You are on page 1of 13

Study of the degradation of typical HVAC materials, filters

and components irradiated by UVC energy--part I: literature


search.
INTRODUCTION
The literature search presented herein describing the different degradation mechanisms of the
materials and nonmetallic components typically subjected to long-term UVC [254 nanometers (nm),
0.01 mil] exposure in HVAC systems is a summary of the project final report (Kauffman 2011). Two
books published by J. Rabek (1995 and 1996) were the basis for the conducted literature search and
contain thousands of references on research performed to study polymer
photodegradation mechanisms. Although the majority of the reported UV studies were performed
with UVA [320-400 nm (0.013-0.016 mil)] and UVB (280-320 nm (0.011-0.013 mil)], a significant
number of the studies explored the effects of UVC on polymer degradation. The results of the
conducted literature search were broken down into the following subcategories for discussion.
1. UVC absorption by the fresh polymeric surface
2. Photodegradation reaction mechanism
3. Effects of UV wavelength on degradation products
4. Effects of contaminants and additives on degradation rate
5. Effects of UV irradiance on degradation rate (Reciprocity Law)
6. Correlation between laboratory and HVAC UVC results
7. Identification of laboratory photoreactors
8. Identification of photodegradation monitoring techniques
UVC ABSORPTION BY THE FRESH POLYMERIC SURFACE
In order for UV to degrade a polymeric surface, chromophoric groups (chemical groups capable of
absorbing UV light) must be contained in the chemical structure of the polymer or in compounds
dispersed in the polymer matrix. Examples of chromophoric groups present in common polymers
are carbonyls (polyesters, nylon, acrylates, polyimide, etc.) and phenyl rings [polystyrene, epoxy,
poly (ethylene terephthalate), etc].
Since they do not contain chromophoric groups, polyethylene andpolypropylene do not absorb UV
light above 220 nm (0.0087 mil) (Rabek1996), and therefore, should be totally resistant to sunlight
(UVA andUVB) as well as UVC. Poly (vinyl chloride) does not absorb light above290 nm (0.011 mil)
(Rabek 1996), and therefore, should only be slightlysusceptible to UVA and UVB degradation.
However, in practice, polyethylene, polypropylene and poly (vinyl chloride) are all highly susceptible
to UVA-UVC degradation (Denizligil and Schnabel 2003, Hamid 2000, Rabek 1995 and 1996,
Schnabel 1981, Scott 1965 and Hamid 2000). The UV reactivities of the polymers are attributed to

the presence of UV absorbing additives and impurities formed during polymerization, processing
and/or storage. The impurities can be both internal (polymerization) and external
(processing/storage) and include hydroperoxides, carbonyl and unsaturated bonds, catalyst
residues, additives, metal traces, etc. (Rabek 1995 and 1996). Consequently, the capability of a
polymer to absorb UV energy is affected by both the chemical structure of the polymer as well as
the presence and location of the chromophoric impurities.
PHOTODEGRADATION REACTION MECHANISM
Once the organic compound absorbs the UV light, the produced excited state must be capable of
producing radicals and other reactive species for the UV degradation mechanism to continue. The
reaction scheme in Figure 1 is basically the same whether the polymer free radicals (P*) are
generated by thermal or UV energy (photolysis).
Figure 1 Photodegradation reaction scheme of polymeric materials. Photolysis PH (polymer) + UV
[right arrow] P + P (backone) or P + H (side chain) Propagation P + [O.sub.2] [right arrow] POO
(peroxy radical) (Step 1) POO + PH [right arrow] POOH (hydroperoxide) + P (back to Step 1)
H + [O.sub.2] [right arrow] HOO (hydrogen peroxy radical) HOO + PH [right arrow] HOOH
(hydrogen peroxide) + P (back to Step 1) Scission HOOH + UV [right arrow] HO + OH (hydroxide
radical) PH + OH [right arrow] [H.sub.2]O + P (back to Step 1) POOH + UV [right arrow] PO
(polymer oxy radical) + OH PO [right arrow] rearrangement to ketone/aldehyde and P (back to
Step 1) PH + OH [right arrow] [H.sub.2]O + P (back to Step 1) Cross-linking P + P [right arrow]
P - P (inactive products, dominant in vacuum) P + PO [right arrow] P - OP (rearrangement to
inactive products) P + POO [right arrow] P - OOP (rearrangement to inactive products) POO +
POO [right arrow] P - OOP + [O.sub.2] (dominant in air)
In the presence of oxygen, the scission steps are generallydominant with regard to the cross-linking
steps (Rabek 1995 and 1996).If the scission step occurs in the polymer back bone, the UV
exposurereduces the tensile strength of the exposed polymer and the carbonylgroups
(ketone/aldehyde) attached to the end of the severed polymermolecules increase the UV absorption
of the degraded polymer products.Alternately, if the scission step occurs at the end or in a short
sidechain of the original polymer, the UV exposure produces volatilecarbonyl products
(ketones/aldehydes, carbon dioxide, carbon monoxide)and a shortened polymer chain with a free
radical end. In contrast toscission, the cross-linking steps cause the molecular weight of thepolymer
to increase and the flexibility of the polymer to decrease.
In addition to the general reaction scheme in Figure 1, several polymers have reaction schemes
specific to their chemical structure. For example, poly (vinyl chloride) also undergoes
dehydrochlorination when exposed to UV (Denizligil and Schnabel 2003, Kaczmarek 2009 and
Rabek 1996) as shown in Figure 2.
[FIGURE 2 OMITTED]
The resulting double bonds (C = C) are chromophoric (increase the UVC reactivity of the reaction
product) and are responsible for the yellow-red coloring of the exposed polymer. The fact that the
poly (vinyl chloride) and any other chlo-rinated polymers would outgas hydrochloric acid gas
indicates UVC exposure would have the potential to promote corrosion of surrounding metallic
surfaces during polymer degradation.
Polycarbonates, polyurethanes, poly (phenyl acrylate) and epoxy resins undergo the photo-Fries
rearrangement (responsible for yellowing of aged polymers) shown in Figure 3 when exposed to UV

(Rabek 1995 and 1996). The rear-ranged polymers produced in Figure 3 are more susceptible to
UV degradation than the original polycarbonate polymer.
[FIGURE 3 OMITTED]
EFFECT OF UV WAVELENGTH ON DEGRADATION PRODUCTS
Although the focus of the literature search was on the specific UVC wavelength of 254 nm (0.01 mil)
[253.7 nm (0.00999 mil) to be precise] photodegradation of polymers, the majority of the UV
experiments in the identified references were performed in the UVA and UVB region or with an
unspecified wavelength. However, several authors (Rabek 1995) have noted that the reaction
products obtained with UVA/UVB light are lower in quantity (produced radicals have less energy)
but similar in composition to those produced with 254 nm (0.01 mil). For instance, both poly (vinyl
chloride) and polycarbonate produce yellow compounds (Figure 2 and 3) when exposed to either
UVA/UVB or UVC light (Rabek 1996). Under 254 and ~300 (UVB) nm (0.01 and 0.012 mil) light,
polypropylene decomposition produced similar products such as ketones, peroxy acids, peroxy
esters, etc. (Aslan-zadeh and Kish 2005).
Any differences in the UVA/UVB and 254nm (0.01 mil) (UVC)photodegradation products arise from
the higher energy of the UVC light(able to break bonds stable to lower energy of UVA/UVB).
Polyacetalundergoes depolymerization to only produce formaldehyde when exposed to360 nm (0.014
mil) UV light but produces ethyl alcohol and ethyleneglycol as well as formaldehyde when exposed
to 254 nm (0.01 mil) (Chiangand Huang 1999). In addition to the Photo-Fries rearrangement
causingthe polycarbonate to yellow, the higher energy of UVC is able todirectly break (C-O) bonds in
the original polycarbonate structure toform reactive free radicals and volatile products (Hamid
2000).
Regardless of the UV energy, once the photodegradation is initiated, the scission degradation
products increase the UV absorption of the polymer surface due to the presence of carbonyl groups
(C=O) and double bonds (C=C). Based on the experimental results reported in the identified
references (Rabek 1995 and 1996, Rabney and Rabek 1975 and Hamid 2000), Table 1 was compiled
to compare the susceptibilities of different materials to photo-initiated scission and gas
production [dependent on the efficiencies of both the photolysis (UV absorption) and scission steps
in Figure 1]. The materials are listed in order of decreasing yield, i.e., decreasing scission (gas
production) with same level of UV exposure indicates increasing resistance to UV degradation.
The results in Table 1 indicate that the susceptibility of the different polymers to scission are more
dependent on polymer composition than on the UV wavelength used in the experiment. Many of
the researchers reported that the quantum yields of scission and other reactions (cross-linking,
gaseous products, etc.) were independent of the wavelength of UV irradiance.
Regardless of the wavelength used in the studies, the primary volatile product of the UV irradiation
was water originating from the hydroxyl radical abstraction of a hydrogen atom from the polymer
(Figure 1). Carbon dioxide and carbon monoxide were also common degradation products resulting
from the scission reaction in Figure 1 when the carbonyl group was at the end of the polymer
radical. Other reported volatile products such as hydrochloric acid (Figure 2),
formaldehyde (polyacetal), formic acid [poly (ethylene terephthalate)], etc. were more dependent
on the composition of the polymer being irradiated than the wavelength of the UV light.
In addition to the type of degradation products produced, the wavelength of the UV light affects the
depth of the polymer surface undergoing photodegradation. For low absorbing polyethylene

and polypropylene, UV light in the 290-360 nm (0.011-0.014 mil) range caused significant scission
reactions to depths of 1.5-0.4 mm (59-16 mil) (depth decreased with increasing wavelength)
(Shyichuk 2005). For high absorbing polystyrene, polycarbonate, acrylonitrile-styrene and
poly (methyl methacrylate), (Nagai 2004) reported that the chemical changes due to exposure to
UVA/UVB was less than 5 microns (0.20 mil) deep. Comparing the depth penetration of an acrylic
resin, 50% of the incident light penetrated to a depth of 10 mm (390 mil) for 364 nm(0.014
mil) light, to 1 mm (39 mil) for 313 nm (0.012 mil) light and to only 0.01 mm [10 microns (0.39 mil)]
for 254 nm (0.1 mil) light (Feller 1994).
EFFECTS OF CONTAMINANTS AND ADDITIVES ON DEGRADATION RATE
As previously discussed, the photodegradation of many non-absorbing polymers are attributed to
the presence of UV absorbing impurities formed during polymerization, processing and/or storage.
The impurities can be both internal (synthesis) and external (processing/storage) and include
hydroperoxides, carbonyl and unsaturated bonds, catalyst residues, additives, metal traces, etc.
(Rabek 1995 and 1996). Consequently, the capability of a polymer to absorb UV energy is affected
by both the chemical structure of the polymer as well as the presence and location of chromophoric
impurities. For instance, the susceptibility of polystyrene to photodegradation is highly dependent
on how it was synthesized (Rabek 1995). Radically prepared polystyrenes are more susceptible to
UV degradation than anionically prepared polystyrenes since they contain double bonds, in-chain
peroxide linkages, and other oxygen containing groups, all of which increase the UV
absorption/reactivity of the internal and external polystyrene matrix.
In contrast to polystyrene, the UV susceptibilities of poly-ethylene and other extruded polymers are
dependent on their thermal history. The hydroperoxides formed at low temperatures [5090[degrees]C (122-194 degrees]F)] are isolated and have minimal photo-initiating effects (radical
produced by impurity is not transferred to polymer). Whereas, the hydroperoxides formed at high
[135-160 degrees]C (275-320 degrees]F)], prolonged temperatures on the surfaces of the
extruded polymers are associated and have significant photo-initiating capabilities (Rabek 1995).
Consequently, low density polyethylene would be expected to be more resistant to UV
photodegradation than high density polyethylene due to its lower extrusion temperatures.
In addition to unwanted impurities, additives such as plasticizers are added intentionally to
polymers to aid processing and improve the flexibility of the final product. Commercial plasticizers
such as di-n-octyl adipate/phthalate esters are capable of acting as photo-initiators for poly (vinyl
chloride). Consequently, the degradation products of commercial poly (vinyl chloride) products
can originate from both the plasticizer as well as the polymeric matrix (Deni-zligil and Schnabel
2003).
In contrast to plasticizers, carbon black has been reported to inhibit the UV photodegradation of
different polymers such as polyacetal (Chiang and Huang 1999) and polyethylene (Scott 1965). All
of the references identified during the literature search were concerned with the protective effects
of carbon black with regard to weathering (UVA/UVB), not UVC exposure. The presence of other
inorganic fillers capable of forming a protective coating (chalking) on the surface of the polymer or
acting as an energy sink for the UV energy absorbed by the polymer would be expected to have an
inhibitory effect on the 254 nm (0.01 mil) photodegradation of polymers.
EFFECTS OF UV IRRADIANCE ON THE DEGRADATION RATE (RECIPROCITY LAW)
One of the primary purposes of the literature search was to identify research that indicated that
short-term, high irradiance UVC experiments could be used to accurately predict the performance

of polymers exposed to long-term, low irradiance UVC, i.e., UVC exposure obeys the reciprocity
law. The reciprocity law is obeyed if the photodegradation of the polymer is dependent only on the
total energy of UV exposure (irradiance x time) and is independent of the time or irradiance level
taken separately. In other words, the degree of photodegradation would be the same whether
produced by 10,000 microwatts per square centimeter ([micro]W/[cm.sup.2]) [64,500 microwatts
per square inch ([micro]W/[in.sup.2])] for 100 hours or 500[micro]W/[cm.sup.2] (3220
[micro]W/[in.sup.2]) for 2000 hours.
When dealing with materials without fillers such as poly-ethylene, researchers have reported that
the rate at which UV (wavelength unspecified) irradiated polyethylene absorbs oxygen (oxidation)
is proportional to the square root of the light irradiance (Scott 1965). Also, the researchers
determining the quantum yields for different polymers in Table 1 reported that the quantum yields
of scission and other reactions (cross-linking, gaseous products, etc.) were independent of the
irradiance.
Table 1. Quantum Yields (a) of Chain Scission and Gas Evolution for UV Irradiation of Different
Polymers Polymer UV Wavelength Quantum Yield (nm) of Scission (Gas) Poly(phenyl isopropyl
254 0.17 - 0.22 ketone) Poly(methyl vinyl ketone) 254 0.025 Poly(methyl methacrylate) 254 0.02
- 0.04 300 0.016 - 0.005 (0.0002) Poly(methyl phenyl 313 0.017 silane) Poly(vinyl chloride) 254
- 400 (0.015) (loss of HCl) Poly(methyl acrylate) 254 0.013 Poly(a-methyl styrene) 254
0.007 Poly(vinyl acetate) 254 0.005, 0.05 (0.0065) EPDM Rubber < (b) 0.003 Poly(ethylene
254 0.0016 terephthalate) (0.0002) 280 - 360 0.0005 (0.0002) Polystyrene 254 0.0015 -0.0005
(0.000045) Cellulose 254 0.001 - 0.0007 Poly(vinyl pyrrolidone) 254 0.00043 Natural Rubber
254 0.0004 (cis-1,4-polyisoprene) (0.001) Polysulphones 254 0.00084 Nylon 6 254
0.0007 Polycarbonate 260 -300 0.0007 - 0.0001 Polyacrylonitrile 254 0.0002 0.00077 Polyurethanes 254 (0.00014) Mixed phenyl - methyl 254 (0.000026) Polysiloxanes (a)
Quantum Yield = Number of molecules decomposed/Number of photons absorbed b
Unspecified (b)Unspecified
More to the point of this project, research with UV light in the wavelength and irradiance ranges of
290 to 400 nm (0.011 to 0.015 mil) and 3,600 to 32,200 [micro]W/[cm.sup.2](23,200 to 207,700
[micro] W/[cm.sup.2]), respectively, demonstrated that the photo-degradation of acrylic-melamine
coatings obeyed the reciprocity law (Chin 2005). Fourier Transform Infrared (FTIR) analyses of the
irradiated coatings determined that the rates of scission, photo-oxidation and mass loss were
directly proportional to dosage regardless the time of irradiance. The references identified by Chin
(2005) also stated that the photodegradation of acrylic coatings, poly (vinyl
chloride), polycarbonate, poly-?-methylstyrene, acrylonitrile butadiene-styrene and poly (butylene
terephthalate) obeyed the law of reciprocity (wavelength unspecified).
CORRELATION BETWEEN LABORATORY AND HVAC UVC RESULTS
The main goal of this project is to allow HVAC designers to select the best material for use in an
UVC application and to allow maintenance personnel to estimate the useful life of an in-service
HVAC component based on its UVC dosage. For the laboratory test results to be the most useful for
the designer or maintenance personnel, the law of reciprocity must be obeyed by both the
laboratory test and HVAC results. The reciprocity law is obeyed if the photodegradation of the
material is dependent only on the total energy of UV exposure (irradiance x time) and is
independent of the irradiance time, geometry, chamber, environment, etc.
Various factors such as:

* wind, particle erosion, rain,


* large temperature variances between day and night,
* air polluants (ozone, nitrogen oxides, etc.),
* wavelengths/angle of sunlight versus UVA/UVB used in lab test,
* light and dark cycles, etc
are listed in ASTM G151 and ISO 4892-1 to explain the poor correlation between accelerated
laboratory (ASTM G154) and standardized outside (D1435) weathering tests (Scott 1965 and Rabek
1995 and 1996). Even with their lack of correlation, studies have shown that the susceptibilities of
different materials to photodegradation are ranked in similar orders by the accelerated UVA/UVB
laboratory and standardized outside weathering tests (Hamid 2000).
The surface photodegradation results produced by UVC lighting in the accelerated laboratory tests
and HVAC systems are expected to have better correlation than the results of the UVA/UVB
laboratory tests and weathering tests since the exposure environments of the
laboratory photoreactors and HVAC systems are controlled, surface erosion is minimal and sun
light is not involved. The sample vibration (rotating plat-form) and the frequent air flow (maintain
constant temperature) during the accelerated laboratory tests are expected to help simulate the
effects of HVAC operation on the photodegradation rates (enlarge microcracks, remove particles,
etc.) of exposed materials. The expected UVC laboratory test correlation with the HVAC systems is
further improved when the exposed materials are at a controlled temperature in the HVAC
system.
To further increase the probability of reciprocity between the photodegradation rates produced by
the laboratory and HVAC systems, the UVC irradiances of the accelerated laboratory tests were
selected to be with in an order of magnitude of those used in HVAC systems (Kauffman 2011). For
instance, the UVC irradiances for air and cooling coil treatments range from 1 to 20,000
[micro]W/[cm.sup.2] (6.4 to 129,000 [micro] W/[in.sup.2]) (IUVA 2005) and 10 to 500
[micro]W/[cm.sup.2] (64.5 to 3220 [micro]W/[in.sup.2]) (Kowalski 2009),
respectively. Consequently, the irradiance range of the developed accelerated laboratory tests was
between 900 and 14,000 [micro]W/[cm.sup.2] (5800 and 90,300 [micro]W/[in.sup.2]). Even for the
deeper penetrating UVA/UVB, in which the rate of photodegradation is limited by oxygen diffusion,
the laboratory tests obeyed reciprocity (Chin 2004) when the lower UV irradiance level was within
10x of the upper irradiance level used. The range of UVC irradiance levels that obey reciprocity
should be even wider since UVC mainly produces scission reactions (Rabek 1995 and 1996) and its
depth of penetration is reported to be only 10 microns (0.39 mil) (Feller 1994 and Shyichuk 2005).
Consequently, the UVC scission step (Figure 1) will always occur in the presence of air, regardless
of the UVC irradiance level, so that the rate of photodegradation will only depend on the level of
UVC irradiance (criteria for reciprocity) independent of the oxygen diffusion rate.
One factor that could lead to large differences in the laboratory and HVAC results is the wavelength
spectrum of the UVC light used by the HVAC system. According to the IUVA Draft Guideline IUVAG01A-2005, the two most common types of UV germicidal lamps are the medium and low pressure
mercury lamps. The emission spectra of the low and medium pressure lamps are quite different as
shown in Figure 4. The emission spectrum of the UVC lamps to be used in the accelerated
laboratory tests matched the spectrum of the low pressure lamp in Figure 4.

[FIGURE 4 OMITTED]
Consequently, the photodegradation rates of the UVC laboratory tests and HVAC systems are
expected to have good correlation and reciprocity when the HVAC systems employ low pressure
mercury lamps. Poor correlation is expected between the photodegradation rates of UVC laboratory
tests and HVAC systems employing medium pressure lamps [depth of penetration and accelerated
reactions dependent on wave-length (Feller 1994, Nagai 2004 and Shyichuk 2005)].
One other significant wavelength that is produced by a low pressure lamp is the 185 nm (0.0073
mil) line. The quartz envelope of the UVC lamp in Figure 4 absorbs the 185 nm (0.0073 mil) so it is
not transmitted to the environment. If the lamp envelope is produced from synthetic quartz
(Heraeus 2009) or has a defect, then both the 185 and 254 nm (0.0073 and 0.01 mil) lines are
transmitted resulting in ozone production [185 nm (0.0073 mil) line]. A large portion of the
generated ozone is converted into diatomic oxygen ([O.sub.2]) and a singlet oxygen (O*) (Jones and
Wayne 1969 and Qu 2005) by the 254 nm (0.01) line. The quantum yield of ozone photolysis by 254
nm (0.01 mil) is nearly unity, i.e., ratio (number of molecules converted/number of photons
absorbed) is almost 1 for ozone compared to <0.02 for most polymer photodegradation (Rabek
1995 and Table 1). In the presence of water [water treatment (Spartan water treatment 2009)] or
water vapor [surface cleaning systems (Kim 1996)], the ozone photolysis by 254 nm (0.1 mil) results
in the production of highly reactive hydroxide radicals (Figure 1) and oxygen.
IDENTIFICATION OF LABORATORY PHOTODEGRADATION CHAMBERS
In order to compare the photodegradation rates of different type materials in the irradiation range
of 900 and 14,000 [micro]W/[cm.sup.2] (5800 and 90,300 [micro]W/[in.sup.2]) photoreactors
designed to perform chemical and biochemistry experiments involving UVA-UVC exposure
were identified. The photoreactors are instrumented to monitor the irradiance reaching the surface
of the test solutions and the test chambers are temperature regulated to ensure repeatable
experimental conditions. The LZC-ICH2 photoreactor (Luzchem Research, Inc. Ottawa, Ontario) has
an inside chamber that is 30 cm wide, 30 cm deep and 22 cm high) (12 in. x 12 in. x 8.5 in.) that is
lined with an aluminum alloy (Al 5052-H32) to maximize UVC reflections and is equipped with 16
UVC lamps (Figure 5) to ensure the UVC irradiance is consistent throughout the reaction zone.
A rotating octagon turntable 21.6 cm wide with 8.9 cm sides (8.5 in., 3.5 in. sides) can be added to
the photoreactor to ensure all of the samples receive equal levels of UVC exposure to allow the
photodegradation rates of successive tests to be compared. The spectral output of the UVC
lamps employed by the ICH2 photoreactor is well characterized with ~ 96% of the UV energy
(92.5% UV, 7.2% visible and 0.5% infrared) produced by the 254 nm (0.01 mil) line [similar to 254
nm (0.01 mil) lamps in Figure 4].
[FIGURE 5 OMITTED]
Even though the ICH2 lamps are designed to produce only 254 nm (0.01 mil), ozone measurements
were made inside the photoreactor to ensure ozone was not being produced by a lamp defect [185
nm (0.0073 mil) transmitted through the defect]. Measurements determined that the ozone level
inside the photoreactor was below the instrument detection limit of 0.03 parts per million by volume
confirming the proper functioning of the photoreactors. Since varying ozone concentration is one of
the factors contributing to poor correlation between accelerated laboratory tests and weathering
tests (Scott 1965), ozone destruction by the UVC lamps in both the accelerated laboratory tests and
HVAC systems should further aid in their mutual reciprocity.
IDENTIFICATION OF PHOTODEGRADATION MONITORING TECHNIQUES

The final topic of the literature search was to identify the analytical technique(s) best suited to
monitor the photodegradation of the non-metallic samples exposed to 254 nm (0.01 mil) UV light.
Although 254 nm (0.01 mil) exposure causes the irradiated molecules to undergo free radical
induced oxidation (Figure 1) followed by scission reactions, and to a lesser extent cross-linking
reactions (Rabek 1995 and 1996), the affected molecules are concentrated close to the
surface (Feller 1994, Rivaton 2002). Also, the oxidation of the surface molecules (C=O and C=C
bonds increase) increases the 254 nm (0.01 mil) absorption by the surface further decreasing the
depth of the UVC penetration into the material (Rabek 1995 and 1996). Consequently, even though
photodegradation strongly affects the mechanical properties and chemical composition of the
exposed surface, the bulk properties and composition of the polymer remain unaffected so that
structural integrity tests such as tensile strength, impact and flexibility are considered impractical
for monitoring UVC photodegradation during the accelerated laboratory tests.
Since the identified literature was not focused on HVAC applications, several scrap samples
(Witham 2009) from HVAC material UVC testing were obtained for initial study (dosage levels
unknown). The scrap samples were analyzed with different analytical techniques to identify the
type of surface photo-oxidation mechanisms occurring and to identify the analytical test(s) with the
best potentials for monitoring the accelerated 254 nm (0.01 mil) photo-oxidation tests. Since
the original, unexposed material was not available for comparison, a portion of each sample's
surface was scraped away to reveal unexposed material for analysis.
For polymers with fillers, the FTIR spectra of the unexposed materials (inside) had distinctive
organic peaks (C-H, C=O, C-O, etc.) not present in the spectra of the UVC exposed surfaces as
illustrated by the representative spectra in Figure 6, i.e., UVC volatilization of polymer surface left
behind protective layer of inorganic filler.
[FIGURE 6 OMITTED]
The Energy Dispersive Spectrometric (EDS) elemental analyses of the polymer samples containing
fillers also indicated that the UVC exposed surface contained a much higher inorganic content than
the interior material (Kauffman 2011).
As opposed to Figure 6, the FTIR and EDS spectra of polymers not containing fillers were virtually
identical for the interior material and UVC exposed surface (Kauffman 2011). Surface analyses
indicated that the UVC exposed organic surface had a distinctive pattern, i.e., mass loss of clean
surface concentrated at susceptible locations (grain structure, crystallinity, etc.) as previously
discussed (Rabek 1995 and 1996).
Consequently, all of the samples underwent organic mass loss during UVC exposure in agreement
with Figure 1 indicating that photolysis (direct bond breaking) and photooxidation scission reactions
producing volatile products (mass loss) occur at a much higher rate than cross-linking reactions
(harden surface). The polymers without fillers would be expected to undergo constant mass loss
regardless of dosage while the polymers containing fillers would lose mass until an exterior layer of
inorganic fillers formed to protect the interior molecules from further UVC exposure and
photodegradation reactions (Figure 6). Several references were identified which stated that
surfaces exposed short-term to 254 nm (0.01 mil) had mass loss and other changes in
morphology (Kaczmarek 2005 and 2009, Rabek 1996 and Soto-Oviedo 2002). Longer-term 254 nm
(0.01 mil) tests with linear polyethylene with out fillers (Ranby and Rabek 1975) demonstrated that
mass loss occurred for the entire 800 hour test as indicated by the constant rate of oxygen
adsorption by the exposed surface (fresh material exposed by mass loss reacts with oxygen).

In addition to the HVAC scrap material analyses, additional polymers without fillers were exposed to
UVC for different lengths of time and then visually inspected for color changes as performed in
ASTM methods such as D6290 or International Standard ISO 4582. The surface of a polyurethane
sealer turned light orange after only 10 hours of UVC exposure [~4,000 [micro]W/[cm.sup.2]
(25,800 [micro]W/[in.sup.2]) with UVC pen light]. After 50 and 200 hours of UVC exposure, the
surface remained light orange in color, i.e., extended UVC exposure did not further darken the
exposed areas since the oxidized species were not accumulating due to mass loss. In contrast to the
polyurethane sealer, the polished surface of a polyacetal http://ac-repair.com/ rod did not change
color (exposed area remained white but became dull) even after 200 hours of UVC exposure.
Based on these initial results and the literature search, surface compositional changes were not
pursued for monitoring UVC photodegradation damage. Although the FTIR spectrum peak heights
are well-suited for quantification, the produced spectra did not change with dosage due to surface
mass loss (no fillers) and became independent of the polymer (filler accumulation due to polymer
loss). Color changes, although simple to perform with a colorimeter, are strongly affected
by surface finish and do not change with exposure time (surface craters observed), and
consequently, do not correlate well with degree of photodegradation. Finally even though
monitoring the rates of oxygen absorption (Ranby and Rabek 1975) and/or volatile compounds
produced during mass loss could be used to monitor the rate of UVC degradation of a polymer
surface, the large numbers of samples to be irradiated in the accelerated UVC laboratory tests
made the monitoring of oxygen consumption and/or produced gases from individual samples
impractical.
Consequently, an analytical technique capable of monitoring surface mass loss was selected as the
best technique for monitoring surface UVC photodegradation. The most direct technique for
monitoring mass loss from a surface is measuring the sample weight loss using an
analytical balance [sensitive to 0.1 milligram (3.5e-06 ounce) changes]. The primary drawback of
measuring weight loss is that small changes in the water content of the sample (due to heating,
room air humidity changes, etc.) will affect the sample weight, and consequently, the accuracy
of the UVC mass loss calculations. The second technique of monitoring surface mass loss is
masking the surface of the sample so that the surface exposed to the UVC suffers mass loss/height
reduction compared to the original surface protected from the UVC (crater formed in exposed area
compared to protected area). Scanning electron micro-scope microphotographs, which are
informative as to the morphology changes of the exposed surface, are limited by expense and
inability to measure crater depth. Optical micro-scopes using multiple focusing depths are able to
monitor crater depths but are laborious and time-consuming. Consequently, surface profilometers
were used to quantitate the depths of the UVC formed craters by sequentially scanning across the
protected then exposed surface areas. The primary drawback of the surface profilometers is that
the sample surface has to be polished to a level finish prior to masking and UVC exposure.
SUMMARY

The results of the literature search indicate that even though the rates of UVC photodegradation are
affected by a wide range of factors (processing parameters, contaminants, fillers, additives, etc.)
and vary by several orders of magnitude for different polymers (Table 1), the primary
photodegradation mechanisms are photolysis (direct bond breakage) and photooxidative scission
(Figure 1). The results of the literature search and initial UVC irradiated samples further
indicate that the primary change in the surface of the UVC irradiated samples is mass loss as
opposed to the compositional changes observed during UVA and UVB tests. For polymers without
fillers, the mass loss due to UVC exposure is expected to obey the law of reciprocity,
i.e., photodegradation of the polymer is dependent only on the total energy of UV exposure
(irradiance x time) and is independent of the time or irradiance level taken separately. For polymers
with fillers, the mass loss is expected to obey the law of reciprocity until the layer of protective
inorganic particles/fibers forms on the exposed surface inhibiting further mass loss. Analytical
balances for measuring sample weight loss or surface profilometers for measuring the depths of
craters formed in masked sample surfaces were identified as the best techniques for monitoring the
degree of surface UVC photodegradation.
The literature search also identified photoreactors with a rotating turntable and intermittent air
flow (simulate HVAC environment) that were capable of irradiating multiple small samples under
controlled temperature conditions as required by the accelerated UVC laboratory test to be
developed. According to the literature search, selecting irradiance levels within an order of
magnitude of those used in HVAC systems would further increase the probability of reciprocity
between the photodegradation rates produced by the developed laboratory tests in Parts 2 and 3 of
this project and the rates experienced in HVAC systems.
REFERENCES
ASHRAE Handbook. 2008. 2008 HVAC Systems and Equipment. Chapter 16. Ultraviolet Lamp
Systems. Atlanta: American Society of Heating, Refrigerating and Air-Conditioning Engineers,
Inc.
Aslanzadeh, S. and M. Kish. 2005. Photodegradation of polypropylene thermal bonded non-woven
fabric. Polymer Degradation and Stability 90 (3): 461-470.
ASTM D1435. 2005. Standard Practice for Outdoor Weathering of Plastics. Philadelphia: American
Society for Testing and Materials.
ASTM D6290. 2005. Standard Test Method for Color Determination of Plastic Pellets. Philadelphia:
American Society for Testing and Materials.
ASTM G151. 2006. Standard Practice for Exposing Nonmetallic Materials in Accelerated Test

Devices that Use Laboratory Light Sources. Philadelphia: American Society for Testing and
Materials.
ASTM G154. 2006. Standard Practice for Operating Light Apparatus for UV Exposure of
Nonmetallic Materials. Philadelphia: American Society for Testing and Materials.
Chiang, W. and C. Huang. 1999. Polymer Blends and Alloys. 1st Ed. Chapter 2. Compounding and
compatibilization of high-performance polymer alloys and blends. G. Shonaike and G. Simon, ed.
New York: Marcel Dekker.
Chin, J., Nguyen, T., Byrd, E. and J. Martin. 2005. Validation of the reciprocity law for coating
photodegradation. JCT Research 2 (7): 499-508.
Denizligil, S. and W. Schnabel. 2003. Photooxidation of polyvinyl chloride at 254 nm in the absence
and presence of phthalate plasticizers, Die Angewandte Makro-molekulare Chemie 229(1): 73-92.
Feller, R. 1994. Accelerated aging: photochemical and thermal aspects, The J. Paul Getty
Trust. http://www.getty.edu/conservation/publicatins/pdf_publications/aging.pdf
Grossetete, T., et al. 2000. Photochemical degradation of poly(ethylene terephthalate)-modified
copolymer, Polymer 41(10):3541-3554.
Hamid, S. 2000. Handbook of Polymer Degradation, 2nd Ed. Chapter 15: Wavelength sensitivity of
photodegradation of polymers and Chapter 18: Fundamental and technical aspects of the
photooxidation of polymers. New York: Marcel Dekker.
Heraeus. 2009. http://www.hi-techlamps.com/pdfs/uv/hLow%20Pressure%20UVC.pdf
ISO 4582. 2007. Plastics-Determination of changes in color and variations in properties after
exposure to day-light under glass, natural weathering or laboratory light sources. Geneva:
International Organization for Standardization.
ISO 4892-1. 1999. Plastics-Methods of exposure to laboratory light sources, Part 1: general
guidance. Geneva: International Organization for Standardization
IUVA. 2005. IUVA Draft Guideline IUVA-G01A-2005. Scottsdale: International Ultraviolet
Association.
Jones, I. and R. Wayne. 1969. Photolysis of ozone by 254-, 313- and 334-nm radiation. J. Chemical
Physics, 51:3617-3618.
Kaczmarek, H., Felczak, A. and D. Bajer. 2009. Photoxidative degradation of carboxylated poly(vinyl
chloride). Polymer Bulletin 62: 503-510.

Kaczmarek, H. et al. 2005. Effect of short wavelength UV-irradiation on ageing of


polypropylene/cellulose compositions. Polymer Degradation and Stability 88(2):189-198.
Kauffman, R. 2011. Study the degradation of typical HVAC materials, filters and components
irradiated by UVC energy (1509-RP). Final Report. American Society of Heating, Refrigerating and

Air-Conditioning Engineers, Inc.


Kim, C., Chung, C. and S. Moon. 1996. Removal of organic impurities from the silicon surface by
oxygen and UV cleaning. Korean J. of Chemical Engineering 13(3): 328-330.IL
Kowalski, W. 2009. ULGI for Cooling Coil Disinfection, Air Treatment, and Hospital Infection
Control. Report prepared for American Air & Water, Hilton Head,
SC http://www.americanairandwater.com/UV-pdf/UVGI-Report.pdf
Kowalski, W.J. and W.P. Bahnfleth. 2002. UVGI design basics for air and surface disinfection. HPAC
72 (1):100-110. http://www.engr.psu.edu/ae/wjk/uvh-pac.html
Kowalski, W.J., and Dunn, C.E. 2002. Current trends in UVGI air and surface disinfection.
INvironment Professional 8(6):4-6.
Kuzina, S. and A. Mikhailov. 2001. Photo-oxidation of polymers 4. The dual mechanism of
polystyrene photo-oxidation: a hydroperoxide and a photochain one. European Polymer Journal 37
(11):2319-2325
Levetin, E., et al. 2001. Effectiveness of germicidal UV radiation for fungal contamination within airhandling units. Applied and Environmental Microbiology 67 (8): 3712-3715.
Nagai, N., Matsunoe, T., and T. lmai. 2005. Infrared analysis of depth profiles in UV-photochemical
degradation of polymers. Polymer Degradation and Stability 88(2): 224-233.
Qu, Z. et al. 2005. The photodissociation of ozone in the Hartley band: A theoretical analysis. J.
Chem. Phys., Vol. 123 (7):074305.
Rabek, J. 1995. Polymer Photodegradation: Mechanisms and Experimental Methods. 1st Ed.
London: Chapman Hall.
Rabek, J. 1996. Photodegradation of Polymers: Physical Characteristics and Applications. 1st Ed.
Berlin: Springer-Verlag.
Ranby, B. and Rabek, J. 1975. Photodegradation, Photo-oxidation and Photostabilization of
Polymers: Principles and Applications. 1st Ed. New York: John Wiley and Sons.
Rivaton, A. et al. 2002. Comparison of the photochemical and thermal degradation of bisphenol-A
polycarbonate and trimethylcyclohexane-polycarbonate. Polymer Degradation and Stability 75: 1733.
Schnabel, W. 1981. Polymer Degradation: Principles and Practical Applications. Chapter 4:
Photodegradation. Hanser International. Macmillan Publishing Co., Inc., New York.
Scott, G. 1965 Atmospheric Oxidation and Antioxidants, 1st Ed. Chapter 7, Oxidative deterioration
of saturated oil and polymers. New York: Elsevier Publishing Company.
Shyichuk, A. et al. 2005. Comparison of UV-degradation depth-profiles in polyethylene,
polypropylene and an ethylene-propylene copolymer. Polymer Degradation and Stability, 88 (3):
415-419.

Soto-Oviedo, M. and M. De Paoli. 2002. Photo-oxidative degradation of poly(epichlorohydrin-c-ethylene oxide) elastomer at 254 nm. Polymer Degradation and Stability 76 (2): 219-225.
Spartan Water Treatment. 2009. http://www.spartanwater-treatment.com/index.html
Witham, D. 2009. Telephone conversation with author and sample submission April 14, 2009.
This paper is based on findings resulting from ASHRAE Research Project RP-1509.
Robert E. Kauffman is a utube.com/watch?v=YtKPS5FXsWk distinguished research chemist at
the University of Dayton Research Institute, Dayton, OH.

You might also like