Quick viewing(Text Mode)

Chromium Catalysts for Ethylene Polymerization and Oligomerization

Chromium Catalysts for Ethylene Polymerization and Oligomerization

CHAPTER THREE

Chromium Catalysts for Polymerization and Oligomerization

Zhen Liu*, Xuelian He*, Ruihua Cheng*, Moris S. Eisen†, Minoru Terano{, Susannah L. Scott}, Boping Liu* *State Key Laboratory of Chemical Engineering, East China University of Science and Technology, Shanghai, P.R. China †Schulich Faculty of Chemistry, Technion-Israel Institute of Technology, Technion City, Haifa, Israel { School of Materials Science, Japan Advanced Institute of Science and Technology, 1-1 Asahidai, Nomi, Ishikawa, Japan } Department of Chemical Engineering, University of California, Santa Barbara, California, United States

Contents 1. Introduction 129 2. Phillips Catalysts for Ethylene Polymerization 131 2.1 Brief overview on Phillips chromium catalysts 131 2.2 Characterization of microstructures of polyethylene chains 145 2.3 Polyethylene-based nanocomposites 149 3. Phillips Chromium Catalysts for Alkyne Cyclotrimerization 154 4. Molybdenum Catalysts for Ethylene Polymerization 162 5. Chromium Catalysts for Selective Ethylene Oligomerization 167 5.1 Cr-DME-mediated ethylene trimerization 169 5.2 Cr-SNS-mediated ethylene trimerization 175 6. Summary and Outlook 179 Acknowledgments 181 References 181

Abstract Chromium-based catalysts are the most important ethylene polymerization and oligo- merization catalysts widely applied for industrial production of polyethylene and 1-hexene. Phillips chromium catalyst is a well-known heterogeneous catalyst for com- mercial production of HDPE products, which accounts for more than 40% of world pro- duction annually. The Chevron-Phillips Cr-based homogeneous catalyst system is the first commercialized catalyst for the production of 1-hexene through selective ethylene oligomerization. Although a great success with these Cr-based catalysts has been achieved in industrial applications, there are still many debates in the academic field concerning the precise structure of active chromium species, the oxidation states of

# Advances in Chemical Engineering, Volume 44 2014 Elsevier Inc. 127 ISSN 0065-2377 All rights reserved. http://dx.doi.org/10.1016/B978-0-12-419974-3.00003-8 128 Zhen Liu et al.

chromium center, the effects of cocatalysts/ligands and the catalytic mechanisms. Dur- ing the last decades, a step-forward mechanistic understanding has been achieved through extensive and successive investigations on these Cr-based catalysts for ethyl- ene polymerization/oligomerization. In addition, the progress in mechanistic under- standing on alkyne cyclotrimerization by the same Phillips catalyst and ethylene polymerization over Mo-based catalyst are also covered. The later might be served as an alternative green catalyst for the industrial production of polyethylene.

ABBREVIATIONS AFM atomic force microscope CB black DFT density functional theory DME dimethyl ether DRS diffuse reflectance spectroscopy DSC differential scanning calorimetry EDS energy dispersive spectrometer EPMA electron probe microanalysis EPR electron paramagnetic resonance ESCR environmental stress-cracking resistance FTIR Fourier transform infrared HDPE high-density polyethylene HLMI High load melt index LA-MS laser ablation-mass spectrometry LDI-MS laser desorption-ionization mass spectrometry LLDPE linear low-density polyethylene MAO methylaluminoxane MECP minimum energy crossing point MWD molecular weight distribution NMR nuclear magnetic resonance PES potential energy surface PIBAO partially hydrolyzed tri-isobutylaluminum PIXE proton induced X-ray emission RBS Rutherford backscattering spectrometry SC step crystallization SCB short-chain branch SCBD short-chain branch distribution SEM scanning electron microscopy SIMS secondary ion mass spectroscopy SSA successive self-nucleation and annealing TG-DTA thermogravimetry-differential thermal analysis TMB trimethylbenzene TOF turnover of frequency TPD-MS temperature-programmed desorption-mass spectrometry TPR temperature-programmed reduction TREF temperature rising elution fractionation Ethylene Polymerization and Oligomerization 129

UV–vis DRS ultraviolet–visible diffuse reflectance spectroscopy XAS X-ray absorption spectroscopy XPS X-ray photoelectron spectroscopy XRD X-ray diffraction

1. INTRODUCTION

In the 1950s, the world had witnessed two kinds of important cat- alysts successfully applied in industrial production of polyolefins including Ziegler-Natta catalyst and Phillips chromium catalyst (Groppo et al., 2013). After about 60 years of intensive researches and continuous innovations, these catalysts are widely used in a large scale in the commercial production of polyolefins. Nowadays, Phillips chromium catalyst is currently produc- ing more than 10 million tons of high-density polyethylene (HDPE) prod- ucts annually throughout the world (McDaniel, 2010). Since the discovery in 1951, Phillips chromium catalyst was soon patented in 1958 (Hogan and Banks, 1958) and has been attracting tremendous researches from both industrial and academic fields during the last 50 years. The catalyst is famous for its high activity for ethylene polymerization without using any organometallic cocatalyst. This self-alkylation characteristic of the Phil- lips chromium catalyst is often described as “unique” when compared to the other important Ziegler-Natta and catalysts (McDaniel, 2013). Although Phillips chromium catalyst has achieved a great success in diverse commercial applications, there are still many debates in the aca- demic field in elucidation of the precise structure of active sites, the active oxidation states of chromium center, and the initiation mechanism for eth- ylene polymerization (Groppo et al., 2005a; McDaniel, 1985, 2008, 2010). The difficulties for fundamental studies of the Phillips chromium catalyst are mainly derived from the following aspects: (a) the low percentage of active chromium species, (b) the complexity of the amorphous silica sup- port, (c) the multiple valence states of chromium center, (d) the instant encapsulation of active sites by produced polymer, (e) the super-fast poly- merization rate, (f ) the existence of many side reactions like active sites deactivation and various chain transfer reactions, etc. As a general agreement is far from being reached, much deeper and clearer basic understanding on the Phillips chromium catalyst is still highly expected (McDaniel, 1985). 130 Zhen Liu et al.

The HDPE products by the Phillips chromium catalyst usually have the following characteristics: (a) ultra-broad molecular weight distribution (MWD) with a typical polydispersity index larger than 10, (b) a small amount of long chain branches, and (c) a vinyl end-group for each polyeth- ylene chain (McDaniel, 2010). These special features bring its HDPE prod- ucts good mechanical properties and high melt strength, which are of key importance in blow molding process. In the past decades, the market demand of the HDPE products made by the Phillips chromium catalyst shows a dramatic increase in many diverse fields including gasoline tanks of automobile industry, ultra large size plastic containers, high-grade pipe materials like PE80 and PE100, and so on. The increasing market of the HDPE, medium density polyethylene (MDPE), and linear low-density polyethylene (LLDPE) products requires large amount of short a-olefins as comonomer for copolymerization with ethylene. Although copolymer- ization with 1-hexene could bring the HDPE products much improved mechanical properties, 1-butene had been dominant in the polyethylene market in the past few decades because of the high cost for conventional 1-hexene production. Only until 2003, the first plant established by Chevron-Phillips started the commercial production of the comonomer grade 1-hexene with a relative low cost through selective ethylene trimerization (Dixon et al., 2004). This technology was originated from the first discovery of Cr(2-EH)3 (EH, ethylhexanoate) system for ethylene polymerization with a small amount of trimerization product reported by Manyik et al. (1977). Recently, the newly invented catalysts for selective ethylene oligomerization including trimerization and tetramerization are mainly based on chromium catalysts, including bi- and tridentate chromium complexes with a ligand providing N, S, O, or P coordination (Agapie, 2011; Dixon et al., 2004; McGuinness, 2011). There are several reviews in the field of the Phillips chromium catalyst that have been published during the past decades including, to name a few, the review by Zecchina and coworkers in 2005 (Groppo et al., 2005a) and the reviews by McDaniel in 2008 and 2010. In the field of ethylene trimerization, Morgan and coworkers have written a review as early as in 2004 (Dixon et al., 2004), and McGuinness published another review very recently (McGuinness, 2011). In this contribution, we will present a short overview on the Phillips chromium catalyst for ethylene polymerization concerning spectroscopic characterizations, kinetic studies, model catalysts investigations, and molecular modeling simulations. Then, we will include recent progresses in the field of Phillips chromium catalyst with particular Ethylene Polymerization and Oligomerization 131 emphasis on the recent studies from the authors’ groups, including the microstructure characterization of the polymer chains and the grafting of HDPE onto carbon black (CB) focused on high-grade pipe materials of improved long-term mechanical and ultraviolet resistance properties, the mechanistic investigation on the alkyne cyclotrimerization catalyzed by the same Phillips chromium catalyst, and the mechanistic studies of environmental-friendly nonchromium (molybdenum) catalyst for ethylene polymerization and two important chromium-based catalyst systems for ethylene-selective trimerization.

2. PHILLIPS CHROMIUM CATALYSTS FOR ETHYLENE POLYMERIZATION 2.1. Brief overview on Phillips chromium catalysts Phillips chromium catalyst is usually prepared by impregnation of chromic(III) acetate onto a silica gel carrier and a subsequent drying around 150C followed by a calcination between 500 and 900 C in dry air. During the cal- cination, chromic(III) acetate could be oxidized to chromic(VI) trioxide (CrO3) followed by subsequent anchoring of CrO3 onto the silica surface resulting in the formation of chromate species including monochromate, dichromate, or even polychromate. By contacting with ethylene mono- mer, the chromate species on the calcined Phillips Cr(VI)Ox/SiO2 catalyst could be reduced to lower valence states, usually Cr(II), showing high eth- ylene polymerization activity without using any organometallic cocatalyst (McDaniel, 1985). This unique feature of the Phillips chromium catalyst brings us a long-standing question: how is the first CrdC bond formed on the naked chromium active site? That is to say, the initiation mecha- nism of ethylene polymerization in terms of the formation of the first poly- mer chain over each chromium active site on the Phillips catalyst is the key problem waiting for elucidation. As shown in Scheme 3.1, three kinds of typical initiation mechanisms have been proposed in the literatures for the Phillips chromium catalyst: Cossee mechanism (Cossee, 1964), carbene mechanism (Ghiotti et al., 1979, 1988; Ivin et al., 1978; McDaniel and Cantor, 1983), and metallacycle mechanism (Ghiotti et al., 1991; Groppo et al., 2006a). Cossee mechanism was originally proposed by Cossee in 1964 for the Ziegler-Natta catalyst (Cossee, 1964). As for metal-alkyl-free catalysts, such as the Phillips chromium catalyst, Cossee mechanism could not rationalize the origin of an extra hydrogen atom, 132 Zhen Liu et al.

Scheme 3.1 Literature proposed initiation mechanisms for ethylene polymerization over the Phillips chromium catalyst. which is needed in the formation of the first metal-ethyl center. Carbene mechanism was proposed by Rooney et al. in 1978 for the stereospecific polymerization of olefins by Ziegler-Natta catalysts (Ivin et al., 1978). Zecchina et al. reported a very weak infrared band for carbene species on the Phillips chromium catalyst (Ghiotti et al., 1979). However, McDaniel et al. found no hydrogen scrambling taking place during polymer- ization on the Phillips chromium catalyst using deuterium-labeled ethylene, which was against the chain growing through carbene mechanism (McDaniel and Cantor, 1983). Metallacycle mechanism was proposed by Zecchina et al. in 1991 in the investigation of CO/C2H4 coadsorption and reaction on the Phillips chromium catalyst (Ghiotti et al., 1991). The subsequent in situ Fou- rier transform infrared (FTIR) spectroscopic studies by the same group showed the absence of methyl end-group during the initial stage of ethylene polymerization suggesting a metallacycle initiation on the Phillips chromium catalyst (Groppo et al., 2006a). Further investigations with more conclusive evidence are indispensable for a complete elucidation of the initiation mech- anism of ethylene polymerization over the Phillips chromium catalysts. In the past decades, Phillips chromium catalyst has been attracting much attention from the academic and industrial fields with various techniques including spectroscopic characterizations, kinetic studies, model catalysts investigations, and molecular modeling simulations. The extensive experimental Ethylene Polymerization and Oligomerization 133 and theoretical studies on the Phillips chromium catalyst for ethylene polymer- ization will be shortly reviewed following these aspects and a more detailed review can be found in Cheng et al. (2013).

2.1.1 Spectroscopic and kinetic investigations Spectroscopy and kinetics are the most important traditional methods for investigations in the field of heterogeneous catalysis. As summarized in Fig. 3.1, many modern analytical techniques have been applied to

Magnetic SEM/EDS susceptibility UV-vis DRS EPR (ESR) AFM EPMA IR/FTIR/DRIFT In situ FTIR Solid-state NMR

Raman Phillips TPR In situ Raman Catalyst TPD-MS

XPS (ESCA) LA-MS TG-DTA XAS (XANES/EXAFS) In situ XAS PIXE RBS LDI-MS

SIMS XRD (XRPD)

Figure 3.1 Various methods for the characterization of the Phillips chromium catalyst: AFM (Thune et al., 2007), EPMA (Liu and Terano, 2001), EPR (Bensalem et al., 1997; Cimino et al., 1991a; Ellison et al., 1993; Groeneveld et al., 1979; Qiu et al., 2011, 2012), FTIR (Groppo et al., 2005a,b,c, 2006b, 2007, 2011; Demmelmaier et al., 2009; Zhong et al., 2012a; Rebenstorf and Larsson, 1981; Nishimura and Thomas, 1993; Barzan et al., 2012), LA-MS (Aubriet et al., 2006), LDI-MS (Aubriet et al., 2006), PIXE (Rahman et al., 1995), magnetic susceptibility measurement (Groeneveld et al., 1979), Raman (Damin et al., 2006; Dines and Inglis, 2003; Groppo et al., 2005d, 2011; Hardcastle and Wachs, 1988; Moisii et al., 2006; Richter et al., 1988; Vuurman et al., 1993; Zaki et al., 1986), RBS (van Kimmenade et al., 2004), SEM/EDS (Schmidt et al., 1996), SIMS (Ellison and Overton, 1993), solid state NMR (Cheng et al., 2010; Ellison and Overton, 1993; Xia et al., 2010), TG-DTA (Qiu et al., 2009), TPD-MS (Liu et al., 2002), TPR (Bensalem et al., 1997; Jozwiak and Dalla Lana, 1997), UV–vis DRS (Bensalem et al., 1997; Groeneveld et al., 1979; Groppo et al., 2011; Zaki et al., 1986), XAS (XANES-EXAFS) (Agostini et al., 2007; Demmelmaier et al., 2009; Ellison et al., 1988; Groppo et al., 2005e; Moisii et al., 2006; Zhong et al., 2012a), XPS (Best et al., 1977; Cheng et al., 2010; Cimino et al., 1976; Fang et al., 2005a; Liu and Terano, 2001; Liu et al., 2002, 2004a, 2005; Merryfield et al., 1982; Okamoto et al., 1976; Rahman et al., 1995), XRD (Jozwiak and Dalla Lana, 1997; McDaniel, 1981; Vuurman et al., 1993; Wang et al., 2000). 134 Zhen Liu et al. characterize the chromium states on the silica surface of the Phillips chro- mium catalyst (Groppo et al., 2005a; Weckhuysen et al., 1996). The char- acterizations based on these methods aim to provide a basic understanding of the surface chromate species formed during the catalyst calcination, and to further clarify the reaction mechanism for ethylene polymerization catalyzed by the Phillips chromium catalyst. Through long-term investigations from both industry and academia, it will be demonstrated that a valuable under- standing has been achieved through (1) spectroscopic investigations on acti- vation of the Phillips chromium catalyst by thermal calcination or reducing agents including CO, Al-alkyl, Al-alkoxy, ethylene, etc. and (2) kinetic studies of slurry or gas-phase polymerizations.

2.1.1.1 Spectroscopic investigations on activation of the Phillips chromium catalyst The chromium species of the Phillips catalyst were anchored onto silica sur- face to form various surface-stabilized chromates during catalyst preparation through thermal activation. However, the hexavalent chromate species must be first reduced to lower valance states before showing activity for ethylene polymerization. This process can be easily fulfilled through activation by ethylene monomer itself or using reducing agents, such as CO, Al-alkyl, Al-alkoxy, etc. During thermal activation, a highly dispersed chromate species, includ- ing monochromate, dichromate, and polychromate, could be generated through a redispersion cycle of sublimation, volatilization, spreading, depo- sition, and stabilization of bulk CrO3 onto the silica surface (McDaniel, 1985). McDaniel suggested that the initially formed species was mon- ochromate at 200 C, the dichromate became dominant at 500 C, and polychromate might exist above 800 C(McDaniel, 1981). Panchenko et al. performed a diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) and diffuse reflectance spectroscopy (DRS) study of the Phillips chromium catalyst and confirmed that the reaction of CrO3 with silica dehydroxylated at 250, 400, and 800 C yielded monochromate, mono- and dichromate, and polychromate, respectively (Panchenko et al., 2006). Liu et al. studied the thermal activation of the Phillips chromium catalyst by means of X-ray photoelectron spectroscopy (XPS) measurements and suggested that two unfavorable situations might occur in the calcination pro- cess: the calcination induced reduction of Cr(VI) species into lower valence state (þ5, þ4, or þ3) and the formation of aggregated Cr2O3 even in the presence of oxygen or dry air (Fang et al., 2005a; Liu and Terano, 2001; Ethylene Polymerization and Oligomerization 135

Liu et al., 2002, 2003, 2004a,b). The irreversible formation of a-Cr2O3 is a common phenomenon for the Phillips chromium catalyst with chromium loading higher than 1 wt% or in the presence of trace amounts of moisture. The previous studies revealed that the formation of aggregated Cr2O3 on the silica surface with low-level chromium loading usually occurred in the later stage of calcination (Cimino et al., 1976, 1991b; Groeneveld et al., 1979; Rahman et al., 1995). Initially, explanation on the formation of aggregated Cr2O3 was considered as a thermal decomposition/reduction of bulk CrO3. However, according to the XPS results (Liu and Terano, 2001), all the Cr(III)Ox,surf species and one-seventh of the Cr(VI)Ox,surf species trans- formed into aggregated Cr2O3 at high temperature in the presence of trace amounts of moisture regardless of the oxygen or inert atmosphere. Consid- ering the trace amount of moisture from the simultaneous dehydroxylation of residual hydroxyl groups on silica surface, the formation of aggregated Cr2O3 microcrystals might be induced by trace amounts of moisture through cleavage of the Cr(III)Ox,surf species during calcination. In laboratory research, the hexavalent chromate species were usually reduced by CO or Al-alkyl cocatalyst in a separated preactivation step, or by ethylene monomer itself during the initial stage of polymerization (McDaniel, 1985). The activation by ethylene monomer is widely used in commercial processes because of the low production cost. Phillips chromium catalyst reduced by CO at 350 C shows instantaneous polymerization activity upon contacting with ethylene monomer. The CO prereduced catalyst is widely employed in spectroscopic characterization of chromium active sites of the Phillips catalyst (McDaniel, 1985; Groppo et al., 2005a). Through a comparison of the chromium oxidation states before and after the reduction by CO, the XPS study showed that about 63% of surface chromate species were reduced to Cr(II) species (Liu et al., 2004b). Furthermore, the chro- mium active sites are fully available for CO reduction at higher temperature at 600 C, as suggested by the DRS results (Weckhuysen et al., 1993). The activation of the Phillips chromium catalyst by Al-alkyl cocatalyst was also systematically studied by XPS and solid state nuclear magnetic resonance (NMR) (Liu et al., 2005; Xia et al., 2006). As indicated by XPS characterizations, there are four oxidation states including þ2, þ3, þ5, and þ6 of surface chromium species of triethylaluminum-modified Phillips catalyst. The observations from above experiments suggested that the active precursor of the chromium site may have the flowing form 2þ 6þ Cr 2Cr composed of one Cr(II)Ox,surf species and two Cr(VI)Ox,surf species, in which Cr(II)Ox,surf species acted as the real active chromium 136 Zhen Liu et al.

precursor and the residual Cr(VI)Ox,surf species acted as the neighboring ligand providing the electronic and steric environment (Liu et al., 2005). Concerning the activation by ethylene monomer, Liu et al. reported an extensive investigation on reduction of the hexavalent chromate by means of XPS, temperature-programmed desorption-mass spectrometry (TPD- MS) methods and found that surface chromium species might exist in three oxidation states: (1) surface chromate Cr(VI)Ox,surf species, (2) surface- stabilized trivalent Cr(III) species, and (3) surface-stabilized Cr(II) species (Liu et al., 2002). Some short alkenes including propylene and butylene as well as the reduction by-product of formaldehyde were confirmed based on TPD-MS characterizations (Liu et al., 2003). Formaldehyde is experimentally observed as a by-product of the redox reaction between ethylene and hexavalent chromate, which is believed to be reduced to Cr(II) species (Baker and Carrick, 1968; Liu et al., 2002, 2004c). Subsequently, the Cr(II) species coordinated with formaldehyde might act as the active precursor at lower temperature to produce the new short olefins with odd or even number of carbon atoms. These experimental evidences obtained in the early stage of ethylene polymerization cannot be rationalized by the classic Cossee-Arlman mechanism. It is worthy of note that the con- version of ethylene into higher olefins with odd or even number of carbon atoms is well understood through ethylene metathesis reaction (O’Neill and Rooney, 1972), indicating that the coordination of formaldehyde on the Cr(II) center might generate an active precursor for olefin metathesis rather than polymerization.

2.1.1.2 Kinetic studies of slurry or gas-phase polymerizations Kinetic investigation through either experiments or mathematic modeling both for slurry and gas-phase polymerization is one of the most important ways to investigate the catalytic mechanisms, and thus to provide basic data for design of the polymerization reactor and developing of new process. Mathematic modeling of ethylene polymerization kinetics over the Phillips chromium catalyst has been demonstrated as a powerful tool for the precise evaluation of the basic kinetic parameters and to establish equations for structure-property regulation of polyethylene products through control of process parameters (Choi et al., 2004; Choi and Tang, 2004; Matos et al., 2004; Kissin et al., 2008). Polymerization kinetics of the Phillips chromium catalyst could be significantly affected by the reductive activation process for ethylene polymerization using different activators, such as ethylene, CO, Al-alkyl cocatalyst, or other reducing agents. Ethylene Polymerization and Oligomerization 137

The kinetics of the Phillips chromium catalyst using ethylene monomer as an activator for ethylene polymerization has been systematically investi- gated (McDaniel, 1985, 2010). Typically, a linearly built-up type kinetic curve would be presented with an induction period depending on the poly- merization temperature and ethylene pressure. Reductive activation by CO only diminishes the induction period without changing the character of the built-up type kinetic curve. In recent years, activation of the Phillips chro- mium catalyst by Al-alkyl cocatalysts becomes one of the most important ways to improve the catalyst performance and to regulate the microstruc- tures of the polyethylene products. The Al-alkyl cocatalysts could act as reducing agent, alkylation, poison scavenger, and thus have a significant impact on the polymer microstructures by control of the chain transfer and stereospecificity. Additionally, excess amount of Al-alkyl cocatalyst could deactivate the catalyst through over-reduction of the chromium active species. Ethylene polymerization with the Phillips chromium catalyst with- out using any organometallic cocatalyst is taken as strong evidence to sup- port the monometallic active site mechanism. Therefore, Al-alkyl cocatalyst can be safely excluded as the active site contributor for the Phillips chro- mium catalyst. During the past decades, experimental reports on the combination of Al-alkyl cocatalyst with the Phillips chromium catalyst are very limited. Spitz et al. reported a significant effect of triethylaluminum on the Phillips chromium catalyst for the activity, kinetic, and 1-hexene incorporation dur- ing the ethylene/1-hexene copolymerization (Spitz et al., 1979). McDaniel et al. studied the effects of triethylborane on the polymerization kinetics of the Phillips chromium catalyst with different supports, such as AlPO4, SiO2, and Al2O3 (McDaniel and Johnson, 1986, 1987). Tait et al. investigated the effects of triisobutylaluminum on kinetics of the Phillips chromium catalyst and polymer morphology (Wang et al., 1991). Liu et al. studied the Phillips chromium catalyst combined with Al-alkyl cocatalyst and revealed that the polymerization kinetics could be significantly affected by the type of Al-alkyl cocatalysts as well as the timing for introducing cocatalyst in both slurry and gas-phase ethylene polymerization (Fang et al., 2005b, 2006; Li et al., 2013; Liu et al., 2004b, 2005; Xia et al., 2006, 2009). As shown in Fig. 3.2A, the kinetic curve is hybrid type kinetics and could be deconvoluted into two basic types of typical kinetic curves including one type with fast activation followed by fast decay and the other type with slow activation followed by slow decay, which might be derived from two different types of active sites. The kinetic curve (Fig. 3.2B) is only one single type kinetics with slow 138 Zhen Liu et al.

AB Polymerization rate Polymerization rate Polymerization

Reaction time Reaction time

Phillips chromium catalyst + AIR3 Phillips chromium catalyst + AIR3 Before slurry polymerization During slurry polymerization Phillips chromium catalyst + AIR2OR Before gas-phase polymerization Before/during slurry polymerization Before gas-phase polymerization Figure 3.2 Two types of kinetic curves for ethylene polymerization over Phillips chro- mium catalysts: (A) a typical hybrid kinetic curve and (B) a typical single kinetic curve.

activation followed by slow decay, which might be derived from one type of active sites. In the cases of activation of the Phillips chromium catalyst, the AlR3 cocatalyst added during slurry polymerization or before gas-phase polymerization (in catalyst preparation) showed a hybrid type kinetics (curve in Fig. 3.2A), while the AlR3 cocatalyst added before slurry polymerization (in catalyst preparation) showed a single type kinetics (curve in Fig. 3.2B). The activation of the Phillips chromium catalyst by AlR2OR cocatalyst could only present a single type kinetics (curve in Fig. 3.2B) regardless of the timing for introducing cocatalyst in both slurry and gas-phase ethylene polymerization. A basic understanding concerning the thermal activation during catalyst preparation, activation by CO or Al-alkyl cocatalyst, and activation by eth- ylene monomer during the induction period had been achieved through various spectroscopic methods and kinetic studies. Further emerging of new techniques applied in recent studies of the Phillips chromium catalyst could be observed, such as time-/temperature-resolved FTIR spectroscopy (Groppo et al., 2007), pressure-/temperature-resolved FTIR spectroscopy under in situ/operando conditions (Barzan et al., 2012; Lamberti et al., 2010), in situ XAS spectroscopy (Bordiga et al., 2013), laser ablation-mass spectrometry (LA-MS) and laser desorption-ionization mass spectrometry (LDI-MS) (Aubriet et al., 2006), etc. The characterization under close to the actual industrial conditions is still a challenge as well as an opportunity to shed Ethylene Polymerization and Oligomerization 139 some light on the related mechanisms. Investigation of the polymerization kinetics over the Phillips type catalyst combined with Al-alkyl cocatalyst could provide valuable information to guide the design and development of ethylene polymerization processes. More efforts should be devoted into the investiga- tionsofgas-phasepolymerizationkineticsthroughcombinationofexperiments with mathematical kinetic modeling and microkinetic modeling based on molecular simulations in the near future.

2.1.2 Model catalysts and molecular modeling Experimental studies on the traditional Phillips catalyst are hardly to give a clear image on the active sites and to unravel the mechanisms for ethylene polymerization. The difficulties are primarily induced by surface complexity of the Phillips catalyst including the heterogeneity of the porous amorphous silica support, the coexistence of mono-, di-, and polychromate species, the formation of Cr2O3 microcrystal, the low fraction of active chromium spe- cies in the total chromium loading, and so on (Groppo et al., 2005a). A step- forward understanding on the behavior of the active sites and the reaction mechanisms for the Phillips catalyst could be achieved through investiga- tions on well-defined model catalysts as well as molecular modeling simulations. 1. Model catalysts. Model catalysts usually feature with well-defined struc- tures and could provide much clearer information of the active site to understand the Phillips catalyst. Scheme 3.2 depicts some typical silica-supported heterogeneous model catalysts for the Phillips chromium catalyst. S-2 catalyst prepared by wet impregnation of bis(triphenylsilyl) chromate onto thermally pretreated silica gel could be considered as a

Scheme 3.2 Heterogeneous model catalysts for Phillips chromium catalyst. 140 Zhen Liu et al.

commercial heterogeneous model (1A) for the Phillips chromium cat- alyst (Carrick et al., 1972). The S-2 catalyst shows an increased activity after supporting onto silica gel compared with bis(triphenylsilyl)chro- mate and produces HDPEs with even broader MWD compared to that of the Phillips chromium catalyst. McDaniel (1982) reported a hetero- geneous model catalyst (2A) via mild grafting of CrO2Cl2 at 200 C onto thermally pretreated silica, which generated a surface monochromate structure and showed activity for ethylene polymerization similar to those of the Phillips chromium catalyst. Scott and coworkers (Demmelmaier et al., 2008, 2009) prepared the similar catalysts via ambi- ent anhydrous grafting of CrO2Cl2 onto silica pretreated at 200, 450, and 800C, respectively. The IR, XANES, and EXAFS spectroscopic results suggested that the higher polymerization activity over CrO2Cl2 grafted onto silica pretreated at 800 C was related to a more strained six- membered chromasiloxane ring (2A). A recent study by Scott et al. on the extremely air-sensitive divalent Phillips model catalysts showed that the coordination of the surface siloxane ligands on the divalent active site precursor and the calcination temperature were crucial for determination of the precise microstructures and coordination environ- ment of the active chromium species of the Phillips catalyst (Zhong et al., 2012a). Tonosaki et al. (2011a) utilized two different starting materials 3 3 (Cr( -allyl)3 and Cr2( -allyl)4) to vary the surface chromate structures of the catalysts. It was found that the surface dichromate model catalyst (3A) produced more methyl branches in the polyethylene products compared with the monochromate model catalyst (2A). Model catalyst (4A) is a supported alkylidene complex reported by Scott et al. (Ajjou and Scott, 1997, 2000; Ajjou et al., 1998a,b; Scott and Ajjou, 2001; Scott et al., 2008) which may initiate ethylene polymerization through Green-Rooney alkylidene mechanism (Ivin et al., 1978). Monoi et al. (2003) and Ikeda et al. (2003) reported a trivalent model catalyst by the supporting of Cr(CH(SiMe3)2)3 onto silica pretreated at 200 C and suggested that model catalyst (5A) is the most plausible active site for ethylene polymerization. Thu¨ne et al. (1997) prepared a flat surface model catalyst (6A) by impregnating aqueous CrO3 on a flat Si(100) sub- strate covered by amorphous silica layer. Then, the obtained model cat- alyst with monochromate supported on the silica surface showed ethylene polymerization activity at 160 C, while the prereduced surface Cr(II) species failed to polymerize ethylene due to its extreme sensitivity to air and moisture (Agostini et al., 2007). Ethylene Polymerization and Oligomerization 141

Heterogeneous model catalysts with more uniform and well-defined struc- ture of surface chromium species supported on silica gel had been demon- strated as a powerful strategy for the basic study of the Phillips chromium catalyst. However, the complexity is still coming from the heterogeneity of the porous amorphous silica support with more than 99% of the active sites existing on the inner surface within the micro- and mesopores of the silica support. In this regards, well-characterized homogeneous model catalysts have been developed in order to simplify such complexity of the traditional Phillips catalyst originating from the silica support. Theopold gave a microreview on the homogenous chromium catalysts for olefin polymerization mostly based on cyclopentadienyl chromium catalysts (Theopold, 1998). These compounds, however, exhibited little similarity compared with the traditional Phillips catalyst. Scheme 3.3 showed some typical homogeneous models of the Phillips chromium catalyst. Bis(triphenylsilyl)chromate (1B) was first proposed as a typical homogeneous model catalyst, which was able to polymerize ethylene without adding any organometallic cocatalyst at high temperatures (T>130C) and high pressures (P¼300–1500 atm) (Baker and Carrick, 1970). A polyhedral oligomeric silsesquioxanes (POSS) supported chro- mium complex [(cyclo-C6H11)7Si7O11(OSiMe3)]CrO2 (2B) developed by Feher et al. seemed to be a more realistic homogeneous model for the

Scheme 3.3 Homogeneous model catalysts for Phillips chromium catalyst. 142 Zhen Liu et al.

Phillips chromium catalyst. This compound polymerized ethylene under mild conditions (20 C, 1 atm) in the presence of AlMe3 as cocatalyst (Feher and Blanski, 1990, 1993). Some bimetallic chromium siloxane com- plexes were also synthesized to model the Phillips catalyst. Abbenhuis pointed out that a 12-memberedinorganic heterocycle [Cr(═O)2{(OSiPh2)2O}]2 (3B) was a potential homogeneous catalyst for ethylene polymerization, but no further experiment was conducted on this bimetallic Cr-based compound (Abbenhuis et al., 1997). Recently, Qiu et al. reported a novel homoge- neous triphenylsiloxy complex of chromium(II) model catalyst [(Ph3SiO) Cr(THF)]2(m-OSiPh3)2 (4B), which polymerized ethylene at low Al/Cr molar ratios (Al/Cr100) with methylaluminoxane (MAO) as cocatalyst. With further increasing Al/Cr molar ratio (Al/Cr>200), an interesting transformation of ethylene polymerization into ethylene nonselective olig- omerization was observed (Qiu et al., 2011). A similar transformation was also presented over bis(triphenylsilyl)chromate combined with MAO as cocatalyst, but the critical point of Al/Cr molar ratio for the transformation was much higher (Qiu et al., 2012). 2. Molecular modeling. To simulate the behavior of the real heterogeneous catalyst, a reasonable molecular model must be first built to mimic the active sites anchored on the support. Figure 3.3 shows some typical molecular models for the active sites of the Phillips chromium catalyst. Espelid and Børve had done a series of systematic density functional the- ory (DFT) investigations on the active sites of the Phillips chromium

Figure 3.3 Molecular models for Phillips chromium catalyst. Ethylene Polymerization and Oligomerization 143

catalyst (Espelid and Børve, 2000, 2001, 2002a, b, c). Accordingly, the six-membered chromasiloxane ring (1C) was proposed as a key model of the active site of the Phillips chromium catalyst. 1C was then adopted as a model of active chromium species in many other theoretical investiga- tions (Cheng et al., 2010; Damin et al., 2009; Demmelmaier et al., 2009; Tonosaki et al., 2011b). Demmelmaier et al. confirmed the validity of 1C as an ideal molecular model rather than the model with larger ring size for the Phillips chromium catalyst through a combination of exper- iments and theoretical calculations (Demmelmaier et al., 2008, 2009). Recently, Zecchina et al. reported the adsorption of probe molecules (CO, N2) on the cluster model 1C and found a good agreement between the experimental IR observations and the calculated vibrational frequen- cies by increasing the percentage of Hartree—Fock exchange in the hybrid density functional B3LYP (Damin et al., 2009). Interestingly, a clear mechanistic understanding on the transformation of a metathesis site into a polymerization site during the induction period of the Phillips chromium catalyst was achieved through DFT calculations with a comparison to the experimental findings (Liu et al., 2003; Zhong et al., 2012b). Furthermore, Liu et al. studied the effects of Ti-modification on the Phillips chromium catalyst using the six- membered chromacycles (1C, 2C, and 3C) and clarified some effects of Ti-modification over the Phillips chromium catalyst, such as the pro- motion of the polymerization activity, extension of MWD to the low molecular weight region, and improvement of the distribution of inserted comonomers (Cheng et al., 2010). Recently, Tonosaki et al. found that both the calculated activation energies for ethylene insertion and chain transfer were in good agreement using model 1C and an extended larger cluster (Tonosaki et al., 2011b). It was pointed out that the intrinsic origin of the broad MWD of the polyethylene pro- duced by the Phillips chromium catalyst might be derived from the mul- tiple coordination environments around the active chromium site on silica surface. It has long been recognized that the silica support is not an inert compo- nent of the catalyst which simply directs polymer particle morphology. The neglect of the real silica surface could introduce some artificial effects and provide unrealistic environment for the adsorption of monomer on active chromium center (Sautet and Delbecq, 2010). Nowadays, with the improvement of the computing resources and the development of quan- tum methodologies, high-level calculations using a large surface supported 144 Zhen Liu et al. model or a periodic model of silica gel surface can be performed. Very recently, Zhong et al. developed a surface model containing 37 Si atoms through anchoring of a six-membered chromasiloxane ring onto silica surface cutting from the b-cristobalite crystal structure (4C), and the modeling results were in good agreement with the experimental IR spectra concerning CO adsorption over prereduced Phillips chromium catalyst (Zhong et al., 2012a). Guesmi and Tielens (2012) reported an amorphous silica surface slab containing 120 atoms (Si27O5413H2O, 5C) represented the amorphous character of the hydroxylated silica surface containing different silanol types. Through a periodic DFT calculation, a higher sta- bility of mono-oxo and di-oxo chromium species was confirmed in com- parison with chromium-hydroxyl species. The main conclusion of their study came with a strong support of the six-membered chromasiloxane ring as a valuable molecular model on the amorphous silica surface for the Phillips chromium catalyst. Thus far, all the related theoretical calcu- lations mentioned above agreed that the six-membered chromasiloxane ring could be served as a reasonable cluster model for the Phillips chro- mium catalyst, but the effects of the silica support should be considered as well. Based on rational design of these active sites, the theoretical studies on the mechanisms of ethylene polymerization by the Phillips catalyst was reviewed by us very recently (Cheng et al., 2013). The initi- ation mechanism of ethylene polymerization in terms of the formation of the first CrdC bond over the active site of the Phillips catalyst and the transformation mechanism for the reactions during the induction period of the Phillips chromium catalyst was elucidated by means of molecular modeling simulations (Espelid and Børve, 2000, 2001, 2002b,c; Zhong et al., 2012b). After all, it is still very difficult to obtain direct evidence for the real active sites of the Phillips catalyst and for the ethylene polymerization mechanisms through traditional kinetic studies. The rational design and utilization of novel model catalysts for the Phillips chromium catalyst could be expected to allow further progress with better understanding of the real and complex catalyst system. A step forward in this field requires the combination of model catalysts with more advanced and multiple characterization techniques especially in situ/operando techniques as well as molecular modeling. With fast growing of computing power and in-depth development of quantum packages, one could perform theo- retical calculations based on more realistic surface models resembling the real Phillips chromium catalyst. The combination of experiments Ethylene Polymerization and Oligomerization 145 and theoretical calculations is crucial and would result in more profound and interesting findings.

2.2. Characterization of microstructures of polyethylene chains Phillips chromium catalyst could produce polyethylene with broad MWD, long-chain branches, and short-chain branches (SCBs). These special fea- tures bring its HDPE products with good mechanical properties, such as environmental stress-cracking resistance (ESCR) and yield stress. One of the most important application areas of its HDPE products is manufacture of high-grade polyethylene pipe materials like PE80 and PE100. The crucial factor to manipulate the long-term mechanical properties of these high- grade pipe materials is to control the short-chain branch distribution (SCBD) of the HDPE products made from Phillips chromium catalyst through copolymerization of ethylene with a-olefin, such as 1-butene and 1-hexene. Therefore, characterization of microstructures of polyethyl- ene chains in terms of SCBD is very important in order to correlate catalyst behavior with structure properties of polyethylene products. The HDPE pipe materials are widely used for the transportation of nat- ural gas, syngas, water, etc. It is generally agreed that the MW, MWD, and SCBD of ethylene/a-olefin copolymers are the most important key factors determining the mechanical properties of the HDPE pipe materials. The relationship of MW and MWD to the mechanical properties had been fully investigated in early time by many researchers such as Kennedy et al. (1994), Failla et al. (1994), and Jordens et al. (2000), etc. In this aspect, a relatively high MW and a very broad MWD are necessary for high-grade polyethylene pipe materials. Later on, a number of studies showed that the long-term mechanical properties of the pipe materials are mainly determined by the SCBD of the polyethylene chains (Fan et al., 2009; Garcia et al., 2008; Luruli et al., 2007; Mortazavi et al., 2010; Razavi-Nouri, 2006; Stadler et al., 2007). The report from the research groups, such as Soares et al. (2000), Hubert et al. (2001, 2002), and McDaniel (Deslauriers and McDaniel, 2007; Deslauriers et al., 2005) confirmed the important effect of SCBD of ethylene/a-olefin copolymers on the long-term mechanical properties of HDPE pipe materials. In general, the more SCBs on the high MW parts and the less SCBs on the low MW parts are beneficial for the high performance of high-grade polyethylene pipe materials (Krishnaswamy et al., 2008). In this regards, two different methods namely TREFþ 13C NMR and TREFþstep crystallization (SC) as effective methods to 146 Zhen Liu et al. characterize the SCBD of ethylene/a-olefin copolymers for HDPE pipe materials were successfully developed (Zhang et al., 2012a). Two HDPE pipe material samples PE-1 and PE-2 with different long- term mechanical properties were fully characterized by means of the above- mentioned two combined methods. The physical and mechanical properties of the two HDPE samples are listed in Table 3.1. Apparently, PE-2 sample showed a much larger ESCR value than that of PE-1. In order to analyze the SCBD of these two samples, the first step is to use temperature rising elution fractionation (TREF) to physically separate the two HDPE samples and get their fractions for the subsequent characteriza- tions. The weight distribution of fractions versus crystalline temperature of the two HDPE samples is shown in Fig. 3.4. Both TREF curves are narrow as typical HDPE samples with very small amount of low and high temper- ature fractions. When the fractions were obtained by TREF method, 13C NMR was applied to characterize 1-hexene content in six selected fractions of each sample (PE-1 and PE-2). High temperature GPC characterization of the six selected fractions of each sample demonstrated that the fraction at higher crystallization temperatures with higher crystallinity showed higher MW.

Table 3.1 The physical and mechanical properties of PE-1 and PE-2 samples Density HLMI Yield stress Breaking Crystallinity ESCRa Sample (g/cm3) (g/10 min) (MPa) elongation (%) (%) (h) PE-1 0.9413 23.6 18.3 830 63.1 1.78 PE-2 0.9430 10.3 21.4 702 63.9 >4000 aAt 80 C and 4.6 MPa with F323 mm pipe.

Figure 3.4 Weight distribution curveP of fractions after TREF fractionation of the HDPE samples (A) PE-1 and (B) PE-2. a: Wi b: Wi/(△T). Ethylene Polymerization and Oligomerization 147

3 PE-1 PE-2

2

1 1-Hexene incorporation (mol%) 0 0.0 2.0 × 105 4.0 × 105 6.0 × 105 M W Figure 3.5 The 1-hexene content distribution curves of PE-1 and PE-2.

A B

a a

b b

c c

d d Heat flow (W/g) Heat flow (W/g) e e f f

60 70 80 90 100 110 120 130 140 150 60 70 80 90 100 110 120 130 140 150 Temperature (°C) Temperature (°C) Figure 3.6 DSC curves of different fractions of (A) PE-1 and (B) PE-2 after SC. a: 113 C (A)/112 C (B), b: 105 C (A)/103 C (B), c: 87 C, d: 70 C, e: 60 C, f: 30 C.

As shown in Fig. 3.5, the 1-hexene comonomer content of PE-2 at low crystallinity and low MW fractions was less than that of PE-1 sample and kept even higher comonomer content at the high crystallinity and high MW parts. The SC method could be used to indirectly characterize the SCBD of different polymers. The differential scanning calorimetry (DSC) curves of the selected fractions of PE-1 and PE-2 are shown in Fig. 3.6. After the treat- ment of SC, the DSC curves showed a similar profile for each of the high temperature fractions above 87 C. Moreover, a strong single peak was 148 Zhen Liu et al. observed at about 132 C with a broad shoulder. The main difference appeared in the low temperature fractions, especially for the 30 C fraction. It was clear that the low temperature fractions of PE-1 sample showed more low temperature multiple peaks in the DSC curves, which suggested that there were much thinner lamella in low temperature fractions of PE-1 sam- ple than those of PE-2. This result was consistent with the comonomer con- tent distribution of PE-1 with higher 1-hexene incorporated in the low crystalline temperature and low MW parts as previously shown in Fig. 3.5. According to Thomas Gibbs equation, we could calculate the lamella thickness distribution from those endothermal peaks in the DSC curves through multiple peaks fitting method. The calculated lamella thickness dis- tribution of the selected fractions of PE-1 and PE-2 are shown in Fig. 3.7. Generally speaking, the more incorporated comonomer would result in a thinner lamella thickness. Comparing the fractions of PE-1 and PE-2 obtained at 30 C, the thinner lamella thickness content of fractions of PE-1 is higher than that of PE-2 indicating that more incorporated 1-hexene in PE-1 sample in low temperature and low MW fractions. Although the DSC curves showed a similar profile of high temperature frac- tions of PE-1 and PE-2, the calculated lamella thickness distribution of the high temperature fractions of PE-2 was slightly thinner than that of PE-1. Therefore, it indicated that the SCBs in the low temperature fractions of

AB 112 °C 112 °C

105 °C 103 °C

87 °C 87 °C

70 °C 70 °C

60 °C

Relative percentage (%) Relative percentage (%) 60 °C

30 °C 30 °C

4 6 8 10121416182022 4 6 8 10121416182022 L (nm) L (nm) Figure 3.7 Lamella thickness distribution of different fractions of (A) PE-1 and (B) PE-2. Ethylene Polymerization and Oligomerization 149

PE-1 was slightly higher than that of PE-2, but the opposite situation occurred in the high temperature fractions of these two samples. It was dem- onstrated that lower the content of comonomer in the low MW parts and increase the content of comonomer in the high MW parts would be crucial in order to achieve much better long-term mechanical properties of the HDPE pipe materials. Two efficient combined methods have been developed and applied to analyze the SCBD of the HDPE pipe materials. Although 13C NMR could precisely and quantitatively characterize comonomer contents of each HDPE fraction, it is very time consuming and expensive, while SC method combined with TREF could effectively and qualitatively describe the rela- tive comonomer contents and the SCBD of the HDPE pipe materials with low cost and much shorter time. More recently, a more convenient but sim- ilar method using TREF combined with successive self-nucleation and annealing (SSA) was successfully utilized to investigate the SCBD of the eth- ylene/1-hexene copolymers made by a novel inorganic and organic hybrid Cr-based catalyst (Zhang et al., 2012b, 2013). The relative SCBs contents for each PE fraction from TREF was qualitatively obtained from the lamella thickness measured by SSA. Characterization methodologies of the micro- structures of polyolefins will surely be playing more and more important role in catalyst innovation in the near future.

2.3. Polyethylene-based nanocomposites HDPE pipe materials produced by the Phillips chromium catalyst have the following distinct advantages: low cost, lightness, good corrosion resistance, high ductility, and excellent mechanical properties (Galli and Vecellio, 2004). However, when long-term exposure to sunlight and air, ultraviolet resistance becomes very important characteristic for these HDPE pipe mate- rials. Incorporation of CB into polyolefin matrix can enhance the ultraviolet resistance, weather adaptability, and therefore prolong the service life time (Hubert et al., 2001). The resistance to ultraviolet degradation is usually related to the morphology, particle size, and surface properties of the CB. However, the perfect dispersion of CB is hampered by the strong inter- particle forces of CB and weak polymer–CB interfacial interaction. On one hand, the ultraviolet resistance property of HDPE pipe materials will be determined by the homogeneous dispersion of CB. On the other hand, the long-term mechanical properties will be determined by not only the dis- persion state of CB but also a strong polymer–CB interfacial interaction. 150 Zhen Liu et al.

The graft of polymers onto carbon nanoparticles is an effective method to improve polymer–carbon interfacial interaction and the dispersion quality of carbon in polymer matrix, and consequently enhance the physical prop- erties and ultraviolet resistance (Paiva et al., 2004; Wang et al., 2005; Xu et al., 2006). Traditionally, there are three different approaches for grafting polymers onto CB surface: (a) “grafting from” method: the surface grafting of polymer chains was initiated from initiating groups introduced onto the CB surface; (b) “grafting onto” method: grafting onto the surface during the polymerization, initiated by a conventional radical initiator in the presence of CB; and (c) polymer reaction method: polymers having terminal func- tional groups were reacted with functional groups on the CB surface (Tsubokawa, 1992; Tsubokawa et al., 1995). However, all of these methods have similar limitations: (1) rigorous and complicated pretreatment; (2) can- not be applied effectively to the grafting of polymers onto with few functional groups (e.g., furnace black and acetylene black). We recently reported the use of HDPE matrix-grafted CB (HDPE-g- CB) prepared by thermal mechanical technique to reinforce commercial HDPE pipe materials (He et al., 2012). The CdC a-bonds of the methine groups of the polyethylene chains encapsulated around the CB surface could be cleaved to produce chain radicals under the intensive thermal and mechanical effects (Hoang et al., 2006). The polymer chain radicals could be terminated either by other chain radicals or by the CB surface as a strong radical scavenger because of its polycondensed aromatic rings. The HDPE- g-CB pipe material was prepared by mixing the HDPE and CB at 140 C for 15 min to graft the molten nonpolar HDPE chains on the larger nonpolar surface of CB. The FTIR spectroscopy is a very useful measurement to study the functional groups chemically attached to the surface of the CB. Figure 3.8 shows the FTIR spectra of (A) pristine CB and (B) HDPE-g- CB after extraction. A peak at 1718 cm-1 occurred in the spectrum of HDPE-g-CB, indicating the presence of carbonyl groups (C]O) on the CB surface. It has been reported that the carbonyl (C]O, 1718 cm-1) groups were formed by an oxidation reaction in the presence of O2 atmo- sphere (Hoang et al., 2006). The peaks in the 2850–2912 cm-1 region are the stretching of CdH bond and the intensities were greatly enhanced after the grafting reaction. Other peaks, such as 477, 707, and 804 cm-1, were the fin- gerprint peaks of HDPE. Due to the stain derived from the strong entanglement of the polyeth- ylene attached on the CB surface, the CB aggregates could be more easily broken down into smaller particles. In addition, the grafted polymer layer Ethylene Polymerization and Oligomerization 151

Figure 3.8 FTIR spectra of (A) pristine CB, and (B) HDPE-g-CB.

Figure 3.9 TEM micrographs of (A) pristine CB and (B) HDPE-g-CB. protected the new generated CB from van der Waal’s attraction to other CB particles. As a result, the HDPE-g-CB could retain nanodimensions at about 100 nm with visible textures of the CB specimens as shown in Fig. 3.9.In contrast, the particles in pristine CB retained a much larger agglomerate at the dimension of about 1–10 mm. The load of grafted polyethylene can be calculated quantitatively according to the TG measurement. The pristine CB and HDPE-g-CB showed 1.2 and 11.2 wt% total weight loss at 500 C, respectively. The weight percentage of grafted polyethylene in the HDPE-g-CB after extrac- tion was about 10.1%. The HDPE/CB and HDPE/HDPE-g-CB composites were thus prepared by mixing the HDPE pipe materials with pristine CB and 152 Zhen Liu et al.

Figure 3.10 SEM micrographs of HDPE/pristine CB (A1, A2) and HDPE/HDPE-g-CB (B1, B2) with the CB content at 1.0 and 10.0 wt%, respectively.

HDPE-g-CB, respectively. The content of CB or HDPE-g-CB in the poly- mer composites was set to be 1.0, 5.0, 10.0, and 15.0 wt%. As an example, the scanning electron microscopy (SEM) images of the fracture surfaces of HDPE/CB and HDPE/HDPE-g-CB composites with CB or HDPE-g- CB content of 1.0 and 10.0 wt% are shown in Fig. 3.10. In the SEM photos of the HDPE/CB composites (Fig. 10A1, A2), there is quite a broad distri- bution of the particle size and large aggregates and agglomerates were clearly observed, indicating a large number of unbroken CB and a poor polymer-CB adhesion. On the other hand, the grafting process efficiently broke down the large aggregates and agglomerates of CB, leading to the homogeneous disper- sion of the CB during composites blending. In the SEM photos of the HDPE/ HDPE-g-CB composites (Fig. 10B1, B2), the carbon particles were uni- formly dispersed in the HDPE matrix, and the particle size was remarkably decreased. The results are in good agreement with respect to the transmission electron microscopy (TEM) images (Fig. 3.9) and indicate that polyethylene- grafted CB was effective in improving the dispersion states of the CB. More- over, it could be inferred that the HDPE matrix grafted onto the CB surface improved the compatibility between CB and the polyolefin matrix. Ethylene Polymerization and Oligomerization 153

It was found that the HDPE/HDPE-g-CB composites generally showed better tensile properties than that of the HDPE/CB composites. For the HDPE/HDPE-g-CB composite with CB content of 5.0 wt%, the tensile strength and ultimate strain were improved by 17% and 30%, respectively, as compared to HDPE/CB composite. The improvement of the tensile properties of the HDPE/HDPE-g-CB composites might be due to the enhanced crystallization, the improved dispersion state of CB in the matrix, and a more efficient load transfer from the pipe materials matrix to the CB nanoparticles induced by the HDPE-g-CB (Tang et al., 2003). The impact strengths of HDPE/HDPE-g-CB composites were improved by 13.4%, 4.3%, and 6.5% when the content of CB was 1.0, 2.5, and 10.0 wt%, respectively, compared with that of HDPE/CB compos- ites. This is possibly caused by the smaller particle size and much better dis- persion of the modified CB, which results in stronger interfacial interaction between CB and PE. It is well known that stress dissipation at the tip of a growing crack by a viscoelastic process results in strain resistance of the poly- mers. The strong interfacial interaction between HDPE-grafted CB and polymer chains renders efficient dissipation of the applied stress, and thereby higher resistance to external force. It is also noted that for composites con- taining HDPE-g-CB, there was a downturn in mechanical properties when the CB content was increased to 15.0 wt%. Therefore, even when the nanoparticles are grafted with polymer chains, it can still be difficult to homogeneously disperse the nanoparticles in the polymer matrix at higher nanoparticle loadings. The ultraviolet resistance after incorporation of CB into polyethylene matrix for both the HDPE/HDPE-g-CB and the HDPE/CB composites was further compared. As shown in Fig. 3.11, at the CB load of 1.0 wt%,

Figure 3.11 Absorption of ultraviolet and visible light of (A) HDPE/CB and (B) HDPE/ HDPE-g-CB composites with the content of 1.0 and 5.0 wt%, respectively. 154 Zhen Liu et al. the pipe materials mixed with HDPE-g-CB exhibited slightly higher adsorption of ultraviolet radiation than that adsorbed by HDPE/CB com- posite. As the CB load increased to 5.0 wt%, the ultraviolet adsorption of HDPE/HDPE-g-CB composites increased significantly. In addition, the HDPE-g-CB absorbed more ultraviolet and visible radiation with higher CB content. The polyethylene chains grafted onto carbon nanoparticles reduce the tendency of CB to aggregate and the nanoparticles can be dis- persed more homogeneously in the polymer matrix (see Fig. 3.10). In the presence of HDPE-g-CB, ultraviolet light is more likely to be absorbed rather than scattered and the polyethylene ultraviolet degradation protective effect was enhanced. As a result, the ultraviolet resistance of composites con- taining HDPE matrix-grafted CB were superior to composites containing only pristine CB. The tensile strength, ultimate strain and toughness of the HDPE were all improved upon the incorporation of proper amount of HDPE-g-CB, and the incremental improvements of CB dispersion enhanced the ultraviolet energy absorption and thus would positively influence the expected lifespan of the HDPE pipe materials. The encapsulated polymer layer was effective in the reduction of the surface free energy of CB and improving the compat- ibility between CB and polymer matrix. The HDPE-g-CB was uniformly dispersed in the HDPE pipe materials, and the particle size was remarkably decreased. With much enhanced mechanical properties, such as tensile strength, elongation at break and impact strength, and the ultraviolet resis- tance, HDPE/HDPE-g-CB composite could possibly be applied for making high grade of HDPE pipe materials.

3. PHILLIPS CHROMIUM CATALYSTS FOR ALKYNE CYCLOTRIMERIZATION

Phillips chromium catalyst has been primarily used as ethylene poly- merization catalyst in a large industrial scale. However, the same catalyst could cyclotrimerize acetylene and methylacetylene into benzene and tri- methylbenzene (TMB) rather than polyacetylene and polymethylacetylene, respectively (McDaniel, 2010). The mechanism has never been studied up to now, which also own particular academic interests in this field. Hogan and coworkers first reported alkyne cyclotrimerization catalyzed by the Phillips chromium catalyst only a few years later after they invented the Phillips chromium catalyst (Clark et al., 1959). In their report, the acetylene was found primarily cyclotrimerized into benzene, while a ratio of 0.18 of Ethylene Polymerization and Oligomerization 155

1,3,5-TMB to 1,2,4-TMB was obtained for the methylacetylene cyclo- trimerization. A decade ago, Zecchina et al. revisited this catalytic system and found that 1,3,5-TMB is the only product of the cyclotrimerization of methylacetylene (Zecchina et al., 2003). Recently, we performed a the- oretical investigation on the mechanism of acetylene and methylacetylene cyclotrimerization catalyzed by the Phillips chromium catalyst (Liu et al., 2012, 2013). The ground spin state of chromium(acetylene) adducts is known to be of quintet, and the most plausible reaction pathway on quintet surface needs to overcome two activation barriers to finish a single catalytic cycle, as depicted in Scheme 3.4. However, the reaction on quintet surface is prohibited by presenting a free-energy barrier of 31.1 kcal/mol that transforms two coordinated acetylene into the key intermediate 54D. Thus, the turnover of frequency (TOF) for the catalytic cycle on the quintet surface is 1.3610 9 h 1, which rules out the quintet reaction mechanism for acet- ylene cyclotrimerization by Cr(II)/SiO2 model catalyst. Next, the reaction crosses to the adjacent triplet surface via a minimum energy crossing point (MECP) 5-3CPI and the following reaction proceeds on the triplet surface. As shown in Fig. 3.12, a chromacyclopropene species 31D is generated immediately after 5-3CPI on the triplet surface without showing any transition state. The key intermediate chromacyclopentadiene species 34D is generated through a rather facile insertion of a second

Scheme 3.4 Catalytic cycle for acetylene cyclotrimerization by Cr(II)/SiO2 on the quintet surface. 156 Zhen Liu et al.

Figure 3.12 Gibbs free energy profiles for acetylene cyclotrimerization by Cr(II)/SiO2 cluster model. The Gibbs free energies are calculated at 298.15 K, 1 atm as default in Gaussian09. Also shown are the total energies in parentheses. The triplet reaction path- way is depicted in gray, while the quintet parts are in black. Energies are in kcal/mol and relative to 51C plus the corresponding number of acetylenes.

acetylene into the three-membered ring in 31D. The adsorption of a third acetylene through formation of a hydrogen bond and the subsequent [4þ2] cycloaddition with a moderate free-energy barrier of 13.7 kcal/mol leads to a coordinated cyclohexadiene-like species 36D. The catalytic cycle is finally finished on the quintet surface through another MECP 3-5CPII. The detachment of the coordinated benzene ring from 56D requires 7.4 kcal/mol of Gibbs free energy. Thus, the conclusion is reached with a two-state reactivity following the pathway 51C! 51D! 5-3CPI! 31D ! 32D! 34D! 35D! 36D! 3-5CPII! 56D! 51C0 (Liu et al., 2012). The proposed two-state mechanism for acetylene cyclotrimerization requires two spin-inversion processes, as shown in Fig. 3.12. The first MECP 5-3CPI is crucial for initiation of the acetylene cyclotrimerization. In this regards, the insertion of a second acetylene molecule into a three- membered ring on the triplet surface was found to be much more facile than that proceeds through oxidative coupling on the quintet surface. Rather than crossover to the quintet surface through 3-5CPII, the displacement of benzene ring by acetylene on the triplet surface was predicted to be ther- modynamically favorable with an exergonicity of 18.9 kcal/mol. Therefore, the acetylene cyclotrimerization initiates with coordination of an acetylene molecule on the quintet surface. After a spin-flipping at the quintet chromium(acetylene) complex, the following catalytic cycle favors a triplet [4þ2] cycloaddition pathway as shown in Scheme 3.5. The calculated TOF Ethylene Polymerization and Oligomerization 157

Scheme 3.5 Two proposed mechanisms for acetylene cyclotrimerization by Cr(II)/SiO2 cluster model: two-state reactivity versus a triplet catalytic cycle.

for the above-mentioned two-state catalytic cycle is about 53 h 1, which is much lower than that for the reaction pathway on a single triplet surface with a TOF of 728 h 1. The spin crossover phenomenon has also been reported in other transition-metal catalyzed alkyne cyclotrimerization reac- tions (Agenet et al., 2007; Gandon et al., 2006; Martinez et al., 2005; Xu et al., 2008). We further studied methylacetylene cyclotrimerization by Cr(II)/SiO2 model catalyst. Similar to acetylene cyclotrimerization, all the quintet reac- tion pathways for methylacetylene cyclotrimerization need to overcome much higher free-energy barriers leading to very low TOFs for the catalytic reactions. The feasibility of spin-flipping reaction was examined at the naked cluster model 51C, the most stable mono-methylacetylene-chromium com- plex 51Ea, and the most stable di-methylacetylene-chromium complex 52Ea, respectively. As shown in Fig. 3.13, the spin flipping reaction was predicted to take place at 51Ea through an MECP 5-3CPIII to its triplet ana- log 31E. After this transition, a raised reactivity for the catalytic cycle on the triplet surface is highly expected. As shown in Scheme 3.6, the first reaction pathway from 31E lead to the cyclic product 1,2,4-TMB through either an intermolecular [4þ2] cyclo- addition pathway on potential energy surface (PES)–T1a or an insertion and reductive elimination pathway PES–T1b. The Gibbs free energy profiles are depicted in Fig. 3.14. The key intermediate 34Ea is formed through direct insertion of a second methylacetylene into the three-membered ring in 32Ea 158 Zhen Liu et al.

Figure 3.13 Gibbs free energy profile at 298.15 K for the spin crossover at 51Ea. The quintet parts are shown in black, while the triplet complexes are in gray. The MECPs are marked with solid cycle. Energies are in kcal/mol and relative to 51C plus the corresponding number of methylacetylenes.

Scheme 3.6 Mechanisms for cyclotrimerization of methylacetylene on the PES–T1. with a Gibbs free energy barrier of 5.7 kcal/mol. This is an exergonic process by 26.5 kcal/mol. A reductive elimination of 34Ea to generate a dimeriza- tion product 35Ea is predicted to be prohibitive by the presence of a high Gibbs free energy barrier of 26.6 kcal/mol. Alternatively, 36Ea is produced by adsorbing a third methylacetylene molecule in a methyl-group-down orientation through the formation of a weak hydrogen bond. This process requires Gibbs free energy of 4.0 kcal/mol. The rate-determining transition-state 3TS[6Ea–9Ea] on the concerted pathway PES–T1a was Ethylene Polymerization and Oligomerization 159

Figure 3.14 Gibbs free energy profile at 298.15 K of the triplet reaction pathway (PES–T1) for methylacetylene cyclotrimerization over the Cr(II)/SiO2 cluster model. The reaction pathway via intermolecular [4þ2] cycloaddition is depicted in black, while the stepwise pathway is in gray. The reaction to generate 1,3-dimethyle-cyclobutadiene 35Ea is in light black. Energies are in kcal/mol and relative to 51C plus the corresponding number of methylacetylenes. found to be 2.2 kcal/mol lower in Gibbs free energy than that of the transition-state 3TS[6Ea–7Ea] on the stepwise pathway. Therefore, the intermolecular [4þ2] cycloaddition is the most favorable pathway to pro- duce 1,2,4-TMB on the PES–T1. The first catalytic cycle is then finished through the displacement of the 1,2,4-TMB in the complex 39Ea by a methylacetylene molecule. After the thermal replacement, a more stable complex 31E0 is regenerated by releasing a free 1,2,4-TMB arene and is ready for the next turn of the catalytic cycle. The coordination of an unsymmetrical methylacetylene molecule in a different orientation plays a key role in determining the regioselectivities of the cycloaddition products. As shown in Scheme 3.7, four kinds of di-methylacetylene-chromium complexes could be generated through the coordination of a second methylacetylene molecule on a Cs symmetric 31E. As discussed above, 32Ea leads to the first [4þ2] cycloaddition path- way PES–T1a. The other three [4þ2] cycloaddition pathways PES–T2a, PES–T3a, and PES–T4a are also depicted in Scheme 3.7. The pathway PES–T4a leads to production of 1,3,5-TMB, while the other three pathways generate 1,2,4-TMB. Figure 3.15 shows the full intrinsic reaction coordinate trajectories for the four [4þ2] cycloaddition pathways. The four pathways 3 proceed smoothly with two moderate activation barriers TS[2E–4E](a–d) 3 and TS[6E–9E](a–d). The reaction coordinate is finished within about 170 steps for PES–T2a and PES–T4a, and about 220 steps for PES–T1a and PES–T3a with a stepsize of 0.2 amu1/2Bohr, respectively. 160 Zhen Liu et al.

Scheme 3.7 Various reaction pathways for cyclotrimerization of methylacetylene on the triplet surface. The Gibbs free energies for the intermediates are listed in parenthe- ses and the Gibbs free energies for the transition states are shown above the arrow. Energies are in kcal/mol and relative to 51C plus the corresponding number of methylacetylenes.

Figure 3.15 Intrinsic reaction coordinates for methylacetylene cyclotrimerization to give a cyclic product through a concerted [4þ2] cycloaddition. Also shown are the opti- mized geometries for the transition states 3TS[6E–9E].

Table 3.2 listed the calculated TOFs for the four [4þ2] cycloaddition pathways. As shown in Table 3.2, the temperature affects the magnitude of the TOF enormously with a slightly different selectivity between 1,3,5-TMB and 1,2,4-TMB. The ratio of the TOF for producing 1,3,5- TMB and 1,2,4-TMB is 0.48 at 363 K, which is larger than the experimental value of 0.18 (Clark et al., 1959). We further investigated the effect of silica support through building a large silica-supported cluster model, denoted as silica-A, for the most plausible reaction pathways. The TOF for each of the Ethylene Polymerization and Oligomerization 161

Table 3.2 The TOFs for the most plausible reaction pathwaysa via an intermolecular [4þ2] cycloaddition of methylacetylene into TMB 250 K (I/II/III)b 298 K (I/II/III) 363 K (I/lI/III) PES–T1a 0.0020/0.022/ 0.093/0.60/3.3106 140/570/1.0108 1.9106 PES–T2a 0.0037/0.034/ 0.16/0.85/6.3106 170/670/1.5108 4.9106 PES–T3a 0.012/0.20/ 0.50/4.3/3.0106 440/3200/9.6107 3.6105 PES–T4a 0.0082/0.092/ 0.30/2.0/2.8106 360/1700/1.1108 1.6 106 Ratio of TOF 0.46/0.36/0.22 0.40/0.35/0.22 0.48/0.38/0.32 (1,3,5-/1,2,4- TMB) aThe Gibbs free energy profiles employed for the calculation of TOFs are calculated at 250 K and l atm, 298.15K and l atm, and 363 K and 40.8 atm, respectively. TOFs are in h 1. b(I) Gibbs free energies calculated using model A; (II) Gibbs free energies calculated using model silica— A; (III) Gibbs free energies calculated using model silica—A with consideration of the dispersion inter- action in the energetics. reaction pathways increased by a factor 9–17 at 250 K, 5.3–8.6 at 298.15 K, and 3.9–7.3 at 363 K, respectively. For instance, the TOFs calculated at 363 K using the six-membered cluster model A for the [4þ2] cycloaddition pathways PES–T1a, PES–T2a, PES–T3a, and PES–T4a are 140, 170, 440, and 360 h 1, respectively. When using the extended model silica-A, the cal- culated TOFs increased to 570, 670, 3200, and 1700 h 1, respectively. Interestingly, the ratio of 1,3,5-TMB to 1,2,4-TMB in the product decreased at all the three conditions indicating that 1,2,4-TMB is a major product on a silica-supported active site. Moreover, the dispersion correc- tions are considered for all the silica-supported species and the TOFs calcu- lated using the corrected Gibbs free energies are also given in Table 3.2. 3 3 Since the transition states TS[2E–4E](a–d) and TS[6E–9E](a–d) are highly stabilized relative to starting complexes by dispersion corrections, the corresponding TOFs for all the four reaction pathways increased by about 4–8 orders of magnitude calculated at three different conditions. The ratio of 1,3,5-TMB to 1,2,4-TMB decreases to 0.22, 0.22, and 0.32 at 250, 298.15, and 363 K, respectively. The TOF is extremely sensitive to the calculated Gibbs free energies and depends exponentially on the activation energies (Kozuch, 2012; Kozuch and Shaik, 2006, 2008, 2011; Uhe et al., 2011). 162 Zhen Liu et al.

Although the calculated ratio of 1,3,5-TMB to 1,2,4-TMB of 0.32 at 363 K is larger than the experimental value, it is predictive for the selectivity. The 1,2,4-TMB is preferred as a dominant product for the cyclotrimerization of the methylacetylene catalyzed by the Phillips chromium catalyst. It is worthy of note that the titanium-catalyzed methylacetylene cyclotrimerization gives a similar ratio of 0.33 of 1,3,5-TMB to 1,2,4-TMB (Pierce and Barteau, 1994). The acetylene and methylacetylene cyclotrimerization follow [4þ2] cycloaddition mechanism, which rules out the proposed [2þ2þ2] mech- anism by Zecchina et al. (2003). After reduction of chromium to divalent state, the model catalyst and all the chromium(alkyne) adducts showed a quintet ground spin state. However, the following reaction on the quintet surface is inhibited by the oxidative coupling of the two coordinated alkynes to yield a five-membered ring. The spin-flipping to the triplet surface is hardly to occur at the naked chromium(II) in the model catalyst as the MECP lies much higher in energy. The coordination of an alkyne on the cluster model lowers the energy gap of the two adjacent states of the chromium(alkyne) complex and the spin-flipping reaction is thus deter- mined with great feasibility. After the spin transition, the CrdC bond is formed immediately in triplet chromacyclopropene species and the follow- ing insertion of a second alkyne is easy to proceed. Therefore, the first CrdC bond could not be formed on the quintet surface, but is formed immediately after the spin-flipping to the triplet surface at the chromium (alkyne) complex. As it can be seen, how is the spin surface crossing phenomena affect the catalytic reactivity, which has been frequently encountered in Cr-catalyzed alkene polymerization and alkyne cyclo- trimerization, is still worthy of great research attention in the field of transition metal-catalyzed coordination chemistry.

4. MOLYBDENUM CATALYSTS FOR ETHYLENE POLYMERIZATION

The traditional Phillips chromium catalyst was prepared with highly toxic CrO3. Although a lot of efforts had been devoted in replacing CrO3 with much less poisonous chromate(III) acetate, the obtained catalyst after calcination was still in the form of highly toxic hexavalent chromate species (McDaniel, 1985, 2010). Furthermore, the chromium compounds contained in wastewater and solid dust produced during the catalyst prepa- ration and the chromium residues in polyethylene products might cause Ethylene Polymerization and Oligomerization 163 significant damage to the environment and were very harmful to human’s health. With increasing concerns in the Cr-induced human health and envi- ronment problems, a green catalyst is highly expected as a potential alternate of the Phillips chromium catalysts in the future (McDaniel, 2013). Sited with chromium in the same group 6B in the Periodic Table of Elements, molybdenum attracts many efforts in the study of ethylene poly- merization. As early as in 1950s, Indiana Standard Oil Company discovered supported molybdenum oxide catalyst was active for ethylene polymeriza- tion (Field and Feller, 1957). However, the catalyst was then abandoned because of its poor catalytic performance. Our preliminary experiments on the MoOx/SiO2 catalyst also showed very low reactivity with compar- ison to the CrOx/SiO2 catalyst for ethylene polymerization. In order to improve the activity of Mo-based catalyst for ethylene polymerization, the active valence states of molybdenum sites, and the mechanism of the cat- alytic reaction should be first elucidated. We performed a detailed theoret- ical study combined with experiments to investigate the active oxidation states of molybdenum and the effects of surface hydroxyl on the polymer- ization activity of supported Mo-based catalysts (Cao et al., 2010). In the experiments, Al/Mo molar ratio was changed from 2.5 to 30 to explore the optimum condition for the highest activity of MoOx/SiO2 catalyst for ethylene polymerization. The results showed that the highest catalyst activity of 2.4 g-PE/g-cat/h was obtained at Al/Mo¼5. Figure 3.16 shows the DSC and FTIR profiles of the polymer produced by the MoOx/SiO2 catalyst. The melting point of the polymer is 134.8 C indicating that the polymer might be polyethylene. The FTIR spectroscopy further confirmed that the obtained polymer is polyethylene.

Figure 3.16 DSC and FTIR profiles of the polymer produced by MoO3/SiO2 catalyst (The catalyst residue was not removed from the polymer.) 164 Zhen Liu et al.

Two typical symmetry and anti-symmetry vibration frequencies of CH2 group were observed at 2848 and 2920 cm 1, respectively. Thus, MoOx/SiO2 catalyst was confirmed to be active for ethylene polymerization although the activity was relatively low. The activity of MoOx/Al2O3 catalysts for ethylene polymerization was 179 g-PE/g-cat/h, which was much higher than that of the MoOx/SiO2 catalyst prepared in this work. It was concluded that MoOx/Al2O3 and MoOx/SiO2 catalysts were all active for ethylene polymerization, but a direct comparison of these two catalyst systems was quiet difficult only using experimental techniques. In order to develop green and highly efficient Mo-based polyethylene catalysts, the corresponding models of the active sites for MoOx/Al2O3 and MoOx/SiO2 catalysts were established in the fol- lowing computational modeling work. As reported in the literature, Mo species mainly existed in an isolated form in the low molybdenum content catalyst. In our work, four kinds of isolated molybdenum models of the active sites were built with a consid- eration of different valence states, as shown in Scheme 3.8. Models F and G represent molybdenum centers attached to two Al atoms that were bridged

Scheme 3.8 Mo active site models with different valence states supported on Al2O3 or SiO2. For the model catalysts F, H and J,X¼OH; for the model catalysts G, I, and K,X¼H. Ethylene Polymerization and Oligomerization 165 by two coordinated hydroxyl groups, while H and I were the simplified models of F and G with one H2O molecule eliminated, and two Al atoms were connected by one oxygen atom. Models J and K were built for rep- resenting active sites of the MoOx/SiO2 catalysts. In the models of F, H, and J, the support atoms (Si or Al) were saturated by hydroxyl (OH) groups to stand for fully hydroxylated support surface, while the support atoms in G, I, and K were saturated by hydrogen (H) atoms to represent the dehydroxylated support surface. The oxidation state of the molybdenum was noted in the label of each molecular model. For instance, 5F–5K rep- resent the Mo active sites with the oxidation states of þ5. According to Cossee mechanism (Scheme 3.1)(Cossee, 1964), the metal-C center was proposed as the active site for olefin polymerization. The energy barriers of ethylene insertion into Mo centers with different valence states supported on Al2O3 or SiO2 are listed in Table 3.3. For all the models of the Mo active sites, a general tendency of the increasing of polymerization activity on Mo valence states was found to be as follows: þ þ þ þ Mo4

Table 3.3 Energy barriers (kcal/mol) for ethylene insertion into the ModC bond in different catalyst models Model Insertion Model Insertion Model Insertion Model Insertion 5F 26.1 4F 29.9 3F 27.1 2F 22.7 5G 26.3 4G 28.3 3G 27.0 2G 22.4 5H 25.3 4H 30.5 3H 26.0 2H 22.4 5I 25.7 4I 26.3 3I 26.0 2I 22.3 5J 23.7 4J 27.6 3J 24.8 2J 20.7 5K 23.3 4K 27.9 3K 24.6 2K 21.7 166 Zhen Liu et al.

Scheme 3.9 Ethylene coordination on the Mo2þ active site models (2F0, 2H0, and 2J0) with hydroxyl coordination. Energies in kcal/mol. catalysts, hydroxyl group was usually considered as poison to the active sites in commercial polymerization process over the Phillips chromium catalyst (Augustine and Blitz, 1996). Surface hydroxyl on support was eliminated during catalyst preparation, for example, by fluorination, calcination, to þ obtain highly active catalysts. Because Mo2 models of the active sties showed the highest activity for ethylene polymerization, we further inves- tigated the effects of surface hydroxyl using the divalent models 2F–2K.As shown in Scheme 3.9, three new models of the active site, namely 2F0, 2H0, and 2J0, were obtained with the hydroxyl group coordinated on the Mo cen- ter through the formation of a weak ModO bond. The ethylene insertion energy barriers for 2F0, 2H0, and 2J0 models with hydroxyl coordination were 25.3, 28.4, and 26.9 kcal/mol, respectively. These activation energies are higher than their corresponding models (2F, 2H, and 2J) without hydroxyl coordination. On the basis of the experiments, we built 24 kinds of molybdenum active sites with the Mo oxidation state ranging from þ5toþ2. The effects of sur- face hydroxyl groups were also considered in the theoretical work. DFT results showed that ethylene insertion barrier for SiO2-supported catalysts were slightly lower than that for alumina-supported catalysts. For the same þ kind of support, Mo2 active sites showed the lowest energy barrier of eth- þ ylene insertion, while Mo4 sites presented a larger activation energy. In addition, the coordination of hydroxyl on Mo center could decrease the electron deficiency of molybdenum center leading to an increase in the eth- ylene insertion energy barrier. It had been demonstrated that pre-reduction of hexavalent Mo into lower valence state þ2 and elimination of surface hydroxyl groups during catalyst preparation were the key factors to obtain highly efficient Mo-based catalysts for ethylene polymerization. Through the combination of experiments and theoretical calculations, a better Ethylene Polymerization and Oligomerization 167 understanding on the Mo-based supported polyethylene catalysts has been achieved so far. The highly toxic Cr-based Phillips type catalysts could be possibly substituted by Mo-based catalyst through further successive modi- fications in the future.

5. CHROMIUM CATALYSTS FOR SELECTIVE ETHYLENE OLIGOMERIZATION

The polyethylene products produced by the Phillips chromium cata- lysts account for a large share of the polyolefin consumer market and hence require a large-scale industrial production (McDaniel, 2010). As a kind of highly valuable linear alpha olefin, 1-hexene is greatly demanded as como- nomer for application in the production of HDPE and LLDPE products. The HDPE and LLDPE products produced with 1-hexene as comonomer showed much improved physical properties in diverse applications com- pared with that produced with 1-butene as comonomer. However, 1-butene copolymer had been dominant in the polyethylene market in the past few decades because of the high production cost of 1-hexene. Tra- ditionally, 1-hexene is produced by metal catalyzed ethylene non-selective oligomerization (Shell, BP Amoco, Chevron Phillips, et al.) or by Fischer- Tropsch technology (Sasol, et al.), along with a broad range of linear alpha olefins. A successive distillation processes for separation leads to a relatively high cost of 1-hexene production (Dixon et al., 2004). In 1967, Manyik et al. in UCC first discovered 1-hexene as a major ethylene oligomer by-product along with the polymerization of ethylene catalyzed by Cr(III)(2-EH)3 with hydrolyzed tri-isobutylaluminum as an activator (Manyik et al., 1967). This is the first report on transition-metal catalyzed ethylene trimerization. After about 40 years of intensive research from both industry and academia, the first commercial plant established by Chevron-Phillips finally came on stream for producing comonomer grade 1-hexene by selective ethylene trimerization using Cr(III)(2-EH)3 homoge- neous catalyst system (Dixon et al., 2004). The main improvements of the Phillips ethylene trimerization catalyst system are mainly concerning the aluminum-alkyl cocatalysts, such as triethylaluminum and diethyl aluminum chloride, together with some promoters like 2,5-dimethylpyrrole. In the past decades, the selective olefin oligomerization has been exten- sively reviewed (Agapie, 2011; Dixon et al., 2004; McGuinness, 2011). It is well-known that there are three alternative mechanistic pathways leading to 168 Zhen Liu et al. the production of 1-hexene from ethylene. The first pathway is through a b-H transfer from the metal-alkyl linear propagating chain based on Cossee- type mechanism. This mechanistic pathway is believed to be able to produce a full range of linear a-olefins with even number of carbon atoms including 1-hexene with a Schulz-Flory or Poisson distribution. The second pathway is through ethylene metathesis based on Chauvin-type mechanism. This mechanistic pathway could only make a full range of olefins with both even and odd number of carbon atoms including 1-hexene. The above- mentioned two mechanistic pathways are not applicable for selective ethyl- ene trimerization. The third pathway is through b-H transfer from the metallacyclic propagating chain also based on Cossee-type mechanism. This so-called metallacyclic intermediates mechanism is most popularly accepted nowadays for the transition-metal catalyzed selective ethylene trimerization to give 1-hexene. There are two key aspects for the metallacyclic mecha- nism: (a) during the metallacyclic chain propagation, the insertion step must be faster than the decomposition of the metallacyclopentane to 1-butene; (b) the liberation of 1-hexene must be faster enough than further ethylene inser- tion into the metallacycloheptane to give a larger nine-membered ring. Regarding the above-mentioned metallacyclic mechanism, there are also three different routes proposed based on the specific b-H transfer from the metallacyclic propagating chain. These three pathways are depicted in Scheme 3.10 as path (a), (b), and (c), respectively. Path (a) is proposed by Manyik indicating a direct b-H transfer to ethylene monomer from the metallacyclopentane species. Thus, 1-hexene is generated through a

Scheme 3.10 Metallacyclic mechanisms for selective ethylene trimerization to 1-hexene proposed by Manyik et al. (1977) (path (a)), Briggs (1989) (path (b)), and Hessen (Deckers et al., 2001, 2002) (path (c)). Ethylene Polymerization and Oligomerization 169 subsequent reductive elimination (Manyik et al., 1977). Path (b) is proposed by Briggs supposing a one-step 1-hexene formation through a b-H transfer directly to the other (a–C) in the metallacycloheptane (Briggs, 1989). Path (c) is proposed by Hessen stating a b-H transfer to the chromium center from the metallacycloheptane (Deckers et al., 2001, 2002). Therefore, 1-hexene is released in a two-step manner rather than a direct hydrogen transfer com- paring to path (b). According to the experimental evidence of metallacycle species and C2H4 isotope experimental findings, path (b) is the most plau- sible mechanistic pathway for selective ethylene trimerization to generate 1-hexene. The current understanding on the mechanism of selective ethylene trimerization has been approaching to a relatively clear image although many key points regarding the oxidation states of the chromium center and the role of the ligands are still unclear. As a first reported catalyst for eth- ylene trimerization, Cr(2-EH)3/partially hydrolyzed tri-isobutylaluminum (PIBAO) system is of most interests, which transforms from a polymeriza- tion catalyst into a trimerization catalyst after adding dimethyl ether (DME). The main challenge for designing and developing new catalyst for ethylene trimerization is to make it clear for the triggering mechanisms for those active sites transformation between ethylene polymerization and trimerization. More importantly, the competition between two oxidative coupling cycles Cr(II)/Cr(IV) or Cr(I)/Cr(III) also attracted much attention of many researchers. The mechanism of ethylene trimerization regarding catalyst initiation, effect of activators, chromium oxidation states, and reac- tion mechanism was investigated by experiments and molecular modeling. The contributions from our group to understand the active sites transforma- tion of the Cr(2-EH)3/PIBAO system and the effects of ligand deprotonation of the Cr-SNS (SNS¼RS(CH2)2N(H)(CH2)2SR) system will be shortly discussed and reviewed in the following two subsections.

5.1. Cr-DME-mediated ethylene trimerization

Manyik et al. first patented Cr(2-EH)3/PIBAO system as ethylene poly- merization catalyst with low ethylene trimerization reactivity (Manyik, 1967). Later on, the research from the same group showed that the activity and selectivity could be improved by adding DME to the Cr(2-EH)3/ PIBAO system (Manyik et al., 1977). In 1989, Briggs reported that the acti- vity could be drastically increased to 2086 g/gCr per hour with much improved selectivity of 73% for ethylene trimerization catalyzed by 170 Zhen Liu et al.

Cr(2-EH)3/PIBAO/DME system (Briggs, 1989). In 2003, the Cr(2-EH)3/ 2,5-dimethylpyrrole/Et3Al/Et2AlCl catalyst system has been finally com- mercialized by Phillips Petroleum Company for producing comonomer- grade 1-hexene (Dixon et al., 2004). The most interesting thing for the Cr(2-EH)3/PIBAO catalytic system is the transformation from a polymerization catalyst into a trimerization cata- lyst after adding DME. Our recent work reported a theoretical study on the mechanism of transformation from ethylene polymerization into ethylene trimerization on the Cr(2-EH)3/PIBAO catalyst with or without DME coordination (Qi et al., 2010). Since the industrial homogeneous chromium-based catalyst for selective ethylene trimerization contains three components including the precursor Cr(2-EH)3, the ligand DME, and the cocatalyst PIBAO, we built five molecular models (L–P) representing the possible active species of the Cr(2-EH)3/PIBAO catalyst system without adding DME ligands and another five models (L0–P0) for the active species of the Cr(2-EH)3/PIBAO/DME catalyst system by consideration of the possible oxidation states of the chromium center, as listed in Table 3.4. According to the metallacycle mechanism for selective ethylene trimerization first proposed by Briggs (1989), the catalytic cycle including side reactions for ethylene dimerization and polymerization was depicted in Scheme 3.11. The reaction starts with coordination of two ethylene monomers to give 1X species, and then the key intermediates chro- macyclopentane species 2X was generated by an oxidative-coupling reac- tion. The 1-butene and 1-hexene products could be released in two different ways: (1) one-step path via an agostic-assisted b-hydride shift to give 4X, or (2) two-step path through a Cr-H intermediate followed by reductive elimination. The formation of nine-membered chromacycle

Table 3.4 Molecular models for active chromium species of the Cr(2-EH)3/PIBAO catalytic system (R denotes isobutyl) Oxidation states Model Without DME Model With DME þ þ Cr(I)/(III) L Cr L0 Cr /DME M CrOR M0 CrOR/DME N CrR N0 CrR/DME þ þ Cr(II)/(IV) O Cr OR O0 Cr OR/DME þ þ P Cr R P0 Cr R/DME Ethylene Polymerization and Oligomerization 171

Scheme 3.11 Proposed catalytic cycle for ethylene trimerization over the Cr(2-EH)3/PIBAO catalyst system. (X¼L.Cr(I)þ; L0.Cr(I)þ/DME; M.Cr(I)OR;M0. Cr(I)OR/DME; N.Cr(I)R;N0.Cr(I) R/DME; O.Cr(II)þOR; O0.Cr(II)þOR/DME; P.Cr(II)þR; P0.Cr(II)þR/DME).

Figure 3.17 Gibbs free energy profile of model Cr(I)þ at 298.15 K. The solid line in black shows the metallacycle growth pathway; the solid line in gray shows 1-butene elimina- tion pathway; the dotted line in gray shows b-hydrogen transfer pathway to give chro- mium hydride species 7L; the dotted line in black shows the b-agostic hydrogen shift pathway to give 1-hexene. Energy differences (kcal/mol) are expressed with respect to 1L corrected for the corresponding number of ethylene molecules. Energy barriers are indicated in italics. species 10X represents a ring expansion pathway leading to the production of oligomers and polymers. The Gibbs free energy profile for the reaction pathways by model L,a þ simple monovalent cationic species Cr(I) , is shown in Fig. 3.17. The lib- eration of 1-butene was prohibited by showing two successive activation 172 Zhen Liu et al. barriers of 14.7 and 11.6 kcal/mol. Alternatively, an ethylene coordinated chromacyclopentane species 5L was formed immediately through the exo- ergic coordination of a third ethylene. Therefore, the chromacycloheptane species 6L was generated with a relatively low energy barrier of 15.4 kcal/mol. Similarly, further coordination of a fourth ethylene on þ Cr(I) was also a fast exergonic process by releasing 12.2 kcal/mol. There- þ fore, site L Cr(I) without DME coordination preferred ethylene polymer- ization rather than ethylene trimerization. The ethylene polymerization catalyzed by Cr(2-EH)3/PIBAO would be transformed into selective trimerization by adding DME component reported by Manyik et al. (1977) and Briggs (1989). In order to investigate the role of DME for the transformation mechanism between ethylene poly- merization and trimerization, we further studied the possible reaction using þ the model L0, a cationic model of Cr(I) with DME coordinated on the chromium center. As shown in Fig. 3.18, the liberation of 1-butene can also be safely excluded by shown two high activation barriers on the Gibbs free energy profile. Interestingly, the steric effects of DME ligand prevent the

Figure 3.18 Gibbs free energy profile of model Cr(I)þ/DME at 298.15 K. The solid line in black shows the metallacycle growth pathway; the dotted line in gray shows 1-butene elimination pathway; the solid line in gray shows 1-hexene elimination pathway via two- step route; the dotted line in black shows 1-hexene elimination pathway via b-agostic hydrogen shift. Energy differences (kcal/mol) are expressed with respect to 1L0 corrected for the corresponding number of ethylene molecules. Energy barriers are indi- cated in italics and heat absorption is shown in parentheses. Ethylene Polymerization and Oligomerization 173 further coordination of a fourth ethylene molecule, which results in an end- ergonic process by 8.7 kcal/mol. 1-Hexene liberation would follow a one- step path of agostic-assisted-hydrogen transfer rather than a two-step route. The direct reductive liberation of 1-hexene only requires overcoming an activation barrier of 16.7 kcal/mol, which is lower than the ring expansion þ with a barrier of 19.2 kcal/mol. Therefore, site L0 Cr(I) /DME with DME coordination preferred ethylene trimerization rather than ethylene polymerization. Table 3.5 lists energy differences of ethylene coordination and insertion for metallacycle expansion and 1-hexene liberation via agostic-assisted b-hydrogen transfer (TS[6X–8X]) in each model. The role of DME on the transformation of ethylene polymerization into ethylene trimerization can be understood readily. Before adding DME, the coordination of extra ethylene molecules occurred spontaneously with exoergic effect (cationic models L, O, and P) or with a negligible endoergic effect for both neutral models M and N. As a result, the formation of metallacycle was determined

Table 3.5 Energy differences (kcal/mol) of crucial steps in catalytic cycle of each model Ring5!Ring7 Ring7!Ring9

a b c d e Models Coord Insertion Coord Insertion 1-C6 þ Cationic L Cr 17.2 15.4 12.2 16.4 15.5 þ L0 Cr /DME 1.2 9.0 8.7 10.5 16.7 þ O Cr OR 15.9 12.5 8.3 15.5 10.6 þ O0 Cr OR/DME 20.9 10.2 24.8 11.6 13.4 þ P Cr R 5.5 10.7 0.6 13.8 14.3 þ P0 Cr R/DME 17.2 10.6 20.9 11.8 17.2 Neutral M CrOR 5.1 13.2 5.7 18.4 18.2 M0 CrOR/DME 24.4 5.2 29.1 7.4 29.8 N CrR 4.3 20.0 7.8 22.8 29.8 N0 CrR/DME 29.4 4.1 34.9 5.7 32.8 aHeat absorption required for the coordination of a third ethylene molecule, 2X!5X. bEnergy barrier required for the insertion of a third ethylene to form chromium seven-membered ring, TS[5X–6X]. cHeat absorption required for the coordination of a fourth ethylene molecule, 6X!9X. dEnergy barrier required for the insertion of a fourth ethylene to form chromium nine-membered ring, TS[9X–10X]. eEnergy barrier of 1-hexene liberation via agostic-assisted b-hydrogen transfer, TS[6X–8X]. 174 Zhen Liu et al.

Table 3.6 Apparent energy barriers (kcal/mol) of nine-membered ring formation (P) and 1-hexene liberation via agostic-assisted b-H transfer (T) and the differences between the two barriers for each cationic model Without DME With DME Pa Tb D(TP) Pc Td D(PT) þ þ L Cr(I) 16.4 27.7 11.3 L0 Cr(I) /DME 19.2 16.7 2.5 þ þ O Cr(II) OR 15.5 18.9 3.4 O0 Cr(II) OR/DME 36.4 13.4 23.0 þ þ P Cr(II) R 13.8 14.9 1.1 P0 Cr(II) R/DME 32.7 17.2 15.5 aEnergy barrier of TS[9X–10X]. bEnergy difference between TS[6X–8X] and 9X. cEnergy difference between 6X and TS[9X–10X]. dEnergy barrier of TS[6X–8X]. by the ethylene insertion step. After adding DME, the energy required for ethylene coordination increased enormously by more than 18 kcal/mol in all cases, which became a major determining factor for further metallacycle expansion to large ring size. The neutral sites (M: CrOR and N: CrR) could be safely excluded for Cr(2-EH)3/PIBAO catalyst system because ethylene trimerization was favored without adding DME compound, which dis- agreed with experimental findings. Table 3.6 summarizes the apparent energy barriers of nine-membered ring formation (representing ethylene polymerization) and 1-hexene liber- ation (ethylene trimerization) and the difference between the two barriers for each cationic model before and after addition of DME compound. For the cationic models L, O, and P, the path for ethylene polymerization is open by metallacycle growth to give a nine-membered ring, while DME coordination on these models increased the ethylene coordination energy and the reaction transformed from ethylene polymerization into ethylene trimerization. Therefore, cationic chromium models L, O, and P might be the most plausible active sites for the Cr(2-EH)3/PIBAO catalyst system with or without DME coordination. Thus, neutral models could be safely excluded from the active sites of the Cr(2-EH)3/PIBAO/DME system because ethylene trimerized into 1-hexene without adding DME. Before adding DME, the cationic models tend to produce polyethylene through further ring expansion on the chro- macycloheptane species. However, the polymerization reaction on the cat- ionic models was transformed into ethylene trimerization reaction after adding DME due to the steric and electronic effects of the DME ligands. All the cationic active site models might be the most plausible active sites Ethylene Polymerization and Oligomerization 175

for the Cr(2-EH)3/PIBAO catalyst system with or without DME coordination.

5.2. Cr-SNS-mediated ethylene trimerization

Cr(2-EH)3/2,5-dimethylpyrrole/Et3Al/Et2AlCl system is the first commer- cialized industrial catalyst for ethylene trimerization into 1-hexene (Dixon et al., 2004). Many experimental and theoretical works have been conducted in the past decades in order to elucidate the mechanisms of ethylene trimerization (Agapie, 2011; McGuinness, 2011). Although the metallacycle pathway (Briggs, 1989) of ethylene trimerization has been widely accepted in recent years due to the support from experimental (Agapie et al., 2004, 2007; Emrich et al., 1997) and theoretical reports (Bhaduri et al., 2009; Blok et al., 2003; Blom et al., 2007; Budzelaar, 2009; de Bruin et al., 2003; Qi et al., 2010; Tobisch and Ziegler, 2003, 2004a,b, 2005; van Rensburg et al., 2004; Yu and Houk, 2003), some key problems such as the oxidation state of the active site and effects of ligands are still subject to debate. In 2003, McGuinness et al. reported a remarkable ethylene trimerization catalyst SNS-CrCl3 (SNS¼RS(CH2)2N(H)(CH2)2SR) activated by MAO (McGuinness et al., 2003). The oxidation states of the active chromium spe- cies and the deprotonation of the SNS ligands remains controversial prob- lems for this efficient catalyst system (Agapie, 2011; Jabri et al., 2006; McGuinness, 2011; McGuinness et al., 2006; Temple et al., 2006, 2007). McGuinness et al. suggested the SNS ligands would undergo deprotonation in the early stages during the activation by MAO. However, Gambarotta et al. isolated an inactive deprotonated SNS-Cr(II) complex for ethylene trimerization (Jabri et al., 2006). We recently reported a theoretical work on the SNS-Cr system in order to elucidate the oxidation state of the active chromium species with a con- sideration of the deprotonation of the SNS ligands (Yang et al., 2011). As depicted in Scheme 3.12, four molecular models Q–T were constructed for the typical SNS-Cr catalyst system. Cr(I) and Cr(II) were considered as the two most plausible oxidation states of the chromium center. Models Q and R are two active species without ligand deprotonation, while S and T were designed with consideration of SNS ligand deprotonation. The ethylene trimerization was calculated on the basis of the metallacycle mechanism first proposed by Briggs (1989). The catalytic cycle including the formation of 1-butene (7X), 1-hexene (6X), and metallacycle expansion 176 Zhen Liu et al.

Scheme 3.12 Schematic representation of molecular models Q, R, S, and T for the SNS-Cr system (SNS¼MeS (CH2)2N(H) (CH2)2SMe).

Scheme 3.13 Proposed catalytic cycle (X¼Q, R, S, and T).

(8X) was shown in Scheme 3.13. As previously reported in Ti (Blok et al., 2003; de Bruin et al., 2003, 2008; Tobisch and Ziegler, 2003, 2004a,b, 2005), Ta (Yu and Houk, 2003), and Cr (Bhaduri et al., 2009; Budzelaar, 2009; Klemps et al., 2009; Qi et al., 2010; van Rensburg et al., 2004) ethylene trimerization systems, the formation of 1-butene and 1-hexene from the metallacyclopentane and metallacycloheptane species, respectively, could follow two different routes: (1) one-step route that through a direct intramolecular-hydrogen transfer to the opposite a-carbon atoms; (2) two- step route that via a Cr-H intermediate followed by reductive elimination. Ethylene Polymerization and Oligomerization 177

However, the Cr-H species could not be located because of the steric effect of the SNS ligand and the search of the corresponding transition state for the two-step reaction path to generate the dimerization product 1-butene and the trimerization product 1-hexene failed in all cases. The direct b-hydrogen transfer in the more strain chromacyclopentane ring was failed to occur, which completely shutdown the side reaction to give dimerization product 7X. In the meanwhile, the fourth ethylene was repelled from chromium cen- ter by the steric effects from both the SNS ligands and the large metallacycloheptane ring, thereby stopping the route to go further metal- lacycle expansion. The only remaining path is to go through a direct b-hydrogen transfer in the chromacycloheptane spices 5X to give the trimerization product 6X. The Gibbs free energy profile of the reaction pathway by model Q was shown in Fig. 3.19. The coordination of the first ethylene to the unsaturated chromium center in the sextet ground state is slightly exoergic by 1.0 kcal/mol, which is followed by an endoergic coordination process of

Figure 3.19 Calculated Gibbs free energy profile for ethylene trimerization on model Q: MeS(CH2)2N(H)(CH2)2SMe-Cr(I). Energetic barriers are indicated in italics and heat absorption energies are exhibited underlined. Black lines and gray lines illustrate free energy surfaces under quartet and sextet respectively. Round points represent MECPs, and the corresponding free energies are shown in bold together with their optimized geometries. 178 Zhen Liu et al. the second ethylene. The subsequent oxidative coupling (oxidation from Cr(I) to Cr(III)) of the two coordinated ethylene on the sextet surface was prohibited by showing a very high activation barrier of 38.4 kcal/mol. Interestingly, the presence of an MECP between the sextet and quartet sur- faces facilitates the reaction and the whole barrier for this step is about 24.0 kcal/mol. Therefore, the formation of the metallacyclopentane species undergoes a spin state change from sextet to quartet via an MECP lying 9.7 kcal/mol above the sextet 2Q. The final reduction of Cr(III) to Cr(I), generating 1-hexene, experiences another spin state change from quartet back to sextet with an activation barrier of 18.5 kcal/mol. The whole cat- alytic cycle by model Q was thus completed through two readily accessible MECPs, which lowered the activation energies of the transition states presented in the rate-determining step. In the following work, Gibbs free energy surfaces under all possible spin states of the other three models R–T were completely searched. It was found that spin surface crossing occurred between two higher spin states on each model. On models Q, S, and T, the spin surface crossing took place during the formation of metallacyclopentane and the generation of 1-hexene, while on model R, it occurred during the formation of metallacycloheptane. According to the calculated results, the rate-determining steps are the for- mation of metallacyclopentane and metallacycloheptane species on the four models and the corresponding activation barriers without and with consid- ering the spin-crossover phenomenon are listed in Table 3.7. If spin surface crossing is not considered, assuming all reactions take place on the surface of the highest spin state (the resting state of the reactants), the monovalent model S with ligand deprotonation would be the most possible active site

Table 3.7 Influence of spin surface crossing on the activation barriersa (kcal/mol) of metallacyclopentane (MCP) formation and metallacycloheptane (MCH) formation on each model Without spin surface crossing With spin surface crossing Model MCP MCH MCP MCH Q 44.6 28.7 30.2 30.4 R 41.9 22.5 41.9 22.5 S 37.2 37.9 20.8 40.4 T 46.2 27.1 28.3 32.9 aThe activation barrier includes the binding energy of ethylene coordination and the energetic barrier of metallacycle formation. Ethylene Polymerization and Oligomerization 179 model in the SNS-Cr system. However, as a matter of fact, spin surface crossing is definitely revealed by our DFT calculations for the SNS-Cr sys- tem. By taking into account spin surface crossing, model Q turns out to be the most favorable, which means the Cr(I)/Cr(III) catalytic cycle with a nondeprotonated ligand is the most plausible active site in the SNS-Cr sys- tem. The Cr(I)/Cr(III) cycle is also reported to be responsible for selective ethylene oligomerization in some other Cr-based systems (Albahily et al., 2011a,b; Jabri et al., 2008; Licciulli et al., 2010; Skobelev et al., 2010; Vidyaratne et al., 2009). The nondeprotonation of the SNS ligand by MAO for this catalyst system is also supported by the experimental report on a similar SNS-Cr system by Gambarotta, Duchateau, et al. (Jabri et al., 2006). Cr(I)/Cr(III) catalytic cycle with a nondeprotonated ligand might be the most plausible active site for ethylene trimerization catalyzed by the SNS-Cr system. A plausible spin surface crossing between Cr(I)/sextet and Cr(III)/ quartet was found to play a spin acceleration effect by lowering the activa- tion energy of the rate-determining step in Cr-based ethylene trimerization catalysis. These theoretical results provided much deeper insight into under- standing the highly selective trimerization mechanism and for further devel- opment of new catalysts with high performance as well. The spin crossover phenomenon was also found to be very important in the Cr-catalyzed ethylene selective oligomerization similar to the findings in Cr-catalyzed ethylene polymerization and alkyne cyclotrimerization. In recent years, more and more attention from both academic and industrial circles is starting to be put on the Cr-mediated ethylene tetramerization into 1-octene, a more important comonomer for production of value-added polyolefins compared with 1-hexene (Bollmann et al., 2004; van Leeuwen et al., 2011).

6. SUMMARY AND OUTLOOK

In the past decades, extensive investigations on the Phillips chromium catalyst have been conducted through spectroscopic and kinetic character- izations, model catalysts, and molecular modeling, usually through combi- nation of experiments and theoretical calculations. A step-forward understanding of the nature of active sites and polymerization mechanisms has been achieved with particular interests in the following aspects con- cerning catalyst activation by thermal calcination or reducing agents (like CO, Al-alkyl cocatalysts, and the ethylene monomer itself, etc.), promotion effects of Ti-modification, spin-crossover phenomenon and its effects on the 180 Zhen Liu et al. catalytic reactivity, analysis of the microstructures of polyethylene chain in terms of SCBD, improvements of the mechanical and ultraviolet-resistant properties of CB-reinforced HDPE pipe materials through synthesis of HDPE-g-CB and so on. In spite of the progress achieved so far, the long-standing key question concerning the precise structure of the active sites and the initiation mechanism in terms of the formation of the first CrdC bond on the Phillips chromium catalysts have not been completely elucidated yet. Although the theoretical calculations suggested that the formation of the first CrdC bond was accelerated by the spin-crossover phenomenon, the suggested mechanism still need the support from experiments. Since selective ethylene trimerization was first reported more than 40 years ago, the researches in the field of selective ethylene oligomerization have been conducted tremendously during the last decade. The increased research activity mainly originated from the fast growth of industrial demand for more highly-valuable comonomers, in particular 1-hexene and 1-octene. Although the chromium-based system for the production of 1-hexene was commercialized 10 years ago, a clearer mechanistic under- standing and further cost-effective improvements of the catalyst system is highly demanded. Meanwhile, the industrial production of 1-octene and mechanistic understanding on the Cr-ligand-mediated ethylene tetramerization is still in its infancy. The understanding on the structure and oxidation states of active chromium species, the role of organic ligands and the corresponding reaction mechanisms for various chromium-based catalysts is still facing many uncertainties. A clear elucidation of these basic questions requires the combination of multiple techniques especially those in situ/operando methods as well as theoretical molecular modeling. The continuous development of highly efficient new catalysts is driven by the desire from both industry and academia, including the environmental- friendly nonchromium-like molybdenum catalyst for ethylene polymeriza- tion and novel catalyst systems for selective ethylene oligomerization, in particular ethylene trimerization and tetramerization. The spin cross- over phenomenon, which was found to be generally existing in the Cr-catalyzed ethylene polymerization, acetylene cyclotrimerization, and ethylene trimerization should be further explored in terms of its intrinsic relationship with the catalytic reactivity through both experimental and the- oretical investigations. A state-of-the-art catalyst design with greatly improved efficiency based on experimental and computational high- throughput screening techniques would be highly expected in the Ethylene Polymerization and Oligomerization 181 polyolefin field in the near future. Bridging the two most important struc- ture–property relationships between catalysts and polyolefins will become more and more indispensable for product innovations in the development of novel catalysts and polymers with high performance as well as for academic approaches in the understanding of the nature of active sites and polymer- ization/oligomerization mechanisms, which calls for persistent efforts and tight collaboration of scientists with different expertise from all over the world.

ACKNOWLEDGMENTS We gratefully thank the financial supports by the National Natural Science Foundation of China (No. 21004020, 21104019, 21174037, 21274040, 21304033 and 51003027), the National High Technology Research and Development Program 863 (2012AA040306), and the Shanghai Science and Technology Commission (Key Project for Basic Research 10JC1403700). This work is also financially supported by the Fundamental Research Funds for the Central Universities, research program of Introducing Talents of Discipline of university (B08021).

REFERENCES Abbenhuis HCL, Vorstenbosch MLW, van Santen RA, Smeets WJJ, Spek AL: A 12-membered inorganic heterocycle: synthesis and structural characterization of a bimetallic chromium(VI) siloxane complex, Inorg Chem 36(27):6431–6433, 1997. Agapie T: Selective ethylene oligomerization: recent advances in chromium catalysis and mechanistic investigations, Coord Chem Rev 255(7–8):861–880, 2011. Agapie T, Schofer SJ, Labinger JA, Bercaw JE: Mechanistic studies of the ethylene trimerization reaction with chromium-diphosphine catalysts: experimental evidence for a mechanism involving metallacyclic intermediates, J Am Chem Soc 126(5):1304–1305, 2004. Agapie T, Labinger JA, Bercaw JE: Mechanistic studies of olefin and alkyne trimerization with chromium catalysts: deuterium labeling and studies of regiochemistry using a model chromacyclopentane complex, J Am Chem Soc 129(46):14281–14295, 2007. Agenet N, Gandon V, Vollhardt KPC, Malacria M, Aubert C: Cobalt-catalyzed cyclo- trimerization of alkynes: the answer to the puzzle of parallel reaction pathways, JAm Chem Soc 129(28):8860–8871, 2007. Agostini G, Groppo E, Bordiga S, et al: Reactivity of Cr species grafted on SiO2/Si(100) surface: a reflection extended X-ray absorption fine structure study down to the sub- monolayer regime, J Phys Chem C 111(44):16437–16444, 2007. Ajjou JAN, Scott SL: Reactions of tetraalkylchromium(IV) with silica: mechanism of grafting and characterization of surface organometallic complexes, Organometallics 16(1):86–92, 1997. Ajjou JAN, Scott SL: A kinetic study of ethylene and 1-hexene homo- and copolymerization catalyzed by a silica-supported Cr(IV) complex: evidence for propagation by a migratory insertion mechanism, J Am Chem Soc 122(37):8968–8976, 2000. Ajjou JAN, Scott SL, Paquet V: Synthesis and characterization of silica-stabilized chromium(IV) alkylidene complexes, JAmChemSoc120(2):415–416, 1998a. 182 Zhen Liu et al.

Ajjou JAN, Rice GL, Scott SL: Kinetics and mechanisms of thermally induced alkane elim- inations from silica-supported bis(alkyl)chromium(IV) and -vanadium(IV) complexes, J Am Chem Soc 120(51):13436–13443, 1998b. Albahily K, Fomitcheva V, Gambarotta S, et al: Preparation and characterization of a reduced chromium complex via vinyl oxidative coupling: formation of a self-activating catalyst for selective ethylene trimerization, J Am Chem Soc 133(16):6380–6387, 2011a. Albahily K, Shaikh Y, Sebastiao E, et al: Vinyl oxidative coupling as a synthetic route to cat- alytically active monovalent chromium, J Am Chem Soc 133(16):6388–6395, 2011b. Aubriet F, Muller JF, Poleunis C, et al: Activation processes and polyethylene formation on a Phillips model catalyst studied by laser ablation, laser desorption, and static secondary ion mass spectrometry, J Am Soc Mass Spectrom 17(3):406–414, 2006. Augustine SM, Blitz JP: A proposed mechanism for silica supported chromium HDPE cat- alyst activation, J Catal 161(2):641–650, 1996. Baker LM, Carrick WL: Oxidation of olefins by supported chromium oxide, J Org Chem 33(2):616–618, 1968. Baker LM, Carrick WL: Bistriphenylsilyl chromate. Oxidation of olefins and use in ethylene polymerization, J Org Chem 35(3):774–776, 1970. Barzan C, Groppo E, Quadrelli EA, Monteil V, Bordiga S: Ethylene polymerization on a SiH4-modified Phillips catalyst: detection of in situ produced alpha-olefins by operando FTIR spectroscopy, Phys Chem Chem Phys 14(7):2239–2245, 2012. Bensalem A, Weckhuysen BM, Schoonheydt RA: Nature of adsorbed species during the reduction of CrO3/SiO2 with CO. In situ FTIR spectroscopic study, J Chem Soc Faraday Trans 93(22):4065–4069, 1997. Best SA, Squires RG, Walton RA: The X-ray photoelectron spectra of heterogeneous cat- alysts. II. The chromia-silica catalyst system, J Catal 47(3):292–299, 1977. Bhaduri S, Mukhopadhyay S, Kulkarni SA: Density functional studies on chromium cata- lyzed ethylene trimerization, J Organomet Chem 694(9–10):1297–1307, 2009. Blok ANJ, Budzelaar PHM, Gal AW: Mechanism of ethene trimerization at an ansa-(arene) (cyclopentadienyl) titanium fragment, Organometallics 22(13):2564–2570, 2003. Blom B, Klatt G, Fletcher JCQ, Moss JR: Computational investigation of ethene trimerization catalyzed by cyclopentadienyl chromium complexes, Inorg Chim Acta 360(9):2890–2896, 2007. Bollmann A, Blann K, Dixon JT, et al: Ethylene tetramerization: a new route to produce 1-octene in exceptionally high selectivities, J Am Chem Soc 126(45):14712–14713, 2004. Bordiga S, Groppo E, Agostini G, van Bokhoven JA, Lamberti C: Reactivity of surface spe- cies in heterogeneous catalysts probed by in situ X-ray absorption techniques, Chem Rev 113(3):1736–1850, 2013. Briggs JR: The selective trimerization of ethylene to hex-l-ene, J Chem Soc Chem Commun 11:674–675, 1989. Budzelaar PHM: Ethene trimerization at CrI/CrIII—a density functional theory (DFT) study, Can J Chem 87(7):832–837, 2009. Cao X, Cheng R, Liu Z, et al: DFT and PIO study of the influences of Mo valance state and surface hydroxyl on supported-MoOx catalysts for ethylene polymerization, J Mol Catal A Chem 321(1–2):50–60, 2010. Carrick WL, Turbett RJ, Karol FJ, et al: Ethylene polymerization with supported bis(triphenylsilyl) chromate catalysts, J Polym Sci A Polym Chem 10(9):2609–2620, 1972. Cheng R, Xu C, Liu Z, et al: High-resolution spectroscopy (XPS, H1 MAS solid-state NMR) and DFT investigations into Ti-modified Phillips CrOx/SiO2 catalysts, J Catal 273(2):103–115, 2010. Cheng R, Liu Z, Zhong L, et al: Phillips Cr/silica catalyst for ethylene polymerization, Adv Polym Sci, 2013http://dx.doi.org/10.1007/12_2013_222. Ethylene Polymerization and Oligomerization 183

Choi KY, Tang S: Polymerization rate modeling of ethylene polymerization with supported chromium oxide catalysts, J Appl Polym Sci 91(5):2923–2927, 2004. Choi KY, Tang S, Yoon WJ: Modeling of ethylene polymerization kinetics over supported chromium oxide catalysts, Macromol Theory Simul 13(2):169–177, 2004. Cimino A, de Angelis BA, Luchetti A, Minelli G: The characterization of chromia (CrOx)/ silica catalysts by photoelectron spectroscopy (XPS), X-ray and optical measurements, J Catal 45(3):316–325, 1976. Cimino A, Cordischi D, de Rossi S, et al: Studies on chromia/zirconia catalysts. II. ESR of chromium species, J Catal 127(2):761–776, 1991a. Cimino A, Cordischi D, de Rossi S, et al: Studies on chromia/zirconia catalysts. I. Prepara- tion and characterization of the system, J Catal 127(2):744–760, 1991b. Clark A, Hogan JP, Witt DR, Lanning WC: The polymerization of acetylene and homologs, Proc World Pet Congr 4:267–273, 1959. Cossee P: Ziegler-Natta catalysis. I. Mechanism of polymerization of alpha-olefins with Ziegler-Natta catalysts, J Catal 3(1):80–88, 1964. Damin A, Bonino F, Bordiga S, et al: Vibrational properties of Cr-II Centers on reduced Phillips catalysts highlighted by resonant Raman spectroscopy, ChemPhysChem 7(2):342–344, 2006. Damin A, Vitillo JG, Ricchiardi G, et al: Modeling CO and N2 adsorption at Cr surface spe- cies of Phillips catalyst by hybrid density functionals: effect of Hartree-Fock exchange percentage, J Phys Chem A 113(52):14261–14269, 2009. de Bruin TJM, Magna L, Raybaud P, Toulhoat H: Hemilabile ligand induced selectivity: a DFT study on ethylene trimerization catalyzed by titanium complexes, Organometallics 22(17):3404–3413, 2003. de Bruin TJM, Raybaud P, Toulhoat H: A DFT chemical descriptor to predict the selectivity in a-olefins in the catalytic metallacyclic oligomerization reaction of ethylene according to the (hemi)labile ligand coordinating to titanium, Organometallics 27(19):4864–4872, 2008. Deckers PJW, Hessen B, Teuben JH: Switching a catalyst system from ethene polymerization to ethene trimerization with a hemilabile ancillary ligand, Angew Chem Int Ed 40(13):2516–2519, 2001. Deckers PJW, Hessen B, Teuben JH: Catalytic trimerization of ethene with highly active cyclopentadienyl-arene titanium catalysts, Organometallics 21(23):5122–5135, 2002. Demmelmaier CA, White RE, van Bokhoven JA, Scott SL: Nature of (SiO)CrO2Cl and (¼SiO)2CrO2 sites prepared by grafting CrO2Cl2 onto silica, J Phys Chem C 112(16):6439–6449, 2008. Demmelmaier CA, White RE, van Bokhoven JA, Scott SL: Evidence for a chromasiloxane ring size effect in Phillips (Cr/SiO2) polymerization catalysts, J Catal 262(1):44–56, 2009. Deslauriers PJ, McDaniel MP: Short chain branching profiles in polyethylene from the Phil- lips Cr/silica catalyst, J Polym Sci A Polym Chem 45(15):3135–3149, 2007. Deslauriers PJ, McDaniel MP, Rohlfing DC, et al: A comparative study of multimodal vs. bimodal polyethylene pipe resins for PE-100 applications, Polym Eng Sci 45(9):1203–1213, 2005. Dines TJ, Inglis S: Raman spectroscopic study of supported chromium(VI) oxide catalysts, Phys Chem Chem Phys 5(6):1320–1328, 2003. Dixon JT, Green MJ, Hess FM, Morgan DH: Advances in selective ethylene trimerization— a critical overview, J Organomet Chem 689(23):3641–3668, 2004. Ellison A, Overton TL: Characterization of Cr/silica catalysts. Part 2. Ti- and Mg-modified catalysts, J Chem Soc Faraday Trans 89(24):4393–4395, 1993. Ellison A, Diakun G, Worthington P: EXAFS-XANES studies of rhenium oxide-alumina and chromium oxide-silica catalysts, J Mol Catal 46(1–3):131–149, 1988. 184 Zhen Liu et al.

Ellison A, Overton TL, Bencze L: Characterization of chromium-silica polymerization cat- alysts, J Chem Soc Faraday Trans 89(5):843–849, 1993. Emrich R, Heinemann O, Jolly PW, Krueger C, Verhovnik GPJ: The role of metallacycles in the Cr-catalyzed trimerization of ethylene, Organometallics 16(8):1511–1513, 1997. Espelid , Børve KJ: Theoretical models of ethylene polymerization over a mononuclear chromium(II)/silica site, J Catal 195(1):125–139, 2000. Espelid , Børve KJ: Theoretical analysis of d-d transitions for the reduced Cr/silica system, Catal Lett 75(1–2):49–54, 2001. Espelid , Børve KJ: Theoretical analysis of CO adsorption on the reduced Cr/silica system, J Catal 205(1):177–190, 2002a. Espelid , Bøtrve KJ: Molecular-level insight into Cr/silica Phillips-type catalysts: polymerization-active dinuclear chromium sites, J Catal 206(2):331–338, 2002b. Espelid , Bøtrve KJ: Molecular-level insight into Cr/silica Phillips-type catalysts: polymerization-active mononuclear chromium sites, J Catal 205(2):366–374, 2002c. Failla MD, Lucas JC, Mandelkern L: Supermolecular structure of random copolymers of eth- ylene, Macromolecules 27(6):1334–1337, 1994. Fan Y, Xue Y, Nie W, Ji X, Bo S: Characterization of the microstructure of bimodal HDPE resin, Polym J 41(8):622–628, 2009. Fang Y, Liu B, Terano M: Photo-stability of surface chromate species on Phillips CrOx/SiO2 catalysts isothermally calcined at various temperatures, probed by high resolution X-ray photoelectron spectroscopy, Appl Catal A Gen 279(1–2):131–138, 2005a. Fang Y, Liu B, Hasebe K, Terano M: Ethylene and 1-hexene copolymerization with CO-prereduced Phillips CrOx/SiO2 catalyst in the presence of Al-alkyl cocatalyst, J Polym Sci A Polym Chem 43(19):4632–4641, 2005b. Fang Y, Liu B, Terano M: Various activation procedures of Phillips catalyst for ethylene poly- merization, Kinet Catal 47(2):295–302, 2006. Feher FJ, Blanski RL: Polyhedral oligometallasilasesquioxanes as models for silica-supported catalysts: chromium attached to two vicinal siloxy groups, J Chem Soc Chem Commun 22:1614–1616, 1990. Feher FJ, Blanski RL: Polyhedral oligometallosilsesquioxanes as models for silica-supported catalysts, Macromol Symp 66(1):95–107, 1993. Field E, Feller M: Promoted molybdena-alumina catalysts in ethylene polymerization, Ind Eng Chem 49(11):1883–1884, 1957. Galli P, Vecellio G: Polyolefins: the most promising large-volume materials for the 21st cen- tury, J Polym Sci A Polym Chem 42(3):396–415, 2004. Gandon V, Agenet N, Vollhardt KPC, Malacria M, Aubert C: Cobalt-mediated cyclic and linear 2:1 cooligomerization of alkynes with alkenes: a DFT study, J Am Chem Soc 128(26):8509–8520, 2006. Garcia RA, Carrero A, Aroca M, Prieto O, Dominguez C: Slow crack growth resistance in resin blends of chromium and metallocene catalyzed ethylene-hexene copolymers for pipe applications, Polym Eng Sci 48(5):925–933, 2008. Ghiotti G, Garrone E, Zecchina A: An infrared study of carbon monoxide/ethylene coadsorption and reaction on silica-supported chromium(II) ions, J Mol Catal 65(1–2):73–83, 1991. Ghiotti G, Garrone E, Coluccia S, Morterra C, Zecchina A: Evidence for alkylidenic con- figuration of polymethylene chains on a Phillips catalyst, J Chem Soc Chem Commun 22:1032–1033, 1979. Ghiotti G, Garrone E, Zecchina A: IR investigation of polymerization centers of the Phillips catalyst, J Mol Catal 46(1–3):61–77, 1988. Groeneveld C, Wittgen PPMM, van Kersbergen AM, et al: Hydrogenation of olefins and polymerization of ethene over chromium oxide/silica catalysts. I. Preparation and struc- ture of the catalyst, J Catal 59(2):153–167, 1979. Ethylene Polymerization and Oligomerization 185

Groppo E, Lamberti C, Bordiga S, Spoto G, Zecchina A: The structure of active centers and the ethylene polymerization mechanism on the Cr/SiO2 catalyst: a frontier for the char- acterization methods, Chem Rev 105(1):115–184, 2005a. Groppo E, Lamberti C, Bordiga S, et al: FTIR investigation of the H2,N2, and C2H4 molec- ular complexes formed on the Cr(II) sites in the Phillips catalyst: a preliminary step in the understanding of a complex system, J Phys Chem B 109(31):15024–15031, 2005b. Groppo E, Lamberti C, Spoto G, et al: Tuning the structure, distribution and reactivity of polymerization centres of Cr(II)/SiO2 Phillips catalyst by controlled annealing, J Catal 236(2):233–244, 2005c. Groppo E, Damin A, Bonino F, et al: New strategies in the Raman study of the Cr/SiO2 Phillips catalyst: observation of molecular adducts on Cr(II) sites, Chem Mater 17(8):2019–2027, 2005d. Groppo E, Prestipino C, Cesano F, et al: In situ, Cr K-edge XAS study on the Phillips cat- alyst: activation and ethylene polymerization, J Catal 230(1):98–108, 2005e. Groppo E, Lamberti C, Bordiga S, Spoto G, Zecchina A: In situ FTIR spectroscopy of key intermediates in the first stages of ethylene polymerization on the Cr/SiO2 Phillips catalyst: solving the puzzle of the initiation mechanism? J Catal 240(2): 172–181, 2006a. Groppo E, Lamberti C, Cesano F, Zecchina A: On the fraction of Cr-II sites involved in the C2H4 polymerization on the Cr/SiO2 Phillips catalyst: a quantification by FTIR spec- troscopy, Phys Chem Chem Phys 8(21):2453–2456, 2006b. Groppo E, Estephane J, Lamberti C, Spoto G, Zecchina A: Ethylene, propylene and ethylene oxide in situ polymerization on the Cr(II)/SiO2 system: a temperature- and pressure- dependent investigation, Catal Today 126(1–2):228–234, 2007. Groppo E, Damin A, Arean CO, Zecchina A: Enhancing the initial rate of polymerisation of the reduced phillips catalyst by one order of magnitude, Chem Eur J 17(40): 11110–11114, 2011. Groppo E, Seenivasan K, Barzan C: The potential of spectroscopic methods applied to het- erogeneous catalysts for olefin polymerization, Catal Sci Technol 3(4):858–878, 2013. Guesmi H, Tielens F: Chromium oxide species supported on silica: a representative periodic DFT model, J Phys Chem C 116(1):994–1001, 2012. Handzlik J, Ogonowski J: Theoretical study on ethene metathesis proceeding on MoVI and MoIV methylidene centres of heterogeneous molybdena-alumina catalyst, J Mol Catal A Chem 175(1–2):215–225, 2001. Hardcastle FD, Wachs IE: Raman spectroscopy of chromium oxide supported on alumina, titania, and silica: a comparative study, J Mol Catal 46(1–3):173–186, 1988. He X, Cai Y, Wang Q, et al: Improvement of mechanical properties and ultraviolet resistance of polyethylene pipe materials using high density polyethylene matrix grafted carbon black, J Macromol Sci Phys 51(2):298–312, 2012. Hoang EM, Allen NS, Liauw CM, Fontan E, Lafuente P: The thermo-oxidative degradation of metallocene polyethylenes: part 2: thermal oxidation in the melt state, Polym Degrad Stab 91(6):1363–1372, 2006. Hogan JP, Banks RL: Polymers and production thereof, US patent 2825721, 1958. Hubert L, David L, Seguela R, et al: Physical and mechanical properties of polyethylene for pipes in relation to molecular architecture. I. Microstructure and crystallization kinetics, Polymer 42(20):8425–8434, 2001. Hubert L, David L, Seguela R, et al: Physical and mechanical properties of polyethylene for pipes in relation to molecular architecture. II. Short-term creep of isotropic and drawn materials, J Appl Polym Sci 84(12):2308–2317, 2002. Ikeda H, Monoi T, Sasaki Y: Performance of the Cr[CH(SiMe3)2]3/SiO2 catalyst for ethyl- ene polymerization compared with the performance of the Phillips catalyst, J Polym Sci A Polym Chem 41(3):413–419, 2003. 186 Zhen Liu et al.

Ivin KJ, Rooney JJ, Stewart CD, Green MLH, Mahtab R: Mechanism for the stereospecific polymerization of olefins by Ziegler-Natta catalysts, J Chem Soc Chem Commun 14:604–606, 1978. Jabri A, Temple C, Crewdson P, et al: Role of the metal oxidation state in the SNS-Cr cat- alyst for ethylene trimerization: isolation of di- and trivalent cationic intermediates, JAm Chem Soc 128(28):9238–9247, 2006. Jabri A, Mason CB, Gambarotta S, Burchell TJ, Duchateau R: Isolation of single-component trimerization and polymerization chromium catalysts: the role of the metal oxidation state, Angew Chem Int Ed 47(50):9717–9721, 2008. Jordens K, Wilkes GL, Janzen J, Rohlfing DC, Welch MB: The influence of molecular weight and thermal history on the thermal, rheological, and mechanical properties of metallocene-catalyzed linear polyethylenes, Polymer 41(19):7175–7192, 2000. Jozwiak WK, Dalla Lana IG: Interactions between the chromium oxide phase and support surface; redispersion of alpha-chromia on silica, alumina and magnesia, J Chem Soc Far- aday Trans 93(15):2583–2589, 1997. Kennedy MA, Peacock AJ, Mandelkern L: Tensile properties of crystalline polymers: linear polyethylene, Macromolecules 27(19):5297–5310, 1994. Kissin YV, Brandolini AJ, Garlick JL: Kinetics of ethylene polymerization reactions with chromium oxide catalysts, J Polym Sci A Polym Chem 46(16):5315–5329, 2008. Klemps C, Payet E, Magna L, et al: PCNCP ligands in the chromium-catalyzed oligomer- ization of ethylene: tri- versus tetramerization, Chem Eur J 15(33):8259–8268, 2009. Kozuch S: A refinement of everyday thinking: the energetic span model for kinetic assess- ment of catalytic cycles, WIREs Comput Mol Sci 2(5):795–815, 2012. Kozuch S, Shaik S: A combined kinetic-quantum mechanical model for assessment of cat- alytic cycles: application to cross-coupling and heck reactions, J Am Chem Soc 128(10):3355–3365, 2006. Kozuch S, Shaik S: Kinetic-quantum chemical model for catalytic cycles: the Haber-Bosch process and the effect of reagent concentration, J Phys Chem A 112(26):6032–6041, 2008. Kozuch S, Shaik S: How to conceptualize catalytic cycles? The energetic span model, Acc Chem Res 44(2):101–110, 2011. Krishnaswamy RK, Yang Q, Fernandez-Ballester L, Kornfield JA: Effect of the distribution of short-chain branches on crystallization kinetics and mechanical properties of high- density polyethylene, Macromolecules 41(5):1693–1704, 2008. Lamberti C, Zecchina A, Groppo E, Bordiga S: Probing the surfaces of heterogeneous cat- alysts by in situ IR spectroscopy, Chem Soc Rev 39(12):4951–5001, 2010. Li L, Wu Y, Dong Q, et al: Effect of Al-alkyl cocatalyst on ethylene polymerization behavior using silyl chromate (SC)/SiO2 catalyst, Asia Pac J Chem Eng 8(4):539–546, 2013. Licciulli S, Thapa I, Albahily K, et al: Towards selective ethylene tetramerization, Angew Chem Int Ed 49(48):9225–9228, 2010. Liu B, Terano M: Investigation of the physico-chemical state and aggregation mechanism of surface Cr species on a Phillips CrOx/SiO2 catalyst by XPS and EPMA, J Mol Catal A Chem 172(1–2):227–240, 2001. Liu B, Nakatani H, Terano M: New aspects of the induction period of ethene polymerization using Phillips CrOx/SiO2 catalyst probed by XPS, TPD and EPMA, J Mol Catal A Chem 184(1–2):387–398, 2002. Liu B, Nakatani H, Terano M: Mechanistic implications of the unprecedented transforma- tions of ethene into propene and butene over Phillips CrOx/SiO2 catalyst during induc- tion period, J Mol Catal A Chem 201(1–2):189–197, 2003. Liu B, Fang Y, Terano M: High resolution X-ray photoelectron spectroscopic analysis of transformation of surface chromium species on Phillips CrOx/SiO2 catalysts isothermally calcined at various temperatures, J Mol Catal A Chem 219(1):165–173, 2004a. Ethylene Polymerization and Oligomerization 187

Liu B, Fang Y, Nakatani H, Terano M: Surface physico-chemical state of CO-prereduced Phillips CrOx/SiO2 catalyst and unique polymerization behavior in the presence of Al-alkyl cocatalyst, Macromol Symp 213(1):37–46, 2004b. Liu B, Fang Y, Terano M: Theoretical study of the monomer reaction mechanism on Phillips CrOx/SiO2 catalyst using density functional theory (DFT) and paired interacting orbitals (PIO) methods, Mol Simul 30(13–15):963–971, 2004c. Liu B, Sindelar P, Fang Y, Hasebe K, Terano M: Correlation of oxidation states of surface chromium species with ethylene polymerization activity for Phillips CrOx/SiO2 catalysts modified by Al-alkyl cocatalyst, J Mol Catal A Chem 238(1–2):142–150, 2005. Liu Z, Cheng R, He X, Wu X, Liu B: DFT functional benchmarking on the energy splitting of chromium spin states and mechanistic study of acetylene cyclotrimerization over the Phillips Cr(II)/silica catalyst, J Phys Chem A 116(28):7538–7549, 2012. Liu Z, Cheng R, He X, Liu B: Reactivity and regioselectivity of methylacetylene cyclo- trimerization over the Phillips Cr/silica catalyst: a DFT study, ACS Catal 3(6):1172–1183, 2013. Luruli N, Pijpers T, Brull R, et al: Fractionation of ethylene/1-pentene copolymers using a combination of SEC-FTIR and SEC-HPer DSC, J Polym Sci Polym Phys 45(21):2956–2965, 2007. Manyik RM, Albans St., Walker WE, Wilson TP, Charleston, WV: Continuous processes for the production of ethylene polymers and catalysts suitable therefor, US patent 3300458, 1967. Manyik RM, Walker WE, Wilson TP: Soluble chromium-based catalyst for ethylene trimerization and polymerization, J Catal 47(2):197–209, 1977. Martinez M, Michelini MDC, Rivalta I, Russo N, Sicilia E: Acetylene cyclotrimerization by early second-row transition metals in the gas phase. A theoretical study, Inorg Chem 44(26):9807–9816, 2005. Matos I,ZhangY, LemosM,etal: Kinetic modelingstudies ofethylene polymerization reactions using supported chromium catalysts, J Polym Sci A Polym Chem 42(14):3464–3472, 2004. McDaniel MP: The state of chromium(VI) on the chromium/silica polymerization catalyst, J Catal 67(1):71–76, 1981. McDaniel MP: The state of chromium(VI) on the Phillips polymerization catalyst. II. The reaction between silica and chromyl chloride, J Catal 76(1):17–28, 1982. McDaniel MP: Supported chromium catalysts for ethylene polymerization, Adv Catal 33:47–98, 1985. McDaniel MP: Polymerization reactions. In Ertl G, Knozinger H, Schu¨th F, Weitkamp Jedi- tors: Handbook of heterogeneous catalysis, Weinheim, 2008, Wiley-VCH. McDaniel MP: A review of the Phillips supported chromium catalyst and its commercial use for ethylene polymerization, Adv Catal 53:123–606, 2010. McDaniel MP: Some reflections on the current state of Cr-based polymerization catalysts, MRS Bull 38(3):234–238, 2013. McDaniel MP, Cantor DM: Hydrogen transfer during propagation of the Phillips catalyst, J Polym Sci Polym Chem Ed 21(4):1217–1221, 1983. McDaniel MP, Johnson MM: A comparison of chromium/silica and chromium/aluminum phosphate polymerization catalysts. I. Kinetics, J Catal 101(2):446–457, 1986. McDaniel MP, Johnson MM: A comparison of Cr/SiO2 and Cr/AlPO4 polymerization cat- alysts. 2. Chain transfer, Macromolecules 20(4):773–778, 1987. McGuinness DS: Olefin oligomerization via metallacycles: dimerization, trimerization, tetramerization, and beyond, Chem Rev 111(3):2321–2341, 2011. McGuinness DS, Wasserscheid P, Keim W, et al: First Cr(III)-SNS complexes and their use as highly efficient catalysts for the trimerization of ethylene to 1-hexene, J Am Chem Soc 125(18):5272–5273, 2003. 188 Zhen Liu et al.

McGuinness DS, Brown DB, Tooze RP, et al: Ethylene trimerization with Cr-PNP and Cr-SNS complexes: effect of ligand structure, metal oxidation state, and role of activator on catalysis, Organometallics 25(15):3605–3610, 2006. Merryfield R, McDaniel MP, Parks G: An XPS study of the Philips chromium/silica poly- merization catalyst, J Catal 77(2):348–359, 1982. Moisii C, Deguns EW, Lita A, et al: Coordination environment and vibrational spectroscopy of Cr(VI) sites supported on amorphous silica, Chem Mater 18(17):3965–3975, 2006. Monoi T, Ikeda H, Sasaki Y, Matsumoto Y: Ethylene polymerization with silica-supported Cr[CH(SiMe3)2]3 catalyst. Effect of silica calcination temperature and Cr content, Polym J 35(7):608–611, 2003. Mortazavi SMM, Arabi H, Zohuri G, et al: Copolymerization of ethylene/a-olefins using bis(2-phenylindenyl)zirconium dichloride metallocene catalyst: structural study of comonomer distribution, Polym Int 59(9):1258–1265, 2010. Nishimura M, Thomas JM: IR study of the interaction of hydroxyl groups of silica gel with chromium species of Phillips-type catalysts, Catal Lett 21(1–2):149–155, 1993. O’Neill PP, Rooney JJ: Direct transformation of ethylene to propylene on an olefin metath- esis catalyst, J Am Chem Soc 94(12):4383–4384, 1972. Okamoto Y, Fujii M, Imanaka T, Teranishi S: X-ray photoelectron spectroscopic studies of catalysts. Chromia-alumina catalysts, Bull Chem Soc Jpn 49(4):859–863, 1976. Paiva MC, Zhou B, Fernando KAS, et al: Mechanical and morphological characterization of polymer-carbon nanocomposites from functionalized carbon nanotubes, Carbon 42(14):2849–2854, 2004. Panchenko VN, Zakharov VA, Paukshtis EA: DRIFTS and DRS studies of Phillips ethylene polymerization catalysts, Appl Catal A Gen 313(2):130–136, 2006. Pierce KG, Barteau MA: Cyclotrimerization of alkynes on reduced TiO2(001) surfaces, J Phys Chem 98(14):3882–3892, 1994. Qi Y, Dong Q, Zhong L, et al: Role of 1,2-dimethoxyethane in the transformation from ethylene polymerization to trimerization using chromium tris(2-ethylhexanoate)-based catalyst system: a DFT study, Organometallics 29(7):1588–1602, 2010. Qiu P, Li X, Zhang S, et al: Supporting mechanism of non-toxic chromium(III) acetate on silica for preparation of Phillips ethylene polymerization catalysts, Asia Pac J Chem Eng 4(5):660–665, 2009. Qiu P, Cheng R, Liu B, et al: A triphenylsiloxy complex of chromium(II) as a switchable catalyst for ethylene polymerization and nonselective oligomerization, Organometallics 30(8):2144–2148, 2011. Qiu P, Cheng R, Liu Z, et al: Switching from ethylene polymerization to nonselective olig- omerization over a homogeneous model catalyst: a triphenylsiloxy complex of chromium(VI), J Organomet Chem 699:48–55, 2012. Rahman A, Mohamed MH, Ahmed M, Aitani AM: Characterization of chromia/alumina catalysts by X-ray photoelectron spectroscopy, proton induced X-ray emission and ther- mogravimetric analysis, Appl Catal A Gen 121(2):203–216, 1995. Razavi-Nouri M: Studies of comonomer distributions and molecular segregations in metallocene-prepared polyethylenes by DSC, Polym Test 25(8):1052–1058, 2006. Rebenstorf B, Larsson R: Why do homogeneous analogs of Phillips (chromium(VI) oxide/ silica) and Union Carbide (chromocene/silica) polyethylene catalysts fail? Some answers from IR investigations, J Mol Catal I1(2–3):247–256, 1981. Richter M, Reich P, Oehlmann G: Evidence for crystalline chromium trioxide on chro- mium/silicon dioxide impregnation catalysts by laser Raman spectroscopy, J Mol Catal 46(1–3):79–86, 1988. Sautet P, Delbecq F: Catalysis and surface organometallic chemistry: a view from theory and simulations, Chem Rev 110(3):1788–1806, 2010. Ethylene Polymerization and Oligomerization 189

Schmidt H, Riederer W, Krauss HL: On surface compounds of transition metals. Part 40. Characterization of chromium-doped silica catalyst systems of the Phillips type by scan- ning electron microscopy with energy dispersive analyzing system (SEM/EDS), J Prakt Chem 338(7):627–633, 1996. Scott SL, Ajjou JAN: Surface organometallic investigation of the mechanism of ethylene polymerization by silica-supported Cr catalysts, Chem Eng Sci 56(13):4155–4168, 2001. Scott SL, Fu A, MacAdams LA: Effect of silica type and grafting method on the reactivity of tetraneopentylchromium(IV) towards and on silica, Inorg Chim Acta 361(11):3315–3321, 2008. Skobelev IY, Panchenko VN, Lyakin OY, et al: In situ EPR monitoring of chromium spe- cies formed during Cr-pyrrolyl ethylene trimerization catalyst formation, Organometallics 29(13):2943–2950, 2010. Soares JBP, Abbott RF, Kim JD: Environmental stress cracking resistance of polyethylene: the use of CRYSTAF and SEC to establish structure-property relationships, J Polym Sci Polym Phys 38(10):1267–1275, 2000. Spitz R, Florin B, Guyot A: Ethylene-hexene-1 copolymers through modified Phillips catal- ysis, Eur Polym J 15(5):441–444, 1979. Stadler FJ, Gabriel C, Muenstedt H: Influence of short-chain branching of polyethylenes on the temperature dependence of rheological properties in shear, Macromol Chem Phys 208(22):2449–2454, 2007. Tang W, Santare MH, Advani SG: Melt processing and mechanical property characterization of multi-walled carbon nanotube/high density polyethylene (MWNT/HDPE) compos- ite films, Carbon 41(14):2779–2785, 2003. Temple CN, Jabri A, Crewdson P, et al: The question of the Cr oxidation state in the Cr-(SNS) catalyst for selective ethylene trimerization: an unanticipated re-oxidation pathway, Angew Chem Int Ed 45(42):7050–7053, 2006. Temple CN, Gambarotta S, Korobkov I, Duchateau R: New insight into the role of the metal oxidation state in controlling the selectivity of the Cr-(SNS) ethylene trimerization catalyst, Organometallics 26(18):4598–4603, 2007. Theopold KH: Homogeneous chromium catalysts for olefin polymerization, Eur J Inorg Chem 1998(1):15–24, 1998. Thu¨ne PC, Verhagen CPJ, van den Boer MJG, Niemantsverdriet JW: Working surface sci- ence model for the Phillips ethylene polymerization catalyst: preparation and testing, J Phys Chem B 101(42):8559–8563, 1997. Thu¨ne PC, Loos J, Chen X, et al: Visualization of local ethylene polymerization activity on a flat CrOx/SiO2/Si(100) model catalyst, Top Catal 46(1–2):239–245, 2007. Tobisch S, Ziegler T: Catalytic linear oligomerization of ethylene to higher a-olefins: insight into the origin of the selective generation of 1-hexene promoted by a cationic cyclopentadienyl-arene titanium active catalyst, Organometallics 22(26):5392–5405, 2003. Tobisch S, Ziegler T: Catalytic oligomerization of ethylene to higher linear a-olefins pro- moted by cationic group 4 cyclopentadienyl-arene active catalysts: a DFT investigation exploring the influence of electronic factors on the catalytic properties by modification of the hemilabile arene functionality, Organometallics 23(17):4077–4088, 2004a. Tobisch S, Ziegler T: Catalytic oligomerization of ethylene to higher linear a-olefins pro- 5 II þ moted by the cationic group 4 [( -Cp-(CMe2-bridge)-Ph)M (ethylene)2] (M ¼ Ti, Zr, Hf ) active catalysts: a density functional investigation of the influence of the metal on the catalytic activity and selectivity, J Am Chem Soc 126(29):9059–9071, 2004b. Tobisch S, Ziegler T: Catalytic oligomerization of ethylene to higher linear a-olefins pro- moted by cationic group 4 cyclopentadienyl-arene active catalysts: toward the compu- tational design of zirconium- and hafnium-based ethylene trimerization catalysts, Organometallics 24(2):256–265, 2005. 190 Zhen Liu et al.

Tonosaki K, Taniike T, Terano M: Comprehensive investigation of catalyst structure and polymerization conditions for chain branching in ethylene polymerization with Phillips-type catalysts, Macromol React Eng 5(9–10):332–339, 2011a. Tonosaki K, Taniike T, Terano M: Origin of broad molecular weight distribution of poly- ethylene produced by Phillips-type silica-supported chromium catalyst, J Mol Catal A Chem 340(1–2):33–38, 2011b. Tsubokawa N: Functionalization of carbon black by surface grafting of polymers, Prog Polym Sci 17(3):417–470, 1992. Tsubokawa N, Hosoya M, Kurumada J: Grafting reaction of surface carboxyl groups on car- bon black with polymers having terminal hydroxyl or amino groups using N, N-dicyclohexyl-carbodiimide as a condensing agent, React Funct Polym 27(1):75–81, 1995. Uhe A, Kozuch S, Shaik S: Automatic analysis of computed catalytic cycles, J Comput Chem 32(5):978–985, 2011. van Kimmenade EME, Kuiper AET, Tamminga K, Thu¨ne PC, Niemantsverdriet JW: Sur- face science model for the Phillips ethylene polymerization catalyst: thermal activation and polymerization activity, J Catal 223(1):134–141, 2004. van Leeuwen PWNM, Clement ND, Tschan MJL: New processes for the selective produc- tion of 1-octene, Coord Chem Rev 255(13–14):1499–1517, 2011. van Rensburg WJ, Grove C, Steynberg JP, et al: A DFT study toward the mechanism of chromium-catalyzed ethylene trimerization, Organometallics 23(6):1207–1222, 2004. Vidyaratne I, Nikiforov GB, Gorelsky SI, et al: Isolation of a self-activating ethylene trimerization catalyst, Angew Chem Int Ed 48(35):6552–6556, 2009. Vuurman MA, Hardcastle FD, Wachs IE: Characterization of CrO3/Al2O3 catalysts under ambient conditions: influence of coverage and calcination temperature, J Mol Catal 84(2):193–205, 1993. Wang S, Tait PJT, Marsden CE: Phillips-type polymerization catalysts: kinetic behavior and active center determination, J Mol Catal 65(1–2):237–252, 1991. Wang S, Murata K, Hayakawa T, Hamakawa S, Suzuki K: Dehydrogenation of ethane with carbon dioxide over supported chromium oxide catalysts, Appl Catal A Gen 196(1):1–8, 2000. Wang M, Pramoda KP, Goh SH: Enhancement of the mechanical properties of poly(styrene- co-acrylonitrile) with poly(methyl methacrylate)-grafted multiwalled carbon nanotubes, Polymer 46(25):11510–11516, 2005. Weckhuysen BM, de Ridder LM, Schoonheydt RA: A quantitative diffuse reflectance spec- troscopy study of supported chromium catalysts, J Phys Chem 97(18):4756–4763, 1993. Weckhuysen BM, Wachs IE, Schoonheydt RA: Surface chemistry and spectroscopy of chro- mium in inorganic oxides, Chem Rev 96(8):3327–3349, 1996. Xia W, Liu B, Fang Y, Hasebe K, Terano M: Unique polymerization kinetics obtained from simultaneous interaction of Phillips Cr(VI)Ox/SiO2 catalyst with Al-alkyl cocatalyst and ethylene monomer, J Mol Catal A Chem 256(1–2):301–308, 2006. Xia W, Tonosaki K, Taniike T, et al: Copolymerization of ethylene and cyclopentene with the Phillips CrOx/SiO2 catalyst in the presence of an aluminum alkyl cocatalyst, J Appl Polym Sci 111(4):1869–1877, 2009. Xia W, Liu B, Fang Y, et al: Multinuclear solid-state NMR study of the coordinative nature of alkylaluminum cocatalyst on Phillips CrOx/SiO2 catalyst, Appl Catal A Gen 389(1–2):186–194, 2010. Xu M, Zhang T, Gu B, Wu J, Chen Q: Synthesis and properties of novel polyurethane-urea/ multiwalled carbon nanotube composites, Macromolecules 39(10):3540–3545, 2006. Xu R, Winget P, Clark T: The cyclooligomerisation of acetylene at metal centres, Eur J Inorg Chem 2008(18):2874–2883, 2008. Ethylene Polymerization and Oligomerization 191

Yang Y, Liu Z, Zhong L, et al: Spin surface crossing between Chromium(I)/Sextet and Chromium(III)/Quartet without deprotonation in SNS-Cr mediated ethylene trimerization, Organometallics 30(19):5297–5302, 2011. Yu Z, Houk KN: Why trimerization? Computational elucidation of the origin of selective trimerization of ethene catalyzed by [TaCl3(CH3)2] and an agostic-assisted hydride trans- fer mechanism, Angew Chem Int Ed 42(7):808–811, 2003. Zaki MI, Fouad NE, Leyrer J, Knoezinger H: Physicochemical investigation of calcined chromia-coated silica and alumina catalysts: characterization of chromium-oxygen spe- cies, Appl Catal 21(2):359–377, 1986. Zecchina A, Bertarione S, Damin A, et al: Temperature resolved FTIR spectroscopy of 2þ Cr /SiO2 catalysts: acetylene and methylacetylene oligomerisation, Phys Chem Chem Phys 5(20):4414–4417, 2003. Zhang S, Zhao N, Wu Y, et al: Short chain branches distribution characterization of ethyl- ene/1-hexene copolymers by using TREF þ 13C-NMR and TREF þ SC methods, Macromol Symp 312(1):63–71, 2012a. Zhang S, Dong Q, Cheng R, et al: A novel SiO2-supported inorganic and organic hybrid chromium-based catalyst for ethylene polymerization, J Mol Catal A Chem 358:10–22, 2012b. Zhang S, Cheng R, Dong Q, et al: Ethylene/1-hexene copolymerization with a novel SiO2- supported inorganic and organic hybrid chromium-based catalyst, Macromol React Eng 7(6):254–266, 2013. Zhong L, Lee MY, Liu Z, et al: Spectroscopic and structural characterization of Cr(II)/SiO2 active site precursors in model Phillips polymerization catalysts, J Catal 293:1–12, 2012a. Zhong L, Liu Z, Cheng R, et al: Active site transformation during the induction period of ethylene polymerization over the Phillips CrOx/SiO2 catalyst, ChemCatChem 4(6):872–881, 2012b.