the acetophenone radical anion: a computational

61
The Pennsylvania State University The Graduate School THE ACETOPHENONE RADICAL ANION: A COMPUTATIONAL INVESTIGATION OF PREFERRED GEOMETRIES A Thesis in Materials Science and Engineering by Maximiliano Aldo Burgess © 2020 Maximiliano Aldo Burgess Submitted in Partial Fulfillment of the Requirements for the Degree of Master of Science December 2020

Upload: others

Post on 22-Nov-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

The Pennsylvania State University

The Graduate School

THE ACETOPHENONE RADICAL ANION:

A COMPUTATIONAL INVESTIGATION OF PREFERRED

GEOMETRIES

A Thesis in

Materials Science and Engineering

by

Maximiliano Aldo Burgess

© 2020 Maximiliano Aldo Burgess

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Master of Science

December 2020

ii

The thesis of Maximiliano A Burgess was reviewed and approved by the following:

Ismaila Dabo

Associate Professor of Materials Science and Engineering

Thesis Co-Advisor

Adri CT van Duin

Professor of Mechanical and Nuclear Engineering

Thesis Co-Advisor

Michael T Lanagan

Professor of Engineering Science and Mechanics

John C Mauro

Professor of Materials Science and Engineering

Chair, Intercollegiate Graduate Degree Program

iii

ABSTRACT

As the world becomes increasingly interconnected, the demand for high-quality

telecommunications infrastructure rises. One of the most important tools for long-

distance power transmission is the high-voltage direct-current cable, which always

includes a layer of insulating material. Since the 1960s, polymers have been the pre-

dominant materials used for high-voltage insulation and more recently, crosslinked

polyethylene has come to the forefront due to its environmental resistance and ease

of manufacturing. The most popular method of crosslinking, initiation with dicumyl

peroxide, produces a few undesirable byproducts that compromise the integrity of

the insulating layer, through mechanisms such as the formation of space charge. One

of these byproducts is the acetophenone radical anion. This work explores the inter-

action between the 17 constituent atoms of the neutral acetophenone molecule and

an extra electron. Using a variety of computational techniques, this work attempted

to determine where on the geometry of the molecule the electron prefers to situate

itself. If this can be understood, predictions can be made about chemistry involving

the acetophenone radical anion and steps can be taken to mitigate its effects on the

performance and durability of insulating materials.

iv

TABLE OF CONTENTS

LIST OF FIGURES ............................................................................................ v

LIST OF TABLES ............................................................................................ vi

ACKNOWLEDGEMENTS .............................................................................. vii

1 BACKGROUND AND MOTIVATION ........................................................... 1

1.1 The Need for High-Quality High-Voltage Cables ...................................... 1

1.2 The Rise of Polymers as Insulating Materials .......................................... 4

1.3 The Process of Crosslinking Polyethylene ................................................ 8

1.4 The Complications of Crosslinking Polyethylene ..................................... 11

1.5 A Survey of Experimental and Computational Studies ........................... 14

2 METHODS .................................................................................................... 16

2.1 The Essentials of Density-Functional Theory .......................................... 16

2.2 Constrained Density-Functional Theory .................................................. 21

2.3 Choosing a Basis Set ................................................................................ 22

2.4 Tuning the Hybrid Functionals ............................................................... 24

2.5 Calculating an Equilibrium Mixture ....................................................... 26

2.6 Atomic Orbital-Based Population Schemes ............................................. 27

2.7 The Essentials of Reactive Force Field Calculations ............................... 29

2.8 Improvements in the Treatment of Electrons and Holes ......................... 34

2.9 Calculating an Equilibrium Mixture, Again ............................................. 36

3 RESULTS ...................................................................................................... 38

3.1 Constrained DFT Results ....................................................................... 38

3.2 eReaxFF-MD Results .............................................................................. 40

3.3 Mulliken Population Analysis Results ..................................................... 41

3.4 Discussion ................................................................................................ 43

3.5 Future Directions ..................................................................................... 45

REFERENCES ................................................................................................. 46

v

LIST OF FIGURES

1 Ultra-high-voltage DC lines in China ........................................................... 2

2 The components of a power cable ................................................................ 3

3 A comparison of paper and polymer insulation ............................................ 4

4 Catalysts used for coordination polymerization ........................................... 5

5 Branching in polyethylene ........................................................................... 7

6 The polyethylene crosslinking process ........................................................ 11

7 Byproducts of crosslinking polyethylene .................................................... 12

8 The structure of acetophenone ................................................................... 21

9 Determining the best composition of the hybrid functionals ....................... 25

10 The ReaxFF calculation process .............................................................. 33

11 The eReaxFF calculation process ............................................................. 36

12 The structure of acetophenone, again ....................................................... 38

13 Equilibrium mixture obtained from CDFT .............................................. 40

14 Equilibrium mixture obtained from eReaxFF-MD ................................... 41

15 The structure of acetophenone, again ....................................................... 41

vi

LIST OF TABLES

1 A comparison of XLPE and EPR ................................................................ 9

2 The FTIR peaks of polyethylene ............................................................... 14

3 Calculated ionization potentials for various basis sets ............................... 23

4 Calculated electron affinities for various basis sets ..................................... 23

5 Values of atom parameters .......................................................................... 29

6 Values of bond parameters .......................................................................... 31

7 Values of valence angle parameters ............................................................ 32

8 Values of torsion angle parameters ............................................................ 32

9 Values of non-bonded interaction parameters ............................................ 33

10 Constrained DFT energies ........................................................................ 39

11 Mulliken charges of neutral molecule and radical anion ............................ 42

vii

ACKNOWLEDGEMENTS

I acknowledge support and training provided by the Computational Materials

Education and Training (CoMET) NSF Research Traineeship (Grant No. DGE-

1449785). Any opinions, findings, and conclusions or recommendations expressed in

this publication are those of the author and do not necessarily reflect the views of

the NSF. Computations for this research were performed on the Pennsylvania State

University’s Institute for Computational and Data Sciences’ Roar supercomputer.

I would like to thank my thesis advisors Dr. Ismaila Dabo and Dr. Adri van Duin

for their guidance through the entire thesis writing process.

I would like to thank all of the members of the Dabo research group for their

support and friendship during my time at Penn State.

I would like to thank Kate Penrod and Dooman Akbarian of the van Duin research

group for their assistance with the research presented in this thesis.

I would like to thank my partner Alexandra, my parents Michael and Maria, and

my brother Benjamin for their love and support throughout my time at Penn State

and my time writing this thesis.

-M.A.B.

1 BACKGROUND AND MOTIVATION

Why This Work Matters

This first section serves as an introduction to the work as a whole and explains why

the research presented is necessary and novel. This section begins by discussing the

need for high-voltage cables of quality, chronicles the rise of crosslinked polyethylene

(XLPE) as an insulating material, describes some of the pitfalls of using XLPE,

introduces the acetophenone molecule at the center of this research, and concludes by

discussing previous work, experimental and computational, done in this area. These

points will all be addressed according to their relevance to the methodology and

results sections of this work.

1.1 The Need for High-Quality High-Voltage Cables

Globalization has become the norm in all aspects of daily proceedings and interac-

tions. While all industries are on board, energy transmission and storage are among

the industries leading this charge. As an example, in 2014, up to 90% of the power

transmitted in Belgium was not for use by Belgians, but rather was being sent through

the country from Germany to France [1]. To meet these demands, the world must

rely on efficient and highly reliable methods of transmitting information and energy.

As such, the function of transmission, achieved by an assortment of cables, wires,

and antennae, requires the best equipment available, which has led to the push for

improvements in power cables. Alternating current (AC) and direct current (DC)

options exist for the high-voltage cables that are needed for modern electrical grids,

where high-voltage in this context is defined as any voltage exceeding 1000 volts,

although most of the cables in question are used for voltages exceeding 10,000 volts

[2].

The main drawback to high-voltage alternating current cables is that they cannot

be used for distances greater than 100 kilometers due to appreciable power loss from

inductive and capacitive effects, pushing high-voltage direct current (HVDC) cables

to the forefront [1]. This advantage becomes more pronounced when looking at the

renewable energy sources that promise to power the future. Most sustainable power

sources tend to be very far away from the population centers they serve. A telling

example is the Three Gorges Dam that spans the Yangtze River in China, which

generates 300 petajoules annually and sends a large portion of that energy to major

2

Figure 1: An illustration of the UHV direct current cable network, planned and extant, inChina circa 2015. Note the Xiangjaba Dam, China’s third-biggest hydropower dam, witha direct line to Shanghai roughly 1500 km away, an even greater distance than power fromthe aforementioned Three Gorges Dam must travel. Figure taken from [3].

cities such as Shanghai and Guangdong, both around 1000 kilometers away, thanks

to an HVDC link provided by Swedish-Swiss electrical engineering and robotics firm

ABB [1].

In Figure 2, the primary components of a high-voltage cable are shown. Note

that the structures of alternating current and direct current cables are the same, only

the physics and testing procedures are different. At the center of the cable is the

conductor, which is made of either aluminum or copper and which can be solid or

made of strands, as seen in the figure. Cables can have one conductor or multiple

conductors in their core, but if multiple cores are used, care must be taken to place

them so as to avoid electric stress, that is, excessive pressure on one conductor from

the net electric field of the multiple conductors [4]. Every high-voltage cable has

two semiconductive shields, the conductor screen between the conductor and the

insulation, which is typically made of crosslinked polyethylene and will be the focus

of a later part of this introduction, and the insulation shield between the insulation

and the sheath.

The primary purpose of the shields is to prevent partial discharges, dielectric

breakdowns that could deteriorate the insulating material [4]. When the electric

3

Figure 2: A diagram of the co-axial structure of a power cable. The principal componentsare, from A to F, the conductor, the conductor screen, the insulation, the insulation shield,the metal sheath, and the oversheath. Figure adapted from [1].

field strength around a conductor is not high enough for an electric breakdown but

still high enough to form a conductive region around a conductor, the system is

susceptible to corona discharges, which lead to economically significant power losses

[5]. The semiconductive shields ameliorate this by reducing the potential gradient

over the surface of the conductors and inside the metal sheath, regulating the electric

field around the insulation [4]. Carbon black (CB) is a necessary ingredient in these

shields for achieving appropriate conduction, so CB-filled ethylene copolymers, such

as ethylene vinyl acetate and ethylene ethyl acrylate are common choices for shielding

materials [4].

One or more sheaths form the exterior of a high-voltage cable. The inner sheath

that forms a contact with the insulation, often made of lead, copper, aluminum, or

combinations of two or three of those, serves as a grounded layer and will conduct

leakage currents [2]. Depending on the cable, the sheath in contact with the insulation

may also act as a moisture barrier, important for the many cables used underground

and underwater. An exterior sheath, such as the one seen in Figure 2, may be used as

additional protection for the inner workings of the cable against various environmental

hazards. Polyvinyl chloride (PVC) has historically been used for the outer jacket, but

the 21st century has seen the rise of halogen-free, low-fire-hazard compounds such as

ethylene vinyl acetate with aluminum trihydrate and other additives.

4

1.2 The Rise of Polymers as Insulating Materials

No component of high-voltage cables has gone through as many notable changes

or is as germane to the work presented here as the insulating component. Before

the modern era of polymer insulation began in the early 1960s, underground power

cables were insulated with oil and paper that ran through a steel pipe or an aluminum

sheath similar to the outer jackets used in modern cables. The oil had to be kept

under pressure to avoid the formation of voids that could lead to partial discharges

[2]. While this kind of insulation is still in use the world over, Figure 3 illustrates the

superiority of polymer-based insulation and the room available for future improvement

after catching up so quickly to the paper-based design.

Figure 3: The evolution of insulating materials for high-voltage cables. Note the jump inworking voltage for polymer-based insulation in the early 1960s. This is due to the use ofcrosslinked polyethylene, which will be discussed at length in this chapter. Figure takenfrom [6].

The first polymer to be widely used for insulation was polyethylene. Polyethylene

is a polymer chain produced by the polymerization of ethylene gas, C2H4. This

polymerization can be done in a few ways, each of which lead to different properties

of the resultant polymer. Coordination polymerization uses metal chlorides or metal

oxides as catalysts, which leads to a linear structure called high-density polyethylene.

For polyethylene production, the most commonly-used catalysts are the Phillips-

type catalyst, chromium oxide deposited on silica, and the Zeigler-Natta catalysts,

which generally have two components, such as TiCl4 and Al(C2H5)2Cl [7]. When the

Phillips-type catalyst is used, the polymerization process works by introducing carbon

5

dioxide to reduce Cr(VI) to Cr(II) which become organo Cr(III) sites upon exposure

to ethylene. This initiates the polymerization process, which continues until chain

termination occurs by the addition of molecular hydrogen or until the center of activity

is transferred to another molecule via beta-hydride elimination, or the conversion of an

alkyl group bonded to the chromium center to the corresponding chromium hydride

and an alkene. The exact mechanism for the polymerization process is the subject of

active research [8] [9] [10].

When the Zeigler-Natta catalysts are used, one component serves as the initia-

tor, or precatalyst, and one serves as the activator, or cocatalyst. In the case of

polyethylene, TiCl4 is the initiator and Al(C2H5)2Cl is the activator in the most

popular configuration [8]. The activator has three main roles in polymerization: ac-

tivation, stabilization, and scavenging. Activation ionizes of the metal by removal of

the halide and replacement of the alkyl group in the transition metal, the conjugate

base stabilizes the cationic species, and any excess activator scavenges the polymer-

ization medium, which is typically made of popular solvents like toluene [8]. Even

though use of the Phillips-type catalyst does not require an activator, alkylaluminum

complexes are sometimes used to scavenge the polymerization medium [11]. Figure 4

illustrates the two aforementioned catalysts.

Figure 4: The two primary types of catalysts used for coordination polymerization of ethy-lene, the Phillips-type on the left and the titanium chloride initiator bonded to the aluminumethylchloride on the right. Note the open coordination site on the titanium, which could beused to bond an ethyl group. This figure adapted from [8].

Free-radical polymerization forms polymers by the continuous addition of free-

radical chain links, chain links that bear an unpaired electron. The three steps of any

free-radical polymerization process are initiation, the formation of radicals followed

by a reaction with a monomer to create an active center for growth, propagation, the

6

rapid addition of monomers to the growing polymer chain without a change of the

active center, and termination, the destruction of the active center, which is inevitable

due to the high reactivity of radical species [12] [13]. The polymerization of ethylene

begins with the dissociation of organic peroxides to create free radicals. Ethylene

is added to a free radical to initiate polymerization and this process is repeated

until termination occurs due to coupling of radicals or disproportionation reactions,

in which a molecule is simultaneously oxidized and reduced [14]. Examples of these

reactions are below. The R’H symbol in the chain transfer reactions represents a chain

transfer agent such as hydrocarbons that may be used to lower molecular weight [14].

Initiation

ROOR −→ 2RO•

2RO• + CH2 = CH2 −→ ROCH2CH•2

Propagation

ROCH2CH•2 + (x+1) CH2 = CH2 −→ ROCH2CH2(CH2CH2)xCH2CH

•2 ≡ Rp•

Termination

Coupling: Rp• + Rp• −→ Rp −Rp

Disproportionation: ∼ CH2CH•2 + •CH2CH2 ∼−→∼ CH = CH2 + CH3CH2 ∼

Chain Transfer

∼ CH2CH•2 + CH2 = CH2 −→ ∼ CH = CH2 + CH3CH

•2

∼ CH2CH•2 + CH2 = CH2 −→∼ CH2CH3 + CH2 = CH•

∼ CH2CH•2 + R′H −→ ∼ CH2CH3 + R′•

Free-radical polymerization results in a branched structure, featuring long branches

and shorter offshoots thereof, referred to as low density polyethylene (LDPE). Long

chain branching comes about due to intermolecular transfer between a radical and

an internal carbon atom of a different chain. Short chain branching is the result

of transfer of a radical from a terminal carbon atom to an internal carbon atom in

the same molecule, which is often called “backbiting” [14]. Most of the short chain

branching in LDPE is done among ethyl, butyl, and 2-ethylhexyl groups [15] [16].

In contrast, coordination polymerization generates a linear structure known as high

density polyethylene (HDPE). A comparison of the degrees of branching seen in the

two structures is depicted in Figure 5.

7

Figure 5: The branching structures seen in LDPE and, scarcely, in HDPE. The branchingin LDPE is a result of free-radical polymerization and the lack of a catalyst that supervisesthe radical sites on the growing chains. As such, ethylene monomers may attach to themiddle of the chain.

The terms high density and low density are not completely accurate, as their

densities are not very different (that of HDPE ranges from 930 to 970 kg/m3 and that

of LDPE ranges from 917 to 930 kg/m3 [17]). Instead, their designations come from

their branching or lack thereof. The absence of branching in HDPE gives it a more

crystalline structure, stronger intermolecular forces, and a higher tensile strength,

while LDPE remains more flexible. As far as using polyethylene for insulation in

high-voltage cables, it is recommended that underground cables use HDPE and all

other cables use LDPE [18].

8

1.3 The Process of Crosslinking Polyethylene

Polyethylene is still widely used, but as far as applications for high-voltage cable

insulation are concerned, it has some severe limitations. The issue with polyethy-

lene, in all its varieties, is that it is a thermoplastic. Thermoplastic polymers become

pliable at elevated temperatures, as the intermolecular forces that characterize poly-

mer chains become weak [19]. Polyethylene initially succeeded because, compared

to the aforementioned paper-based insulation used at the time, it could be produced

at a lower cost and offered better flexibility and dielectric performance. However,

its operating temperature was limited to 70°C, below the temperature rating of the

paper-oil systems. This problem was addressed with the introduction of crosslinked

polyethylene (XLPE), which is a thermoset, a polymer that has been irreversibly

hardened, leading to the formation of rigid bonds in three dimensions [20]. XLPE

provides all the advantages of polyethylene and can be used as an insulating material

at long-term operation temperatures up to 90°C with a short-circuit rating of 250°C,

making it appropriate for use in direct voltage cables that transmit several hundred

kilovolts [21].

To appreciate the ubiquity and performance of XLPE, it is worth comparing it

to another common high-voltage cable insulating material, ethylene propylene rub-

ber (EPR). The most attractive feature of EPR is its resistance to environmental

wear from weather and solvents, issues that sometimes plague XLPE, particularly

wear from moisture [6]. This resistance is primarily due to the fact that ethylene

propylene copolymers (EPMs) are nonpolar and fully saturated, unable to accept any

extra hydrogens because all bonds in the molecule are single bonds [4]. Every EPR

insulating material is made up only of about 50% base polymer, with the rest being

filler of various types. The addition of fillers generally improves mechanical strength

and smoothness of the material, but may compromise electrical properties (the afore-

mentioned carbon black, a common filler in rubber products, cannot be used for this

reason), so they must be chosen carefully. Two common filler materials for high-

voltage cable insulation are treated clay, which improves mechanical strength, and

hydrated alumina, which can absorb heat very well and raise the operating temper-

ature of the cable [4]. Still, XLPE rates as the superior material by most measures,

as can be seen in Table 1.

While the crosslinking process improves the quality of polyethylene for use in high-

voltage cables, it also may have adverse effects on cable performance. The steps of the

9

Properties XLPE EPRDensity (g/cm3) 0.92 1.2-1.4

Tensile strength (MPa) 19 9-12Heat distortion (%) 20 5-8

Dissipation factor (%)at 20°C <0.03 0.16-0.3at 90°C <0.03 0.3-1.0

Volume resistivity (Ω-cm) at 23°C 1016 1013

Table 1: A comparison of some key properites of XLPE and EPR as insulatingmaterials for high-voltage cables. The ranges given for EPR measurements representthe range of compositions available industrially [4].

crosslinking process are similar to the steps involved in the polymerization process:

initiation, propagation, branching, and termination. The initiation step consists of

the generation of free radicals, which bond to create a dense network of polymer chains

in the propagation and branching steps until termination occurs due to the presence

of impurities or additives [22]. The branching necessary for crosslinking to occur is

more likely to happen if some kind of branching is already present, so low-density PE

is more likely to be used than the linear structure of high-density PE [23].

Initiation can be done via irradiation or via chemical reaction. In the former

method, the free radical is created when a radiation source breaks a C-H bond in

polyethylene, which in this case has a dissociation energy of 364 kJ/mol [22]. The

carbon in the free radical that has just lost its bonded hydrogen then seeks out

another hydrogen, which leads to crosslinking. Different applications require different

radiation sources and for cable insulation, the preferred source is an electron beam.

Different types of radiation produce different depths of penetration into the polymer

layer, so specific areas of the sample can be excluded from the crosslinking [24].

Penetration depth is governed by Equation 1, in which r is the rate of crosslinking, k

is a rate constant, c is the concentration of the photoinitiator, and ε is the extinction

coefficient of the photoinitiator [25].

r = k c12 e−1.151εcx (1)

While these reactions are conceptually simple and require no additives, radiation

sources are a significant monetary investment.

10

Initiation via chemical reaction comes in two varieties: organic peroxide-based

and silane-based. Peroxide crosslinking is frequently done via the Engel process, de-

veloped in the 1960s, which begins with polyethylene and the peroxide being mixed

together at low temperatures, below the peroxide decomposition temperature. Then,

the temperature is raised, and the peroxides break apart to become peroxide radi-

cals. These free radicals remove hydrogen atoms from the polymer chain, creating

more radicals. When all these radicals combine, crosslinking occurs until the supply

of peroxide is exhausted [22]. The Engel process is typically done with HDPE at

temperatures between 200 and 250°C, though LDPE could also be used at a lower

range of temperatures [26].

Extensive experimental work has been done to determine which organic perox-

ides are best suited for crosslinking of polyethylene. One of the seminal papers in

this field is the 1993 work done by Bremner and Rudin [27], comparing three dialkyl

peroxides for crosslinking with linear low-density polyethylene (LLDPE), a type of

PE produced by initiation with transition metal catalysts that has many of the prop-

erties of LDPE, but minimal branching [8][28]. Dicumyl peroxide, a popular choice

because of its favorable decomposition rate at typical crosslinking operation tem-

peratures [27], was compared to 2,5-dimethyl-2,5-di (tertiary butylperoxy)-hexane, a

common initiator for the polymerization of styrene, and 2,5-dimethyl-2,5-di (tertiary

butylperoxy)-hexyne-3. The three were evaluated according to their ability to in-

crease the storage modulus of the PE resin during curing, which is a proxy for their

crosslinking effectiveness. This work established a link between the rate of peroxide

decomposition and the rate of increase of the storage modulus of the resin during

curing and confirmed that dicumyl peroxide is effective for crosslinking polyethylene

[27].

The silane-based process begins with a silane molecule such as vinyl trimethoxysi-

lane (C5H12O3Si) grafting to a polyethylene chain [22]. For grafting to occur, a free

radical must appear somewhere on the polyethylene chain. This can be accomplished

by irradiation or in the presence of a small amount of a peroxide, usually dicumyl

peroxide. Once the polymer is Si-functionalized, it is placed in a water bath and

silicon hydroxide groups are formed by hydrolysis. Crosslinks are then formed on the

polyethylene by Si-O-Si bridges. These reactions can be accelerated by the use of a

catalyst such as dibutyltin dilaurate [22] [16]. A comparison of the three methods of

crosslinking is shown in Figure 6.

11

Figure 6: The reactions involved in each of the three crosslinking methods for polyethylene.The peroxide-based method gives the highest degree of crosslinking, up to 90%, with theirradiation-based method between 34 and 75% and the silane-based method between 45 and70%. This figure adapted from [22].

1.4 The Complications of Crosslinking Polyethylene

The primary drawback of XLPE is the formation of byproducts during the crosslinking

process that can compromise its electrical properties. Most studies have observed the

formation of acetophenone, cumyl alcohol, and methyl styrene, with at least one study

observing the formation of 2,4-diphenyl-4-methyl-1-pentene [29]. The steps involved

in crosslinking polyethylene with dicumyl peroxide have been described in detail and

Figure 7 illustrates those reactions, along with the secondary reactions that produce

the aforementioned byproducts. It can be seen in Figure 7 that acetophenone is

produced by the beta scission that breaks a methyl group off of the free radical,

cumyl alcohol is produced when the free radical interacts with the polyethylene, and

methyl styrene is produced upon the dehydration of cumyl alcohol.

The presence of these byproducts is known to lead to space charge accumulation,

in which excess electric charge creeps into the dielectric and distorts the electric

field. This can only happen in dielectrics, including vacuum, because in a conductive

material, the charge would be rapidly neutralized. The two types of space charge

accumulation are hetero charge, in which the polarity of the excess charge is opposite

that of the neighboring electrode, and homo charge, in which the inverse situation is

true [30]. Hetero charge near an electrode has been shown to be destructive in high-

voltage applications- it stresses the insulation around the electrode area by a factor of

12

Figure 7: Reaction map for the crosslinking of polyethylene illustrating how acetophenone,cumyl alcohol, methyl styrene, and methane are formed. This figure adapted from [29].

two over what would occur in the absence of excess charge and lowers the breakdown

voltage of the polymer (homo charge would increase it), making an insulating material

more susceptible to conductivity [30] [31].

In a semiconducting or insulating medium, the current is dominated by charge-

carriers injected through the electrodes and not the charge-carriers present in ohmic

materials like metals. As such, space charge current can be described by the Mott-

Gurney law (Equation 2), in which J is the current density, ε is the permittivity

of the solid, µ is the charge-carrier mobility, Va is the applied voltage and L is the

thickness of the sample. While the Mott-Gurney law is conceptually advantageous

because the only unknown is the charge-carrier mobility, it rests on many assumptions

that may not be valid for all applications [32] [33]. The most relevant to crosslinked

polyethylene is the assumption that the current cannot be limited by energetic traps,

but a few studies have proposed that the structures of acetophenone and methyl

styrene, namely the benzene ring and double bond present in both, present energetic

traps, making the Mott-Gurney law a less-than-ideal description [34] [35]. Since

13

acetophenone is the primary subject of the work presented in this document, this

idea will be revisited.

J =9

8εµV 2a

L3(2)

In early studies of space charge accumulation, effects on XLPE were separated

into four categories: the crosslinked polymer, a non-crosslinked reference, which is

essentially the base polymer, additives such as antioxidants, and residual byprod-

ucts [35] [36]. To examine charge buildup in each of these sections, charge carriers

were inserted into the XLPE sample via two electrodes and were allowed to diffuse

throughout the body of the polymer. After some time, spatial charge density profiles

were examined for evidence of a buildup of charges of the same sign (homo charge) or

charges of the opposite sign (hetero charge) around the cathode and anode. According

to charge density measurements, the non-crosslinked part shows some development

of heterocharge, which would be at odds with the theory that hetero charge is in-

troduced by the byproducts of crosslinking, but homo charge mostly takes over in

the high-voltage regime. These same results support the idea that the formation of

heterocharge is discouraged by the presence of the antioxidant and encouraged by

the presence of byproducts, which the authors of that study posit is due to the low

molecular weight of the products, which leads to them being more easily charged and

more likely to drift towards the counter electrode under DC stress [36].

The issue of space charge accumulation can be addressed by degassing, the re-

moval of dissolved gases, often via heating. Degassing should be done to eliminate

the flammable methane byproduct and it has been shown that the removal of gasses

can relieve some pressure on the cable insulation that might build up when an exter-

nal sheath is in place, as is often the case for high-voltage cables [21]. This pressure

relief in turn increases the stress necessary for space charge to begin leaking into the

insulating material [35]. Still, even though the operating threshold of the cable might

be improved by degassing, care must be taken to ensure that the insulating perfor-

mance of the material is not seriously compromised by the presence of crosslinking

byproducts.

14

1.5 A Survey of Experimental and Computational Studies

The last section of this introductory chapter is a lead-in to the methodology chapter of

this document, as it gives a brief overview of previous experimental and computational

studies that pertain to most of the concepts and phenomena described in this section.

An important experimental task is to be able to distinguish between different

grades of polyethylene using spectroscopy. It has been shown that Fourier-Transform

Infrared (FTIR) spectroscopy, in which a beam containing several frequencies of light

is shone on a sample, the degree to which the beam is absorbed by that sample is

measured by the displacement of a mirror, and those displacement measurements are

transferred back to the frequency domain, is an appropriate tool for this. Differ-

ent varieties of PE all show the same peaks in a spectrograph, but the degrees of

branching that differentiate them are reflected in the relative intensities of the peaks

[37]. The FTIR peaks of polyethylene are shown in Table 2. The ability of FTIR to

detect such subtleties has also proved useful in analyzing possible byproducts of the

crosslinking process. One study found that the universally utilized dicumyl peroxide

caused partial oxidation of the polymer, which the authors posited could be stopped

by the introduction of the antioxidant Irganox 1081 [38].

Band (cm−1) Assignment2919 CH2 Asymmetric Stretching2851 CH2 Symmetric Stretching

1473 and 1463 Bending Deformation1377 CH3 Symmetric Deformation

1366 and 1351 Wagging Deformation1306 Twisting Deformation1176 Wagging Deformation

731-720 Rocking Deformation

Table 2: The primary absorptions of polyethylene in the IR region and what theycorrespond to [37].

FTIR has been of use in studying the aging and degradation of polyethylene in

high-voltage cables. For this, FTIR can be enhanced by a pairing with attenuated

total reflectance (ATR) sampling, which allows samples to be studied in the liquid

or solid state without any preparation by use of a crystal in contact with the sample

that absorbs the beam from the spectrometer and forms an evanescent wave that

passes into the sample [39]. FTIR can monitor the polymer over time and is capa-

ble of detecting degradation byproducts, mainly ketones, esters, and lactones [40].

15

Similar observations can be made using magic angle spinning hydrogen-based nu-

clear magnetic resonance, or H-MAS NMR, in which the sample is rotated at the

resolution-improving “magic angle” of 57.74 degrees relative to the external magnetic

field generated by a hydrogen nucleus. Perturbations of the nuclear spins of the

sample correspond to radio frequencies, which can be assembled to form a spectrum

[41] [42]. H-MAS NMR studies have been able to detect the aforementioned aging

products, as well as significant quantities of aldehydes [43].

Extensive computational work has been done on all aspects of crosslinked polyethy-

lene. One such study looked at the possibility of using gold atoms in the crosslinking

process, creating carbon-gold-carbon bridges to serve as the backbone of the polymer.

The use of metal atoms leads to dispersion interactions during crosslinking, which were

modeled with density-functional theory (DFT), one of the key methods used in the

work presented in this document. The DFT energetics revealed reasonable binding

energies [44]. Another study imitated the methods of the FTIR- and NMR-based

works by monitoring microstructural changes that occur during crosslinking with a

Monte Carlo simulation. The extensive branching of polyethylene used for crosslink-

ing makes it a good target for the probabilistic nature of Monte Carlo simulations [45].

The use of Monte Carlo has been used to generate branched and crosslinked struc-

tures, which are then subject to DFT energetics calculations to obtain an accurate

picture of electronic states of polyethylene [46].

Electronic states were of interest to a few studies that used molecular dynamics

(MD) to investigate the presence of energy traps in polyethylene, with one study

asserting the acetophenone byproduct as an energy trap based on its influence on

the density of states and another looking at electron affinity in an environment with

various chemical defects [47][48]. MD has also been used to analyze the strain asso-

ciated with crosslinking and scission reactions in several different polymers [49] [50].

Possibilities for improving the crosslinking of polyethylene have been studied with

DFT, optimizing the bond lengths and calculating the energies associated with graft-

ing voltage stabilizers to the polymer chain [51]. The methods used in the original

research presented in this document draw inspiration from all the aforementioned

studies and will be discussed in the next chapter.

16

2 METHODS

How This Work Was Done

This second section serves to explain the methodology used for the research pre-

sented in this work. This section begins by discussing the key equations of density-

functional theory (DFT) and the improvements made by the development of con-

strained density- functional theory (cDFT), then will describe the process of conver-

gence testing necessary to find the right basis set and exchange-correlation functionals,

in order to have an appropriate level of approximation in the DFT calculations. Fol-

lowing this, the process for calculating an equilibrium mixture of electronic states at

finite temperatures based on DFT energetics will be detailed. These steps will all be

repeated for the reactive force field calculations done by the author. These points will

all be addressed according to their relevance to the introduction and results sections

of this work.

2.1 The Essentials of Density Functional Theory

The starting point for the development of density-functional theory (DFT) is the

assertion that the fundamental properties of materials can be determined and pre-

dicted solely on the basis of quantum mechanical considerations, without knowing

any empirical material properties such as specific heat capacity or tensile strength.

To start solving these kinds of problems, one could consider the many-electron time-

independent Schrodinger equation (Equation 3), in which H is the Hamiltonian op-

erator of the many-body system, Ψ is the wavefunction of the system, and E is the

total energy of the system.

H |Ψ〉 = E |Ψ〉 (3)

The central problem associated with this kind of calculation is that it quickly becomes

prohibitively complex. This becomes clear when looking at a more complete version

of the Schrodinger equation, seen in Equation 4, in which m is the electron mass and

the three terms in brackets are, from left to right, the kinetic energy of each electron,

the electrostatic interaction energy between each electron and all the nuclei, and the

electrostatic interaction energy between different electrons [52].[− h2

2m

N∑i=1

∇2i +

N∑i=1

V (ri) +N∑i=1

∑j<1

U(ri, rj)

]|Ψ〉 = E |Ψ〉 (4)

17

For a concrete example, consider the benzene ring featured in the acetophenone

molecule, which has 12 nuclei and 42 electrons. In three dimensions, this would lead

to a wavefunction with (3 x 12) + (3 x 42) = 162 variables. It is known that the

computational complexity necessary to solve an eigenvalue problem increases faster

than the square of the number of coordinates and often at a cubic rate [53]. So, the

best-case scenario for this calculation is that it would require 1622 = 2 × 104 steps.

The number of steps can be reduced to some extent with the Born-Oppenheimer

approximation, which accounts for the substantial differences in mass and time scales

of motion of the electrons and the nuclei by expressing the wavefunction for the

system as the product of an electronic and a nuclear wavefunction, eliminating cross-

terms in the Hamiltonian [54]. This breaks the calculation down into two steps, first

a calculation with the 126 electronic variables (104 steps), then a calculation that

applies this result to each of the 36 nuclear variables (103 steps). This is a marked

improvement, but is still computationally impractical.

Another inconvenience is that wavefunctions cannot be observed directly. Thank-

fully, it is possible to measure the probability that the N electrons in the system are at

particular coordinates. This is further simplified by the fact that it not important to

distinguish between individual electrons, so the quantity of interest is the probability

that a set of N electrons in any order have the coordinates r1. . . rN . The physical

quantity that matches this description well is the electron density as a function of po-

sition, n(r) [52]. Equation 5 shows the electron density in terms of individual electron

wavefunctions. A factor of two arises due to the Pauli exclusion principle.

n(r) = 2∑i

ψ∗i (r)ψi(r) (5)

The first model to use the electron density in this way was the Thomas-Fermi

model, developed independently by Llewellyn Thomas and Enrico Fermi in 1927 [55].

They started with an electron distribution that was uniform across phase space, with

two electrons per h3. For every differential element of coordinate space volume d3r,

they filled a sphere of momentum space up to the Fermi momentum pF to obtain an

electron density in coordinate space that takes the form

n(r) =8π

3h3p3F (r). (6)

Solving this equation for the Fermi momentum and plugging it into the classic expres-

18

sion for kinetic energy, K = 12mv2, leads to kinetic energy as a functional (function

of a function) of the electron density, seen below in Equation 7 [55].

K[n] =3h2

10me

(3

) 23∫n

53 (r) d3r (7)

This equation was combined with classical expressions of electron-nuclear and electron-

interactions to obtain a total energy. However, this kinetic energy functional was only

a rough approximation and density-functional theory did not stand on firm ground

until the work of Walter Kohn and Pierre Hohenberg in the 1960s.

Kohn and Hohenberg were able to lay the foundation for DFT with their proofs of

two fundamental theorems [56]. The first says that the ground-state energy obtained

from the Schrodinger equation is a unique functional of the electron density, which is

equivalent to saying that the ground-state electron density determines all properties

of the ground state. This is important because it reduces the problem of solving

the Schrodinger equation (i.e. finding the ground state energy) from a problem with

potentially thousands of variables to a problem with three variables. The second

theorem gives some more information about the functional, stating that the electron

density that minimizes the overall functional is the true electron density corresponding

to the full solution of the Schrodinger equation. This leads to what is called the

variational principle, the idea that if the true form of the functional were known,

the electron density could be varied until the energy is minimized and the electron

density could be determined in this way [57].

The functional described by the Hohenberg-Kohn theorems can be written in

terms of the single-electron wavefunctions ψi(r) seen in Equation 8 which, again,

define the electron density. This energy functional can be written as

E[ψi] = Eknown[ψi] + EXC [ψi] (8)

in which the “known” terms are those that can be written down in a simple ana-

lytical form and the exchange-correlation (XC) terms are everything else (exchange

correlation functionals will be discussed in detail later in this chapter). The known

terms can be seen below in Equation 9 and are, from left to right, the electron kinetic

energies, the Coulomb interactions between electrons and the nuclei, the Coulomb

interaction between pairs of electrons, and the Coulomb interaction between pairs of

19

nuclei [52].

Eknown[ψi] = −h2

m

∑i

∫ψ∗i∇2ψid

3r +

∫V (r)n(r)d3r

+e2

2

∫ ∫n(r)n(r′)

|r − r′|+ Eion (9)

Even with these terms and a reasonable description of the exchange-correlation

terms, the act of solving the Schrodinger equation is not made any easier. Kohn and

Sham addressed this, showing that the correct electron density could be found by

solving a system of equations in which each equation only involves a single electron

[58]. These Kohn-Sham equations take the form[− h2

2m∇2 + V (r) + VH(r) + VXC(r)

]ψi(r) = εiψi(r). (10)

This is very similar to the Schrodinger equation, except there are no summations

because the solutions of the Kohn-Sham equations are single-electron wavefunctions

that only depend on the three spatial variables. The first of the three potentials, V in

Equation 10 is the same potential as in the full Schrodinger equation and defines the

interaction between an electron and the nuclei surrounding it. The second potential,

VH is called the Hartree potential and is expressed as

VH(r) = e2

∫n(r)′

|r − r′|d3r′. (11)

This potential describes the Coulomb interaction between the electron featured in

a single Kohn-Sham equation and all other electrons in the system. The Hartree

potential includes some degree of self-interaction, since every individual electron being

considered is also one of the electrons in the system. The correction for this self-

interaction is included in the third potential, VXC , which defines the contribution

of exchange interactions, interactions that occur between indistinguishable particles,

and correlation interactions, the influence of all electrons on the movement of one

electron, to single-electron equations. VXC is defined as the functional derivative

(conceptually very similar to a regular derivative) of the exchange-correlation energy:

VXC(r) =δEXC(r)

δn(r). (12)

20

The steps involved in these calculations form a cycle. To solve the Kohn-Sham

equations, the Hartree potential must be defined and to define the Hartree potential,

the electron density must be known. But knowing the electron density requires know-

ing the single-electron wavefunctions, which are obtained by solving the Kohn-Sham

equations. Sholl and Steckel’s seminal DFT book describes the steps necessary for

calculations in the self-consistent cycle [52]:

1. Define a trial electron density n(r)

2. Solve the Kohn-Sham equations using the trial electron density to find the

single-electron wavefunctions, ψi(r)

3. Calculate the electron density defined by the Kohn-Sham single-particle wave-

functions

4. Compare the calculated electron density with the electron density used to solve

the Kohn-Sham equations. If they are the same, then this is the ground-state

electron density that can be used to calculate the total energy. If not, the trial

electron density must be changed in some way and this process starts again at

Step 2.

The next section of this chapter explains how to apply this self-consistent procedure

to the acetophenone system that is the focus of this thesis.

21

2.2 Constrained Density Functional Theory

Since the advent of DFT, many methods have been developed to improve and build

upon it. One of the most versatile is constrained density functional theory (CDFT)

[59]. One CDFT formalism was laid out by Dederichs et al in the 1980s [60], with

the idea of finding the electronic ground state of a system subject to the constraint

that there are N electrons in a volume Ω. This was achieved by adding a Lagrange

multiplier to the familiar DFT energy functional form:

E(N) = E[n] + λ

(∫Ω

n(r) d3r −N)

(13)

The addition of this Lagrange multiplier was enough to ensure that this calculation

would yield the lowest-energy state that contained exactly N electrons in a volume Ω.

This tuning of the functional also helps mitigate the aforementioned self-interaction

error that plagues many DFT functionals [59].

Figure 8: The acetophenone molecule, adapted from [61].

This formalism is the basis for much of the work presented in this thesis. The

previously-discussed acetophenone molecule that forms as a byproduct in peroxide-

based polyethylene crosslinking reactions is the species of interest here. Since it

is well established that the presence of byproducts compromises the integrity of the

cable, it is worth investigating how acetophenone behaves when exposed to free charge

carriers. As such, CDFT has been used to study the acetophenone radical anion, with

the electron being constrained to each of the 17 atoms in the molecule (seen above

in Figure 15). The DFT energies of each of these constrained states were compared

22

to each other to determine the locations on the molecule most likely to attract the

excess electron. This allows inferences to be drawn about possible chemical reactions

involving the acetophenone radical anion. The next several sections of this chapter

describe the details of these energy calculations.

2.3 Choosing a Basis Set

The first step in these calculations is selecting an appropriate basis set, the set of

functions used to represent the electronic wavefunction. The basis sets considered

for the work presented here include Pople basis sets [62] and correlation-consistent

basis sets [63]. Pople basis sets are described as “split-valence” basis sets because

they represent valence orbitals, which contain the valence electrons that primarily

participate in bonding, with multiple basis functions. These basis sets are of the

form X-YZG, in which X represents the number of primitive Gaussians (G being

Gaussian) making up each core orbital basis functions, and the Y and Z indicate that

the valence orbitals are made of two basis functions each, one a linear combination of

Y primitive Gaussians and the other a linear combination of Z primitive Gaussians.

If the valence orbitals are described by three basis functions, the basis set may take

the form X-YZWG.

Additionally, polarization functions (denoted by * or ** if polarization functions

are on hydrogen and heavy atoms) may be added to describe the polarization of the

electron density of a particular atom and diffuse functions (denoted by + or ++ if

diffuse functions are on hydrogen and heavy atoms) may be added to better describe

the “tail” of the orbital function, far away from the nucleus.

The correlation-consistent basis sets were designed for converging Post-Hartree-

Fock calculations, calculations that describe correlation between electrons instead of

simply repulsion. They are typically written as cc-pVNZ, in which cc-p stands for

correlation-consistent polarized, V indicates that they are valence-only, and NZ indi-

cates the number of basis functions that make up each orbital (D=double, T=triple,

etc.). Z stands for “zeta,” which was commonly used as an exponent on early ba-

sis functions [64]. The basis sets considered for this work were 6-31G, 6-31G**,

6-311G**, and cc-pVDZ. No basis sets with diffuse functions were considered as it

is known that CDFT calculations sometimes have issues with diffuse functions (this

will be discussed further in the CDFT results section) [59].

These basis sets were evaluated with a sort of convergence test, seeing how close

23

they could get to the electron affinity (EA) and ionization potential (IP) predictions

of a much larger basis set, aug-cc-pVDZ (augmented with diffuse functions). Note

that these EA and IP values were not obtained by CDFT calculations, but by adding

whole charges to the neutral acetophenone molecule and finding the energy difference

between the -1 charge and neutral states (EA) and the between the +1 charge and

neutral states (IP). The results are shown in Tables 3 and 4 for three select functionals

that will be discussed in the following section, B3LYP, PBE0, and pure HF. Based

on this convergence test, the 6-311G** basis set was chosen.

Basis set B3LYP PBE0 Pure HF6-31G 9.0445 9.1318 7.806

6-31G** 8.9493 9.0295 7.81126-311G** 9.0465 9.1925 7.9178cc-pvdz 9.06 9.1209 7.886

aug-cc-pvdz 9.1777 9.1885 9.1127

Table 3: Values of the ionization potential for different basis sets

Basis set B3LYP PBE0 Pure HF6-31G 0.1542 0.1167 0.7658

6-31G** 0.2818 0.2402 1.03586-311G** -0.0344 -0.0261 0.8492cc-pvdz 0.0922 0.0721 0.9097

aug-cc-pvdz -0.3526 -0.3053 0.7017

Table 4: Values of the electron affinity for different basis sets

24

2.4 Tuning the Hybrid Functionals

The next step in establishing an appropriate level of theory for these calculations is

finding an acceptable exchange-correlation functional. The functionals considered for

this work improve upon Hartree-Fock (HF) exact exchange functionals by hybridizing

them with energy functionals from other sources, which is known to improve predic-

tions of many molecular properties [65]. These other energy functionals depend on

two approximations commonly employed in DFT calculations: the local density ap-

proximation (LDA) and the generalized gradient approximation (GGA).

The LDA formalism is based on the treatment of electrons in the homogeneous

electron gas and an LDA functional depends only on the density at the point in

space where the functional is being evaluated. The LDA scheme assumes that elec-

tron density is uniform, which leads to underestimating the exchange energy and

overestimating the correlation energy [66]. These errors compensate for each other

to a degree, but an improvement over LDA is the GGA scheme, in which the non-

homogeneity of the electron density is captured by writing a functional in terms of

the electron density gradient [67].

The first hybrid functional considered was B3LYP, the constituent parts of which

are seen below in Equation 14, in which EGGAx is a GGA approximation of the Becke 88

exchange functional, EGGAc is a GGA approximation of the Lee, Yang, Parr functional,

and ELDAc is the VWN LDA approximation of the correlation functional [68].

EB3LY Pxc = ELDA

x + 0.20(EHFx − ELDA

x ) + 0.72(EGGAx − ELDA

x )

+ ELDAc + 0.81(EGGA

c − ELDAc ) (14)

The three parameters that give B3LYP its name are borrowed from parameterization

done by Becke [69]. The other hybrid functional considered for this work was PBE0,

which breaks down as seen in Equation 15.

EPBE0xc =

1

4EHFx +

3

4EPBEx + EPBE

c , (15)

where EHFx is the Hartree-Fock exact exchange functional, EPBE

x is the PBE exchange

functional, and EPBEc is the PBE correlation exchange functional [70].

The parameter that needed tuning was the percentage of Hartree-Fock exact ex-

change, which avoids the self-interaction error that results from functionals with a

correlation component and leads to problems such as overestimated electron affini-

25

ties [71]. A range of HF contributions above the base percentages seen in Equations

14 and 15 were evaluated on the basis of compliance with a DFT-adapted version of

Koopmans’ theorem, which states that the negative of the energy of the highest occu-

pied molecular orbital (HOMO) should equal the ionization potential of the molecule

and the negative of the energy of the lowest unoccupied molecular (LUMO) orbital

should be equal to or slightly less than the electron affinity [72]. It can be seen in

Figure 9 that, for both functionals, this condition is met at an HF contribution of

about 65%. Changing the HF contribution means changing how much other terms

contribute to the exact exchange. For the PBE0 functional, this means that the PBE

exchange contribution is only 35% and for the B3LYP functional, since there are three

contributions to the exchange, the Becke 88 local contribution is dropped to 35% and

the Becke 88 nonlocal weight is dropped from 0.72 to 0.315.

Figure 9: Determining the best Hartree-Fock percentage for the hybrid functional by lookingfor Koopmans’ compliance. Figures A and B are the B3LYP hybrid functional and FiguresC and D are for PBE0.

26

2.5 Calculating an Equilibrium Mixture

With all the parameters for the CDFT calculations established, a mixture composition

of electronic states can be obtained. As is generally recommended, the charge and

spin are constrained in each calculation. A charge of -1, representing the electron of

the acetophenone radical anion, and a spin of 1, making this extra electron an alpha

electron, are constrained to each of the 17 atoms that make up acetophenone. A DFT

energy is obtained from each calculation, which was done with the NWChem software

[73]. The difference between this energy and the energy of the “ground state” of an

electron constrained to the entirety of the molecule is calculated and then a baseline

equal to the resulting lowest energy is subtracted, leaving one state at zero energy.

These relative energies are treated as ∆G values and the equation below, taken

from a 1997 paper by van Duin et al [74], is used to calculate the percentage of each

of the 17 states in the equilibrium mixture. Entropic effects are ignored. A mixture

is then calculated for a range of temperatures: 300, 400, 500, 600, 700, 800, 900, and

1000 K.

%Ci = 100 ×exp(

∆Gi−∆G1

RT

)1 +

ns∑n=2

exp(

∆Gn−∆G1

RT

) (16)

27

2.6 Atomic Orbital-Based Population Schemes

This thesis compared the equilibrium mixture composition obtained from CDFT en-

ergies with population analysis schemes that estimate how partial charges distribute

themselves among the atoms of a molecule. Two of the most prominent are the Mul-

liken and Lowdin schemes. The Mulliken scheme is simpler in terms of the underlying

mathematics, if less robust. In the Mulliken scheme [75], if the coefficients for the ba-

sis functions of the atomic orbitals are written as Cµi for basis function µ in molecular

orbital i, then the terms of a charge density matrix can be described as a probability

density and written as

Dµν = 2∑i

CµiC∗νi (17)

for a closed shell system in which each orbital is doubly occupied, hence the factor

of two. The matrix Sµν describes the overlap between the atomic orbitals that make

up the molecular orbitals and the product of Sµν and the density matrix Dµν is the

population matrix

Pµν = DµνSµν . (18)

The sum of all the Pµν terms over all basis functions µ is called the Gross Orbital

Product (GOP) for the atomic orbital ν. The sum of the GOPs of each orbital should

be the total number of electrons. When we sum the GOPs of all orbitals ν associated

with a given atom A, we obtain a quantity known as the Gross Atom Population

(GAP). The Mulliken charge, Q, on that atom A is then the difference between the

charge of a free atom, Z, and the GAP, expressed as

QA = ZA −GAPA. (19)

While the Mulliken population scheme is easy to understand and implement, it

suffers from an explicit dependence on the basis set. If the complete basis set for a

molecule can be spanned by placing functions on a single atom, the Mulliken scheme

will assign every electron in the system to that single atom, which is clearly unphysical

and will yield inconsistent results.

An alternative to the Mulliken scheme that reduces the basis set dependence

while not deviating too far in terms of the underlying theory is the Lowdin pop-

ulation scheme. While the Mulliken scheme does not require anything of its basis

functions, the Lowdin scheme transforms all the atomic orbitals to an orthogonal ba-

28

sis with a symmetric orthogonalization scheme that gives all the wavefunctions equal

weight in calculating linear combinations of wavefunctions to obtain the new basis

set. The transformation found by Lowdin in his seminal 1950 paper [76] to produce

an orthonormal basis set closest to the original basis is the linear transformation

S12 , where S is the overlap matrix. To examine the validity of this transformation

within the established mathematical framework, it is necessary to return to Equation

18. When the aforementioned transformation is applied, the product defining the

population matrix becomes

Pµν = (S12PS

12 )µν . (20)

The following is a well-known property of square matrices: tr(ABC) = tr(BCA) =

tr(CAB). This means that, for any value of λ, the trace of SλPSλ−1 will give the

same total number of electrons, but will give different partial traces for the different

atoms, leading to potential differences in delocalization [77].

While atomic-orbital based population schemes can be unreliable when used in

conjunction with CDFT [59], no issues were found when performing a population

analysis of the “ground state” of the acetophenone anion, that is, the state in which

the free electron is “constrained” to the entire molecule and allowed to delocalize. The

Mulliken scheme was chosen over the Lowdin scheme due to easier convergence of the

CDFT energy calculations. None of the issues with the Mulliken scheme presented

themselves in this work.

29

2.7 The Essentials of Reactive Force Field Calculations

At the turn of the 21st century, established quantum chemistry methods allowed

calculation of the geometry and energy of small molecules, but fell short when it

came to predicting the dynamics of larger molecules [78]. The addition of force fields

grants the ability to accurately describe dynamical properties, but most force fields

could not describe chemical reactivity. The exception was the Brenner potential,

which could describe bond breaking, but did not include van der Waals and Coulomb

interactions [79]. A few bond-order-based methods emerged that improved on this,

but did not fully address the need to be able to fully describe bond formation, bond

breaking, and equilibrium geometries [80] [81].

This void was filled by the development of the ReaxFF method. This reactive

force field method divides the energy of the system into several contributions, each

seen in Equation 21.

Esystem = Ebond + Ecoord + Eval + Etors

+ EvdWaals + ECoulomb + Especific (21)

The bond energy can be calculated from the bond order. In turn, one of the as-

sumptions central to ReaxFF is that the bond order can be calculated directly from

interatomic distances according to the following equation:

BOij = exp

[pbo,1 ·

(rσijr0

)pbo,2]+ exp

[pbo,3 ·

(rπijr0

)pbo,4]+ exp

[pbo,5 ·

(rππijr0

)pbo,6](22)

in which BO is the bond order between atoms i and j, rij is the distance between

atoms i and j, the r0 terms are equilibrium bond lengths, and the pbo terms are

empirically derived. The superscripts σ, π, and ππ refer to sigma bonds, pi bonds,

and second pi bonds [82].

Atom type r0,σ(A) r0,π(A) r0,ππ(A) pover(kcal/mol) punder(kcal/mol)C 1.399 1.266 1.236 52.2 29.4H 0.656 - - 117.5 -

Table 5: Atom parameters used in Equations 22, 24, and 25 [78].

30

This equation is continuous, ensuring continuity across transitions between σ,

π, and ππ bonds. This yields the differentiable potential energy surface necessary

for force field calculations. This equation also accounts for long-distance covalent

interactions in transition state structures, which allows the force field to predict re-

action barriers. This covalent range is typically taken to be 5 angstroms, but can

be expanded if elements with large covalent radii are involved. One danger with this

approach is the inclusion of false bonded character between neighboring, non-bonded

species such as the hydrogens in a methane molecule. To combat this, a correction is

typically added to the bond order [82]. This corrected bond order is used to calculate

bond energy, as seen in Equation 23, in which all the variables are the same as those

used in the bond order equation [78].

Ebond = −De ·BOij · exp[pbe,1

(1−BOpbe,1

ij

)](23)

Even after bond order corrections, the coordination of the molecule may not be

exactly right. For an over-coordinated atom (e.g. bond order of greater than four

for carbon or one for hydrogen), an energy penalty is assessed. In the case of under-

coordination, an energy term is added to account for the contributions from the

resonance of the π-electron between neighboring under-coordinated atoms. This res-

onance contribution and the over-coordination penalty are reflected in the Ecoord term

and the forms of their energies are seen in Equations 24 and 25, respectively.

Eover = pover ·∆i ·(

1

1 + exp(−8.90 ·∆i)

)(24)

Note: The ∆i terms in all of these equations refer to the difference between the bond

orders around an atomic center and its valency.

Eunder = −punder ·1− exp(1.94 ·∆i)

1 + exp(3.47 ·∆i)· f1(BOij,π,∆j) (25a)

f1(BOij,π,∆j) =1

1 + 5.79 · exp

(12.38 ·

neighbors∑j=1

∆j ·BOij,π

) (25b)

The valence angle contribution, captured in the Eval term, must go to zero as the

bond orders of the associated atoms go to zero. Equation 26a describes the energy

associated with deviations of the angle Θijk from the equilibrium angle Θ0. The bond-

31

Bond type De(kcal/mol) pbe,1 pbo,1 pbo,2 pbo,3 pbo,4 pbo,5 pbo,6C-C 145.2 0.318 -0.097 6.38 -0.26 9.37 -0.391 16.87C-H 183.8 -0.454 -0.013 7.65 - - - -H-H 168.4 -0.310 -0.016 5.98 - - - -

Table 6: Bond parameters used in Equations 22 and 23 [78].

order dependent term in Equation 26b ensures that the energy associated with the

valence angle goes to zero and bond dissociation occurs. Equation 26c describes the

effects of under- and over-coordination of the bond’s central atom j. These equations

capture the dependence of the equilibrium valence angle on bond order [78].

Eval = f2(BOij) · f2(BOjk) · f3(∆j) ·(ka − ka exp

[−kb(Θ0 −Θijk)

2])

(26a)

f2(BOij) = 1− exp(−1.49 ·BO1.28

ij

)(26b)

f3(∆j) =2 + exp(−6.30 ·∆j)

1 + exp(−6.30 ·∆j) + exp(pv,1 ·∆j)·[

2.72− 1.72 · 2 + exp(33.87 ·∆j)

1 + exp(−33.87 ·∆j) + exp(−pv,2 ·∆j)

] (26c)

The torsion angle, the angle between the two outer bonds of three consecutive

bonds, whose energy contributions are tied up in the Etors term, is also highly bond

order-dependent. The valence angle-dependent sine terms in Equation 27a ensure

that the torsion angle contribution goes to zero as either of the two valence angles,

Θijk or Θjkl, approaches 180 degrees. Equation 27b ensures that the torsion angle

contribution goes to zero as the bonds in the torsion angle dissociate and Equation

27c ensures that the torsion contribution will not be disproportionate if atoms j and

k are over-coordinated.

Etors = f4(BOij, BOjk, BOkl) · sin Θijk ·Θjkl[1

2V2 · exp

[pl(BOjk − 3 + f5(∆j,∆k))

2]· (1− 2 cosωijkl)

+1

2V3 · (1 + 3ωijkl)

(27a)

f4(BOij, BOjk, BOkl) = [1− exp(−3.17 ·BOij)]·

[1− exp(−3.17 ·BOjk)] · [1− exp(−3.17 ·BOkl)](27b)

f5(∆j,∆k) =2 + exp(−10.00 · (∆j + ∆k))

1 + exp(−10.00 · (∆j + ∆k)) + exp(0.90 · (∆j + ∆k))(27c)

32

Valence angle Θ0 ka(kcal/mol) kb((1/rad)2) pv,1 pv,2C-C-C 71.31 35.4 1.37 0.01 0.77C-C-H 71.56 29.65 5.29 - -H-C-H 69.94 17.37 1.00 - -C-H-C 0 28.5 6.00 - -H-H-C 0 0 6.00 - -H-H-H 0 27.9 6.00 - -

Table 7: Valence angle parameters used in Equation 26 [78].

Torsion angle V2 V3 plC-C-C-C 21.7 0.00 -2.42C-C-C-H 30.5 0.58 -2.84H-C-C-H 26.5 0.37 -2.33

Table 8: Torsion angle parameters used in Equation 27 [78].

Energies associated with noncovalent interactions, namely van der Waals and

Coulomb forces, are also incorporated into the system energy by ReaxFF for all

pairs of atoms. The van der Waals contribution, captured in the EvdWaals term, uses

a distance-corrected Morse potential, seen in Equation 28. The use of a shielded

interaction, seen in Equation 28b, avoids excessive repulsion between bonded atoms

and atoms sharing a valence angle [78].

EvdWaals = Dij ·[exp

(αij

(1− f6(rij)

rvdW

))− 2 · exp

(αij2

(1− f6(rij)

rvdW

))](28a)

f6(rij) =

[r1.69ij +

(1

λw

)1.69]0.592

(28b)

The system’s Coulomb energy uses a shielded potential to account for orbital overlap

between atoms close to each other, seen in Equation 29. The atomic charges qi and

qj are calculated using the electron equilibration method (EEM), which accounts for

geometry and connectivity in assigning atomic charges and, like DFT, is based on the

work of Hohenberg and Kohn [83].

ECoulomb = C · qi · qj(r3ij +

(1γij

)3)1/3

(29)

33

Atom type rvdW (A) α γW (A) γCoulomb(A)C 3.912 10.71 1.41 0.69H 3.649 10.06 5.36 0.37

Table 9: Coulomb and van der Waals parameters used in Equations 28 and 29 [78].

The last term in Equation 21, Especific, captures assorted dynamics that might

arise in particular systems. For example, an allene compound, in which a carbon

atom is double-bonded to each of its two adjacent carbon centers, requires an energy

penalty be assessed to capture its stability [78]. The relationships between these

energy terms and their roles in the ReaxFF calculation process is depicted in Figure

10.

Figure 10: An illustration of the ReaxFF calculation process [82].

34

2.8 Improvements in the Treatment of Electrons and Holes

Electrons are treated implicitly in the ReaxFF formulation, that is, only their wave-

like behavior is captured. The lack of an explicit electron description makes it im-

possible to accurately calculate electron affinities and ionization potentials, making

ReaxFF unsuitable for describing redox chemistry [84]. The eReaxFF method intro-

duces an explicit, or particle-like, electron description into the ReaxFF framework. In

eReaxFF, the electron or hole is treated as a particle that carries a -1 or +1 charge,

respectively. Nuclei are treated as point charges and charge carriers have Gaussian

wavefunctions of the form Ψ = exp(−α(r − r′)2). The electrostatic interaction be-

tween the charge carriers and the nuclei is written as

Eelec =1

4πε0β∑i,j

ZiRij

erf(√

2αRij) (30)

in which Zi is the nuclear charge, Rij is the distance between the electron and nucleus,

and α and β are constants that depend on the atom type. Interactions between

electrons are treated as Coulombic interactions between point charges.

When explicit electron/hole degrees of freedom are introduced, the valency of

and number of lone pairs on an atom must be subject to change. In eReaxFF, an

exponential function is used to describe the number of electrons on a particular atom

to ensure that a single electron can split itself among multiple atom centers. This

function is seen in Equation 31, in which Rij is the distance between the charge carrier

and the atom center and pval is a general parameter in the force field [85].

nel = exp(−pval ·R2

ij

)(31)

As seen in Figure 10, in the ReaxFF method, bonded and non-bonded interac-

tions are calculated separately. In eReaxFF, the explicit charges on an atom are

coupled with valency and the over- and under-coordination energy contributions de-

pend on this corrected valency. A decrease in the valency of an atom increases the

over-coordination in its bonding environment, increasing the over-coordination en-

ergy penalty and reducing the bond order associated with that atom. The following

set of equations (degree of over-coordination in 32a, bond type dependency in 32b,

atom type dependency in 32c, explicit electron correlation in 32d) show how electron-

corrected valencies lead to new expressions for the over- and under-coordination en-

35

ergies (32f and 32g, respectively).

∆i =

neighbor(i)∑j=1

BOij − V ali (32a)

fbond = nel ·

nbond∑j=1

BOij · pxel1ij

neighbor(i)∑j=1

BOij

(32b)

fi = exp(−pxel2i · fbond

)(32c)

∆xeli = ∆i · exp

−pxel2i · nel ·

nbond∑j=1

BOij · pxel1ij

neighbor(i)∑j=1

BOij

(32d)

∆lpcorri = ∆xel

i

− ∆lpi

1 + pcoord3 · exp(pcoord4 ·

(∑neighbors(i)j=1 (∆xel

j −∆lpj ) · (BOπ

ij +BOππij ))) (32e)

Eover =

∑nbondj=1 pcoord1 ·De ·BOij

∆lpcorri + V ali

·∆lpcorri ·

1

1 + exp(pcoord2 ·∆lpcorr

i

) (32f)

Eunder = −pcoord5 ·1− exp

(pcoord6 ·∆lpcorr

i

)1 + exp

(−pcoord2 ·∆lpcorr

i

) ·∆lpcorri

1 + pcoord7 · exp(pcoord8 ·

(∑neighbors(i)j=1 (∆xel

j −∆lpj ) · (BOπ

ij +BOππij ))) (32g)

Note that in the above equations, the pxel1ij term are bond type parameters used to

adjust the number of electrons available to the host atom, the pxel2i term is an atom

type parameter, the lpcorr superscript refers to a corrected number of lone pairs, and

the ∆xeli term represents the explicit electron correction to the degree of over- and

under-coordination ∆i. Figure 11 shows the eReaxFF calculation process.

36

Figure 11: An illustration of the eReaxFF calculation process. The bonded and non-bondedinteractions are coupled by the explicit electron/hole description [82].

2.9 Calculating an Equilibrium Mixture, Again

The first step in using reactive force field methods is training the force field, optimizing

the parameters such as those described in Tables 5, 6, 7, 8, and 9. The training set,

as described in a 1994 paper by van Duin et al [86], must have the kind of data that

the force field is looking to produce. For example, if the force field will be used to

predict enthalpies and entropies, an appropriate training set would be experimentally

determined heats of formation.

For these calculations, since the purpose is to serve as a comparison to the CDFT

data, the CDFT energies are a perfect training set. Once the hydrocarbon force field

is optimized, a series of molecular dynamics (MD) calculations was done to obtain

another equilibrium mixture. The range of temperature values was the same as that

of the CDFT calculations. The eReaxFF software produces a file called xmolout that

keeps track of the coordinates of every atom in the system being analyzed (for the

purposes of these calculations, the free electron can be treated as an atom) as the MD

simulations run. Each MD run produced 5000 frames to be analyzed. A script was

written to calculate the distance between the electron and each of the 17 constituent

atoms of the acetophenone molecule. The number of frames in which a particular

atom was closest to the electron divided by 5000 was interpreted as the “probability”

37

of the electron being found on that atom, to allow for direct comparison with the

CDFT results.

The following section will display the results of all the calculations heretofore

described and offer interpretations in the context of all the underlying principles that

have been explained in this and the preceding section.

38

3 RESULTS

What to Make of the Work Presented Here

This third and final section serves to present the results of the research conducted

by the author and offer an interpretation thereof. This section begins by reviewing

the results of the CDFT energy calculations and the equilibrium mixture composition

obtained by using those values, then will compare that composition to the results of

the Mulliken population analysis. These results will both then be compared to the

mixture composition obtained from eReaxFF-based MD simulations. Finally, some

possibilities for future directions of this research will be addressed. These points will

all be addressed according to their relevance to the introduction and methods sections

of this work.

3.1 Constrained DFT Results

Figure 12: The acetophenone molecule, provided again for reference, adapted from [61].

Table 10 shows the relative energies for each atom state for the two hybrid func-

tionals analyzed in this thesis. The B3LYP data and the PBE0 data are quantitatively

and qualitatively similar enough that the choice of which one to use as the training

set for the reactive force field is arbitrary. The B3LYP data were chosen to train the

force field, but the PBE0 energies still serve as an illustration of the variability of the

CDFT data.

39

Atom B3LYP PBE0C1 0.086187 0.079552C2 0.044706 0.045908C3 0.000982 0.001121O4 0.382458 0.380907C5 0 0C6 0.004038 0.004072C7 0.029844 0.031194C8 0.029504 0.030804C9 0.002 0.001555H10 1.425913 1.454906H11 1.547719 1.578976H12 1.547481 1.578776H13 0.608245 0.623824H14 0.673268 0.711448H15 0.697143 0.732155H16 0.716641 0.732155H17 0.70672 0.7262

Table 10: Relative energies (eV) obtained from constrained density functional theorycalculations

Looking back at Equation 16 reveals that a lower relative energy leads to a greater

probability of the electron being found around that particular atom. When looking

at the energies, it is clear that there are three distinct tiers of electron preference: the

carbons are greatly preferred, particularly the ones that make up the benzene ring,

the lone oxygen is not very likely to host the electron, and the electron wants nothing

to do with any of the hydrogen atoms, particularly the ones outside the benzene ring.

Since the carbon atoms are the only species that show any appreciable contribution

to the mixture, Figure 13 shows the contribution of each of the carbon atoms as the

temperature increases.

The data show that the distribution smoothly becomes more uniform as the tem-

perature increases, which lines up with physical expectations: the electron will have

an easier time transitioning from state to state as thermal energy rises. Indeed, the

percentage contribution of the lone oxygen state rises from less than 10−5 at 300 K

to 0.19 at 1000 K.

40

Figure 13: The percentage chance of the electron being found around each of the particularatoms in the benzene ring at varying temperature values, according to CDFT calculations

3.2 eReaxFF-MD Results

Figure 14 shows the evolution of the equilibrium mixture composition obtained from

MD simulations. As is the case with the CDFT calculations, the only appreciable

contributions to the mixture come from the carbon atoms. Since the analysis of the

MD simulations only accounts for the closest atom to the electron in each frame,

there is no guarantee that every atom will be represented. These data indicate that

the most favorable atom state for the electron is C8, different from the C5 prediction

of the CDFT data, though, again, the MD data is less uniform across the range of

temperatures.

The general trend of the mixture composition becoming more evenly distributed

across the atom states is seen again in this data, though the increases/decreases in

percentage composition are not as uniform. This is perhaps due to the more dynamic

nature of the eReaxFF-MD calculations relative to the CDFT energy calculations,

which are performed at 0 K and then assigned a temperature to be able to calculate

a mixture composition.

41

Figure 14: The percentage chance of the electron being found around each of the particularatoms in the benzene ring at varying temperature values, according to reactive force field-based MD calculations

3.3 Mulliken Population Analysis Results

Figure 15: The acetophenone molecule, adapted from [61].

Table 11 shows the charge density results of the Mulliken population analysis for

both the neutral acetophenone molecule and the radical anion. A set of differential

Mulliken charges is necessary because, although Mulliken charges bear a passing sim-

ilarity to atomic numbers, they should not be interpreted as physical charges without

context. The differential Mulliken charge can be interpreted as the effect that the

42

Atom Neutral Anion DifferenceC1 6.33 6.32 -0.01C2 5.69 5.81 +0.12C3 6.20 6.20 -O4 8.37 8.52 +0.15C5 6.08 6.15 +0.07C6 6.03 6.10 +0.07C7 6.10 6.11 +0.01C8 6.10 6.11 +0.01C9 6.09 6.19 +0.10H10 0.85 0.92 +0.07H11 0.87 0.92 +0.05H12 0.87 0.92 +0.05H13 0.89 0.95 +0.06H14 0.87 0.92 +0.05H15 0.89 0.96 +0.07H16 0.89 0.95 +0.06H17 0.89 0.97 +0.08

Table 11: A comparison of the Mulliken charges of neutral acetophenone and theacetophenone radical anion

introduction of the electron has on the occupancy of each atom. This allows a com-

parison to the CDFT equilibrium mixture, even though the orbitals of each atom are

not completely independent the way that the atom states are in the CDFT mixture

composition calculation.

The charge differentials seen here are comparable to those obtained when different

basis sets are used. For example, the same calculation with the aforementioned cc-

pvdz basis set yields very similar results, with the largest positive increase in Mulliken

charge on the lone oxygen and substantial positive increases on the hydrogens around

the benzene ring. This assuages some fear about the basis set dependence that can

plague Mulliken population schemes.

43

3.4 Discussion

It is encouraging that the CDFT and eReaxFF-MD mixture composition results are

similar, since the CDFT energies were used to train the force field used in the MD

calculations. Their emphatic agreement that the electron prefers to situate itself in

the benzene ring is, however, slightly surprising. When looking at the constituent

pieces of the acetophenone molecule, one might expect that the lone oxygen, which

as an element is far more electronegative than carbon or hydrogen, might attract the

lone electron. However, O4 and its partner in the carbonyl group, C2, are not well-

represented in the mixture compositions. The oxygen takes some electron density

away from carbon, leaving C2 as an electrophile, but for any anion, the excess charge

would like to delocalize, not localize on C2, even if C2 is electron deficient.

The prominence of C2 and O4 is a point of disagreement between the mixture

composition calculations and the Mulliken population, in which the two atom states

have charge differentials of +0.12 and +0.15, respectively. To understand this dispar-

ity, it is necessary to understand the difference in what these two types of calculations

measure. The CDFT mixture (and the CDFT-trained MD mixture) measure the sta-

bility of each state after an extra electron that is already present in the system fixes

itself in a particular state. The Mulliken charge differential considers what happens

when a charge is introduced into the system.

So in the case of the Mulliken charge, it is reasonable that an extra electron would

be drawn to the most electronegative atom and also be drawn to the electron-seeking

carbon attached to that very electronegative atom at the end of an extremely polar

bond. And in the case of the CDFT energies, an electron that makes its way to the

lone oxygen might not stay very long because the slightly-positive C2 might pull it

away.

The hydrogen atoms in the methyl group, H10, H11, and H12, are the most

undesirable locations for the atoms according to the CDFT data. This makes sense

from a stability perspective. The carbon at the center of the methyl group is so

much more electronegative than the hydrogen atoms surrounding it that an electron

that found itself on any of those hydrogen atoms would not last long. The Mulliken

charge data shows substantial increases for this trio of hydrogen atoms after the

introduction of the electron. This may be because Mulliken population schemes

distribute electronic overlap between two atoms equally between the atoms with no

regard for characteristics of the bonded atoms, such as electronegativity. Through

44

this, the hydrogen atoms in the molecule may get a large boost to their Mulliken

charges while the carbons that they are bonded to benefit only marginally. This

same effect is seen with the hydrogen atoms in the benzene ring.

The carbon atom in the methyl group, C1, is not an attractive place for an electron

to land because its valence electrons are all tied up in bonding with the three hydro-

gens around it. The Mulliken charges show a nice symmetry around the benzene ring.

C5 and C6 have the same increase in Mulliken charge, as do C7 and C8. However, C9

sees a substantial increase, while C3 does not change at all. This is perhaps because

of the proximity in the molecular geometry of C3 to the very electron-attractive car-

bonyl group. In the CDFT energies data, C5 and C3 are the most stable atom states,

followed closely by C9 and C6.

45

3.5 Future Directions

The CDFT energies and the Mulliken populations were calculated with basis sets and

exchange-correlation functionals that could be easily changed. While work was done

to find the best combination of parameters for these calculations, changes made can

be used to perform a systematic sensitivity analysis for calculations on this system,

such as those already established by calculating the CDFT energies for the modified

B3LYP and PBE0 hybrid functionals.

As mentioned previously, in standard eReaxFF calculations, the mass of the elec-

tron or hole is taken to be one atomic mass unit, or the mass of a hydrogen atom,

following the lead of a few other computational methods that treat electrons semi-

classically [85]. In the real world, the mass of an electron is much lighter, about 104

times lighter than an atomic mass unit. While the mass of the charge carrier does

not show up in any of the equations describing the terms of the reactive force field

energies [78] [85], varying this mass could have interesting effects on how the electron

distributes itself across the molecule over the course of an MD run. At a small enough

mass, tunneling may arise, but there has been work done on evaluating the behavior

of lighter particles in molecular dynamics simulations [87].

The force field that has been developed for energy calculations on the acetophe-

none radical anion system can be put to myriad uses beyond what has been discussed

in this thesis.

46

REFERENCES

[1] J. C. Fothergill, “City Research Online City , University of London Institutional

Repository The Coming of Age of HVDC Extruded Power Cables,” Electrical

Insulation Conference (EIC), pp. 124–137, 2014.

[2] F. H. Kreuger, Industrial High Voltage, Vol 1. Delft, NL: Delft University Press,

1991.

[3] “China powers ahead with a new direct-current infrastructure,” 2017.

[4] V. Vahedy, “Polymer insulated high voltage cables,” IEEE Electrical Insulation

Magazine, vol. 22, no. 3, pp. 13–18, 2006.

[5] K. Lahti and K. Nousiainen, “Transmission Line Corona Losses under Hoar Frost

Conditions,” IEEE Transactions on Power Delivery, vol. 12, no. 2, pp. 928–933,

1997.

[6] N. Hampton, R. Hartlein, H. Lennartsson, H. Orton, and R. Ramachandran,

“Long-Life Xlpe Insulated Power Cable,” 7th International Conference on Insu-

lated Power Cable, Jicable, France, pp. 1–6, 2007.

[7] E. Y. Chen, “Coordination polymerization of polar vinyl monomers by single-site

metal catalysts,” Chemical Reviews, vol. 109, no. 11, pp. 5157–5214, 2009.

[8] O. Perez and J. B. Soares, “Coordination Polymerization,” Handbook of Polymer

Synthesis, Characterization, and Processing, pp. 85–104, 2013.

[9] C. Elschenbroich and A. Salzer, Organometallics: A Concise Introduction. New

York: VCH Verlag, 1992.

[10] C. Brown, A. Lita, Y. Tao, N. Peek, M. Crosswhite, M. Mileham, J. Krzystek,

R. Achey, R. Fu, J. K. Bindra, M. Polinski, Y. Wang, L. J. Van De Burgt,

D. Jeffcoat, S. Profeta, A. E. Stiegman, and S. L. Scott, “Mechanism of Initiation

in the Phillips Ethylene Polymerization Catalyst: Ethylene Activation by Cr(II)

and the Structure of the Resulting Active Site,” ACS Catalysis, vol. 7, no. 11,

pp. 7442–7455, 2017.

[11] HOGAN JP, “Ethylene polymerization catalysis over chromium oxide,” J Polym

Sci Part A-1 Polym Chem, vol. 8, no. 9, pp. 2637–2652, 1970.

47

[12] P. Flory, “Polymerization of Unsaturated Monomers by Free Radical Polymer-

ization,” in Principles of Polymer Chemistry, pp. 106–174, Ithaca: Cornell Uni-

versity Press, 1953.

[13] J. Cowie, “Free Radical Addition Polymerization,” in Polymers: Chemistry and

Physics of Modern Materials, pp. 52–82, London: Chapman and Hall, 2 ed.,

1991.

[14] D. B. Malpass, “Free Radical Polymerization of Ethylene,” Introduction to In-

dustrial Polyethylene, pp. 23–32, 2010.

[15] A. Peacock, Handbook of Polyethylene. New York: Marcel Dekker, 2000.

[16] K. S. Whiteley, T. G. Heggs, H. Koch, R. L. Mawer, and W. Immel, “Polyolefins,”

in Ullmann’s Encyclopedia of Industrial Chemistry, Weinheim, Germany: Wiley-

VCH Verlag GmbH & Co. KGaA, 6 2000.

[17] “Polyethyelene (PE) Plastic,” 2020.

[18] P. P. Institute, “Properly Specifying HDPE Conduit for Power and Communi-

cations Projects,” 2018.

[19] S. A. Baeurle, A. Hotta, and A. A. Gusev, “On the glassy state of multiphase

and pure polymer materials,” Polymer, vol. 47, no. 17, pp. 6243–6253, 2006.

[20] M. Nic, J. Jirat, B. Kosata, A. Jenkins, and A. McNaught, eds., Compendium

of Chemical Terminology. Research Triagle Park, NC: IUPAC, 6 2009.

[21] T. Andrews, R. N. Hampton, A. Smedberg, D. Wald, V. Waschk, and W. Weis-

senberg, “The role of degassing in XLPE power cable manufacture,” IEEE Elec-

trical Insulation Magazine, vol. 22, no. 6, pp. 5–16, 2006.

[22] S. M. Tamboli, S. T. Mhaske, and D. D. Kale, “Crosslinked polyethylene,” Indian

Journal of Chemical Technology, vol. 11, no. 6, pp. 853–864, 2004.

[23] T. O. J. Kresser, “Basic Chemistry of Polyethylene,” in Polyethylene, pp. 40–64,

New York: Reinhold Publishing Corp., 1957.

[24] Y. Badr, Z. I. Ali, A. H. Zahran, and R. M. Khafagy, “Characterization of gamma

irradiated polyethylene films by DSC and X-ray diffraction techniques,” Polymer

International, vol. 49, no. 12, pp. 1555–1560, 2000.

48

[25] A. Brandolin, “An Improved Kinetic Model for the Peroxide Initiated Mod if

icat i on of Po lye,” vol. 39, no. 10, pp. 2085–2095, 1999.

[26] C. Meola, G. M. Carlomagno, A. Squillace, U. Prisco, and R. E. Morace,

“Analysis of composites with infrared thermography,” Macromolecular Symposia,

vol. 228, pp. 273–286, 2005.

[27] T. Bremner, A. Rudin, and S. Haridoss, “Effects of polyethylene molecular struc-

ture on peroxide crosslinking of low density polyethylene,” Polymer Engineering

& Science, vol. 32, no. 14, pp. 939–943, 1992.

[28] S. Hosoda, “Structural Distribution of Linear Low-Density Polyethylenes,” Poly-

mer Journal, vol. 20, pp. 383–397, 1988.

[29] J. Sahyoun, A. Crepet, F. Gouanve, L. Keromnes, and E. Espuche, “Diffusion

mechanism of byproducts resulting from the peroxide crosslinking of polyethy-

lene,” Journal of Applied Polymer Science, vol. 134, no. 9, pp. 1–11, 2017.

[30] R. J. Coelho, “Space Charge-Limited, Injection-Controlled Conduction,” in

Physics of Dielectrics for the Engineer, pp. 123–135, Amsterdam: Elsevier Sci-

entific Pub., 1979.

[31] J. Hornak, P. Kadlec, P. Totzauer, P. Trnka, and M. Gutten, “Dielectric spec-

troscopy and absorption/resorption analysis of composite dielectric with MgO

filler,” in 18th International Scientific Conference on Electric Power Engineer-

ing (EPE), (Kouty nad Desnou, CZE), pp. 1–4, 2017.

[32] R. W. Mott, N. F.; Gurney, Electronic processes in ionic crystals. Oxford: Oxford

University Press, 1 ed., 1940.

[33] P. N. Murgatroyd, “Theory of space-charge-limited current enhanced by Frenkel

effect,” Journal of Physics D: Applied Physics, vol. 3, no. 2, pp. 151–156, 1970.

[34] N. Hussin and G. Chen, “The effect of acetophenone and alpha-methylstyrene

on the space charge properties of low density polyethylene,” Annual Report -

Conference on Electrical Insulation and Dielectric Phenomena, CEIDP, pp. 702–

705, 2008.

[35] M. Fu, G. Chen, A. Dissado, and J. C. Fothergill, “Influence of thermal treat-

ment and residues on space charge accumulation in XLPE for DC power cable

49

application,” IEEE Transactions on Dielectrics and Electrical Insulation, vol. 14,

no. 1, pp. 53–64, 2007.

[36] T. Takada, S. J. Hwang, and J. S. Noh, “Effects of constituents of XLPE on

the formation of space charge,” IEEE Transactions on Dielectrics and Electrical

Insulation, vol. 1, no. 6, pp. 1077–1083, 1994.

[37] J. V. Gulmine, P. R. Janissek, H. M. Heise, and L. Akcelrud, “Polyethylene

characterization by FTIR,” Polymer Testing, vol. 21, no. 5, pp. 557–563, 2002.

[38] M. Uhniat, M. Sudol, and S. Kudla, “Stabilization of LDPE cross-linked in

the presence of peroxides. II. FTIR study of chemical changes taking place in

the LDPE-dicumyl peroxide-Irganox 1081 system,” Polymer Degradation and

Stability, vol. 71, no. 1, pp. 75–82, 2000.

[39] P. Pinkerneil, J. Guldenhaupt, K. Gerwert, and C. Kotting, “Surface-attached

polyhistidine-tag proteins characterized by FTIR difference spectroscopy,”

ChemPhysChem, vol. 13, no. 11, pp. 2649–2653, 2012.

[40] J. V. Gulmine and L. Akcelrud, “FTIR characterization of aged XLPE,” Polymer

Testing, vol. 25, no. 7, pp. 932–942, 2006.

[41] M. Bydder, A. Rahal, G. D. Fullerton, and G. M. Bydder, “The magic angle

effect: A source of artifact, determinant of image contrast, and technique for

imaging,” Journal of Magnetic Resonance Imaging, vol. 25, no. 2, pp. 290–300,

2007.

[42] J. M. Tyszka, S. E. Fraser, and R. E. Jacobs, “Magnetic resonance microscopy:

Recent advances and applications,” Current Opinion in Biotechnology, vol. 16,

no. 1 SPEC. ISS., pp. 93–99, 2005.

[43] J. Paul, E. W. Hansen, and J. Roots, “Application of single-pulse solid-state

1H-MAS NMR to probe the oxidation products in crosslinked polyethylene: A

detailed spectral analysis procedure,” Polymer Degradation and Stability, vol. 98,

no. 1, pp. 408–415, 2013.

[44] M. Blasko, P. Mach, A. Antusek, and M. Urban, “DFT Modeling of Cross-

Linked Polyethylene: Role of Gold Atoms and Dispersion Interactions,” Journal

of Physical Chemistry A, vol. 122, no. 5, pp. 1496–1503, 2018.

50

[45] M. Golriz, H. A. Khonakdar, J. Morshedian, S. H. Jafari, Y. Mohammadi, and

U. Wagenknecht, “Microstructural evolution in linear low density polyethylene

during peroxide modification: A Monte Carlo simulation study,” Macromolecular

Theory and Simulations, vol. 22, no. 8, pp. 426–438, 2013.

[46] A. Moyassari, M. Unge, M. S. Hedenqvist, U. W. Gedde, and F. Nilsson, “First-

principle simulations of electronic structure in semicrystalline polyethylene,”

Journal of Chemical Physics, vol. 146, no. 20, 2017.

[47] S. Iwata, H. Uehara, and T. Takada, “Computational study on acetophenone in

amorphous polyethylene,” Journal of Molecular Modeling, vol. 23, no. 10, 2017.

[48] M. Meunier, N. Quirke, and A. Aslanides, “Molecular modeling of electron traps

in polymer insulators: Chemical defects and impurities,” Journal of Chemical

Physics, vol. 115, no. 6, pp. 2876–2881, 2001.

[49] D. R. Rottach, J. G. Curro, J. Budzien, G. S. Grest, C. Svaneborg, and R. Ev-

eraers, “Permanent set of cross-linking networks: Comparison of theory with

molecular dynamics simulations,” Macromolecules, vol. 39, no. 16, pp. 5521–

5530, 2006.

[50] D. R. Rottach, J. G. Curro, J. Budzien, G. S. Grest, C. Svaneborg, and R. Ev-

eraers, “Molecular dynamics simulations of polymer networks undergoing se-

quential cross-linking and scission reactions,” Macromolecules, vol. 40, no. 1,

pp. 131–139, 2007.

[51] H. Zhang, Y. Shang, M. Li, H. Zhao, X. Wang, and B. Han, “Theoretical study

on the radical reaction mechanism in the cross-linking process of polyethylene,”

RSC Advances, vol. 5, no. 110, pp. 90343–90353, 2015.

[52] D. Sholl and J. Steckel, “What is Density Functional Theory?,” in Density Func-

tional Theory: A Practical Introduction, pp. 1–34, Hoboken, NJ, USA: John

Wiley and Sons, 2009.

[53] T. Cormen, C. Leiserson, R. Rivest, and C. Stein, “Strassen’s algorithm for

matrix multiplication,” in Introduction to Algorithms, pp. 735–742, Cambridge,

MA: MIT Press, 2009.

[54] M. Born and J. Oppenheimer, “On the Quantum Theory of Molecules (Trans-

lated from German by S.M. Blinder),” Ann. Physik, vol. 84, pp. 457–484, 1927.

51

[55] L. Thomas, “The calculation of atomic fields,” Mathematical Proceedings of the

Cambridge Philosophical Society, vol. 23, no. 5, pp. 542–548, 1927.

[56] P. Hohenberg and W. Kohn, “Inhomogeneous Electron Gas,” Physical Review,

vol. 136, no. 3B, 1964.

[57] E. Kryachko and E. Ludena, “A Rigorous Formulation of the Variational Prin-

ciple in Density Functional Theory,” in Energy Density Functional Theory of

Many-Electron Systems, p. 506, Dordrecht, NL: Kluwer Academic, 1990.

[58] W. Kohn and L. Sham, “Self-Consistent Equations Including Exchange and Cor-

relation Effects,” Physical Review, vol. 140, no. 4A, 1965.

[59] B. Kaduk, T. Kowalczyk, and T. Van Voorhis, “Constrained density functional

theory,” Chemical Reviews, vol. 112, no. 1, pp. 321–370, 2012.

[60] P. H. Dederichs, S. Blgel, R. Zeller, and H. Akai, “Ground states of constrained

systems: Application to cerium impurities,” Physical Review Letters, vol. 53,

no. 26, pp. 2512–2515, 1984.

[61] “Calculated Geometries available for CH3COC6H5 (acetophenone) 1A’ Cs,”

2019.

[62] R. Ditchfield, W. J. Hehre, and J. A. Pople, “Self-consistent molecular-orbital

methods. IX. An extended gaussian-type basis for molecular-orbital studies of

organic molecules,” The Journal of Chemical Physics, vol. 54, no. 2, pp. 720–

723, 1971.

[63] D. E. Woon and T. H. Dunning, “Gaussian basis sets for use in correlated molecu-

lar calculations. V. Core-valence basis sets for boron through neon,” The Journal

of Chemical Physics, vol. 103, no. 11, pp. 4572–4585, 1995.

[64] E. R. Davidson and D. Feller, “Basis Set Selection for Molecular Calculations,”

Chemical Reviews, vol. 86, no. 4, pp. 681–696, 1986.

[65] J. P. Perdew, M. Ernzerhof, and K. Burke, “Rationale for mixing exact exchange

with density functional approximations,” Journal of Chemical Physics, vol. 105,

no. 22, pp. 9982–9985, 1996.

52

[66] A. D. Becke, “Perspective: Fifty years of density-functional theory in chemical

physics,” Journal of Chemical Physics, vol. 140, no. 18, 2014.

[67] J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J.

Singh, and C. Fiolhais, “Erratum: Atoms, molecules, solids, and surfaces: Appli-

cations of the generalized gradient approximation for exchange and correlation

(Physical Review B (1993) 48, 7, (4978)),” Physical Review B, vol. 48, no. 7,

p. 4978, 1993.

[68] S. H. Vosko, L. Wilk, and M. Nusair, “Accurate spin-dependent electron liq-

uid correlation energies for local spin density calculations: a critical analysis,”

Canadian Journal of Physics, vol. 58, no. 8, pp. 1200–1211, 1980.

[69] A. D. Becke, “Density-functional thermochemistry. III. The role of exact ex-

change,” The Journal of Chemical Physics, vol. 98, no. 7, pp. 5648–5652, 1993.

[70] J. P. Perdew, K. Burke, and M. Ernzerhof, “Generalized gradient approximation

made simple,” Physical Review Letters, vol. 77, no. 18, pp. 3865–3868, 1996.

[71] M. C. Kim, E. Sim, and K. Burke, “Communication: Avoiding unbound anions

in density functional calculations,” Journal of Chemical Physics, vol. 134, no. 17,

pp. 4–7, 2011.

[72] R. Kar, J. W. Song, and K. Hirao, “Long-range corrected functionals satisfy

Koopmans’ theorem: Calculation of correlation and relaxation energies,” Journal

of Computational Chemistry, vol. 34, no. 11, pp. 958–964, 2013.

[73] E. Apra, E. J. Bylaska, W. A. de Jong, N. Govind, K. Kowalski, T. P. Straatsma,

M. Valiev, H. J. van Dam, Y. Alexeev, J. Anchell, V. Anisimov, F. W. Aquino,

R. Atta-Fynn, J. Autschbach, N. P. Bauman, J. C. Becca, D. E. Bernholdt,

K. Bhaskaran-Nair, S. Bogatko, P. Borowski, J. Boschen, J. Brabec, A. Bruner,

E. Cauet, Y. Chen, G. N. Chuev, C. J. Cramer, J. Daily, M. J. Deegan, T. H.

Dunning, M. Dupuis, K. G. Dyall, G. I. Fann, S. A. Fischer, A. Fonari, H. Fruchtl,

L. Gagliardi, J. Garza, N. Gawande, S. Ghosh, K. Glaesemann, A. W. Gotz,

J. Hammond, V. Helms, E. D. Hermes, K. Hirao, S. Hirata, M. Jacquelin,

L. Jensen, B. G. Johnson, H. Jonsson, R. A. Kendall, M. Klemm, R. Kobayashi,

V. Konkov, S. Krishnamoorthy, M. Krishnan, Z. Lin, R. D. Lins, R. J. Little-

field, A. J. Logsdail, K. Lopata, W. Ma, A. V. Marenich, J. Martin Del Campo,

53

D. Mejia-Rodriguez, J. E. Moore, J. M. Mullin, T. Nakajima, D. R. Nascimento,

J. A. Nichols, P. J. Nichols, J. Nieplocha, A. Otero-de-la Roza, B. Palmer, A. Pa-

nyala, T. Pirojsirikul, B. Peng, R. Peverati, J. Pittner, L. Pollack, R. M. Richard,

P. Sadayappan, G. C. Schatz, W. A. Shelton, D. W. Silverstein, D. M. Smith,

T. A. Soares, D. Song, M. Swart, H. L. Taylor, G. S. Thomas, V. Tipparaju,

D. G. Truhlar, K. Tsemekhman, T. Van Voorhis, Vazquez-Mayagoitia, P. Verma,

O. Villa, A. Vishnu, K. D. Vogiatzis, D. Wang, J. H. Weare, M. J. Williamson,

T. L. Windus, K. Wolinski, A. T. Wong, Q. Wu, C. Yang, Q. Yu, M. Zacharias,

Z. Zhang, Y. Zhao, and R. J. Harrison, “NWChem: Past, present, and future,”

The Journal of chemical physics, vol. 152, no. 18, p. 184102, 2020.

[74] A. C. Van Duin, J. M. Baas, B. Van De Graaf, J. W. De Leeuw, T. P. Bastow,

and R. Alexander, “Comparison of calculated equilibrium mixtures of alkylnaph-

thalenes and alkylphenanthrenes with experimental and sedimentary data; the

importance of entropy calculations,” Organic Geochemistry, vol. 26, no. 3-4,

pp. 275–280, 1997.

[75] R. S. Mulliken, “Electronic Population Analysis on LCAO,” J. Chem. Phys,

vol. 23, no. January 1955, p. 1833, 1955.

[76] P. O. Lowdin, “On the non-orthogonality problem connected with the use of

atomic wave functions in the theory of molecules and crystals,” The Journal of

Chemical Physics, vol. 18, no. 3, pp. 365–375, 1950.

[77] G. Bruhn, E. R. Davidson, I. Mayer, and A. E. Clark, “Lowdin Population

Analysis With and Without Rotational Invariance,” International Journal of

Quantum Chemistry, vol. 106, pp. 2065–2072, 2006.

[78] A. C. Van Duin, S. Dasgupta, F. Lorant, and W. A. Goddard, “ReaxFF: A

reactive force field for hydrocarbons,” Journal of Physical Chemistry A, vol. 105,

no. 41, pp. 9396–9409, 2001.

[79] D. W. Brenner, “Empirical potential for hydrocarbons for use in simulating the

chemical vapor deposition of diamond films,” Physical Review B, vol. 42, no. 15,

pp. 9458–9471, 1990.

[80] H. S. Johnston and C. Parr, “Activation Energies from Bond Energies. I. Hy-

drogen Transfer Reactions,” Journal of the American Chemical Society, vol. 85,

no. 17, pp. 2544–2551, 1963.

54

[81] D. M. Root, C. R. Landis, and T. Cleveland, “Valence Bond Concepts Applied

to the Molecular Mechanics Description of Molecular Shapes. 1. Application to

Nonhypervalent Molecules of the P-Block,” Journal of the American Chemical

Society, vol. 115, no. 10, pp. 4201–4209, 1993.

[82] T. P. Senftle, S. Hong, M. M. Islam, S. B. Kylasa, Y. Zheng, Y. K. Shin,

C. Junkermeier, R. Engel-Herbert, M. J. Janik, H. M. Aktulga, T. Verstraelen,

A. Grama, and A. C. Van Duin, “The ReaxFF reactive force-field: Develop-

ment, applications and future directions,” npj Computational Materials, vol. 2,

no. November 2015, 2016.

[83] W. J. Mortier, S. K. Ghosh, and S. Shankar, “Electronegativity Equalization

Method for the Calculation of Atomic Charges in Molecules,” Journal of the

American Chemical Society, vol. 108, no. 15, pp. 4315–4320, 1986.

[84] D. Bedrov, G. D. Smith, and A. C. Van Duin, “Reactions of singly-reduced ethy-

lene carbonate in lithium battery electrolytes: A molecular dynamics simulation

study using the ReaxFF,” Journal of Physical Chemistry A, vol. 116, no. 11,

pp. 2978–2985, 2012.

[85] M. M. Islam, G. Kolesov, T. Verstraelen, E. Kaxiras, and A. C. Van Duin,

“EReaxFF: A Pseudoclassical Treatment of Explicit Electrons within Reactive

Force Field Simulations,” Journal of Chemical Theory and Computation, vol. 12,

no. 8, pp. 3463–3472, 2016.

[86] A. C. Van Duin, J. M. Baas, and B. Van De Graaf, “Delft molecular mechanics:

A new approach to hydrocarbon force fields. Inclusion of a geometry-dependent

charge calculation,” Journal of the Chemical Society, Faraday Transactions,

vol. 90, no. 19, pp. 2881–2895, 1994.

[87] M. Ceriotti, M. Parrinello, T. E. Markland, and D. E. Manolopoulos, “Effi-

cient stochastic thermostatting of path integral molecular dynamics,” Journal of

Chemical Physics, vol. 133, no. 12, 2010.