throbber

`
`!!.! Dielectric Behavior
`
`$%,
`
`U
`Ic
`
`Ueff
`
`NB = I** Ueff*sin(cid:160)
`
`I* * sin(cid:160)
`
`I*
`
`I* * cos(cid:160)
`
`(cid:160)
`
`(cid:98)
`
`NW = I*
`
` * Ueff*cos(cid:160)
`
`f
`
`Ir
`Figure 11.8  Current-voltage diagram or power-indicator diagram of electric alternating
`currents
`
`a)
`
`U
`
`I
`
`(cid:160) = 90°
`
`Period = 360°
`
`a)
`
`U
`
`t
`
`(cid:160)
`(cid:98)
`(cid:160) = 90°
`
`I'
`
`I
`Period = 360°
`
`t
`
`Figure 11.9 Current and voltage in a condenser. I ≡ current, U ≡ voltage, t ≡ time a) Without
`dielectric losses (ideal condition), current and voltage are displaced by the phase angle
`°90 or /π 2 ; b) With dielectric loss, the current curve I ´ is delayed by the loss angle δ
`ϕ=
`
`If the condenser has losses, when tanδ > 0, a resistive current Ir is formed leading
`to a heating energy rate in the dielectric of
`= 1
`E
`UI
`tanδ
`(11.27)
`h
`eff
`2
`where Ieff represents the total current or the magnitude of the vector in Fig. 11.8.
`Using Eq. 11.25 for capacitance leads to
`*
`’
`’
`’’
`
`C
`
`=
`
`C
`
`−
`
`=
`
`C
`
`−
`
`iC
`
`1ω
`
`(11.28)
`
`R
`where C* is the complex capacitance, with C’ as the real component defined by
`
`Ad
`
`’
`
`C
`
`C
`
`’’
`
`=
`
`’
`= εε0
`
`r
`and C’’ as the imaginary component described by
`=1
`εε
`’’
`r
`0ω
`R
`Using the relationship in Eq. 11.5 we can write
`−(
`) =
`*
`’
`*
`C
`C
`C
`ε iε

`=
`’’
`r
`r
`r
`0
`0
`
`(11.29)
`
`(11.30)
`
`(11.31)
`
`Ad
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 515
`
`

`

`$%-
`
` 11 Electrical Properties of Polymers
`
`where εr* is called the complex dielectric coefficient. According to Eqs. 11.25 and
`11.31, the phase angle difference or dielectric dissipation factor can be defined by
`=I
`’
`’
`I
`If we furthermore consider that electric conductivity is determined by
`σ = 1
`(11.33)
`
`R
`then the imaginary component of the complex dielectric coefficient can be rewritten
`as
`
`εε
`
`r r’
`
`r c
`
`tanδ
`
`=
`
`dA
`
`(11.32)
`
`=
`

`’
`r
`
`tan
`

`
`
`
`ε σ
`=
`’’
`r
`ωε
`0
`Typical ranges for the dielectric dissipation factor of various polymer groups are
`shown in Table 11.2. Figures 11.10 [1] and 11.11 [1] present the dissipation factor
`tan δ as a function of temperature and frequency, respectively.
`
`(11.34)
`
`Table 11.2 Dielectric Dissipation Factor (tan δ) for Various Polymers
`tan δ
`Material
`Non-polar polymers (PS, PE, PTFE)
`< 0.0005
`Polar polymers (PVC and others)
`0.001 – 0.02
`Thermoset resins filled with glass, paper, cellulose
`0.02 – 0.5
`
`PA66
`PVC+40% TCP
`
`PVC
`
`PMMA
`
`PET
`
`PI
`B-Glass
`
`PTFE
`
`PE-HD
`
`PSU
`
`PS
`
`120
`
`°C
`
`160
`
`PF 31.5
`
`PF HP
`
`PC
`
`101
`
`tan (cid:98)
`
`102
`
`103
`
`104
`
`Dissipation factor
`
`105
`
`0
`
`40
`
`80
`Temperature
`Figure 11.10 Dielectric dissipation factor as a function of temperature for various polymers
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 516
`
`

`

`!!.! Dielectric Behavior
`
`$%%
`
`PF
`
`PA66
`
`PCTFE
`
`PVC
`
`PC
`
`B-Glass
`PS
`
`PE-HD
`
`
`
`tan (cid:98)
`
`PMMA
`
`101
`
`102
`
`103
`
`(cid:98)(cid:9)
`
`Dissipation factor
`
`Dissipation factor (tan
`104
`
`106
`Frequency
`Figure 11.11 Dielectric dissipation factor as a function of frequency for various polymers
`
`105
`100
`
`102
`
`104
`
`108
`
`1010
`
`Hz
`
`1012
`
`11.1.4 Implications of Electrical and Thermal Loss in a Dielectric
`
`The electric losses through wire insulation running high frequency currents must
`be kept as small as possible. Insulators are encountered in transmission lines or in
`high-frequency fields such as the housings of radar antennas. Hence, we would
`select materials that exhibit low electrical losses for these types of applications.
`On the other hand, in some cases we want to generate heat at high frequencies.
`Heat sealing of polar polymers at high frequencies is an important technique used
`in the manufacturing of so" PVC sheets, such as the ones encountered in auto-
`mobile vinyl seat covers.
`To assess whether a material is suitable for either application the loss properties of
`the material must be determined and the actual electrical loss calculated. To do
`this, we can rewrite Eq. 11.27 as
`
`E
`
`h = 2ω δtan
`U C
`
`or as
`
`(11.35)
`
`(11.36)
`
`C
`fU d
`E


`εε
`= 2
`’ tan
`2 2
`r
`h
`0
`0
`The factor that is dependent on the material and indicates the loss is the loss factor
`r’ tan , called εr’’ in Eq. 11.34. As a rule, the following should be used:


`−10 3 for high-frequency insulation applications, and


`r’ tan <
`−10 2 cfor heating applications.


`r’ tan >
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 517
`
`

`

`*&&
`
` 11 Electrical Properties of Polymers
`
`In fact, polyethylene and polystyrene are perfectly suitable as insulators in high-
`frequency applications. To measure the necessary properties of the dielectric, the
`standard DIN 53 483 and ASTM D 150 tests are recommended.
`
` (cid:132) 11.2 Electric Conductivity
`
`11.2.1 Electric Resistance
`
`The current flow resistance, R, in a plate-shaped sample in a direct voltage field is
`defined by Ohm’s law as
`R U
`=
`I
`
`(11.37)
`
`or by
`
`dA
`
`R
`

`
`(11.39)
`
`= 1
`(11.38)

`where σ is known as the conductivity and d and A are the sample’s thickness and
`surface area, respectively. The resistance is o"en described as the inverse of the
`conductance, G,
`= 1
`R G
`and the conductivity as the inverse of the specific resistance, ρ,
`= 1
`(11.40)
`

`The simple relationship found in Eq. 11.37–39 is seldom encountered because the
`voltage, U, is rarely steady and usually varies in cyclic fashion between 10-1 to 1011
`Hz [3].
`Current flow resistance is called volume conductivity and is measured one minute
`a"er direct voltage has been applied using the DIN 53 482 standard test. The time
`definition is necessary, because the resistance decreases with polarization. For
`some polymers we still do not know the final values of resistance. However, this has
`no practical impact, because we only need relative values for comparison. Figure
`11.12 compares the specific resistance, ρ, of various polymers and shows its
`dependence on temperature. Here, we can see that similar to other polymer proper-
`ties, such as the relaxation modulus, the specific resistance not only decreases with
`time but also with temperature.
`The surface of polymer parts o"en shows different electric direct-current resist-
`ance values than their volume. The main cause of this phenomenon is surface
`contamination (e. g., dust and moisture). We therefore have to measure the surface
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 518
`
`

`

`
`
`!!.$ Electric Conductivity
`
`*&’
`
`resistance using a different technique. One common test is DIN 53 482, which uses
`a contacting sample. Another test o"en used to measure surface resistance is DIN
`53 480. With this technique, the surface resistance is tested between electrodes
`placed on the surface. During the test, a saline solution is dripped on the elec-
`trodes causing the surface to become conductive, thus heating up the surface and
`causing the water to evaporate. This leads not only to an increased artificial con-
`tamination but also to the decomposition of the polymer surface. If during this
`process conductive derivatives such as carbon form, the conductivity quickly
`increases to eventually create a short circuit. Polymers that develop only small
`traces of conductive derivatives are considered resistant. Such polymers are poly-
`ethylene, fluoropolymers, and melamines.
`
`Polyethylene
`
`Poly vinylchloride
`
`Cellulose acetate
`PE type 31 resin
`
`Polyacrylnitrile
`Organic semiconductor
`
`Pressed graphite
`
`Constantan
`Pt
`
`-200
`
`0
`200
`Temperature, T
`
`Cu
`ºC
`
`400
`
`1020
`Ohm*m
`
`1016
`
`1012
`
`108
`
`104
`
`1
`
`10-4
`
`Specific electric resistance
`
`Figure 11.12  Specific electric resistance of polymers and metals as a function
`of temperature
`
`11.2.2 Physical Causes of Volume Conductivity
`
`Polymers with a homopolar atomic bond, which leads to pairing of electrons, do not
`have free electrons and are not considered to be conductive. Conductive polymers
`in contrast, allow for movement of electrons along the molecular cluster, because
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 519
`
`

`

`*&(
`
` 11 Electrical Properties of Polymers
`
`they are polymer salts. The classification of polymer families and a comparison to
`other conductive materials is given in Fig. 11.13.
`Potential uses of electric conductive polymers in electrical engineering include
`flexible electric conductors of low density, strip heaters, anti-static equipment,
`high-frequency shields, and housings. In semi-conductor engineering, some appli-
`cations include semi-conductor devices (Schottky-Barriers) and solar cells. In elec-
`trochemistry, applications include batteries with high energy and power density,
`electrodes for electrochemical processes, and electrochrome instruments.
`Because of their structure, polymers cannot be expected to conduct ions. Yet the
`extremely weak electric conductivity of polymers at room temperature and the fast
`decrease of conductivity with increasing temperatures is an indication that ions do
`move. They move because engineering polymers always contain a certain amount
`of added low-molecular constituents that act as moveable charge carriers. This is a
`diffusion process that acts in field direction and across the field. The ions “jump”
`from potential hole to potential hole as activated by higher temperatures (Fig.
`11.12). At the same time, the lower density speeds up this diffusion process. The
`strong decrease of specific resistance with the absorption of moisture is caused by
`ion conductivity.
`
`Metal
`
`Semi-conductors
`
`Silver, copper
`Iron
`Bismuth, mercury
`
`In, Sb
`
`Germanium
`
`Insulators
`
`Silicon
`
`Glass
`DNA
`Diamond
`Sulfur
`Quartz
`
`S cm-1
`106
`104
`102
`10
`10-2
`10-4
`10-6
`10-8
`10-10
`10-12
`10-14
`10-16
`10-18
`(cid:49)-1
`Figure 11.13  Electric conductivity of polyacetylene (trans-(CH)x) in comparison to other
`materials
`
`Polyacetylene
`
`(SN)X
`TTF (cid:37)TCNQ
`NMP (cid:37)TCNQ
`KCP
`
`Trans-(CH)x
`
`Molecular
`
`crystals(cid:35)
`
`cm-1
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 520
`
`

`

`
`
`!!.$ Electric Conductivity
`
`*&)
`
`M2
`M3
`M1
`M= Median contact points per particle
`200
`300
`100
`
`T= 25°C
`
`T= 75°C
`
`T= 200°C
`
`1012
`
`Ohm
`
`108
`
`106
`
`104
`
`102
`
`Resistance, R
`
`100
`
`0
`
`10
`
`20
`
`30
`
`Vol %
`
`50
`
`V2
`V3
`Iron particle content
`Figure 11.14 Resistance R of a polymer filled with metal powder (iron)
`
`V1
`
`Conducting polymers are useful for certain purposes. When we insulate high-
`energy cables, for example, as a first transition layer we use a polyethylene filled
`with conductive filler particles such as carbon black. Figure 11.14 demonstrates
`the relationship between filler content and resistance. When contact tracks
`develop, resistance drops spontaneously. The number of inter-particle contacts,
`M, determines the resistance of a composite. At M1 or M = 1 there is one contact
`per particle. At this point, the resistance starts dropping. When two contacts per
`particle exist, practically all particles participate in setting up contact and the
`resistance levels off. The sudden drop in the resistance curve indicates why it is
`difficult to obtain a medium specific resistance by filling a polymer.
`Figure 11.15 [4] presents the resistance of epoxy resins filled with metal flakes
`or powder. The figure shows how the critical volume concentration for the epoxy
`systems filled with copper or nickel flakes is about 7 % concentration of filler, and
`the critical volume concentration for the epoxy filled with steel powder is approx.
`15 %.
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 521
`
`

`

`*&$
`
` 11 Electrical Properties of Polymers
`
`Steel particles
`Nickel flakes
`Copper flakes
`
`1012
`
`Ohm
`
`106
`
`104
`
`102
`
`Resistance, R
`
`0
`
`0
`
`10
`
`20
`Filler, (cid:113)
`Figure 11.15 Resistance of epoxy systems filled with metal flakes or powder
`
`30
`
`%
`
`40
`
` (cid:132) 11.3 Application Problems
`
`11.3.1 Electric Breakdown
`
`Because the electric breakdown of insulation may lead to failure of an electric com-
`ponent or may endanger people handling the component, it must be prevented.
`Hence, we have to know the critical load of the insulating material to design the
`insulation for long continuous use with the appropriate degree of confidence. One
`of the standard tests used to generate this important material property data for
`plate or block-shaped specimens is DIN 53 481. This test neglects the effect of
`material structure and of processing conditions. From the properties already
`described, we know that the electric breakdown resistance or dielectric strength
`must depend on time, temperature, material condition, load application rate, and
`frequency. It is also dependent on electrode shape and sample thickness. In prac-
`tice, however, it is very important that the upper limits measured on the experi-
`mental specimens in the laboratory are never reached. The rule of thumb is to use
`long-term load values of only 10 % of the short-term laboratory data. Experimental
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 522
`
`

`

`
`
`!!.% Application Problems
`
`*&*
`
`evidence shows that the dielectric strength decreases as soon as crazes form in a
`specimen under strain and continues to decrease with increasing strain. This is
`demonstrated in Fig 11.16 [5].
`
`PP(T= 40°C)
`Ed= f((cid:161)mech )
`
`U= 100 kV
`
`min
`
`230
`
`kV/mm
`
`210
`
`190
`
`170
`
`150
`
`130
`
`Dielectric strength, Ed
`
`6
`Strain, (cid:161)
`Figure 11.16 Drop of the dielectric strength of PP films with increasing strain
`
`8
`
`%
`
`12
`
`0
`
`2
`
`4
`
`On the other hand, Fig. 11.17 [5] demonstrates how the dielectric dissipation factor,
`tan δ, rises with strain. Hence, one can easily determine the beginning of the vis-
`coelastic region (begin of crazing) by noting the starting point of the change in
`tan δ. It is also known that amorphous polymers act more favorably to electric
`breakdown resistance than partly crystalline polymers. Semi-crystalline polymers
`are more susceptible to electric breakdown as a result of breakdown along inter-
`spherulitic boundaries as shown in Fig. 11.18 [6]. Long-term breakdown of semi-
`crystalline polymers is either linked to “treeing”, as shown in Fig. 11.19, or occurs
`as a heat breakdown, burning a hole into the insulation, such as the one in Fig.
`11.18. In general, with rising temperature and frequency, the dielectric strength
`continuously decreases.
`Insulation materials – mostly LDPE – are especially pure and contain voltage
`stabilizers. These stabilizers are low-molecular cyclic aromatic hydrocarbons.
`Presumably, they diffuse into small imperfections or failures, fill the empty space,
`and thereby protect the material from breakdown.
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 523
`
`

`

`*&+
`
` 11 Electrical Properties of Polymers
`
`Table 11.3 [7] provides dielectric strength and resistivity data for selected polymeric
`materials.
`
`PP
`tan (cid:98)=f((cid:161)mech,E)
`
` kV
`E= 30
` mm
`
`24
`
`20
`
`81
`
`6
`12
`
`4
`
`First registered mechanical damage
`
`3.0
`tan (cid:98)
`
`2.6
`
`2.2
`
`1.8
`
`1.4
`
`1.0
`
`Dielectric dissipation factor
`
`4
`Strain, (cid:161)
`Figure 11.17 Increase of dielectric dissipation with increased strain in PP foils
`
`6
`
`%
`
`0
`
`2
`
`Figure 11.18 Breakdown channel around a polypropylene spherulitic boundary
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 524
`
`

`

`
`
`!!.% Application Problems
`
`*&,
`
`Figure 11.19 Breakdown channel in a structure-less, finely crystalline zone of poly-propylene
`
`Table 11.3 Dielectric Strength and Resistivity for Selected Polymers
`Polymer
`Dielectric strength
`Resistivity
`(MV/m)
`(Ohm-m)
`25
`1014
`20
`1013
`20
`1013
`11
`1013
`11
`109
`10
`109
`16
`1013
`22
`1015
` 8
`1013
`15
`1012
`19
`1014
`17
`1013
`50
`1014
`12
`109
`23
`1015
`28
`1015
`20
`1014
`27
`1014
`22
`1015
`45
`1016
`14
`1012
`30
`1011
`25
`1014
`
`ABS
`Acetal (homopolymer)
`Acetal (copolymer) acrylic
`Acrylic
`Cellulose acetate
`CAB
`Epoxy
`Modified PPO
`Polyamide 66
`Polyamide 66 + 30 % GF
`PEEK
`PET
`PET + 36 % GF
`Phenolic (mineral filled)
`Polycarbonate
`Polypropylene
`Polystyrene
`LDPE
`HDPE
`PTFE
`uPVC
`pPVC
`SAN
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 525
`
`

`

`*&-
`
` 11 Electrical Properties of Polymers
`
`11.3.2 Electrostatic Charge
`
`An electrostatic charge is o"en a result of the excellent insulation properties of
`polymers – the very high surface resistance and current-flow resistance. Because
`polymers are bad conductors, the charge displacement of rubbing bodies, which
`develops with mechanical friction, cannot equalize. This charge displacement
`results from a surplus of electrons on one surface and a lack of electrons on the
`other. Electrons are charged positively or negatively up to hundreds of volts. They
`release their surface charge only when they touch another conductive body or a
`body that is inversely charged. O"en the discharge occurs without contact, as the
`charge arches through the air to the close-by conductive or inversely charged body,
`as demonstrated in Fig. 11.20. The currents of these breakdowns are low. For
`example, there is no danger when a person suffers an electric shock caused by a
`charge from friction of synthetic carpets or vinyls. There is danger of explosion,
`though, when the sparks ignite flammable liquids or gases.
`
`Discharges
`
`Flow of charge
`through polymer part
`Figure 11.20 Electrostatic charges in polymers
`
`As the current-flow resistance of air is generally about 109 Ωcm, charges and flash-
`overs only occur, if the polymer has a current-flow resistance of > 109 to 1010 Ωcm.
`Another effect of electrostatic charges is that they attract dust particles on polymer
`surfaces.
`Electrostatic charges can be reduced or prevented by the following means:
` (cid:131) Reduce current-flow resistance to values of < 109 Ωcm, for example by using con-
`ductive fillers such as graphite.
` (cid:131) Make the surfaces conductive by using hygroscopic fillers that are incompatible
`with the polymer and surface. It can also be achieved by mixing-in hygroscopic
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 526
`
`

`

`
`
`!!.% Application Problems
`
`*&%
`
`materials such as strong soap solutions. In both cases, the water absorbed from
`the air acts as a conductive layer. It should be pointed out that this treatment
`loses its effect over time. Especially, the rubbing-in of hygroscopic materials has
`to be repeated over time.
` (cid:131) Reduce air resistance by ionization through discharge or radioactive radiation.
`
`11.3.3 Electrets
`
`An electret is a solid dielectric body that exhibits permanent dielectric polari-
`zation. Some polymers can be used to manufacture electrets by solidifying them
`under the influence of an electric field, by bombarding them with electrons, or
`sometimes through mechanical forming processes.
`Applications include films for condensers (polyester, polycarbonate, or fluoropoly-
`mers).
`
`11.3.4 Electromagnetic Interference Shielding (EMI Shielding)
`
`Electric fields surge through polymers as shown schematically in Fig. 11.20. Because
`we always have to deal with the influence of interference fields, signal-sensitive
`equipment such as computers cannot operate in polymer housings. Such housings
`must therefore have the function of Faradayic shields. Preferably, a multilayered
`structure is used – the simplest solution is to use one metallic layer. Figure 11.21
`classifies several materials on a scale of resistances. We need at least 102 Ωcm for a
`material to fulfill the shielding purpose. The best protective properties are achieved
`with carbon fibers or nitrate coated carbon fibers used as fillers. The shielding prop-
`erties are determined using the standard ASTM ES 7-83 test.
`
`Polymers
`
`4
`
`6
`
`8
`
`12
`
`log ((cid:49)*cm)
`
`16
`
`Metal-Plastics
`
`Metal
`
`-6
`
`-4
`-2
`0
`2
`Electric resistance
`
`a)
`
`Al
`
`Ceramic
`
`Glass
`
`Metal-Plastics
`
`Polymers
`
`Insulators
`
`b)
`
`-5
`
`-4
`
`log [cal/(cm*s)]
`
`0
`
`-2
`-3
`Thermal resistance
`Figure 11.21 Comparison of conductive polymers with other materials: a) Electric resistance
`ρ of metal-plastics compared to resistance of metals and polymers b) Thermal resistance λ of
`metal-plastics compared to other materials
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 527
`
`

`

`*’&
`
` 11 Electrical Properties of Polymers
`
` (cid:132) 11.4 Magnetic Properties
`
`External magnetic fields have an impact on substances that are exposed to them
`because the external field interacts with the internal fields of electrons and atomic
`nuclei.
`
`11.4.1 Magnetizability
`
`Pure polymers are diamagnetic; that is, the external magnetic field induces mag-
`netic moments. However, permanent magnetic moments, which are induced on
`ferromagnetic or paramagnetic substances, do not exist in polymers. This magnet-
`izability M of a substance in a magnetic field with a field intensity H is computed
`with the magnetic susceptibility, χ as
`(11.41)
`H= χ
`M
`The susceptibility of pure polymers as diamagnetic substances has a very small and
`negative value. However, in some cases, we make use of the fact that fillers can
`alter the magnetic character of a polymer completely. The magnetic properties of
`polymers are o"en changed using magnetic fillers. Well-known applications are
`injection molded or extruded magnets or magnetic profiles, and all forms of elec-
`tronic storage devices.
`
`11.4.2 Magnetic Resonance
`
`Magnetic resonance occurs when a substance, in a permanent magnetic field,
`absorbs energy from an oscillating magnetic field. This absorption develops as a
`result of small paramagnetic molecular particles stimulated to vibration. We use
`this phenomenon to a great extent to clarify structures in physical chemistry.
`Methods to achieve this include electron spinning resonance (ESR) and, above all,
`nuclear magnetic resonance (NMR) spectroscopy.
`Electron spinning resonance becomes noticeable when the field intensity of a static
`magnetic field is altered and the microwaves in a high frequency alternating field
`are absorbed. Because we can only detect unpaired electrons using this method,
`we use it to determine radical molecule groups.
`When atoms have an odd number of nuclei, protons and neutrons, the magnetic
`fields caused by self-motivated spin cannot equalize. The alignment of nuclear
`spins in an external magnetic field leads to a magnetization vector that can be
`measured macroscopically as is schematically demonstrated in Fig. 11.22. This
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 528
`
`

`

`
`
`References
`
`*’’
`
`method is of great importance for the polymer physicist to learn more about mole-
`cular structures.
`
`2
`
`5
`
`Nuclear resonance
`signal
`
`3
`
`4
`
`1
`
`1
`
`Figure 11.22 Schematic of the operating method of a nuclear spin tomograph: 1) magnet
`producing a high, steady magnetic field; 2) radio wave generator; 3) high-frequency field,
`produced by 2, when switch 5 is in upper position; 4) processing nucleus, simulated by high
`frequency field; 5) switch; in this position the decrease of relaxation of the nucleus’ vibrations
`is measured
`
` (cid:132) References
`
`[1]
`
`[3]
`
`[4]
`
`Domininghaus, H., Elsner, P., Eyerer, P., and Hirth, T., Kunststoffe – Eigenscha#en
`und Anwendungen, Springer, Munich, (2012).
`[2] Menges, G., Haberstroh, E., Michaeli, W., and Schmachtenberg, E., Menges Werk-
`stoffkunde Kunststoffe, Hanser Publishers, Munich, (2011).
`Baer, E., Engineering Design for Plastics, Robert E. Krieger Publishing Company,
`(1975).
`Reboul, J.-P., Thermoplastic Polymer Additives, Chapter 6, J. T. Lutz, Jr., Ed., Marcel
`Dekker, Inc., New York, (1989).
`Berg, H., Ph.D Thesis, IKV, RWTH-Aachen, Germany, (1976).
`[5]
`[6] Wagner, H., Internal report, AEG, Kassel, Germany, (1974).
`Crawford, R. J., Plastics Engineering, 2nd ed., Pergamon Press, (1987).
`[7]
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 529
`
`

`

`
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 530
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 530
`
`

`

` 12
`
`Optical Properties
`of Polymers
`
`Because some polymers have excellent optical properties and are easy to mold and
`form into any shape, they are o!en used to replace transparent materials such as
`inorganic glass. Polymers have been introduced into a variety of applications, such
`as automotive headlights, signal light covers, optical fibers, fashion jewelry, chan-
`deliers, toys, and home appliances. Organic materials, such as polymers, are also
`an excellent choice for high-impact applications where inorganic materials, such
`as glass, would easily shatter. However, due to the difficulties encountered in main-
`taining dimensional stability, they are not suitable for precision optical applica-
`tions. Other drawbacks include lower scratch resistance, when compared to
`inorganic glasses, making them still impractical for applications such as automo-
`tive windshields.
`In this section, we will discuss basic optical properties, which include the index of
`refraction, birefringence, transparency, transmittance, gloss, color, and behavior of
`polymers in the infrared spectrum.
`
` (cid:132) 12.1 Index of Refraction
`
`cv
`
`n
`
`As rays of light pass through one material into another, the rays are bent due to the
`change in the speed of light from one media to the other. The fundamental material
`property that controls the bending of the light rays is the index of refraction, n. The
`index of refraction for a specific material is defined as the ratio between the speed
`of light in a vacuum to the speed of light through the material under consideration
`=
`(12.1)
`where c and v are the speeds of light through a vacuum and transparent media,
`respectively. In more practical terms, the refractive index can also be computed as
`a function of the angle of incidence, θi, and the angle of refraction, θr, as follows:
`= sin
`n
`sin
`
`(12.2)
`
`
`
`θ θ
`
`i r
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 531
`
`

`

`$%&
`
` 12 Optical Properties of Polymers
`
`where θi and θr are defi ned in Fig. 12.1.
`
`Angle of
`incidence
`
`(cid:101)i
`
`(cid:101)r
`
`Angle of
`refraction
`
` 
`
` Figure 12.1 Schematic of light
`refraction
`
`The index of refraction for organic plastic materials can be measured using the
`standard ASTM D 542 test. It is important to mention that the index of refraction is
`dependent on the wavelength of the light under which it is being measured. Figure
`12.2 shows plots of the refractive index for various organic and inorganic materials
`as a function of wavelength. One of the signifi cant points of this plot is that acrylic
`materials and polystyrene have similar refractive properties as inorganic glasses.
`
`Glass
`
`Polystyrene
`
`Quartz
`
`Acrylic
`
`1.75
`
`1.70
`
`1.65
`
`1.60
`
`1.55
`
`1.50
`
`Index of refraction, n
`
`1.45
`200
`
`600
`400
`Wavelength, (cid:104)
`
`nm 800   Figure 12.2 Index of refraction
`as a function of wavelength for various
`materials
`
`An important quantity that can be deduced from the light’s wavelength dependence
`on the refractive index is the dispersion, D, which is defi ned by
`D dn
`=

`d
`
`(12.3)
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 532
`
`

`

`
`
`!".! Index of Refraction
`
`$%$
`
`01234567
`
`104 (
`dn
`d(cid:104)
`
`Dispersion, D
`
`Polystyrene
`
`Glass
`
`Acrylic
`
`Quartz
`
`200
`
`400
`
`600
`Wavelength, (cid:104)
`
`nm
`
`800
`
` 
`
` Figure 12.3 Dispersion as a
`function of wavelength for various
`materials
`
`Figure 12.3 shows plots of dispersion as a function of wavelength for the same
`materials shown in Fig. 12.2. The plots show that polystyrene and glass have a
`high dispersion in the ultra-violet light domain.
`It is also important to mention that since the index of refraction is a function of
`density, it is indirectly affected by temperature. Figure 12.4 shows how the refrac-
`tive index of PMMA changes with temperature. A closer look at the plot reveals the
`glass transition temperature.
`
`Tg = 105 ºC
`
`1.51
`
`1.50
`
`1.49
`
`1.48
`
`Index of refraction, n
`
`1.47
`
`0
`
`20
`
`40
`
`60
`80
`Temperature, T
`Figure 12.4 Index of refraction as a function of temperature for PMMA (λ = 589.3 nm)
`
`100
`
`ºC
`
`140
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 533
`
`

`

`$%(
`
` 12 Optical Properties of Polymers
`
` (cid:132) 12.2 Photoelasticity and Birefringence
`
`(12.4)
`
`(12.5)
`
`Photoelasticity and flow birefringence are applications of the optical anisotropy of
`transparent media. When a transparent material is subjected to a strain field or a
`molecular orientation, the index of refraction becomes directional; the principal
`strains ε1 and ε2 are associated with principal indices of refraction n1 and n2 in a
`two-dimensional system. The difference between the two principal indices of
`refraction (birefringence) can be related to the difference of the principal strains
`using the strain-optical coefficient, k, as
`−(
`)
`n
`n
`k
`ε ε
`−
`=
`1
`2
`1
`2
`or, in terms of principal stress, as
`(
`)
`n
`n
`C
`σ σ
`−
`=
`−
`1
`2
`1
`2
`where C is the stress-optical coefficient.
`Double refractance in a material is caused when a beam of light travels through a
`transparent medium in a direction perpendicular to the plane that contains the
`principal directions of strain or refraction index, as shown schematically in Fig.
`12.5 [1]. The incoming light waves split into two waves that oscillate along the two
`principal directions. These two waves are out of phase by a distance δ defined by
`−(
`)
`n
`n t
`1
`2
`
`δ =
`
`
`
`(12.6)
`
`To observer
`
`(cid:98)= (n1-n2)t
`
`(cid:161)2
`
`n2
`
`(cid:161)1
`
`n1
`
`M
`
`t
`
`Direction of
`vibration
`
`Direction of
`propagation
`Figure 12.5 Propagation of light in a strained transparent medium
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 534
`
`

`

`
`
`!"." Photoelasticity and Birefringence
`
`$%)
`
`where t is the thickness of the transparent body. The out-of-phase distance, δ,
`between the oscillating light waves is usually referred to as the retardation.
`In photoelastic analysis, the magnitude of the stresses is determined by measuring
`the direction of the principal stresses or strains and the retardation. The technique
`and apparatus used to perform such measurements is described in the ASTM D
`4093 test. Figure 12.6 shows a schematic of such a set-up, composed of a narrow
`wavelength band light source, two polarizers, two quaterwave plates, a compen-
`sator, and a monochromatic fi lter. The polarizers and quaterwave plates must be
`perpendicular to each other (90°). The compensator is used to measure retarda-
`tion, and the monochromatic fi lter is needed when white light is not suffi cient to
`perform the photoelastic measurement. The set-up presented in Fig. 12.6 is gener-
`ally called a polariscope.
`The parameter used to quantify the strain fi eld in a specimen observed through a
`polariscope is the color. The retardation in a strained specimen is associated with
`a specifi c color. The sequence of colors and their respective retardation values and
`fringe order are shown in Table 12.1 [1]. The retardation and color can also be
`associated to a fringe order using
`Fringe order = δ
`

`
`(12.7)
`
`Monochromator
`
`Quarterwave plates
`
`Specimen
`
`Compensator
`Polarizers
`
`Light source
`
`Figure 12.6 Schematic diagram of a polariscope
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 535
`
`

`

`$%*
`
` 12 Optical Properties of Polymers
`
`Table 12.1 Retardation and Fringe Order Produced in a Polariscope
`Color
`Retardation (nm)
`Fringe order
`Black
`   0
`0
`Gray
` 160
`0.28
`White
` 260
`0.45
`Yellow
` 350
`0.60
`Orange
` 460
`0.79
`Red
` 520
`0.90
`Tint of passage
` 577
`1.00
`Blue
` 620
`1.06
`Blue-green
` 700
`1.20
`Green-yellow
` 800
`1.38
`Orange
` 940
`1.62
`Red
`1050
`1.81
`Tint of passage
`1150
`2.00
`Green
`1350
`2.33
`Green-yellow
`1450
`2.50
`Pink
`1550
`2.67
`Tint of passage
`1730
`3.00
`Green
`1800
`3.10
`Pink
`2100
`3.60
`Tint of passage
`2300
`4.00
`Green
`2400
`4.13
`
`A black body (fringe order zero) represents a strain free body, and closely spaced
`color bands represent a component with high strain gradients. The color bands
`are generally called the isochromatics. Figure 12.7 shows the isochromatic fringe
`pattern in a stressed notched bar. The fringe pattern can also be a result of mole-
`cular orientation and residual stresses in a molded transparent polymer compo-
`nent. Figure 12.8 shows the orientation induced fringe pattern in a molded part.
`The residual stress-induced birefringence is usually smaller than the orientation-
`induced pattern, making them more difficult to measure.
`Flow induced birefringence was explored by several researchers [2–4]. Likewise,
`the flow induced principal stresses can be related to the principal refraction indi-
`ces. For example, in a simple shear flow this relation can be written as [5]
`C
`C
`2
`2
`ηγ⋅
`−(
`) =
`n
`x
`x
`sin
`2
`2
`sin
`where x is the orientation of the principal axes in a simple shear flow.
`Figure 12.9 [6] shows the birefringence pattern for the flow of linear low-density
`polyethylene in a rectangular die.
`
`n
`
`2
`
`1
`

`12
`
`=
`
`(12.8)
`
`MacNeil Exhibit 2178
`Yita v. MacNeil IP, IPR2020-01139, Page 536
`
`

`

`
`
`!"." Photoelasticity and Birefringence
`
`$%+
`
`Yellow
`Red
`White
`Purple
`Grey
`Blue
`
`Lowest stress
`
`Highest stress
`
`Pink
`Pale green
`Pink
`Pink Pale green
`White
`Green
`Green
`Yellow
`
`Red
`
`Red
`
`Pale yellow
`Emerald
`Red
`Yellow
`Green
`
`Figure 12.7 Fringe pattern on a notched bar under tension
`

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket