Skip to main content
Scientific Reports logoLink to Scientific Reports
. 2015 Dec 8;5:17800. doi: 10.1038/srep17800

Thermal catalytic oxidation of octachloronaphthalene over anatase TiO2 nanomaterial and its hypothesized mechanism

Guijin Su 1,a, Qianqian Li 1, Huijie Lu 1, Lixia Zhang 1, Linyan Huang 1, Li Yan 1, Minghui Zheng 1
PMCID: PMC4672297  PMID: 26643373

Abstract

As an environmentally-green technology, thermal catalytic oxidation of octachloronaphthalene (CN-75) over anatase TiO2 nanomaterials was investigated at 300 °C. A wide range of oxidation intermediates, which were investigated using various techniques, could be of three types: naphthalene-ring, single-benzene-ring, and completely ring-opened products. Reactive oxygen species on anatase TiO2 surface, such as O2−• and O2−, contributed to oxidative degradation. Based on these findings, a novel oxidation degradation mechanism was proposed. The reaction at (101) surface of anatase TiO2 was used as a model. The naphthalene-ring oxidative products with chloronaphthols and hydroxyl-pentachloronaphthalene-dione, could be formed via attacking the carbon of naphthalene ring at one or more positions by nucleophilic O2−. Lateral cleavage of the naphthalene ring at different C1-C10 and C4-C9, C1-C2 and C4-C9, C1-C2 or and C3-C4 bond positions by electrophilic O2−• could occur. This will lead to the formation of tetrachlorophenol, tetrachloro-benzoic acid, tetrachloro-phthalaldehyde, and tetrachloro-acrolein-benzoic acid, partially with further transformation into tetrachlorobenzene-dihydrodiol and tetrachloro-salicylic acid. Unexpectedly, the symmetric half section of CN-75 could be completely remained with generating the intricate oxidative intermediates characteristically containing tetrachlorobenzene structure. Complete cleavage of naphthalene ring could produce the ring-opened products, such as formic and acetic acids.


As a new type of persistent organic pollutants, polychlorinated naphthalenes (PCNs) was proposed into Annexes A and C of the Stockholm Convention (SC) on POPs in 20151. There are 75 possible PCN congeners, in eight homolog groups, with one to eight chlorine atoms substituted around the planar aromatic naphthalene molecule. PCNs have been widely used in many commercial products, e.g., for wood preservation, as additives to paints and engine oils, for cable insulation, and in capacitors. Because of the structural similarities between PCNs and polychlorinated biphenyls (PCBs), PCNs are present in technical PCB formulations2,3. Yamashita et al.4 examined the concentrations and profiles of tri- through octa-chloro-substituted congeners in 18 technical PCB mixtures, and detected concentrations ranging from 5.2 to 730 μg/g. PCNs are also unintentionally generated during high-temperature industrial processes in the presence of chlorine. Of the known releases, waste incineration is considered to be the significant current source5, with similar formation mechanism to that of polychlorinated dibenzo-p-dioxins and dibenzofurans (PCDD/Fs)6. The production and use of PCNs were banned in the United States and Europe in the 1980 s, because of their toxicity and environmental persistence7. Nevertheless, PCNs can be released from past use, products that have not yet been disposed of, devices containing PCBs still in use, and thermal processes such as waste incineration. In accordance with the relevant SC regulations, based on a risk management evaluation and consideration of the management options, the Committee recommended a Conference of the Parties to consider listing and specifying the relevant control measures for PCNs8. The reduction of PCN levels is therefore a matter of public concern in the context of environmental protection.

Catalytic oxidation for the removal of chlorinated aromatic hydrocarbons has attracted much attention as a green technique9,10. TiO2-based catalysts are generally used for the oxidation of chlorinated aromatic compounds11,12,13,14. Lichtenberger et al.12 examined the oxidation of chlorobenzene, and 1,2-, 1,3-, and 1,4-dichlorobenzene over V2O5/TiO2 catalysts. A common reaction mechanism was proposed based on kinetic and in situ fourier transform infrared (FTIR) results. Surface phenolates are formed via nucleophilic attack at the chlorine position in the aromatic ring, followed by electrophilic substitution of the adsorbed partially dechlorinated species in the second step. Krishnamoorthy et al.11 investigated the catalytic oxidations of 1,2-dichlorobenzene over Cr2O3, V2O5, MoO3, Fe2O3, and Co3O4 supported on TiO2 and Al2O3. The TiO2-supported systems were more active than the corresponding Al2O3-supported ones, indicating that the support is significant in the catalytic performance of the catalyst in this reaction. Gannoun et al.15 showed that sulfated TiO2 nanotubes (HNTs) were a promising support for V2O5-based materials in the oxidative elimination of chlorobenzene. The formed bridged bidentate Ti and acidic sites on the HNT surface probably govern chlorobenzene oxidation and decrease the reducibility of vanadium, leading to higher reactivity at redox sites and therefore to higher-efficiency catalysts. Thus far, however, the reports to deeply identify the oxidation products and the associated mechanisms of PCNs as new POPs, are particularly scarce.

TiO2 is an important semiconductor material and has been used in a variety of applications such as photosplitting of water16, photovoltaic devices17, liquid solar cells, surface wettability conversion, and degradation of toxic pollutants18. This wide range of applications can be attributed to its nontoxicity, low cost, photostability, redox efficiency, and availability. TiO2 has three crystal form, i.e., brookite, anatase, and rutile. The crystal form of TiO2 has a decisive effect on its catalytic performance, because the electronic band gaps (EBGs) of the different forms of TiO2 are different. It has been reported that the photocatalytic activity of anatase TiO2 is limited by its small absorption range in the solar spectrum, as a result of its large EBG (Eg = 3.2 eV). However, the larger EBG of anatase TiO2 has attracted great interest in its better oxidation performance. Therefore, it is of significance that the catalytic oxidation of PCNs is performed by anatase TiO2 with illustrating the involved deep oxidation mechanism.

In this study, the reactivity of an anatase TiO2 nanomaterial toward a model compound, i.e., octachloronaphthalene (CN-75), which is fully substituted with chlorine atoms, was evaluated at 300 °C. The degradation products, especially the oxidation products, were comprehensively investigated using gas chromatography–mass spectrometry (GC/MS) combined with silicane derivatization, high-performance liquid chromatography/hybrid quadrupole time-of-flight mass spectrometry (HPLC/Q-TOF-MS/MS), and ion chromatography (IC). Electron spin resonance (ESR) experiments, in combination with X-ray photoelectron spectroscopy (XPS) analysis of the TiO2, were used to study the role of reactive oxygen species in the degradation of CN-75. An oxidative degradation mechanism was proposed based on the findings. The results will be useful in developing methods for eliminating PCN-concentrated wastes.

Results

Kinetic study

The time-dependent degradation behavior of CN-75 (990.1 nmol) over anatase TiO2 at 300 °C was investigated. The black squares in Fig. 1 represent changes in the amount of residual CN-75 with heating time at 300 °C, calculated based on quasi-exponential decay. The amount of CN-75 decreased from 990.1 to 78.28 nmol in 60 min. This suggests that nanosized anatase TiO2 is an effective catalyst for CN-75 degradation. A linear ln(RCN-75/ICN-75) versus time plot corresponding to pseudo-first-order reaction kinetics with an initial rate constant kobs (min−1) of 0.04 was obtained as shown in the inset in Fig. 1(ICN-75 is the initial number of moles of CN-75, and RCN-75 is the number of moles of the remained CN-75 following heating for a given time period). It can be seen from Fig. 1 that only a small amount of 1,2,3,4,5,6,7-heptachloronaphthalene (CN-73) was detected in the hydrodechlorination products from 5 to 60 min. In contrast, in the progress of CN-75 degradation over as-prepared Fe3O4 with the similar dosage for the same reaction phases, a series of hydrodechlorination products from heptachloronaphthalenes to dichloronaphthalenes were detected19. The hydrodechlorination reaction of CN-75 was less favored on anatase TiO2 than on Fe3O4. This may be because the stability of anatase TiO2 is higher than that of Fe3O4, as shown by the higher EBG of anatase TiO2 (3.2 eV) compared with that of Fe3O4 (0.1 eV). Similarly, the weaker hydrodechlorination of decachlorobiphenyl was also found in its degradation over NiFe2O4 with EBG at 2.19 eV than over Fe3O420.

Figure 1. Contents of residual CN-75 and generated CN-73 as function of heating time.

Figure 1

Inset shows pseudo-first-order kinetic plot of the reaction.

GC/MS analysis of oxidation products after derivatization

Competition between hydrodechlorination and oxidation reactions in the degradation of chlorinated benzenes over metal oxides has often been reported9,12,20,21. The reason is that lower chlorinated products and oxidation products, such as phenolate, acetate, and carbon monoxide species, have been detected simultaneously9,22,23. This may be explained by different types of active centers on catalysts. One of the reactions will be the main process, depending on the reaction conditions and reactants. A low level of hydrodechlorination suggests that oxidative degradation occurs preferentially. The oxidation intermediate products formed during catalytic degradation of CN-75 were studied to obtain a better understanding of the degradation pathway. Theurich et al.24 reported that 15 different oxidation intermediates were identified during the photocatalytic degradation of naphthalene in aqueous suspensions of TiO2 under UV irradiation. To evaluate the existence of oxidative intermediates during the reaction, the dosage of CN-75 increased from 990.1 nmol to 4,950.5 nmol. GC/MS is often used to identify unknown substances. However, the response of the oxidative degradation products often with high polarity was poor in GC/MS. Silylation is one of the derivatization procedures widely used to improve GC behavior of polar compounds containing phenolic and carboxylic groups. In this procedure, the active hydrogens could be replaced by trimethylsilyl groups, producing derivatives which are more volatile and thermally stable. Albero et al.25 reported that phenolic and carboxylic compounds in soil, such as parabens, bisphenols and triclosan, were determinated by gas chromatography tandem mass spectrometry with in situ derivatization of N,O-bis(trimethylsilyl)trifluoroacetamide with 1% trimethylchlorosilane (BSTFA:TMCS = 99:1, v/v). Saitta et al.26 also demonstrated 21 phenolic compounds in Italian and Turkish pistachio oil samples by means of the mass spectra of the BSTFA-TMCS derivatives. In present study, the reaction products were derivatized using BSTFA:TMCS (99:1)27, and then analyzed using GC/MS in EI full-scan mode. The main derivatization reactions are as follows:

graphic file with name srep17800-m1.jpg
graphic file with name srep17800-m2.jpg

Figure 2 shows the GC/MS chromatograms of the chemically derivatized samples after CN-75 degradation over anatase TiO2 at 300 °C for 5 min. Analysis of the derivatized products showed that tetrachlorophenol, tetrachlorobenzoic acid, tetrachloroacroleinbenzoic acid, tetrachlorophthalaldehyde, tetrachlorosalicylic acid, and hexachloronaphthols were produced. The list of corresponding oxidation products is given in Table 1. Full-scan MS analysis was performed to identify the structures of the detected oxidation derivatives. Qualitative analysis was performed based on the molecular ions, fragment ions, the ratio between 35Cl and 37Cl, and comparison with data in the NIST02 standard spectral database28. As shown in Fig. 2, clear molecular ions and fragment ions were observed for seven derivatized products. For example, the mass spectrum corresponding to Peak P2 showed the presence of derivatized tetrachlorobenzoic acid. A clear molecular ion [M]+ at m/z 332, and fragmentation clusters at m/z 317 [M−CH3]+, 243 [M−OSi(CH3)3]+, 215 [M−COOSi(CH3)3]+, 178 [M−ClCOOSi(CH3)3]+, and 143 [M−2ClCOOSi(CH3)3]+were observed. The isotope distributions fit a four Cl atom profile (the ratio of the peaks at m/z 332 and 334 was 1:1.3). This information clearly identifies the product as tetrachlorobenzoic acid29. The mass spectrum corresponding to Peak P4 showed the presence of tetrachlorophthalaldehyde30. The mass spectrum showed a molecular ion [M]+ at m/z 272 and a fragmentation cluster at m/z 243 [M−CHO]+. The mass spectrum of Peak P7, corresponding to the derivative of hexachloronaphthol, showed a molecular ion [M]+ at m/z 422 and typical fragmentation clusters at m/z 407 [M−CH3]+ and 372 [M−ClCH3]+. The identification of naphthalene rings and single benzene rings with –OH, –COOH, and –CHO substituents confirmed that oxidation reactions occurred. The presence of oxidation intermediates containing single benzene rings indicated partial splitting of the naphthalene rings during the oxidative degradation reaction. In contrast, the oxidative products only with naphthalene-ring, i.e., tetrachloronaphthols and dihydrodiol, have been determined by GC-MS during the biodegradation of 1,4-dichloronaphthalene31.

Figure 2. GC/MS chromatograms of derivatized products of CN-75 degradation over anatase TiO2 at 300 °C for 5 min.

Figure 2

Table 1. Oxidative products following degradation of CN-75 over anatase TiO2 at 300 °C for 5 min, determined by GC/MS after the derivatization.

graphic file with name srep17800-t1.jpg

HPLC/Q-TOF-MS/MS analysis of oxidation products

LC/MS is a sensitive analytical technique that is widely used for the separation and quantification of highly polar products32,33. During the analytical process, polar oxidation products are efficiently ionized using the ionization techniques associated with LC/MS, enabling their identification34. This technique has been often applied together with GC/MS to comprehensively determine the polar species35. The oxidation process was therefore further investigated by monitoring the formation of oxidation intermediate products during the catalytic degradation of CN-75 over anatase TiO2 using HPLC/Q-TOF-MS/MS. Figure 3a shows the HPLC results for the chemical species following reaction between CN-75 (4,950.5 nmol) and anatase TiO2 (50 mg) at 300 °C for 5 min. Tetra-chlorophenols, tetrachlorobenzenedihydrodiol, hydroxypentachloronaphthalenedione (OH-PeCN-dione), hydroxypentachloronaphthalene (OH-PeCN), and hydroxyhexachloronaphthalene (OH-HxCN) were determined as degradation products (Table 2). However, the isomer patterns of the hydroxyl congeners could not be identified because of limitations associated with the external standards. These results further show that oxidation intermediates with naphthalene rings and single benzene rings were produced during the oxidative degradation reaction. However, the only hydroxyl-oxidative products with naphthalene ring, i.e. hydroxyl-trichloronaphthalene (TrCN), -tetrachloronaphthalene (TeCN), -PeCN, and -HxCN have been determined by HPLC/Q-TOF-MS/MS during the degradation of CN-75 on Fe3O419. This suggests the occurrence of deep oxidative degradation of CN-75 on anatase-type TiO2.

Figure 3.

Figure 3

(a) HPLC spectrum of chemical species obtained by degradation of CN-75 over anatase TiO2 at 300 °C for 5 min and (b) distribution profiles of organic acids formed during degradation of CN-75 over anatase TiO2 at 300 °C.

Table 2. Oxidative products following degradation of CN-75 over anatase TiO2 at 300 °C for 5 min, determined by HPLC/Q-TOF-MS/MS.

graphic file with name srep17800-t2.jpg

Analysis of oxidation products by IC

Literature reports have indicated that chlorinated aromatic compounds containing hydroxyl, aldehyde, and carboxyl groups can be easily ring-cracked to smaller organic molecules such as formate and acetate11. Ma et al.36 detected the formation of surface formate species using in situ FTIR spectroscopy in low-temperature 1,2-dichlorobenzene oxidation over water-resistant Fe–Ca–Ox/TiO2 catalysts. Similar results were reported for the catalytic oxidation of 1,2-dichlorobenzene over Ca-doped FeOx hollow microspheres37. Formic, acetic, and propanoic acids have been detected during degradation of decachlorobiphenyl over Fe3O419. In the current study, ring-cracked products were detected, using IC, in the reaction between CN-75 (990.1 nmol) and anatase TiO2 (50 mg) at 300 °C. Formic and acetic acids were the main ring-cracked degradation products, as shown in Fig. 3b. The amount of acetic acid rapidly increased to a maximum of 140.6 nmol after heating for about 10 min, and then decreased with heating time. In contrast, the formic acid content increased steadily with heating time, with a maximum content of 90.4 nmol at 60 min. These oxidation products indicate that TiO2 also facilitates the ring-cracking oxidation pathway of chlorinated aromatics.

Discussion

The presence of active oxygen species on nanosized anatase TiO2 catalysts is believed to contribute to the occurrence of oxidation reactions during CN-75 degradation19. The O 1s XP spectrum of the anatase TiO2 catalyst is shown in Fig. 4a. The peak at 530.97 eV (denoted by P1) is attributed to surface oxygen and adsorbed oxygen species, and the peak located at 529.15 eV (denoted by P2) is attributed to lattice oxygen38. Similar oxygen species were detected on the surface of Ca-doped FeOx hollow microspheres and CaCO3/α-Fe2O3 composite catalysts37. A high proportion of surface oxygen on the metal oxide catalyst increases the activity in low-temperature oxidative degradation of 1,2-dichlorobenzene.

Figure 4.

Figure 4

(a) O 1s XPS spectrum of TiO2 catalyst, (b) ESR spectra of O2−• (I) and •OH (II) generated by reaction of anatase TiO2 and CN-75 at 300 °C for 10 min, (c) XRD pattern of TiO2 catalyst and (d) Cl 2p XPS spectrum of TiO2 sample after the reaction for 10 min.

Reactive oxygen species such as O2−• and •OH are strong electrophilic oxidants. They can attack organic substrates, leading to their degradation and ultimately to their total mineralization to CO2 and H2O39,40. Its role in a range of photocatalytic oxidative degradation reactions, including those of pathogenic bacteria over NiO/SrBi2O441, rhodamine B over TiO242, and azo dyes over Ag/AgBr/TiO43, have been confirmed by ESR spectroscopy. ESR spectroscopy, with DMPO as the spin-trapping agent, was used to obtain information on the active radicals involved, to determine whether O2−• and •OH were available products in the decomposition of CN-75 over anatase TiO2. A reaction was performed between anatase TiO2 (50 mg) and CN-75 (990.1 nmol) at 300 °C for 10 min. The reaction products were immediately dissolved in dimethyl sulfoxide (DMSO), and then characterized using an ESR analyzer, as shown in Fig. 4b. Four peaks were observed, and the hyperfine constants, i.e., αN = 12.7429 G, αH = 10.0304 G, and g = 2.0103, coincided with those previously reported for DMPO–O2−• (Fig. 4b-I)9. The results identify that the superoxide anion may be involved in CN-75 degradation, resulting in the formation of a series of oxidation products and perhaps even into formic acid and acetic acid. The DMPO–∙OH species were examined under identical conditions, except water was used as the solvent instead of DMSO. No obvious signal was observed, as shown in Fig. 4b-II. This differs from the photocatalytic degradation of many organic molecules, in which ∙OH species are often identified41,42,43.

An oxidative degradation pathway (Fig. 5) is proposed, based on the available oxygen species and the detected oxidation intermediates. The (101) surface is the most stable and frequent surface of anatase TiO2, as shown in Fig. 4c, which was therefore selectively took as a model44,45,46. It has the same periodicity as the bulk truncated surface and exposes undercoordinated pentacoordinated Ti cations (Ti5c) and dicoordinated oxygen anions (O2c), and fully coordinated Ti6c cations and tricoordinated oxygen anions (O3c)47. Coordination theory states that unsaturated ions are prone to bond with ligands23. It is therefore hypothesized that CN-75 molecules are adsorbed on the anatase TiO2 surface via coordination interactions between Lewis acid Ti5c cations and Lewis base Cl21. When CN-75 degraded on the surface of the anatase TiO2 catalyst, firstly, dissociative adsorption of CN-75 on the central Ti5c cations occurs, followed by the attack of carbon atom potential to accepting the electrons by reactive nucleophilic oxygen O2− species. This results in C–Cl bond cleavage and subsequent Ti–Cl bond formation. Association of the free chloride ions with Lewis acid Ti ions occurs during CN-75 degradation over anatase TiO2. This is confirmed by the Cl 2p core-level XP spectrum of the catalyst after heating for 10 min (Fig. 4d). Three peaks (denoted by P1, P2, and P3) are observed. The peak at 197.8 eV corresponds to Cl bonded to Ti4+, with a net charge of −1, indicating possible formation of TiCl4 during degradation of CN-7548. In this reaction pathway, OH-HxCN and OH-PeCN can be formed via nucleophilic attack by basic O2−. Further nucleophilic attack can occur at other positions on PCNs, forming species such as OH-PeCN-dione. The formation of naphthol species was detected during photocatalytic degradation of naphthalene over TiO224.

Figure 5. Possible degradation pathways of CN-75 over anatase TiO2.

Figure 5

Superoxide O2−• species are electrophilic. They have been reported to be formed by transformation of adsorbed O2 molecules19,49. When a subsurface oxygen vacancy is present, it is energetically favorable for O2 to adsorb at a Ti5c site close to this defect. On adsorption, the extra charge associated with the defect is transferred to the O2 molecule, converting it to a superoxide O2−• species. The strongly reactive electrophilic O2−• species can attack the π-electron cloud of the naphthalene ring, which has a highly dense electron population. This leads to the cracking of the naphthalene ring at different positions. The detection of tetrachlorophenol and the resultant tetrachlorobenzenedihydrodiol indicates that one of the rings in the naphthalene ring of CN-75 is first opened through C1–C10 and C4–C9 bond cleavage. Breakage of the C1–C2 and C4–C9 bonds in one ring could result in the formation of tetrachlorobenzoic acid, which is further oxidized to tetrachlorosalicylic acid. Similarly, the breakage of C1–C2 or and C3–C4 bonds could lead to the formation of tetrachloroacroleinbenzoic acid or and tetrachlorophthalaldehyde, respectively. These results show that lateral cleavage of one naphthalene ring at different C–C bond positions by electrophilic O2−• could occur, leading to formation of various single-benzene-ring oxidation products. Unexpectedly, the symmetric half section of CN-75 could be retained along with generation of complex oxidation products containing the tetrachlorobenzene structure.

It is important to note that the reaction pathways via electrophilic and nucleophilic attack by reactive oxygen species such as O2− and O2−• are not independent of each other. The newly formed chlorinated naphthol species can also be attacked by reactive oxygen species such as O2−•. Moreover, oxidation products with both naphthalene and single benzene rings can be further attacked by reactive oxygen species, and completely cracked to small molecules such as formic and acetic acids. A wide range of oxidation products such as naphthols, phenols, hydroxy-diones, benzoic acids, acroleinbenzoic acid, phthalaldehyde, salicylic acid, dihydrodiols, and formic and acetic acids, with chlorinated naphthalene or benzene rings, or without aromatic rings, were detected during the degradation of CN-75 over anatase TiO2. This is different from the previously reported results for CN-75 degradation over Fe3O4 micro/nanomaterials19, in which only chloronaphthol species, and formic acid and acetic acids were detected as the oxidation products under the same experimental conditions. This shows that oxidative degradation of CN-75 on anatase TiO2 was more extensive that on Fe3O4 micro/nanomaterials. Deep oxidative degradation of CN-75 on anatase TiO2 occurs possibly because of the electronic structure with an EBG of 3.2 eV and the reactive oxygen species on its surface.

Methods

Chemical reagents

Anatase TiO2 (nanopowder, diameter <25 nm) was supplied by Sigma-Aldrich (USA). CN-75 (Supelco, USA) was laboratory analytical grade and used without further purification. HPLC-grade ethyl acetate was purchased from Fisher Scientific (Geel, Belgium). Chromatographic-grade methanol, acetonitrile, and hexane were purchased from Dika Technologies (Lake Forest, CA, USA). Derivatization reagents, BSTFA:TMCS = 99:1 were supplied by Supelco (USA).

Degradation experiments

Degradation experiments were performed in sealed glass ampoules. Prior to the reaction, a hexane solution of CN-75 (990.1 or 4,950.5 nmol) was injected into an ampoule and subsequently evaporated to dryness at room temperature, and then mixed with later added 50 mg of anatase TiO2. The samples were heated at 300 °C for an appropriate time. A blank experiment was performed in the absence of TiO2 under the same conditions. All experiments were performed in triplicate to ensure repeatability of the results.

Degradation product analysis

After the decomposition reaction, the ampoule was cooled to room temperature and crushed, and the sample was extracted. The unreacted CN-75 and newly formed PCNs were analyzed using an Agilent 6890 gas chromatograph equipped with a DB-5 MS capillary column (30 m × 0.25 mm i.d., 0.25 μm film thickness) and an Agilent 5973 N mass selective detector. Helium (≥99.999%) at a flow rate of 1 mL/min was used as the carrier gas, and the injector was set at 260 °C. The column temperature was set at 75 °C for 2 min, gradually increased to 150 °C at 20 °C/min, then increased to 205 °C at 1.5 °C/min, and finally increased to 270 °C at 2.5 °C/min. The diluted sample (1.0 μL) was injected in split-less mode. An electron ionization system with an ionization energy of 70 eV was used.

For oxidation product analysis, the reaction products obtained after CN-75 degradation over anatase TiO2 at 300 °C for 5 min were extracted by the mixture solvent of hexane/ methanol/ ethyl acetate (1:1:1, v/v/v). The extract was dehydrated using a column packed with anhydrous sodium sulfate, and then evaporated under stream of nitrogen to dryness. Dry residue was dissolved in 0.2 mL of derivatizing reagent BSTFA:TMCS (99:1) and vortexed. The mixture reacted at room temperature for 60 min, and the derivatization products were analyzed using GC/MS. The column temperature was initially 50 °C, and increased to 180 °C (for 2 min) at 10 °C/min, to 210 °C at 1 °C/min, and to 280 °C at 10 °C/min. The carrier gas was helium at a flow rate of 1 mL/min.

The oxidation products were also analyzed using HPLC/Q-TOF-MS/MS (Micromass Q-TOF micro, Waters, USA). After the degradation of CN-75 (4,950.5 nmol) over anatase TiO2, product samples were extracted using HPLC-grade methanol, filtered through a 0.45 μm mesh membrane, and concentrated to approximately 100 μL. The oxidation products were detected using a Supelcosiltmlc-18 C18 column (Sigma; 4.6 mm × 250 mm; 5 μm particle size). The elution flow rate was 0.5 mL/min with a gradient of 0.1% acetic acid in water–acetonitrile [acetonitrile concentrations 0% (isocratic, 5 min), 70% (isocratic, 5 min), 70–90% (linear, 5 min), 90–100% (linear, 5 min), 100% (isocratic, 5 min), and 0% (isocratic, 4 min)]. MS was performed using a Waters Micromass Quattro Premier XE (triple-quadrupole) detector, equipped with an electrospray ionization (EI) source (Micromass, USA). The mass analyzer was operated in negative ionization (EI) mode and the optimized parameters were source temperature 120 °C, desolvation temperature 200 °C, capillary voltage 2.50 kV, desolvation gas flow rate 600 L/h, and cone gas flow rate 50 L/h.

The organic acid oxidation products such as acetic and formic acid were analyzed using IC. The degradation samples obtained from the reaction of CN-75 (990.1 nmol) and anatase TiO2 (50 mg) were extracted three times with 15 ml deionized water for 10 min each time under ultrasonication. And then the combined extracts were filtered through a 0.45 μm mesh membrane for IC measurements. The employed IC was a DIONEX AS 5000 instrument equipped with an AS-AP automated sampler. A Dionex AS11-HC guard column (50 × 4 mm i.d.) and a Dionex AS11-HC analytical column (250 × 4 mm i.d.) were used for the analyses. The analyses were performed at 30 °C with a potassium hydroxide eluent that was generated from a Dionex EG on line and run with a linear gradient at a flow rate of 1.0 mL min−1.

XPS and ESR

The surface element oxidation states of the TiO2 catalyst, which reacted with CN-75 at 300 °C for 10 min, were investigated using XPS (Escalab 250), with monochromated Al Kα (1,486.6 eV) radiation (200 W, 200 eV) as the X-ray source. The operating pressure was ~1 × 10−8 Torr.

The radical species formed during degradation were investigated using ESR spectroscopy (ESP 300 E electron paramagnetic resonance spectrometer, Bruker) with 5,5-dimethyl-1-pyrroline N-oxide (DMPO; Sigma Chemical Co.) as the spin-trapping agent. Typically, anatase TiO2 (50 mg) and CN-75 (990.1 nmol) reacted at 300 °C for 10 min. A reaction using anatase TiO2 but without CN-75 was also examined under the same conditions for comparison. The settings for the ESR spectrometer were center field, 3,485 G; sweep width, 100.0 G; microwave frequency, 9.8 GHz; and power, 10 mW.

Additional Information

How to cite this article: Su, G. et al. Thermal catalytic oxidation of octachloronaphthalene over anatase TiO2 nanomaterial and its hypothesized mechanism. Sci. Rep. 5, 17800; doi: 10.1038/srep17800 (2015).

Acknowledgments

This study was supported by the National 973 Program (2015CB453103), the Strategic Priority Research Program of the Chinese Academy of Sciences (XDB14020102), and the National Natural Science Foundation of China (21377147, 21177141, 21321004).

Footnotes

Author Contributions G.S. and M.Z. conceived and designed the experiments. L.Z., Q. L. and L.Y. conducted the experimental parts. G.S., Q.L. and H.L. analyzed the data and wrote the manuscript. G.S., Q.L., L.H. and M.Z. reviewed the literature and checked the data. All authors have reviewed the manuscript.

References

  1. Seventh meeting of the Conference of the Parties to the Stockholm Convention. (2015) Available at: http://synergies.pops.int/2015COPs/MeetingDocuments/tabid/4243/language/en-US/Default.aspx. (Accessed: 4th July 2015).
  2. Haglund P. Determination of polychlorinated naphthalenes in polychlorinated biphenyl products via capillary gas chromatography-mass spectrometry after separation by gel permeation chromatography. J. Chromatogr. 634, 79–86 (1993). [Google Scholar]
  3. Seventh meeting of the Persistent Organic Pollutants Review Committee (POPRC7). (2011) Available at: http://chm.pops.int/Convention/POPsReview Committee/POPRCMeetings/POPRC7/POPRC7Documents/tabid/2267/language/en-US/Default.aspx. (Accessed: 5th July 2015).
  4. Yamashita N., Kannan K., Imagawa T., Miyazaki A. & Giesy J. P. Concentrations and Profiles of Polychlorinated Naphthalene Congeners in Eighteen Technical Polychlorinated Biphenyl Preparations. Environ. Sci. Technol. 34, 4236–4241 (2000). [Google Scholar]
  5. Abad E., Caixach J. & Rivera J. Dioxin like compounds from municipal waste incinerator emission: assessment of the presence of polychlorinated naphthalenes. Chemosphere 38, 109–120 (1999). [DOI] [PubMed] [Google Scholar]
  6. Imagawa T. & Lee C. W. Correlation of polychlorinated naphthalenes with polychlorinated dibenzofurans formed from waste incineration. Chemosphere 44, 1511–1520 (2001). [DOI] [PubMed] [Google Scholar]
  7. Crookes M. J. & Howe P. D. Environmental hazard assessment: Ethyl benzene. In: Toxic Substances Division, Directorate for Air, Climate and Toxic Substances. Department of the Environment, Watford, U.K. (1993). [Google Scholar]
  8. Ninth meeting of the Persistent Organic Pollutants Review Committee (POPRC9). (2013) Available at: http://chm.pops.int/TheConvention/POPsReviewCommittee/Meetings/POPRC9/Documents/tabid/3281/Default.aspx. (Accessed: 1th July 2015).
  9. Lin S. J. et al. The degradation of 1,2,4-trichlorobenzene using synthesized Co3O4 and the hypothesized mechanism. J. Hazard. Mater. 192, 1697–1704 (2011). [DOI] [PubMed] [Google Scholar]
  10. Jia M. K., Su G. J., Zheng M. H., Zhang B. & Lin S. J. Development of self-assembled 3D FexOy micro/nano materials for application in hexachlorobenzene degradation. J. Nanosci. Nanotechnol. 11, 2100–2106 (2011). [DOI] [PubMed] [Google Scholar]
  11. Krishnamoorthy S., Rivas J. A. & Amiridis M. D. Catalytic oxidation of 1,2-dichlorobenzene over supported transition metal oxides. J. Catal. 193, 264–272 (2000). [Google Scholar]
  12. Lichtenberger J. & Amiridis M. D. Catalytic oxidation of chlorinated benzenes over V2O5/TiO2 catalysts. J. Catal. 223, 296–308 (2004). [Google Scholar]
  13. Wang J. et al. Catalytic oxidation of chlorinated benzenes over V2O5/TiO2 catalysts: The effects of chlorine substituents. Catal. Today 241, 92–99 (2015). [Google Scholar]
  14. Weber R. & Sakurai T. Low temperature decomposition of PCB by TiO2-based V2O5/WO3 catalyst: evaluation of the relevance of PCDF formation and insights into the first step of oxidative destruction of chlorinated aromatics. Appl. Catal. B 34, 113–127 (2001). [Google Scholar]
  15. Gannoun C. et al. Elaboration and characterization of sulfated and unsulfated V2O5/TiO2 nanotubes catalysts for chlorobenzene total oxidation. Appl. Catal. B 147, 58–64 (2014). [Google Scholar]
  16. Fujishima A. & Kudo K. Electrochemical photolysis of water at a semiconductor electrode. Nature 238, 37–38 (1972). [DOI] [PubMed] [Google Scholar]
  17. Linsebigler A. L., Lu G. Q. & Jr J. T. Y. Photocatalysis on TiO2 Surfaces: principles, mechanisms, and selected results. Chem. Rev. 95, 735–758 (1995). [Google Scholar]
  18. Obata K. et al. Photocatalytic decomposition of NH3 over TiO2 catalysts doped with Fe. Appl.Catal. B 160-161, 200–203 (2014). [Google Scholar]
  19. Su G. J. et al. Thermal degradation of octachloronaphthalene over as-prepared Fe3O4 micro/nanomaterial and its hypothesized mechanism. Environ. Sci. Technol. 48, 6899–6908 (2014). [DOI] [PubMed] [Google Scholar]
  20. Huang L. Y. et al. Degradation of polychlorinated biphenyls using mesoporous iron-based spinels. J. Hazard. Mater. 261, 451–462 (2013). [DOI] [PubMed] [Google Scholar]
  21. Zhang L. F., Zheng M. H., Liu W. B., Zhang B. & Su G. J. A method for decomposition of hexachlorobenzene by γ-alumina. J. Hazard. Mater. 150, 831–834 (2008). [DOI] [PubMed] [Google Scholar]
  22. Ma X. D., Sun H. W., He H. & Zheng M. H. Competitive reaction during decomposition of hexachlorobenzene over ultrafine Ca–Fe composite oxide catalyst. Catal. Lett. 119, 142–147 (2007). [Google Scholar]
  23. Ma X. D. et al. Synergic effect of calcium oxide and iron (III) oxide on the dechlorination of hexachlorobenzene. Chemosphere 60, 796–801 (2005). [DOI] [PubMed] [Google Scholar]
  24. Theurich J. & Bahnemann D. W. Photocatalytic degradation of naphthalene and anthracene: GC-MS analysis of the degradation pathway. Res. Chem. lntermed. 23, 247–274 (1997). [Google Scholar]
  25. Albero B., Sánchez-Brunete C., Miguel E., Pérez R. A. & Tadeo J. L. Determination of selected organic contaminants in soil by pressurized liquid extraction and gas chromatography tandem mass spectrometry with in situ derivatization. J. Chromatogr. A 1248, 9–17 (2012). [DOI] [PubMed] [Google Scholar]
  26. Saitta M., La Torre G. L., Potortì A. G., Di Bella G. & Dugo G. Polyphenols of pistachio (pistacia vera L.) oil samples and geographical differentiation by principal component analysis. J. Am. Oil Chem. Soc. 91, 1595–1603 (2014). [Google Scholar]
  27. Wenclawiak B. W., Jensen T. E. & Richert J. F. O. GC/MS-FID analysis of BSTFA derivatized polar components of diesel particulate matter (NBS SRM-1650) extract. Fresenius J. Anal. Chem. 346, 808–812 (1993). [Google Scholar]
  28. Su G. J. et al. Synergetic effect of alkaline earth metal oxides and iron oxides on the degradation of hexachlorobenzene and its degradation pathway. Chemosphere 90, 103–111 (2013). [DOI] [PubMed] [Google Scholar]
  29. Wasada N. Spectral Database for Organic Compounds SDBS. (2015) Available at: http://sdbs.db.aist.go.jp/sdbs/cgi-bin/direct_frame_disp.cgi?sdbsno=51814. (Accessed: 25th Aprel 2015).
  30. Oh J. A. & Shin H. S. Determination of ortho-phthalaldehyde in water by high performance liquid chromatography and gas chromatography-mass spectrometry after hydrazine derivatization. J. Chromatogr. A 1247, 99–103 (2012). [DOI] [PubMed] [Google Scholar]
  31. Mori T., Nakamura K. & Kondo R. Fungal hydroxylation of polychlorinated naphthalenes with chlorine migration by wood rotting fungi. Chemosphere 77, 1230–1235 (2009). [DOI] [PubMed] [Google Scholar]
  32. Erratico C. A., Szeitz A. & Bandiera S. M. Validation of a novel in vitro assay using ultra performance liquid chromatography-mass spectrometry (UPLC/MS) to detect and quantify hydroxylated metabolites of BDE-99 in rat liver microsomes. J. Chromatogr. B 878, 1562–1568 (2010). [DOI] [PubMed] [Google Scholar]
  33. Mas S. et al. Comprehensive liquid chromatography-ion-spray tandem mass spectrometry method for the identification and quantification of eight hydroxylated brominated diphenyl ethers in environmental matrices. J. Mass Spectrom. 42, 890–899 (2007). [DOI] [PubMed] [Google Scholar]
  34. Sun J. T. et al. Sample preparation method for the speciation of polybrominated diphenyl ethers and their methoxylated and hydroxylated analogues in diverse environmental matrices. Talanta 88, 669–676 (2012). [DOI] [PubMed] [Google Scholar]
  35. Bian W. J., Song X. H., Liu D. Q., Zhang J. & Chen X. H. The intermediate products in the degradation of 4-chlorophenol by pulsed high voltage discharge in water. J. Hazard. Mater. 192, 1330–1339 (2011). [DOI] [PubMed] [Google Scholar]
  36. Ma X. D. et al. Water-resistant Fe–Ca–Ox/TiO2 catalysts for low temperature 1,2-dichlorobenzene oxidation. Appl. Catal. A 466, 68–76 (2013). [Google Scholar]
  37. Ma X. D. et al. Catalytic oxidation of 1,2-dichlorobenzene over Ca-doped FeOx hollow microspheres. Appl. Catal. B 147, 666–676 (2014). [Google Scholar]
  38. Wang X. Y., Kang Q. & Li D. Catalytic combustion of chlorobenzene over MnOx–CeO2 mixed oxide catalysts. Appl. Catal. B 86, 166–175 (2009). [Google Scholar]
  39. Lai T. L., Lai Y. L., Lee C. C., Shu Y. Y. & Wang C. B. Microwave-assisted rapid fabrication of Co3O4 nanorods and application to the degradation of phenol. Catal. Today 131, 105–110 (2008). [Google Scholar]
  40. Hoffmann M. R., Martin S. T., Choi W. & Bahnemannt D. W. Environmental applications of semiconductor photocatalysis. Chem. Rev. 95, 69–96 (1995). [Google Scholar]
  41. Hu C., Hu X. X., Guo J. & Qu J. H. Efficient destruction of pathogenic bacteria with NiO/SrBi2O4 under Visible Light Irradiation. Environ. Sci. Technol. 40, 5508–5513 (2006). [DOI] [PubMed] [Google Scholar]
  42. Zhao J. C. et al. Photoassisted degradation of dye pollutants. 3. degradation of the cationic dye rhodamine B in aqueous anionic surfactant/TiO2 dispersions under visible light irradiation: evidence for the need of substrate adsorption on TiO2 particles. Environ. Sci. Technol. 32, 2394–2400 (1998). [Google Scholar]
  43. Hu C., Lan Y. Q., Qu J. H., Hu X. X. & Wang A. M. Ag/AgBr/TiO2 visible light photocatalyst for destruction of azodyes and bacteria. J. Phys. Chem. B 110, 4066–4072 (2006). [DOI] [PubMed] [Google Scholar]
  44. Liu L. L. et al. O2 adsorption and dissociation on a hydrogenated anatase (101) Surface. J. Phys. Chem. C 118, 3471–3482 (2014). [Google Scholar]
  45. Aschauer U. & Selloni A. Hydrogen interaction with the anatase TiO2 (101) surface. Phys. Chem. Chem. Phys. 14, 16595–16602 (2012). [DOI] [PubMed] [Google Scholar]
  46. Islam M. M., Calatayud M. & Pacchioni G. Hydrogen adsorption and diffusion on the anatase TiO2(101) surface: a first-principles investigation. J. Phys. Chem. C 115, 6809–6814 (2011). [Google Scholar]
  47. Li Y. F., Aschauer U., Chen J. & Selloni A. Adsorption and reactions of O2 on anatase TiO2. Acc. Chem. Res. 47, 3361–3368 (2014). [DOI] [PubMed] [Google Scholar]
  48. Fleutot S., Dupin J. C., Renaudin G. & Martinez H. Intercalation and grafting of benzene derivatives into zinc-aluminum and copper-chromium layered double hydroxide hosts: an XPS monitoring study. Phys. Chem. Chem. Phys. 13, 17564–17578 (2011). [DOI] [PubMed] [Google Scholar]
  49. Li Y. F. & Selloni A. Theoretical study of interfacial electron transfer from reduced anatase TiO2 (101) to adsorbed O2. J. Am. Chem. Soc. 135, 9195–9199 (2013). [DOI] [PubMed] [Google Scholar]

Articles from Scientific Reports are provided here courtesy of Nature Publishing Group

RESOURCES