1 Introduction

In the early 1950s, J. P. Hogan and R. L. Banks of Phillips Petroleum Company made the discovery that ethylene can be converted to polyethylene by a chromium oxide–silica–alumina catalyst [1] that was subsequently commercialized. The Phillips type catalyst system, CrOx supported on an amorphous support, such as silica, is one of three types of catalysts currently used for olefin polymerization. The other two types are Ziegler–Natta and single-site homogeneous catalysts or supported homogeneous catalysts, both of which require an activator. The appeal of the Phillips catalyst lies in its many advantages: (i) yielding over 50 different types of polyethylene, (ii) functioning without activators, and (iii) operating at low temperatures and pressures [13]. The original catalyst system has since been fine-tuned and ethylene polymerization by silica supported CrOx catalysts is now responsible for ~40–50 % of all high-density polyethylene (HDPE) produced [3]. In spite of the extensive research studies that have been performed about the supported CrOx/SiO2 catalyst system over the past six decades, the same fundamental structural and mechanistic questions are still being debated [14]. For example, the initial molecular structure of the oxidized surface Cr+6Ox site has been proposed to be present as isolated surface dioxo CrO4, isolated surface mono-oxo CrO5, and dimeric surface Cr2O7, while the chromia oxidation state during ethylene polymerization has been proposed to be reduced Cr+2 and Cr+3. This perspective focuses on the nature of the surface chromia sites on silica in the different environments [oxidizing (hydrated and dehydrated) and reducing (CO, H2 and C2H4)] to stress what is currently known and what more needs to be done to fully understand the nature of the surface CrOx sites present for the silica-supported chromium oxide catalysts.

2 Advantages and Disadvantages of Characterization Techniques for Supported CrOx/SiO2 Catalysts

Many characterization techniques have been used to investigate supported CrOx/SiO2 catalysts, but each method has its advantages and disadvantages. The typical techniques have been ultraviolet–visible diffuse reflectance spectroscopy (UV–Vis DRS), electron paramagnetic resonance (EPR) spectroscopy, infrared (IR) spectroscopy, Raman spectroscopy, X-ray absorption spectroscopy (XAS), and temperature programmed surface reaction (TPSR) spectroscopy.

UV–Vis spectroscopy probes electronic transitions, allowing for determination of the coordination and oxidation state of charge transfer (CT) Cr+6 bands and d–d Cr+5,+3 transition bands. The broad UV–Vis bands, however, make it difficult to distinguish between multiple species and the d–d transitions are weak and may overlap nearby CT or d–d bands. The UV–Vis edge energy, Eg, provides direct information about the extent of oligomerization of the surface CrO+6 sites (monomer, dimer, etc.).

EPR spectroscopy can detect paramagnetic (containing unpaired electrons) Cr+5 and Cr+3 sites and provide their coordination information. Although Cr+2 sites are also paramagnetic, its EPR signal is too weak to be detected with conventional EPR spectrometers.

IR spectroscopy is a very powerful technique that provides molecular vibrational information and is more sensitive to asymmetric vibrations due to its selection rules. IR has been valuable in determining the anchoring sites of the surface CrOx species since the Si–OH vibrations are rather intense. The strong IR absorption by the SiO2 support, however, only allows the monitoring of Cr–O vibrations in the ~850–970 cm−1 window. IR spectroscopy also provides vibrational information about surface intermediates and reaction products formed on the catalyst surface.

Raman spectroscopy also probes molecular vibrations and is complementary to IR but is more sensitive to symmetric vibrations. The weak Raman bands from the SiO2 support allows monitoring vibrations of surface CrOx sites from 0 to 4000 cm−1 for the supported CrOx/SiO2 catalyst system.

XAS includes X-ray absorption near edge structure (XANES) and extended X-ray absorption fine structure (EXAFS). XANES allows determination of oxidation and coordination states by observing the pre-edge energy and selection rules. The position of the pre-edge energy reflects the oxidation state. For CrO4 coordinated sites, there is a strong pre-edge, but the pre-edge is forbidden for CrO6 coordinated sites with inverse symmetry due to the XANES selection rules. EXAFS allows determination of Cr–O bond distances. The shortcoming of XAS is that it averages over multiple sites if a single site is not present, which complicates analysis since the concentration of sites may not be known. Furthermore, the XAS signal reflects the dominant species if two sites are present with one having a high concentration and the other being a minority species (e.g., CrOx and Cr2O3 or Cr+6 and Cr+3). Thus, it is critical to make sure that Cr2O3 nanoparticles are not present since they will complicate analysis of the surface CrOx species. The presence of multiple Cr oxidation states also complicates the analysis.

Temperature Programmed Surface Reaction (TPSR) spectroscopy chemically probes the reactivity of the supported chromia phase on silica. During TPSR, the catalyst temperature is ramped in the presence of a chemical reactant and the gaseous reaction products are continuously monitored with an online mass spectrometer. TPSR has the capability to discriminate between multiple surface CrOx sites on silica, quantify their relative concentrations and determine their reaction kinetics.

No single characterization technique is sufficient to fully characterize the supported CrOx sites on SiO2 given the limitation of each characterization method. Applications of multiple characterization techniques, however, can tease out the desired fundamental information. Consequently, as will be seen below, the most informative studies employ multiple characterization techniques to yield a comprehensive model of the supported CrOx/SiO2 catalyst system.

3 Hydrated Supported CrOx/SiO2 in Initial Oxidized Catalyst under Ambient Conditions

The supported CrOx on silica catalyst system is prepared by the incipient-wetness impregnation method employing a chromia precursor that is soluble in the solvent being employed. The impregnated catalyst is initially dried overnight at room temperature in air, further dried in air at ~100 °C  and then calcined at elevated temperatures (~400–800 °C) in an oxidizing environment [3]. Upon exposure to ambient air, the oxidized catalyst becomes hydrated by adsorption of moisture, which affects the molecular structures of the oxidized surface CrOx species on SiO2 [5, 6].

The lack of appreciation that the surface CrOx sites on silica are hydrated under ambient conditions and limited application of characterization techniques capable of discriminating between multiple chromia states has also led to much confusion about the nature of the hydrated surface CrOx sites on silica. Raman spectroscopy, however, readily discriminates between different chromia molecular structures since each molecular structure gives rise to a unique vibrational spectrum [7]. The first application of Raman spectroscopy to the examination of hydrated supported CrOx/SiO2 catalysts was reported by Hardcastle and Wachs [5]. The hydrated surface CrOx phase on silica was found to consist of monomers (CrO4), dimers (Cr2O7), trimers (Cr3O10) and tetramers (Cr4O13), with the extent of oligomerization increasing with chromia loading. The oligomerization of the surface CrOx sites on SiO2 with increasing chromia loading is a consequence of the decreasing pH at point of zero charge (PZC) of the thin aqueous film present on the supported CrOx/SiO2 catalyst [8].

Weckhuysen et al. [9] investigated supported chromia catalysts on silica under ambient conditions as a function of Cr loading (0.2–8 % Cr) [9]. The UV–Vis DRS spectra for the lowest Cr loading exhibited two bands dominating at 348 and 240 nm with a shoulder at 429 nm, and the corresponding Raman spectra contained just one band at 895 cm−1. These spectral features indicate hydrated surface dichromate species [7, 10]. As the Cr loading was further increased, the UV–Vis bands broadened and shifted from 348 to 370 nm, and the shoulder at 429 nm became more pronounced. With increasing Cr loading, the Raman spectra contained bands at 380, 850, 900 and 960 cm−1 that are representative of hydrated trichromate and tetrachromate species [7]. The XANES spectra of the hydrated catalyst exhibited a sharp pre-edge at about 5992 eV, which is a dipole-forbidden transition 1s → 3d characteristic of 3d metal oxides not possessing an inversion center such as tetrahedral coordinated CrO4. The corresponding EXAFS spectra of the hydrated catalyst possessed one FT peak at ~1.3 Å with an out-of-phase shift correction. After curve-fitting of the EXAFS spectra, the Cr–O distance and coordination number for the first shell were determined to be 1.58 Å and 3.7, respectively, which are similar to those of the CrO4 unit in the K2CrO4 reference compound (R = 1.62 Å and N = 3.9). The combination of the spectroscopic techniques allowed for a full picture of the hydrated chromia species on silica. It was determined that the nature of the hydrated surface CrOx species only depends on the pH at PZC, which leads to dichromates at low Cr loading and trichromates and tetrachromates at high Cr loading. Crystalline Cr2O3 nanoparticles also form at higher Cr loading.

The molecular structures of the hydrated surface CrOx sites on SiO2 are schematically shown in Fig. 1.

Fig. 1
figure 1

Schematic of molecular structures of "hydrated" CrOx sites on the SiO2 support: (a) monomer CrO4 2-, (b) dimer Cr2O7 2-, (c) trimer Cr3O10 2-, and (d) tetramer Cr4O13 2-

4 Supported CrOx/SiO2 in Initial Oxidized Catalyst under Dehydrated Conditions

The hydrated surface CrOx species on SiO2 become dehydrated and change their molecular structures upon heating to elevated temperatures [5]. Early thermogravimetric analysis (TGA) by Hogan examined the loss of water by silica upon adsorption of chromia on dehydrated SiO2 and led to the conclusion that surface CrOx on silica is mainly present as isolated CrO4 sites, although at high temperatures dimeric Cr2O7 sites cannot be excluded [11]. This conclusion was based on the model that only isolated surface CrO4 and dimeric Cr2O7 sites are present and anchor to the silica support by titrating surface Si–OH hydroxyls. Each chromia structure was assumed to consume two surface Si–OH hydroxyls with a ratio of 2 Si–OH/Cr for the isolated CrO4 and 1 Si–OH/Cr for the dimeric Cr2O7. This model also assumes that the supported chromia phase is (i) 100 % dispersed (crystalline Cr2O3 nanoparticles are not present), (ii) only possesses CrO4 coordination and (iii) the SiO2 surface is not altered by anchoring of the surface chromia species (e.g., siloxane Si–O–Si bonds do not participate). In spite of the absence of direct spectroscopic evidence about the nature of the surface CrOx sites on SiO2, this early publication introduced the concept that the surface CrOx sites on SiO2 are only present as isolated CrO4 and dimeric O3Cr–O–CrO3 structures [11].

Zecchina et al. were the first group to apply in vacuo IR and UV–Vis spectroscopy to examine the dehydrated supported CrOx/SiO2 catalysts [10]. The IR spectra exhibited a broad band at 925 cm−1 that was assigned to surface dichromate sites on silica. Recent DFT and Raman analysis, to be presented below, demonstrate that this IR band is actually the vibration from the bridging Cr–O–Si bond and does not provide any information about the structure of dehydrated the surface CrOx sites on silica. In the silanol stretching region (~3700 cm−1), the intensity of the Si–OH band linearly decreased with increasing chromia loading up to ~5 % Cr and the silanol concentration did not further decrease with increasing Cr loading. The leveling off of the consumption of the surface silanols coincides with the formation of crystalline α-Cr2O3 nanoparticles. The ratio of surface Cr to Si–OH consumed was determined to be ~1, which suggested surface dichromate sites since isolated surface chromate sites would be expected to be ~0.5 according to Hogan’s model. The UV–Vis spectra contained O → Cr+6 ligand-to-metal charge transfer (LMCT) bands below 556 nm that were assigned to surface dichromate and isolated surface chromate sites, with relative intensities suggesting that the surface dichromate dominates. The broad nature of these UV–Vis bands at 256 and 351 nm with the shoulder at 385 nm, however, can be assigned to both isolated surface chromate and surface dichromate since clear distinction between both structures is complicated by their broad and overlapping UV–Vis bands as shown in Fig. 2. The band at 461 nm was assigned to dichromates, but this band is also present for Cr2O3 nanoparticles. A small d–d band above 556 at 714 nm was attributed to chromia in an oxidation state lower than Cr+6 and was thought to be from surface Cr+5O4 sites on silica. This pioneering characterization study, however, could not clearly provide spectroscopic identification of the nature of the dehydrated surface CrOx sites on SiO2 since the UV–Vis bands for the dehydrated supported CrOx/SiO2 catalyst and the CrO4 and Cr2O7 reference compounds are too broad to allow assignment as shown in Fig. 2.

Fig. 2
figure 2

UV–vis of reference compounds under ambient conditions

Fubini et al. also reported in vacuo UV–Vis–NIR spectra of dehydrated supported CrOx/SiO2 catalysts and detected LMCT bands at 476, 345, and 238 nm from surface Cr+6Ox sites. These bands were very close to those observed by Zecchina et al. with the exception of an additional shoulder at ~385 nm. Comparison with K2CrO4, containing isolated CrO4 units, and K2Cr2O7, consisting of dimeric Cr2O7 units, suggested that surface dichromate sites were the dominant surface chromia sites on silica [12]. This group came to the same conclusions as Zecchina et al.

In the past two decades, there was a growing emphasis on investigations with in situ characterization techniques that provided direct observation of the dehydrated surface CrOx sites on silica.

Weckhuysen et al. examined the structure of the dehydrated surface CrOx sites on SiO2 as a function of Cr loading with in vacuo XANES/EXAFS, in situ Raman and in situ UV–Vis spectroscopy [9, 13]. The XANES pre-edge intensity possessed a strong feature at 5992 eV indicative of Cr+6O4 coordination. The corresponding in situ EXAFS spectra contained 2 Cr–O distances at 1.53 Å and 2.05 Å with coordination numbers of 2.2 and 2.1, respectively. An additional Cr–Cr peak was present at ~3.1 Å with a coordination number of ~0.5. Two possible explanations were given. In the first interpretation, polychromate species are present with Cr=O at 1.53 Å and bridging Cr–O–Si at 2.05 Å. Alternatively, the 1.53 Å distance could be related to surface polychromate sites and the 2.05 Å peak is from the presence of Cr+3 oxide species. The Cr–Cr peak at ~3.1 Å is also consistent with Cr2O3 nanoparticles. Weckhuysen et al. favored the latter interpretation because it was more in line with the corresponding UV–Vis spectra that detected Cr+3 [9]. The in situ UV–Vis bands for the supported CrOx/SiO2 catalysts gave bands at 645, 455, 328 and 247 nm, with a weak shoulder at 370 nm. These UV–Vis spectra were deconvoluted into 11 bands [13]: 3 weak bands (645, 465, and 295 nm) assigned to Cr+ 3O6, 4 bands assigned to isolated Cr+6O4 (476, 370, 294, and 268 nm) and 4 bands assigned to dimeric Cr+6 2O7 (455, 323, 282, and 246 nm). The intensity ratio of the 370 nm band for isolated CrO4 and the 323 nm band for dimeric Cr2O7 was employed to estimate the ratio of isolated/dimeric surface chromia species, which was found to be ~0.62 for 0.2 % Cr loading. In situ Raman spectra of the dehydrated catalysts, unfortunately, could not be collected with the 514.5 nm laser due to strong sample fluorescence. More recent successful Raman analysis of dehydrated supported CrOx/SiO2 catalysts confirmed the presence of Cr2O3 nanoparticles at loadings of 4 % Cr and greater [14]. It was concluded that under dehydrated conditions and very high Cr loading on silica, surface polychromates and Cr2O3 nanoparticles dominate [9].

The in situ XANES/EXAFS studies by Groppo et al. agree with the earlier findings of Weckhuysen et al. and demonstrated that the XANES spectra of the dehydrated supported CrOx/SiO2 catalysts matched with that of bulk CrO3, consisting of polymeric dioxo CrO4 units that are compatible with both dioxo isolated CrO4 and dimeric Cr2O7 structures [15]. Groppo et al.’s earlier UV–Vis spectra of the dehydrated catalyst showed three main components (250, 333, and 463 nm) assigned to O → Cr+6 LMCT transitions, which were quite similar to those previously reported by others [15]. Groppo et al. pointed out the limitations of applying UV–Vis to identify surface chromate species because of their very broad and overlapping bands.

Groppo et al. also employed Raman spectroscopy to determine the structure of the surface chromia species [16]. In situ Raman spectroscopy with the 442 nm laser, as opposed to the 514.5 nm laser previously available, allowed for the first time the elimination of sample fluorescence for the supported CrOx/SiO2 catalysts. The Raman spectra gave well-resolved bands at 394 (m), 987 (s) and 1014 (m) cm−1 with the 394 and 1014 cm−1 bands previously not observed due to strong sample fluorescence with lasers using 514 and 532 nm excitation. The band at 987 cm−1 was assigned to the νs(O=Cr=O) stretching, the band at 394 cm−1 to δ(O=Cr=O) bending, and the band at 1014 cm−1 to νas(O=Cr=O) stretching, as suggested by the DFT calculations of Dines et al. [17]. Groppo et al., however, had some concern about the assignment of the 1014 cm−1 band because it was stronger than expected and suggested that there may be some heterogeneity of the surface sites. The Raman spectra also showed an absence of chromia bands at 200–300 and 400–700 cm−1 indicating the lack of any surface polymeric chromia species at low Cr loadings. Comparison of DFT calculations and Raman spectra revealed that the presence of surface Cr+6Ox could also modify the pure silica vibrations and the IR band at 908 cm−1 band previously reported by Zecchina is actually related to a Si–O vibration affected by anchoring of the surface chromate species to the silica support (e.g., bridging Cr–O–Si bonds). The combination of in situ Raman, IR and UV–Vis spectra of the dehydrated supported CrOx/SiO2 catalysts at low chromia loadings on silica led Groppo et al. to conclude that surface Cr+6Ox on SiO2 is anchored as isolated dioxo surface (O=)2CrO2 species [16].

The more recent in situ UV–Vis and Raman studies by Lee and Wachs finally provided the fundamental insights about the molecular structures of the dehydrated surface CrOx sites on SiO2 [14, 18]. The in situ UV–vis absorption bands at ~250, 340 and 460 nm reflect the presence of Cr+6Ox species on silica [14]. The exclusive presence of isolated surface CrOx species on silica was demonstrated by the high UV–Vis edge energy (Eg) value that is consistent with isolated CrOx reference compounds [14]. The corresponding Raman spectrum, with 442 nm laser excitation, exhibited bands at 987 (s) and 1014 (m) cm−1, and the selective reduction of the 987 cm−1 band by H2 showed that both bands originate from two distinct surface CrOx sites on silica [14]. Corresponding isotopic 18O-16O exchange studies of the surface CrOx species on silica exhibited splitting of the 987 cm−1 band into three bands (16O=Cr=16O, 16O=Cr=18O and 18O=Cr=18O) consistent with the dioxo nature of the 987 cm−1 band ((O=)2CrO2). The band at 1014 cm−1 was assigned to mono-oxo O=CrO4 sites by comparison with chromium oxide reference compounds, but also contains a small contribution from the νas(O=Cr=O) stretch of the isolated surface dioxo (O=)2CrO2 species [18]. These molecular assignments to isolated surface dioxo (O=)2CrO2 and isolated mono-oxo O=CrO4 sites on SiO2 were further supported by rigorous DFT calculations that also predicted the greater stability of the dioxo CrO4 species over the mono-oxo CrO5 surface species on SiO2 [19]. This suggests that the surface dioxo CrO4 species should have a higher concentration than the surface mono-oxo CrO5 species on the silica support (as will be experimentally shown below with H2-TPSR). The confusion surrounding the dehydrated molecular structures of surface CrOx sites on SiO2 required the application of multiple in situ spectroscopic approaches and techniques (Raman, H2 reduction, isotopic 18O-16O exchange, XAS, IR, UV–Vis and DFT calculations) for its final resolution. The molecular structures of the two dehydrated surface CrOx sites on SiO2 are depicted in Fig. 3.

Fig. 3
figure 3

Schematic of molecular structures of "dehydrated" isolated surface CrOx sites on SiO2 support (a) dioxo and (b) mono-oxo

5 Activated Supported CrOx/SiO2 Catalysts in Reducing Environments

In industrial ethylene polymerization processes, the supported CrOx/SiO2 catalysts are activated with the ethylene reactant [2]. This results in an induction period during which the surface CrOx catalytic sites are activated by partial reduction to a suboxide for ethylene polymerization. Hogan found during his early studies that reduction in CO could basically eliminate the induction period observed with ethylene, after which activity was immediately observed. It became customary in academic studies to initially activate the supported CrOx/SiO2 catalyst by reduction with CO and occasionally H2. Additionally, pretreating with CO or H2 affects the polyethylene (PE) yields. Pretreatment with CO modestly increases the PE yield while pretreatment with H2 can significantly suppress the PE yield and the difference may be related to the presence of moisture produced with the latter [20]. The application of different reducing agents in the literature has caused some confusion since the final activated state may depend on the specific reducing agent being employed.

5.1 Activation of CrOx/SiO2 Catalysts with CO

Reduction with CO is a preferred activation method by academic researchers since it does not generate water as a reaction product that may coordinate with the surface CrOx sites and the findings from different groups are listed in Table 1. Early studies employed CO-TPR to determine how much oxygen was removed from the supported CrOx/SiO2 catalysts and found that ~2.05–2.15 O/Cr were consumed indicating that the average oxidation state was Cr+2 [9, 13]. This was in agreement with EPR spectra that showed no Cr signal since Cr+2 is EPR silent, although it is d4 and paramagnetic in most chemical environments [21]. Chemical probing of the reduced surface sites with CO-IR studies suggested that two surface Cr+2 and one surface Cr+3 site may be present after CO reduction [22]. Other studies concluded that even more reduced surface Cr sites may be present [9, 12, 13, 2325]. The more recent definitive work of Groppo et al. with in situ Cr K-edge XANES/EXAFS demonstrated with direct monitoring of the surface chromia species on silica that surface Cr+2 sites are present after CO reduction. These studies, however, are somewhat complicated by the presence of 40 % Cr2O3 nanoparticles in their supported CrOx/SiO2 catalyst [15]. The UV–Vis band at 833 nm was assigned to Cr+2, but is quite close to the UV–Vis band of bulk Cr2O3 at 810 nm. Most researchers agree that multiple surface Cr+2 sites are present on silica after activation with CO at elevated temperatures (see Table 1). The molecular structure of the Cr+2 site(s) still needs to be addressed and is complicated by possible presence of two distinct surface Cr+2 sites from the initial isolated dioxo and mono-oxo surface chromates on silica.

Table 1 Summary of CO–Activation Studies

5.2 Activation of CrOx with H2

Activation studies involving H2 are relatively rare and none of the studies applied spectroscopic methods to directly monitor the nature of the reduced surface CrOx sites on silica. All the reported H2 activation studies employed H2-TPR, with or without TGA, to investigate the reduction features of the supported CrOx sites on silica. Hogan found with TGA that oxygen consumption from the supported CrOx/SiO2 was greater with CO than H2 [11]. This was also confirmed by the final colors of the activated CrOx/SiO2 catalysts that showed a gray-green color with H2 reduction and a blue color with CO reduction that are characteristic of Cr+3 and Cr+2, respectively.

Gaspar et al. also applied H2-TPR spectroscopy to investigate the reduction of supported CrOx/SiO2 catalysts [26]. The H2-TPR reduction peak(s) were found to depend on the Cr loading (~479 °C for 0.5 % Cr, ~496 °C for ~1 % Cr, and at ~376 °C and ~470 °C for ~3 % Cr). The H2-TPR peaks at the higher temperatures were ascribed to reduction of surface Cr+6→ Cr+3 and the peak at 649 K for the higher Cr loading was assigned to reduction of large Cr+6O3 particles → Cr+3 that formed because the amount of hydroxyls was not sufficient to stabilize all the chromium on the silica surface [26]. The CrO3 particles, however, would have thermally decomposed to Cr2O3 particles during the 500 °C calcination and would not be present. Although UV–Vis spectra were collected before H2-TPR, the broad nature of the UV–Vis bands prevented clear structural assignments of the surface chromia species on silica as already discussed above. After H2-TPR, all the reduced supported CrOx/SiO2 catalysts possessed UV–Vis bands at 272, 355, 467 and 610 nm that were assigned to Cr+3. In agreement with Hogan, the final oxidation state resulting from H2 reduction was concluded to be Cr+3 [11].

5.3 Activation of CrOx with C2H4

Activation studies with C2H4 are also limited because of complications associated with the presence of multiple hydrocarbons and possibly the H2O reaction product that can interact with the surface CrOx sites on silica [11]. A recent detailed in situ XANES/EXAFS study by Groppo et al. [15] monitored ethylene activation of supported CrOx/SiO2 catalysts and found that the catalyst exhibits features different than the CO-activated catalyst, which exhibits an oxidation state of Cr+2. The C2H4-reduced CrOx/SiO2 catalyst yields a XANES spectrum with a higher intensity of the white line, reflecting an increased average Cr coordination, and the Cr+2 fingerprint is not present, which suggests that all surface isolated CrOx sites are involved. Additionally, as seen in the corresponding EXAFS, these surface sites are not Cr+2 dimers because there is no signal for Cr–Cr in the second coordination shell [15]. The fraction of surface Cr sites interacting with ethylene has previously been proposed to vary from 10 to 55 % depending on the probe method [2729], but these claims were based on indirect measurements and are not supported by this new direct XAS measurement indicating that all the dispersed Cr sites are involved [15]. The difference in the states of the surface CrOx sites on silica upon CO reduction and C2H4 reduction were further demonstrated by adsorbing ethylene on a catalyst that had been pre-reduced with CO. Adsorption of ethylene on the Cr+2 sites increased the white line intensity (slight increase in coordination of Cr+2) and completely removed the Cr+2 finger print in the pre-edge region. These results clearly demonstrate that the surface Cr+2 sites formed with CO reduction are not involved in ethylene polymerization reaction at room temperature (and low ethylene pressure) and that all the reduced surface Cr+2 sites are oxidized upon exposure to ethylene. This work nicely demonstrates that reduction in different environments leads to different activated surface CrOx states on SiO2 and that it is important not to confuse results from one reducing agent with another.

5.4 Activation of Well-Defined Catalysts With C2H4

Recent studies by Copéret et al. have examined the polymerization activities of well-defined model catalysts, dinuclear Cr+2 and Cr+3 model compounds and mononuclear Cr+3 model compounds, anchored on SiO2 [30, 31]. The grafted dinuclear model (≡SiO)4Cr+2 2 and (≡SiO)6Cr+3 2 compounds were characterized by in situ IR, Cr K-edge XAS, and EPR spectroscopy before and after ethylene polymerization. The IR spectrum of (≡SiO)4Cr+2 2 only contained new silanol vibrations and no C–H stretches before exposure to ethylene. X-ray crystallographic analysis of a molecular model closely related to (≡SiO)6Cr2 revealed two five-coordinate Cr+3 centers (distorted trigonal bipyramidal and square-pyramidal geometry). The IR spectrum of (≡SiO)6Cr+3 2 after exposure to ethylene possessed C–H stretches and a white film formed on the pellet indicating polyethylene formation. The XANES spectrum of (≡SiO)4Cr+2 2 did not undergo any changes during the synthesis procedure. The initial ethylene polymerization activity of (≡SiO)6Cr+3 2 was an order of magnitude higher than (≡SiO)4Cr+2 2 suggesting that Cr+3 is the catalytic active site for this reaction. The EPR spectrum of (≡SiO)4Cr+2 2 taken before exposure to ethylene showed a weak signal for Cr+3, suggesting that the minor amount of Cr+3 is most probably responsible for the polymerization activity [30]. Similar high initial polyethylene polymerization activity was also found for the mononuclear (≡SiO)3Cr+3 model compound consistent with the role of Cr+3 sites for ethylene polymerization [28]. Furthermore, the same initial ethylene polymerization rate was also obtained with a traditional CO-activated supported CrOx/SiO2 catalyst consistent with the role of surface Cr+3 sites for ethylene polymerization [32]. These new findings further support the role of Cr+3 sites on silica as the active sites, but may not necessarily be identical to the traditional catalyst.

6 Structure–Activity Relationship for Supported CrOx/SiO2 Catalysts

Developing fundamental structure–activity relationships requires knowing the molecular structure of the catalytic active site(s). This is rather complicated for traditional supported CrOx/SiO2 catalysts since two distinctive isolated surface chromia sites are initially present in the oxidized catalyst: dioxo (O=)2CrO2 and mono-oxo O=CrO4. Thus, developing structure–activity relations requires knowing the concentrations of the two surface chromia sites and their specific activities, which is quite challenging. Consequently, the reported studies to date have assumed that only one surface CrOx site is present on SiO2 and only focused on the average oxidation state, which appears to primarily be surface Cr+3 during ethylene polymerization from the above review of activation of supported CrOx sites by ethylene. The establishment of fundamental molecular structure–activity relationships for ethylene polymerization on each surface chromia site present in supported CrOx/SiO2 catalysts still awaits resolution with future studies.

New insights about the structure–activity relationships of supported CrOx/SiO2 catalysts are provided below from H2-TPR studies. Although H2 reduction kinetics are not ethylene polymerization kinetics, the reduction of the surface chromia sites on silica is the initial step in activating surface chromia for ethylene polymerization.

The electronic structures of the surface CrOx sites on silica during H2-TPR were monitored with in situ UV–Vis DRS and the spectra are presented in Fig. 4. The UV–Vis spectra revealed that the reduced catalyst has a Cr+3 oxidation state and is in an octahedral coordination. The initial dehydrated supported CrOx/SiO2 catalyst possesses LMCT transfer bands at 252, 346 and 460 nm that are characteristic of surface Cr+6Ox sites on silica [14]. During H2-TPR, the UV–Vis Cr+6Ox LMCT bands decrease with increasing temperature and a new UV–Vis d–d band monotonically increases at ~690 nm that corresponds to surface Cr+3O6 sites [33]. The UV–Vis d–d transitions for Cr+3O6 sites should also be accompanied by additional bands at ~288 and 383 nm, but incomplete reduction of the Cr+6Ox LMCT bands in the same region prevents detection of these additional weak d–d bands. The intense green color of the H2 reduced supported CrOx/SiO2 is consistent with the dominance of surface Cr+3 sites on silica. Thus, both the initial fully oxidized surface dioxo (O=)2CrO2 and mono-oxo O=CrO4 sites on the SiO2 support appear to reduce to surface Cr+3O6 sites during H2-TPR.

Fig. 4
figure 4

In situ H2 TP–UV–vis DRS of 3 % CrO3/SiO2 (a) full spectrum and (b) zoomed in to see the Cr+3 band. The 3 % CrO3/SiO2 catalyst was heated to 500 °C at 10 °C/min in 10 % O2/Ar, held at 500 °C for 1 h, and cooled to 150 °C. It was then flushed for 30 min in Ar. Then the temperature was ramped to 500 °C in 4.8 % H2. Spectra were taken every 50 °C

The in situ H2-TP Raman spectra supported for the supported 3 % CrOx/SiO2 catalyst are presented in Fig. 5. The Raman spectrum of the initial dehydrated CrOx/SiO2 catalyst at 150 °C exhibits bands at ~988 and ~1020 cm−1, for the dioxo and mono-oxo CrOx species, respectively, and at ~400 cm−1 from the CrOx bending vibrations. In flowing hydrogen at 200 °C, the Raman band at ~988 cm−1 from the surface dioxo species rapidly diminishes relative to that from the band at 1020 cm−1 from the surface mono-oxo species. The Raman band at ~1020 cm−1 of the surface mono-oxo species continuously decreases with increasing temperature from 200 to 400 °C and is almost completely reduced at 400 °C.

Fig. 5
figure 5

In situ TP–Raman spectroscopy of 3 % CrO3/SiO2 in 4.8 % H2/Ar after subtracting the silica support spectrum. The 3 % CrO3/SiO2 catalyst was heated to 500 °C at 10 °C/min in 10 % O2/Ar, held at 500 °C for 1 h, and cooled to 150 °C. It was then flushed for 30 min in Ar. Then the temperature was ramped to 500 °C in 4.8 % H2/Ar. Spectra were taken every 50 °C with the 442 nm laser and a D1 filter

The relative concentrations of the surface dioxo (O=)2CrO2 and mono-oxo O=CrO4 sites on silica and their H2 reduction kinetics were chemically probed with H2-TPR as shown in Fig. 6. As shown above in Fig. 5 for the H2 TP-Raman spectra, the surface dioxo species reduce prior to the surface mono-oxo chromia species on silica. Deconvolution of the H2-TPR spectrum of the supported 3 % CrOx/SiO2 catalyst found that the surface chromia phase consists of ~2/3 surface dioxo sites and ~1/3 surface mono-oxo sites (exact dioxo:mono-oxo ratio ~1.7). These findings are in agreement with the DFT calculations by Handzlik et al. that predicted that dioxo surface (O=)2CrO2 sites should be much more stable than mono-oxo surface O=CrO4 sites on the SiO2 support [19]. The H2 reduction kinetics for the surface dioxo and mono-oxo chromia sites on silica are also provided from the peak temperatures (Tp) during the H2-TPR experiments by application of the Redhead equation for first-order reactions [34, 35]. The Arrhenius rate constants for H2-TPR reduction of the surface dioxo and mono-oxo chromia sites are 0.30/s and 8.2 × 10−3/s, respectively, at a reference temperature of 500ºC. Thus, the surface dioxo (O=)2CrO2 sites on silica represent about 2/3 of the supported chromia phase and their H2 reduction kinetics are ~37 times faster than that for the surface mono-oxo O=CrO4 sites on silica at 500 °C and much higher at lower temperatures. The reduction kinetics indicate the more facile activation of surface dioxo (O=)2CrO2 than surface mono-oxo O=CrO4 sites on silica. This is the first time that molecular structure–activity relationships have been established for supported CrO3/SiO2 catalysts. Future studies will establish the molecular structure–activity relationships for ethylene polymerization by the dioxo and mono-oxo chromia sites on SiO2 catalysts.

Fig. 6
figure 6

 H2-TPSR of 3 wt% CrO3/SiO2. The 3 % CrO3/SiO2 catalyst was heated to 500 °C at 10 °C/min in 10 % O2/Ar, held at 500 °C for 1 h, and cooled to 150 °C. It was then flushed with UHP He. Then the temperature was ramped to 800 °C in 0.83 % H2/He. H2O production was monitored

7 Conclusions

The nature of the surface CrOx sites strongly depends on the environment to which the supported CrOx/SiO2 catalysts are exposed. Under ambient conditions, hydrated surface chromia species are present (CrO4, Cr2O7, Cr3O10 and Cr4O13) and the extent of oligomerization increases with decreasing surface pH values at point of zero charge. Under dehydrated and oxidizing conditions, the surface chromia sites are present as isolated surface dioxo (O=)2CrO2 and mono-oxo O=CrO4 with a ratio of ~2/1 on silica that is independent of coverage below the maximum dispersion limit. Crystalline Cr2O3 nanoparticles are also present above the maximum dispersion limit, which depends on the Cr precursor, solvent and surface properties of the SiO2 support. Many of the early studies applied UV–Vis spectroscopy to determine the nature of the surface CrOx sites on silica, but it turns out that the broad and overlapping UV–Vis bands prevented clear cut structural assignments. The supported CrOx/SiO2 catalysts are activated for ethylene polymerization by exposure to reducing environments. Activation with CO leads to Cr+2 sites, but activation with H2 and C2H4 results in Cr+3 sites. For reduction by H2, the initial surface dioxo site is ~37 times more easily reduced than the initial surface mono-oxo site reflecting the greater ease of activation of the surface dioxo site than mono-oxo site in reducing environments. Similar fundamental molecular structure–activity relationships still need to be developed for ethylene polymerization by supported CrOx/SiO2 sites.