


Energy contained
Wavelength of at this wavelength Type of bond or Energy to break
light, nm KJ/m2 structure in PE bond, KJ/mol.

189 647 Carbon-carbon (C-C) 347


253 473 Carbon-hydrogen (C-H) 413
315 228 Double bond 607
carbon-carbon (C=C)

1
The shorter the wavelength, the more energy and thus, destructive power it has. Actual UV light is composed of a range of
wavelengths as shown in Table 2. Thus, it is important when doing indoor, accelerated weathering exposure to use a lamp or
light source that best matches natural sunlight.

Table 2: UV light has a range of wavelengths

Type of UV light Wavelength, nanometer (nm)

UV-A 320-380
UV-B 280-320
UV-C 180-280
Natural Sunlight 280-800+

Test protocol for UV exposure


We use both indoor, accelerated weathering testing as well as outdoor exposure testing in Florida. Indoor, accelerated testing
can be done in shorter time periods as the exposure is continuous and irradiation more intense.

The instruments we use for accelerated testing are Xenon Ci-65 and Ci-5000 Weather-O-Meters. Test conditions follow the
guidelines outlined in ASTM G155 cycle 1*. It uses a xenon arc lamp with a borosilicate inner and outer filter, which best
simulates the UV band of natural sunlight. Carbon arc lights are unrealistic as they produce too much of the wavelengths
between 325 to 425 nm. QUV-B testers also are not as good for non-crosslinked polyethylene as they reduce the higher
wavelengths of light.

UV light intensity is measured by irradiance and is usually expressed in watts per square meter at a given wavelength. The
weather-o-meters are run at 0.35 watts/m2 and 340 nanometers (nm), an industry accepted protocol, which matches one of
the conditions in Florida, as illustrated in Figure 2.

Figure 2: Irradiance level vs. Miami, Florida conditions

1.6
Calculated and measured sunlight
+ calculated summer noon
1.4 measured summer noon in Miami. Summer solstice and
“average” annual sunlight are shown.
1.2
Irradiance (W/m2/nm)

x calculated annual average


measured "average optimal"
1.0

0.8

0.6

0.4

0.2

0.0
290 300 310 320 330 340 350 360 370 380 390 400
Wavelength (nm)

Source: J.E. Pickett, K.K. Webb, GE Electric Co., Die Angewandte Makromoldekulare Chemie 252, (1997) 217-236 (4524)

*ASTM test procedures may be modified to accommodate operating conditions or facility limitations. 2
The temperature of exposure is also important, as the higher the temperature, the more degradation will occur. A black panel
in the machine measures this temperature and is set at 63°C. The backing on the sample may change the actual sample
temperature. For this reason, our samples are unbacked both in the Ci-65 and in Florida.

Figure 3 shows a correlation of Xenon testing Figure 3: Correlation between Xenon Ci-65 exposure and Florida
with Florida exposure. It shows that 2000
hours in the weather-o-meter is approxi- 10,000
mately 140 kilolangleys (kly) per year, which
is typical of one year exposure in Miami,

T50 (h) Weather-O-Meter


Florida. A kilolangley is a measure of UV 8000
exposure per square meter of surface area.
Figure 3 was based on PE film samples. We
have found that using thicker molded sam- 6000
ples and running the weather-o-meter on a
wet and dry cycle still gives a correlation of
4000
approximately 2000 hours, being one year
in Miami. Most of our Florida data is done
at a 45 degree south exposure angle. We Comparison Weather-O-Meter/Florida exposure
2000 200 µm LDPE blown films, started March 1979
have over 62 data points done over 15 years T50 (E50): Time (energy) to 50% retained elongation
showing good agreement between our accel-
erated and outdoor test results in developing
this correlation. 200 400 600

E50 plexiglass backing. kly Florida

Source: Dr. Gugumus, Ciba Specialty Chemicals, Polymer Stabilization and Degradation Symposium, Manchester, September 18-20, 1985

UV performance depends on geographic location


Geographic location, along with changes in climate and elevation, affect actual UV performance. Figure 4 shows an updated
chart outlining isolines of equal global UV radiation. This figure has been updated to reflect the increased severity in higher
elevation areas. In fact, for every 0.5 km increase in elevation, UV exposure increases by 3.5%. This chart also reflects changes
in ozone layers in the world. While well publicized, the changes have been relatively minor when considering exposure to
polymers. Sunshine variations remain the key variable when correlating accelerated and outdoor exposure data.

Another question to consider is the UV performance over water relative to land. While definitive studies on PE are not
available to date, the cooler temperature and generally increased cloudiness near coastal areas tend to offset the increased
reflectance off the water surface.

Below is an example on how to use Figure 4 to relate years of exposure to weather-o-meter indoor exposure, knowing the
UV rating of the resin.

Years = (325/your location’s isoline) x UV rating of resin


Example: Resin with UV-8 rating (See definition in next section) in Southern Florida.
Southern Florida = 660 KJ/cm2 per year (from Figure 4)
Years of exposure = (325/660) x 8 = 3.9 years

3
Figure 4: Isolines of UV performance

Annual Global Radiation


units: KJ/cm2 per year
(to get MJ/m2, multiply by 10)

Reference: B. de Jong, Net radiation received by a horizontal surface at the earth, Delft University Press, 1973.

4
Broad Band UV Exposure (295-385 nm)
How are our resins rated for UV?
An often-used industry standard is UV-X, where X is a multiple of 1000 hours of weather-o-meter exposure. For instance,
UV-8 means 8000 hours of weather-o-meter exposure. UV-2 is 2000 hours of exposure. At this point, 50% of the original
break elongation is left. The higher the number, the longer the UV exposure and the better the polymer UV protection.
Figure 5 is a representation of this data.

Some suppliers use qualitative ratings, such as long-term or short-term UV protection, but this tends to be vague. Generally,
most of the industry would currently think of short-term as UV-4 with UV-8 being long-term. Refer to our data sheets to
determine the UV performance of each of our grades.

Figure 5: What does UV-8* mean?

UV Stability
100
2000 hrs. is approximately 1 year Florida
1250 hrs. is approximately 1 year Southern Canada

90

80
Retained Elongation (%)

UV-4 UV-8 UV-15

70

60

Failure Criteria
50

Xenon Weather-O-Meter
test, ASTM D-2565*
40
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10,000 15,000

Hours of Xenon Exposure

*See text for further details. ASTM test procedures may be modified to accommodate operating conditions or facility limitations.

5
How is degradation determined?
UV attack on polyethylene results in loss of physical properties. The property of greatest concern is embrittlement. The trend
towards embrittlement may be tested in many ways. In one study, tensile break elongation decline was more severe than ten-
sile impact decline. It also was found that tensile break elongation decline was more sensitive than tensile strength decline or
the formation of oxidized species at the exposed surface of the part, which may provide evidence of polymer degradation.
Thus, our testing tends to be based on break elongation decline. By convention, we take a 50% decline from the original break
value to be a “failure point” to establish the end of test. It is important to determine your own end point, as this will vary
depending on the applications of your products. At the 50% point, surface crazing (a crosshatched pattern caused when the
sample is bent) is not often evident on our polymers and often field use goes well beyond 50%. An example of this surface
crazing is shown in Figure 6 below.

Figure 6: Surface crazing as a


result of UV damage on
the polyethylene surface

Due to the variability of the molding process, our samples are currently prepared by compression molding to minimize data
scatter. Some people feel that using rotomolded samples and allowing a higher level of break elongational decline is preferred.
Studies conducted at our labs have shown that overcuring of rotationally molded samples can reduce UV life from 25% to
70% of the rated performance. This introduces much variability into reporting UV performance based on rotational molded
samples. We will continue to apply best control capabilities and to try to improve repeatability in rotomolded samples to
some day have confidence in reporting data on this basis.

6
How do pigments affect UV performance?
Figure 7 shows the effect of color pigments on UV stability. As can be seen, the proper choice of pigment is important as it
can have a detrimental or beneficial effect depending on the choice of pigment. All pigments should be melt compounded to
obtain the proper dispersion necessary for UV protection. The amount, particle size and chemical type of pigment, such as its
organic or inorganic nature, affects the UV performance. Generally, carbon black tends to be the best UV performer due to its
high absorption of UV light. Dry blending pigment has minimal effect as the dispersion and thus, absorption characteristics
are not sufficient to protect the base polymer.

It is important to not confuse UV performance with color-fading problems when dealing with pigments. Sometimes, a pigment
may fade while the base polymer remains unaffected by true UV degradation. Thus, impact and tensile properties are unaffected
while the part appearance has changed. Ensure your pigment supplier understands that you want not only a pigment that will
enhance or not hinder UV performance, but also one that will be color-stable when exposed to weathering. These pigments
are generally referred to as UV grade pigments.

Figure 7: Color pigment affect of UV performance

Arizona exposure of pigmented HDPE plaques (1mm)


Failure Criteria: 50% retained
tensile impact strength. Source: Left figure:
Ciba: Stabilization of
Phthalocyanine Green 0.25% 330 Polyolefins – Part 2.
Phthalocyanine Blue
0.25% >700* Polyolefin Thick Sections,
Azo-Red
0.25% 230 Arizona 45° south
Organic Yellow (start November)
(tetrachloroisoindolinone) 0.25% 250
TiO2 (stabilized, coated Rutile) 0.5% 570
Ultramarine Blue 210
0.5%
Iron Oxide
0.5% 500
Cd-Red
0.5% 460
Cd-Yellow
0.5% 230
Unpigmented
175

0 100 200 300 400 500 600 700 800

kly Arizona *no sample left


Polymer – HDPE (Ziegler)
Basic Stabilization: 0.03% IRGANOX 1076 + 0.05% Ca-stearate
Light Stabilization: 0.15% TINUVIN 770

Effect of pigments on UV stabilization of


.96 density unstabilized polyethylene resins
5000
1% Carbon Black

4000 1% Iron Oxide


Tensile Strength, psi

1% Phthalocyanine Green
3000 1% Cadmium Red
1% Cadmium Yellow

2000

1% Phthalocyanine Blue
Source: Right figure: 1000 Natural 1% TiO2 (Rutile)
Marlex HDPE Product Brochure,
Phillips Petroleum, 1% TiO2 (Anatase)
R.J. Martinovich for Plastic Technology 0
November 63. 0 6 12 18 24 30 36 40

Months exposed in Arizona 7


What other factors affect UV performance?
Table 3 outlines qualitatively other factors which may affect UV performance.

Table 3: Other factors which affect UV performance


Factor Directional affect on UV performance

Molding cycle Negative


(higher temperature or longer time)
Thicker part Positive to neutral
Higher density Positive
Lower melt index • Negative due to need for higher time or
temperatures to properly process part

• Positive in the fact that the resin has inherently


higher toughness
Environmental conditions which may Negative
promote particulates depositing on parts
Higher mechanical demands Negative
on the molded part
Higher resin stabilization Positive

Resin formulation is important. A resin must be properly stabilized to minimize degradation during the rotational molding
cycle. Sufficient UV stabilizers must be present throughout the product life cycle to protect the part. These stabilizers should
not migrate out of the part, which could reduce protection as well as potentially affect the contents of the material being
stored in the parts.

As discussed before, the most important parameters which the molder may control is part time and temperature history.
Overheating the part tends to shorten UV performance as it provides the same mechanisms that UV light promotes, and thus,
provides it with a faster starting point for degradation when the part is exposed. The extent of this effect is difficult to quantify
due to the variations of part thickness, and variations in heat history which different melt indexes of polyethylene may require
to produce initial acceptable part properties. A more comprehensive understanding of these effects is the subject of continuing
research.

Thicker parts generally would have better UV performance due to reduced UV penetration as it travels through the part.
Once the part is sufficiently thick, no further benefit would be expected.

Generally, higher density resins, which provide a larger crystal structure and less potential for entrapping oxygen, provide better
UV stability. Melt index (MI) does not directly affect UV stability, though the secondary effect of a lower MI providing a tougher
part and thus, longer life to obtain the same absolute break point, is a factor. Thus, all things being equal, higher density and lower
MI PE’s enhance UV performance. Again these factors are generally small.

Other variables include higher particulate levels in the atmosphere such as found in industrial areas. These may result in dark
material being deposited on the polymer surface, which may shorten UV performance by providing elevated temperatures.
Also, some exposure conditions may interfere with UV wavelengths, thus extending service life of the product.

8
Further references
Many excellent articles exist on this topic. Below are some listings that the reader may find helpful in further understanding
this topic.
• George Wypych, “Handbook of Material Weathering, 2nd Edition,” ChemTec Publishing, 1995.
• F. Gugumus, Ciba Specialty Chemicals, “Examples of use and misuse of accelerated testing of plastics,” Symposium; Polymer
Stabilization and Degradation: Problems, Techniques and Applications, Manchester, September 18-20, 1985.
• K.P. Scott, “Narrow Band vs. Wide Band Control in Accelerated Weathering Tests,” 1st International Symposium on
Weatherability, Tokyo, May 12-13, 1992.
• J.L. Martin, “Test Methods for Determining the Outdoor Durability of Coatings,” DSET Laboratories, Miami, Florida.
• L. Crewdson, “Correlation of Outdoor and Laboratory Accelerated Weathering Tests at Currently Used and Higher Irradiance
Levels – Part 1,” J.B. Atlas Co. Ltd.

©2003 Exxon Mobil Corporation. The user may forward, distribute, and/or photocopy this copyrighted document only if unaltered and complete, including all of its headers, footers, disclaimers, and other information. ExxonMobil does not guaran-
tee the typical (or other) values. Typical values only represent the values one would expect if the property were tested in our laboratories with our test methods on the specified date. Some product properties are not frequently measured, and accord-
ingly typical values are not based upon a statistically relevant number of tests. The information in this document relates only to the named product or materials when not in combination with any other product or materials. We based the information
on data believed to be reliable on the date compiled, but we do not represent, warrant, or otherwise guarantee, expressly or impliedly, the merchantability, fitness for a particular purpose, suitability, accuracy, reliability, or completeness of this infor-
mation or the products, materials, or processes described. The user is solely responsible for all determinations regarding any use and any process. We expressly disclaim liability for any loss, damage, or injury directly or indirectly suffered or incurred
as a result of or related to anyone using or relying on any of the information in this document. There is no warranty against patent infringement, nor any endorsement of any product or process, and we expressly disclaim any contrary implication.
The terms, “we,” “our,” “ExxonMobil Chemical,” or “ExxonMobil” are used for convenience, and may include any one or more of ExxonMobil Chemical Company, Exxon Mobil Corporation, or any affiliates they directly or indirectly steward. The
ExxonMobil Chemical Emblem and the “Interlocking X” Device are trademarks of Exxon Mobil Corporation.

S-ar putea să vă placă și