Acessibilidade / Reportar erro

Titanate Nanotubes as New Nanostrutured Catalyst for Depolymerization of PET by Glycolysis Reaction

Abstract

The final destination of PET packaging is creating economic and environmental concerns. One of the alternatives to minimize this problem would be making use of chemical recycling of this material through glycolysis with the aim to produce bis(hydroxiethyl) terephthalate, BHET monomer. This reaction is well known, but it still presents problems as BHET purity since it makes necessary the development of new catalysts highly selective. In this context, the present work studied the catalytic activity of a nanostructured material, titanate nanotubes (TNT), and compared it to a commercial catalyst (zinc acetate), which is the most used for this glycolysis reaction according to literature researches, and analyzed the influence of PET type (virgin and post-consumer) in the depolymerization for reaction times of 2, 3 and 4 hours. Using TNT as catalyst, BHET production yield and values of turnover number for the evaluated reaction times were higher than the results using Zn(OAc)2 for virgin PET, proving itself as a promising catalyst.

Keywords:
chemical recycling; glycolysis; nanostructure catalysts; titanate nanotubes


1. Introduction

Poly(ethylene terephthalate), PET, is a commodity polymer of great importance due to its excellent physical and chemical properties11 Eshaq G, Elmetwally AE. (Mg-Zn)-Al layered double hydroxide as a regenerable catalyst for the catalytic glycolysis of polyethylene terephthalate. Journal of Molecular Liquids. 2016;214:1-6.. PET also offers excellent barrier property and low weight compared to glass packaging22 Welle F. Twenty years of PET bottle to bottle recycling--An overview. Resources, Conservation and Recycling. 2011;55(11):865-875.. Therefore, this polymer is especially used in the packing industry. However, PET has a high global production and, consequently, consumption and non-biodegradability makes the disposal of post-consumer PET waste a problem as far as economic and environmental considerations are concerned. Among the various possibilities to mitigate this problem, the chemical recycling process would be the best alternative for this material. This process is in accordance with the principles of sustainable development because it leads to the formation of raw materials (monomers) used for producing PET33 Bartolome L, Imran M, Cho BG, Al-Masry WA, Kim DH. Recent Developments in the Chemical Recycling of PET. In: Achilias D, ed. Material Recycling - Trends and Perspectives. Rijeka: InTech; 2012. p. 65-84.. Chemical recycling by glycolysis is one of de most used recycling types. It can be described as a depolymerization process by transesterification between PET ester groups and a diol (generally ethylene glycol, EG)44 Ghaemy M, Mossaddegh K. Depolymerisation of poly(ethylene terephthalate) fibre wastes using ethylene glycol. Polymer Degradation and Stability. 2005;90(3):570-576.,55 Yue QF, Wang CX, Zhang LN, Ni Y, Jin YX. Glycolysis of poly(ethylene terephthalate) (PET) using basic ionic liquids as catalysts. Polymer Degradation and Stability. 2011;96(4):399-403., allowing obtaining of bis(hydroxyethyl) terephthalate (BHET)66 López-Fonseca R, Duque-Ingunza I, de Rivas B, Arnaiz S, Gutiérrez-Ortiz JI. Chemical recycling of post-consumer PET wastes by glycolysis in the presence of metal salts. Polymer Degradation and Stability. 2010;95(6):1022-1028. as main depolymerization product (Figure 1).

Figure 1
Reaction of PET depolymerization by glycolysis.

This reaction depends on some variables such as temperature, molar ratio EG/PET, glycolysis time and amount of catalyst33 Bartolome L, Imran M, Cho BG, Al-Masry WA, Kim DH. Recent Developments in the Chemical Recycling of PET. In: Achilias D, ed. Material Recycling - Trends and Perspectives. Rijeka: InTech; 2012. p. 65-84.. Among these variables, the catalyst type can dramatically influence the reaction conditions such as low temperature and reaction time. Recently various transition metal-based catalysts with good catalytic activities on PET glycolysis have been described in literature, the most used among them being zinc acetate77 Chen JW, Chen LW. The glycolysis of poly (ethylene terephthalate). Journal of Applied Polymer Science. 1999;73(1):35-40.,88 Troev K, Grancharov G, Tsevi R, Gitsov I. A novel catalyst for the glycolysis of poly(ethylene terephthalate). Journal of Applied Polymer Science. 2003;90(4):1148-1152., but they have limitations such as difficult synthesis of the catalyst11 Eshaq G, Elmetwally AE. (Mg-Zn)-Al layered double hydroxide as a regenerable catalyst for the catalytic glycolysis of polyethylene terephthalate. Journal of Molecular Liquids. 2016;214:1-6.,99 Wang S, Wang C, Wang H, Chen X, Wang S. Sodium titanium tris(glycolate) as a catalyst for the chemical recycling of poly(ethylene terephthalate) via glycolysis and repolycondensation. Polymer Degradation and Stability. 2015;114:105-114. or catalyst toxicity or corrosivity, resulting in pollution1010 Al-Sabagh AM, Yehia FZ, Eissa AMF, Moustafa ME, Eshaq G, Rabie AM, et al. Cu- and Zn-acetate-containing ionic liquids as catalysts for the glycolysis of poly(ethylene terephthalate). Polymer Degradation and Stability. 2014;110:364-377.. Thus, the search for new highly selective catalysts under mild conditions is required33 Bartolome L, Imran M, Cho BG, Al-Masry WA, Kim DH. Recent Developments in the Chemical Recycling of PET. In: Achilias D, ed. Material Recycling - Trends and Perspectives. Rijeka: InTech; 2012. p. 65-84..

Among the materials with potential application for catalysis are nanomaterials, such as nanotubes, nanoribbons, nanowires that exhibit unique properties, such as, changes in chemical reactivity and electrical conductivity1111 Ferreira HS, Rangel M do C. Nanotecnologia: aspectos gerais e potencial de aplicação em catálise. Química Nova. 2009;32(7):1860-1870.,1212 Galetti AE, Barroso MN, Monzón A, Abello MC. Synthesis of Nickel Nanoparticles Supported on Carbon Using a Filter Paper as Biomorphic Pattern for Application in Catalysis. Materials Research. 2015;18(6):1278-1283., which, combined, are not found in conventional materials1313 Camargo PHC, Satyanarayana KG, Wypych F. Nanocomposites: synthesis, structure, properties and new application opportunities. Materials Research. 2009;12(1):1-39.. Recent studies showed that titanate nanotubes have Brönsted and Lewis acid sites1414 Camposeco R, Castillo S, Mejia-Centeno I, Navarrete J, Rodriguez-Gonzalez V. Behavior of Lewis and Brönsted surface acidity featured by Ag, Au, Ce, La, Fe, Mn, Pd, Pt, V and W decorated on protonated titanate nanotubes. Microporous and Mesoporous Materials. 2016;236:235-243. formed from lattice distortion due to the scrolling of titanate nanotubes layers. This property makes this nanomaterial an effective catalysts in various applications as in CO2 conversion1515 Parayil SK, Razzaq A, Park SM, Kim HR, Grimes CA, In S Il. Photocatalytic conversion of CO2 to hydrocarbon fuel using carbon and nitrogen co-doped sodium titanate nanotubes. Applied Catalysis A: General. 2015;498:205-213.,1616 Monteiro WF, Vieira MO, Aquino AS, de Souza MO, de Lima J, Einloft S, et al. CO2 conversion to propylene carbonate catalyzed by ionic liquid containing organosilane groups supported on titanate nanotubes/nanowires. Applied Catalysis A: General. 2017;544:46-54., biodiesel synthesis1717 Hernández-Hipólito P, García-Castillejos M, Martínez-Klimova E, Juárez-Flores N, Gómez-Cortés A, Klimova TE. Biodiesel production with nanotubular sodium titanate as a catalyst. Catalysis Today. 2014;220-222:4-11.

18 Hernández-Hipólito P, Juárez-Flores N, Martínez-Klimova E, Gómez-Cortés A, Bokhimi X, Escobar-Alarcón L, et al. Novel heterogeneous basic catalysts for biodiesel production: Sodium titanate nanotubes doped with potassium. Catalysis Today. 2015;250:187-196.
-1919 Salinas D, Guerrero S, Cross A, Araya P, Wolf EE. Potassium titanate for the production of biodiesel. Fuel. 2016;166:237-244., styrene epoxidation2020 Nepak D, Srinivas D. Spectroscopy and catalytic activity study of gold supported on barium titanate nanotubes for styrene epoxidation. Applied Catalysis A: General. 2016;523:61-72., oxidesulfurization process2121 Salmasi M, Fatemi S, Mortazavi Y. Fabrication of promoted TiO2 nanotubes with superior catalytic activity against TiO2 nanoparticles as the catalyst of oxi-desulfurization process. Journal of Industrial and Engineering Chemistry. 2016;39:66-76., nanocomposite polymer2222 Monteiro WF, dos Santos CAB, Einloft S, Oberson M, Carone CLP, Ligabue RA. Preparation of Modified Titanate Nanotubes and Its Application in Polyurethane Nanocomposites. Macromolecular Symposia. 2016;368(1):93-97. among others. However, until present moment, there aren't studies that show the catalytic activity of TNT in glycolysis reactions, representing a yet unexploited area.

In this context, this study aims to evaluate the catalytic activity of TNT compared with zinc acetate, at different reaction times and to evaluate the influence of PET source (virgin and post-consumer bottle-grade) in the PET depolymerization.

2. Material and Methods

The materials used for the production of titanate nanotubes (TNT) and glycolysis reactions were: sodium hydroxide (99% Vetec), titanium dioxide (98% anatase phase, JB Química), zinc acetate, (98%, Fmaia), ethylene glycol (99.5%, Dinâmica), virgin PET (Rhopet S-80 - Rhodia Ster/Mossi and Ghisolfi Group) and post-consumer bottle-grade PET. All reactants were used as received. The filters used to the first and second filtration were Unifil C42 (1-2 µm).

2.1 Titanate nanotubes synthesis

The TNT were synthesized based on the hydrothermal method adapted from Kasuga et al. (1998)2323 Kasuga T, Hiramatsu M, Hoson A, Sekino T, Niihara K. Formation of Titanium Oxide Nanotube. Langmuir. 1998;14(12):3160-3163. according to the following procedure: in a beaker was weighted 1.5 g of titanium dioxide and added 120 mL of 10M NaOH aqueous solution. The mixture was kept under magnetic stirring for 30 minutes at room temperature. After that, the mixture was added to a stainless steel autoclave reactor and remained reacting at 130-140°C for 72h. Finally the formed precipitate was washed with distilled water and centrifuged several times until the pH of washing water reached about 7.

2.2 PET glycolysis

For PET depolymerization reactions by glycolysis, it was used virgin PET in pellets (2,85x2,60 mm). In the case of post-consumer PET bottles, these were from carbonated drinks, exclusively of soda, and transparent, in order to avoid variations relating to colorants and the polymer crystallinity. These bodies of PET bottles were separated from other materials that are used to compose the bottles, like the covers (PP) and labels (LDPE or PP). They were then washed, dried, manually cutted into dimensions of about 5x5mm and again washed and dried, to remove any impurities from the cutting step.

The PET glycolysis was optimized based on literature2424 Viana ME, Riul A, Carvalho GM, Rubira AF, Muniz EC. Chemical recycling of PET by catalyzed glycolysis: Kinetics of the heterogeneous reaction. Chemical Engineering Journal. 2011;173(1):210-219.,2525 Xi G, Lu M, Sun C. Study on depolymerization of waste polyethylene terephthalate into monomer of bis(2-hydroxyethyl terephthalate). Polymer Degradation and Stability. 2005;87(1):117-120. where 15g PET, 60g EG and 50 mg of catalyst were added to a 500 mL round-bottom three-naked flask equipped with magnetic stirrer, thermometer and reflux condenser. The reactions were carried out at 196°C since it is the boiling temperature of EG. The glycolysis reaction was carried out at different times (2, 3 and 4 hours) and after it was added approximately 300 mL of boiling water into flask so the products formed were dissolved. Then it was filtered under reduced pressure and the filtrate was stored at 4-10°C for 72h to forming the BHET crystals. The separation and reuse of the catalyst weren't aim of this work. After, the BHET crystals were filtered under reduced pressure using G4 sintered glass funnel, to ensure that catalysts used in the reaction were retained during filtration. The white BHET crystals were dried at 60°C for 24 hours. The influence of reaction time was then evaluated in PET conversion results and yield of BHET. The conversion (C) of PET in glycolysis reactions was calculated based on the Eq. (1):

(1) C = W i W f W i × 100

where Wi represents the initial weight of PET and Wf represents the weight of undepolymerized PET.

The molar yield (X) of BHET produced by PET depolymerization was calculated according to Eq. (2):

(2) X = W BHET , f / MW BHET W PET , i / MW PET × 100

where WBHET,f represents the final weight of BHET, MWBHET represents the molar weight of BHET (254 g.mol-1), WPET,i represents the initial weight of PET used in the reaction and MWPET represents the molecular weight of the repeat unit of PET chain (192 g.mol-1).

The turn over number (TON) of the used catalyst was calculated according to Eq. (3):

(3) TON = mols of product mols of catalyst

And the turn over frequency (TOF) of the catalyst was calculated based on Eq. (4):

(4) TOF = TON time h

2.3 Characterization

2.3.1 Titanate nanotubes (TNT)

Scanning Electron Microscopy (SEM)

TNT morphology was analyzed by field emission scanning electron microscopy (FEI Inspect F50) in SE mode (secondary electron beam). The samples were analyzed as powder using gold film deposited on the surface by ion-sputter.

Transmission Electron Microscopy (TEM)

The evaluation of the internal structure and numbers of rolled multilayer lamellar walls of TNT was made using transmission electron microscopy (FEI Tecnai G2 T20) using cupper grids with carbon film (300 mesh). The measures of TNT dimensions were obtained by by TEM analysis used Image J software (number of measurements=25).

X-Ray Diffraction (XRD)

The characterization of the crystalline structure of TNT was made by X-ray diffraction analysis (Shimadzu XRD 7000) using radiation Kα of the copper (λ= 1.542 Å), voltage 40kV, 30mA, scanning between 5°-70° 2θ and scan speed of 0.02°/min. In all analysis TNT was in powder form.

2.3.2 PET glycolysis

Differential Scanning Calorimetry (DSC)

BHET formed by PET depolymerization reactions were characterized by thermal analysis of Differential Scanning Calorimetry (DSC) in a calorimeter, Model Q20 from TA Instruments in the range from 45°C to 270°C at a heating rate of 10°C.min-1, under inert atmosphere of N2. Melting enthalpy analyzes were made in triplicate.

Thermogravimetric Analysis (TGA)

The thermal stability of starting virgin and post-consumer PET and BHET formed by PET depolymerization reactions were characterized by thermogravimetric analysis (TGA) in a SDT equipment, Model Q600 from TA Instruments in the range from 50°C to 800°C at a heating rate 20°C.min-1, under inert atmosphere of N2.

Nuclear Magnetic Resonance Spectroscopy (NMR)

In order to confirm the production of BHET, 1H-NMR and 13C-NMR were recorded on a Bruker Ascend 400 NMR spectrometer operating at 400 MHz in deuterated DMSO solution.

3. Results and Discussion

3.1 Titanate Nanotubes (TNT)

The formation of titanate nanotubes (TNT) was confirmed by SEM and TEM analysis (Figures 2a and 2b, respectively). As expected these nanotubes have a tubular morphology but very crowded. This agglomeration clearly observed by the SEM image (Figure 2a) can occur due to some effect on the drying process. Figure 2b shows TEM image of the TNT, where it is possible to observe that the nanotubes are formed by winding at least three titanates multilayer lamellar walls and have an external diameter of 8.6 nm ± 1.4 nm and presents wall thickness of 0.6 nm ± 0.1 nm (calculated using Image J software in Fig. 2). These results are in agreement with those found by Monteiro et al.2222 Monteiro WF, dos Santos CAB, Einloft S, Oberson M, Carone CLP, Ligabue RA. Preparation of Modified Titanate Nanotubes and Its Application in Polyurethane Nanocomposites. Macromolecular Symposia. 2016;368(1):93-97..

Figure 2
Images of TNT obtained by a) SEM (magnification of 100k) and b) TEM (magnification of 410K).

Figures 3 shows the diffraction pattern obtained of TiO2 precursor (anatase phase) and the characteristics peaks are situated at 2θ= 25° (101), 38° (004), 48° (200), 53° (106) and 62° (215) in agreement with the assignments described in literature2626 Leong KH, Monash P, Ibrahim S, Saravanan P. Solar photocatalytic activity of anatase TiO2 nanocrystals synthesized by non-hydrolitic sol-gel method. Solar Energy. 2014;101:321-332.,2727 Wang H, Cao S, Fang Z, Yu F, Liu Y, Weng X, et al. CeO2 doped anatase TiO2 with exposed (001) high energy facets and its performance in selective catalytic reduction of NO by NH3. Applied Surface Science. 2015;330:245-252.. In relation to TNT diffractogram, this shows disappearance of the fine peak situated at 2θ= 25° (101), main peak of anatase phase and appearance of signs of lamellar titanates, 2θ=24° (110), 28° (211), 48° (020) and 62° (422)2828 Wang L, Liu W, Wang T, Ni J. Highly efficient adsorption of Cr(VI) from aqueous solutions by amino-functionalized titanate nanotubes. Chemical Engineering Journal. 2013;225:153-163.,2929 Liu J, Liu Y, Wu Z, Chen X, Wang H, Weng X. Polyethyleneimine functionalized protonated titanate nanotubes as superior carbon dioxide adsorbents. Journal of Colloid and Interface Science. 2012;386(1):392-397. and the peak at around 2θ=10° assigned to the interlayer distance3030 Thennarasu S, Rajasekar K, Balkis Ameen K. Hydrothermal temperature as a morphological control factor: Preparation, characterization and photocatalytic activity of titanate nanotubes and nanoribbons. Journal of Molecular Structure. 2013;1049:446-457..

Figure 3
XRD results of TiO2 and TNT.

3.2 BHET characterization

In order to confirm the BHET production as the main product of PET glycolysis obtained at different times and with catalysts TNT or Zn(OAC)2, DSC, TGA, 1H and 13C-NMR analysis were performed.

Figure 4 shows the DSC curves of the main product of glycolysis from virgin (Figure 4a) and post-consumer PET (Figure 4b), which has an endothermic peak located at 110°C corresponding to the melting point of BHET1010 Al-Sabagh AM, Yehia FZ, Eissa AMF, Moustafa ME, Eshaq G, Rabie AM, et al. Cu- and Zn-acetate-containing ionic liquids as catalysts for the glycolysis of poly(ethylene terephthalate). Polymer Degradation and Stability. 2014;110:364-377.. Moreover, there are not observed peaks corresponding to dimers and/or oligomers (~170°C)3131 Imran M, Kim DH, Al-Masry WA, Mahmood A, Hassan A, Haider S, et al. Manganese-, cobalt-, and zinc-based mixed-oxide spinels as novel catalysts for the chemical recycling of poly(ethylene terephthalate) via glycolysis. Polymer Degradation and Stability. 2013;98(4):904-915. and unreacted residual PET (255-265°C)3232 Awaja F, Pavel D. Recycling of PET. European Polymer Journal. 2005;41(7):1453-1477., indicating that both catalysts were efficient in virgin and post-consumer PET glycolysis.

Figure 4
DSC curves of the BHET obtained from: a) virgin PET and b) post-consumer PET.

The melting enthalpy (ΔHm) for the glycolysis product (BHET) obtained by the use of virgin (Figure 5a) and post-consumer PET (Figure 5b) and catalyzed by TNT or Zn(OAc)2 are shown in Figure 5. As noted, regardless the type of PET or the catalyst used on glycolysis reaction, BHET melting enthalpy values showed no significant difference, indicating that BHET purity in different reaction conditions is similar.

Figure 5
Enthalpies of fusion for BHET from: a) virgin PET and b) post-consumer PET.

Thermogravimetric analysis (Figure 6) of the main product of glycolysis of virgin PET (Figure 6a) and post-consumer PET (Figure 6b) shows small weight loss differences in the degradation first stage that occurs around 250°C (degradation of BHET product). Followed by a second stage which starts at about 400°C that it is related to PET degradation formed by thermal polymerization of BHET during analysis3131 Imran M, Kim DH, Al-Masry WA, Mahmood A, Hassan A, Haider S, et al. Manganese-, cobalt-, and zinc-based mixed-oxide spinels as novel catalysts for the chemical recycling of poly(ethylene terephthalate) via glycolysis. Polymer Degradation and Stability. 2013;98(4):904-915.. Figure 6c and 6d shows in detail the first degradation stage for BHET obtained from virgin and post-consumer PET, respectively. There are included in Figure 6 the degradation curves of virgin PET and post-consumer PET. Degradation of virgin PET (Figure 6a) occurs at one single step starting at approximately 400°C, as well as degradation of post-consumer PET (Figure 6b).

Figure 6
TGA curves for the starting material and BHET obtained from: a) virgin PET, b) post-consumer PET, c) increased image for first stage weight loss for BHET from virgin PET and d) increased image for first stage weight loss for BHET from post-consumer PET.

Table 1 shows the initial weight loss temperature, Ti 5% w. l., (considered the temperature when the sample loses 5% of its weight) of all samples, % weight loss for the first and second stages of BHET and PET degradation, respectively, and also % of ash content. Results of degradation temperature at 5% weight loss showed, with small difference, that BHET obtained using Zn(OAc)2 are more thermally stable for both virgin and post-consumer PET, because it is observed the need of higher temperatures to initiate degradation.

Table 1
Degradation temperature at 5% weight loss, weight loss for the 1º and 2º degradation stages, ash content of TGA results obtained for both starting material and BHET from virgin and post-consumer PET for different reaction times.

Table 1 also shows a weight loss difference in the first stage of degradation, with the TNT catalyst samples being smaller (about 25%) than those with Zn(OAc)2 (about 30%) for BHET obtained from virgin PET. For BHET obtained from post-consumer PET, weight loss of samples using both TNT and Zn(OAc)2 as catalyst remained the same (about 25%). This results are in agreement with what has been showed previously about the purity of the BHET being unaffected by the use of different catalysts.

Samples from virgin PET and the starting material showed little difference in ash content (about 10-12%). Furthermore, to post-consumer, ash content presented higher difference (about 6-13%) that can be associated with impurities, such as additives used during processing of PET3333 Kim H, Gilbert SG, Johnson JB. Determination of Potential Migrants from Commercial Amber Polyethylene Terephthalate Bottle Wall. Pharmaceutical Research. 1990;7(2):176-179.,3434 Coltro L, Buratin AEP. Garrafas de PET para óleo comestível: avaliação da barreira à luz. Polímeros. 2004;14(3):206-211..

Accordingly to Geng et al.3535 Geng Y, Dong T, Fang P, Zhou Q, Lu X, Zhang S. Fast and effective glycolysis of poly(ethylene terephthalate) catalyzed by polyoxometalate. Polymer Degradation and Stability. 2015;117:30-36. and Imran et al.3131 Imran M, Kim DH, Al-Masry WA, Mahmood A, Hassan A, Haider S, et al. Manganese-, cobalt-, and zinc-based mixed-oxide spinels as novel catalysts for the chemical recycling of poly(ethylene terephthalate) via glycolysis. Polymer Degradation and Stability. 2013;98(4):904-915. weight loss about 20% at 250°C on the first step refers to the thermal decomposition of BHET dimers, and the second weight loss about 70% at 400°C refers to the thermal decomposition of the PET produced from the BHET dimer thermal repolymerization during the TGA process. Differently from these results, the TGA analysis in this work showed the first step of degradation refers to basically BHET monomer. This result is corroborated by the results of NMR (Figure 7 and 8), where no peaks were observed regarding BHET dimers. The NMR spectra presented signals referring only to BHET monomer.

Figure 7
1H-NMR spectrum for BHET, the main product of PET glycolysis.

The chemical structure of BHET was analyzed by nuclear magnetic resonance (NMR) spectroscopy. The 1H-NMR spectrum of BHET is shown in Figure 7 with chemical structure illustration.

The signs labelled 1, 2, 3 and 4 (Figure 7) are assigned the protons of the aromatic ring (δH = 8.1 ppm, s, 4H), hydroxyl groups (δH = 4.95 ppm, t, 2H), methylenes (-CH2-) adjacent to the -OH groups (δH = 3.73 ppm, m, 4H), methylenes (-CH2-) adjacent to the -COO groups (δH = 4.33 ppm, t, 4H), respectively. The sign around of 2.5 ppm is DMSO and in 3.3 and 2.0 ppm can be attributed to H2O residual and contamination.99 Wang S, Wang C, Wang H, Chen X, Wang S. Sodium titanium tris(glycolate) as a catalyst for the chemical recycling of poly(ethylene terephthalate) via glycolysis and repolycondensation. Polymer Degradation and Stability. 2015;114:105-114.,3131 Imran M, Kim DH, Al-Masry WA, Mahmood A, Hassan A, Haider S, et al. Manganese-, cobalt-, and zinc-based mixed-oxide spinels as novel catalysts for the chemical recycling of poly(ethylene terephthalate) via glycolysis. Polymer Degradation and Stability. 2013;98(4):904-915.,3636 Imran M, Kim BK, Han M, Cho BG, Kim DH. Sub-and supercritical glycolysis of polyethylene terephthalate (PET) into the monomer bis(2-hydroxyethyl) terephthalate (BHET). Polymer Degradation and Stability. 2010;95(9):1686-1693. The 13C-NMR spectrum of BHET is shown in Figure 8 with chemical structure illustration.

Figure 8
13C-NMR spectrum for BHET, the main product of PET glycolysis.

The sign 1 (δC = 165.65 ppm), 2 (δC = 134.24 ppm), 3 (δC = 129.96 ppm), 4 (δC = 67.48 ppm) and 5 (δC = 59.48 ppm) are assigned to the carbons of the chemical structure of BHET, as shown in Figure 8. The sign of DMSO appears in 40 ppm.

The yield values in % of BHET, turnover number (TON) and turnover frequency (TOF) are shown in Tables 2 (for virgin PET) and 3 (for post-consumer PET). TON and TOF are parameters that indicate the ability of a catalyst to convert reactants to products in molar units before becoming inactive and the number of cycles that this catalyst is able to make per time unit3737 Umpierre AP, de Jesús E, Dupont J. Turnover Numbers and soluble Metal Nanoparticles. ChemCatChem Catalysis. 2011;3(9):1413-1418., respectively.

Table 2
Yield, TON and TOF results obtained for BHET from virgin PET for different reaction times.
Table 3
Yield, TON and TOF results obtained for BHET from post-consumer PET for different reaction times.

Values of PET conversion for all reactions (using both virgin and post-consumer PET) were above 99%. BHET yields were obtained above 70% using the Zn(OAc)2 catalyst and above 80% with the use of TNT for virgin PET (Table 2). However, the reaction time seems to have no significant effect on the BHET yields regardless the catalyst. The main difference is in the results of TON and TOF, where larger values of these catalytic parameters were obtained with TNT, proving itself as a most promising catalyst than the commercial catalyst. For both catalysts there were a decrease in the values of TOF with longer reaction times, indicating that in 2 hours of glycolysis obtains the highest catalytic efficiency (greater number of cycles per time), and thereby a high yield of BHET.

For reactions using post-consumer PET (Table 3) again TNT performance were more effective than zinc acetate, i.e. higher TON and TOF to TNT as compared with Zn(OAc)2, as noted in Table 2. BHET yields remains similar to all reaction times. When comparing values of TOF and the nature of PET it is observed that TNT were more effective when compared with Zn(OAc)2 in all reactions (virgin and post-consumer). Futhermore, TNT decrease their catalytic efficiency (Tables 2 and 3), but this isn't that significant. It can be attributed to additives present in post-consumer PET.

In view of these results, it has been observed that the optimal reaction time is 2 hours, because high yields and high values of TOF are achieved with minimal energy expenditure. Consequently, TNT are the most promising catalysts compared with zinc acetate in all reactions. The glycolysis reaction can occur via acyl-type (AAC2) mechanism according to literature3838 Otton J, Ratton S. Investigation of the formation of poly(ethylene terephthalate) with model molecules: Kinetics and mechanism of the catalytic esterification and alcoholysis reactions. I. Carboxylic acid catalysis (monofunctional reactants). Journal of Polymer Science Part A: Polymer Chemistry. 1988;26(8):2183-2197.. Different metallic-based catalysts are used in this reaction, however depending of metal kind the mechanism will be different. When the catalyst is the Zn(OAc)2, the zinc atom acts as Lewis acid, interacting with oxygen from carbonyl group, corresponding to the ester group from PET chain, favoring nucleophilic attack of EG99 Wang S, Wang C, Wang H, Chen X, Wang S. Sodium titanium tris(glycolate) as a catalyst for the chemical recycling of poly(ethylene terephthalate) via glycolysis and repolycondensation. Polymer Degradation and Stability. 2015;114:105-114.. In the case of TNT as catalyst, a high actives sites quantity is available (Na+ ions and titanium atoms). Sodium ions will activate the nucleophile (EG) to attack the carbon from carbonyl group, corresponding to ester group while the titanium atoms produce a bidentate specie via coordination of the two oxygens of ester group of PET. Thus EG can attack more easily the carbon from carbonyl group that it is more susceptible. This behavior justify the higher activity of TNT when compared with Zn(OAc)2, making it the best catalyst for PET depolymerization by glycolysis.

4. Conclusions

Titanates nanotubes (TNT) were successfully synthesized by the hydrothermal method with 8.6nm ± 1.4nm external diameter. They were applied as catalyst in reactions of glycolysis of PET depolymerization comparing the results with a commercial catalyst, zinc acetate. In the reactions using virgin PET, the TNT were promising having higher yields of BHET compared to the results obtained with Zn(OAc)2, and 2 hours reaction being more efficient. The obtainment of BHET was confirmed by 1H and 13C-NMR analysis. Furthermore, the nature of PET influences decreasing the efficiency of TNT for glycolysis reaction when used post-consumer PET with BHET yields lower than those obtained Zn(OAc)2.

5. Acknowledgements

The authors thank CNPq and CAPES by scholarships. PUCRS and UFRGS for technical support. RL thanks CNPq for Technological Development-DT scholarship.

6. References

  • 1
    Eshaq G, Elmetwally AE. (Mg-Zn)-Al layered double hydroxide as a regenerable catalyst for the catalytic glycolysis of polyethylene terephthalate. Journal of Molecular Liquids 2016;214:1-6.
  • 2
    Welle F. Twenty years of PET bottle to bottle recycling--An overview. Resources, Conservation and Recycling 2011;55(11):865-875.
  • 3
    Bartolome L, Imran M, Cho BG, Al-Masry WA, Kim DH. Recent Developments in the Chemical Recycling of PET. In: Achilias D, ed. Material Recycling - Trends and Perspectives Rijeka: InTech; 2012. p. 65-84.
  • 4
    Ghaemy M, Mossaddegh K. Depolymerisation of poly(ethylene terephthalate) fibre wastes using ethylene glycol. Polymer Degradation and Stability 2005;90(3):570-576.
  • 5
    Yue QF, Wang CX, Zhang LN, Ni Y, Jin YX. Glycolysis of poly(ethylene terephthalate) (PET) using basic ionic liquids as catalysts. Polymer Degradation and Stability 2011;96(4):399-403.
  • 6
    López-Fonseca R, Duque-Ingunza I, de Rivas B, Arnaiz S, Gutiérrez-Ortiz JI. Chemical recycling of post-consumer PET wastes by glycolysis in the presence of metal salts. Polymer Degradation and Stability 2010;95(6):1022-1028.
  • 7
    Chen JW, Chen LW. The glycolysis of poly (ethylene terephthalate). Journal of Applied Polymer Science 1999;73(1):35-40.
  • 8
    Troev K, Grancharov G, Tsevi R, Gitsov I. A novel catalyst for the glycolysis of poly(ethylene terephthalate). Journal of Applied Polymer Science 2003;90(4):1148-1152.
  • 9
    Wang S, Wang C, Wang H, Chen X, Wang S. Sodium titanium tris(glycolate) as a catalyst for the chemical recycling of poly(ethylene terephthalate) via glycolysis and repolycondensation. Polymer Degradation and Stability 2015;114:105-114.
  • 10
    Al-Sabagh AM, Yehia FZ, Eissa AMF, Moustafa ME, Eshaq G, Rabie AM, et al. Cu- and Zn-acetate-containing ionic liquids as catalysts for the glycolysis of poly(ethylene terephthalate). Polymer Degradation and Stability 2014;110:364-377.
  • 11
    Ferreira HS, Rangel M do C. Nanotecnologia: aspectos gerais e potencial de aplicação em catálise. Química Nova 2009;32(7):1860-1870.
  • 12
    Galetti AE, Barroso MN, Monzón A, Abello MC. Synthesis of Nickel Nanoparticles Supported on Carbon Using a Filter Paper as Biomorphic Pattern for Application in Catalysis. Materials Research 2015;18(6):1278-1283.
  • 13
    Camargo PHC, Satyanarayana KG, Wypych F. Nanocomposites: synthesis, structure, properties and new application opportunities. Materials Research 2009;12(1):1-39.
  • 14
    Camposeco R, Castillo S, Mejia-Centeno I, Navarrete J, Rodriguez-Gonzalez V. Behavior of Lewis and Brönsted surface acidity featured by Ag, Au, Ce, La, Fe, Mn, Pd, Pt, V and W decorated on protonated titanate nanotubes. Microporous and Mesoporous Materials 2016;236:235-243.
  • 15
    Parayil SK, Razzaq A, Park SM, Kim HR, Grimes CA, In S Il. Photocatalytic conversion of CO2 to hydrocarbon fuel using carbon and nitrogen co-doped sodium titanate nanotubes. Applied Catalysis A: General 2015;498:205-213.
  • 16
    Monteiro WF, Vieira MO, Aquino AS, de Souza MO, de Lima J, Einloft S, et al. CO2 conversion to propylene carbonate catalyzed by ionic liquid containing organosilane groups supported on titanate nanotubes/nanowires. Applied Catalysis A: General 2017;544:46-54.
  • 17
    Hernández-Hipólito P, García-Castillejos M, Martínez-Klimova E, Juárez-Flores N, Gómez-Cortés A, Klimova TE. Biodiesel production with nanotubular sodium titanate as a catalyst. Catalysis Today 2014;220-222:4-11.
  • 18
    Hernández-Hipólito P, Juárez-Flores N, Martínez-Klimova E, Gómez-Cortés A, Bokhimi X, Escobar-Alarcón L, et al. Novel heterogeneous basic catalysts for biodiesel production: Sodium titanate nanotubes doped with potassium. Catalysis Today 2015;250:187-196.
  • 19
    Salinas D, Guerrero S, Cross A, Araya P, Wolf EE. Potassium titanate for the production of biodiesel. Fuel 2016;166:237-244.
  • 20
    Nepak D, Srinivas D. Spectroscopy and catalytic activity study of gold supported on barium titanate nanotubes for styrene epoxidation. Applied Catalysis A: General 2016;523:61-72.
  • 21
    Salmasi M, Fatemi S, Mortazavi Y. Fabrication of promoted TiO2 nanotubes with superior catalytic activity against TiO2 nanoparticles as the catalyst of oxi-desulfurization process. Journal of Industrial and Engineering Chemistry 2016;39:66-76.
  • 22
    Monteiro WF, dos Santos CAB, Einloft S, Oberson M, Carone CLP, Ligabue RA. Preparation of Modified Titanate Nanotubes and Its Application in Polyurethane Nanocomposites. Macromolecular Symposia 2016;368(1):93-97.
  • 23
    Kasuga T, Hiramatsu M, Hoson A, Sekino T, Niihara K. Formation of Titanium Oxide Nanotube. Langmuir 1998;14(12):3160-3163.
  • 24
    Viana ME, Riul A, Carvalho GM, Rubira AF, Muniz EC. Chemical recycling of PET by catalyzed glycolysis: Kinetics of the heterogeneous reaction. Chemical Engineering Journal 2011;173(1):210-219.
  • 25
    Xi G, Lu M, Sun C. Study on depolymerization of waste polyethylene terephthalate into monomer of bis(2-hydroxyethyl terephthalate). Polymer Degradation and Stability 2005;87(1):117-120.
  • 26
    Leong KH, Monash P, Ibrahim S, Saravanan P. Solar photocatalytic activity of anatase TiO2 nanocrystals synthesized by non-hydrolitic sol-gel method. Solar Energy 2014;101:321-332.
  • 27
    Wang H, Cao S, Fang Z, Yu F, Liu Y, Weng X, et al. CeO2 doped anatase TiO2 with exposed (001) high energy facets and its performance in selective catalytic reduction of NO by NH3 Applied Surface Science 2015;330:245-252.
  • 28
    Wang L, Liu W, Wang T, Ni J. Highly efficient adsorption of Cr(VI) from aqueous solutions by amino-functionalized titanate nanotubes. Chemical Engineering Journal 2013;225:153-163.
  • 29
    Liu J, Liu Y, Wu Z, Chen X, Wang H, Weng X. Polyethyleneimine functionalized protonated titanate nanotubes as superior carbon dioxide adsorbents. Journal of Colloid and Interface Science 2012;386(1):392-397.
  • 30
    Thennarasu S, Rajasekar K, Balkis Ameen K. Hydrothermal temperature as a morphological control factor: Preparation, characterization and photocatalytic activity of titanate nanotubes and nanoribbons. Journal of Molecular Structure 2013;1049:446-457.
  • 31
    Imran M, Kim DH, Al-Masry WA, Mahmood A, Hassan A, Haider S, et al. Manganese-, cobalt-, and zinc-based mixed-oxide spinels as novel catalysts for the chemical recycling of poly(ethylene terephthalate) via glycolysis. Polymer Degradation and Stability 2013;98(4):904-915.
  • 32
    Awaja F, Pavel D. Recycling of PET. European Polymer Journal 2005;41(7):1453-1477.
  • 33
    Kim H, Gilbert SG, Johnson JB. Determination of Potential Migrants from Commercial Amber Polyethylene Terephthalate Bottle Wall. Pharmaceutical Research 1990;7(2):176-179.
  • 34
    Coltro L, Buratin AEP. Garrafas de PET para óleo comestível: avaliação da barreira à luz. Polímeros 2004;14(3):206-211.
  • 35
    Geng Y, Dong T, Fang P, Zhou Q, Lu X, Zhang S. Fast and effective glycolysis of poly(ethylene terephthalate) catalyzed by polyoxometalate. Polymer Degradation and Stability 2015;117:30-36.
  • 36
    Imran M, Kim BK, Han M, Cho BG, Kim DH. Sub-and supercritical glycolysis of polyethylene terephthalate (PET) into the monomer bis(2-hydroxyethyl) terephthalate (BHET). Polymer Degradation and Stability 2010;95(9):1686-1693.
  • 37
    Umpierre AP, de Jesús E, Dupont J. Turnover Numbers and soluble Metal Nanoparticles. ChemCatChem Catalysis 2011;3(9):1413-1418.
  • 38
    Otton J, Ratton S. Investigation of the formation of poly(ethylene terephthalate) with model molecules: Kinetics and mechanism of the catalytic esterification and alcoholysis reactions. I. Carboxylic acid catalysis (monofunctional reactants). Journal of Polymer Science Part A: Polymer Chemistry 1988;26(8):2183-2197.

Publication Dates

  • Publication in this collection
    06 Nov 2017
  • Date of issue
    2017

History

  • Received
    12 July 2017
  • Reviewed
    22 Sept 2017
  • Accepted
    05 Oct 2017
ABM, ABC, ABPol UFSCar - Dep. de Engenharia de Materiais, Rod. Washington Luiz, km 235, 13565-905 - São Carlos - SP- Brasil. Tel (55 16) 3351-9487 - São Carlos - SP - Brazil
E-mail: pessan@ufscar.br