Upcycling of Polyethylene To Gasoline Through A Se
Upcycling of Polyethylene To Gasoline Through A Se
Upcycling of Polyethylene To Gasoline Through A Se
Article https://doi.org/10.1038/s41557-024-01506-z
Received: 19 July 2023 Ziyu Cen 1,2, Xue Han 3 , Longfei Lin 1,2 , Sihai Yang 4,5 ,
Wanying Han6, Weilong Wen1,2, Wenli Yuan1, Minghua Dong1,2, Zhiye Ma1,2,
Accepted: 11 March 2024
Fang Li6, Yubin Ke7, Juncai Dong 8, Jin Zhang8, Shuhu Liu8, Jialiang Li8, Qian Li9,
Published online: xx xx xxxx Ningning Wu9, Junfeng Xiang 9, Hao Wu10, Lile Cai10, Yanbo Hou10,
Yongqiang Cheng 11, Luke L. Daemen11, Anibal J. Ramirez-Cuesta 11,
Check for updates
Pilar Ferrer 12, David C. Grinter 12, Georg Held 12, Yueming Liu6 &
Buxing Han 1,2,6,13
The plastic wastes are projected to exceed 25 billion tonnes by 20501–5, reactor and/or down-stream processing14. Emerging approaches to
requiring urgent developments of approaches to chemically recycling transform polyolefins into the fuel-range hydrocarbons offer great
plastic wastes6–10. The key to convert polyolefins is to break their inert industrial potentials. However, fuel-range alkanes have a higher H/C
C(sp3)–C(sp3) bonds and to place effective control of selectivity to desir- ratio than polyolefins (~2.2–2.3 and 2.0, respectively). Therefore, an
able products. A state-of-the-art toolbox for generating single products external H2 source or hydrogen enriched co-reactants is required to
from polyolefins includes partial dehydrogenation and tandem isomer- promote the conversion. For example, with high-pressure H2, liquid
izing ethenolysis to yield propylene11, an electrified spatio-temporal fuels can be generated from polyolefins through hydrogenolysis or
heating approach to generate monomers in far-from-equilibrium state12 hydrocracking on noble metal catalysts15–21 (Fig. 1a and Supplementary
and pairing chemical oxidation and biological funnelling to form poly- Table 1). Likewise, the addition of hexane (H/C ratio 2.33)22 or isopentane
hydroxyalkanoates13. However, these strategies critically rely on the (H/C ratio 2.40)23 in depolymerization of polyolefins makes fuels by
use of organometallic complexes, noble metal catalysts, a complex alkane metathesis or tandem cracking and alkylation. The absence of
A full list of affiliations appears at the end of the paper. e-mail: xue.han@bnu.edu.cn; linlongfei@iccas.ac.cn; sihai.yang@pku.edu.cn; hanbx@iccas.ac.cn
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
a
Noble metals m Noble metals Noble metals
n
n –m C1–C40
n-alkanes
b
Gasoline in CH2Cl2
Waste PE Yield 81%, selectivity 99%
SSH SSH
SSH SSH
Iso-alkanes
PE LSP zeolite
n β-scis/Isom x x
C4–C12 alkenes C4–C12
m m
R R
Fig. 1 | Representative routes of chemical conversion of polyolefins. under mild conditions via an SSH strategy and the resultant gasoline shows
a, Conversion of PE to n-alkanes through hydrogenolysis over noble metal an unprecedented selectivity of 99% for direct use as a fuel. Habs, hydride
catalysts under high-pressure H2. b, Conversion of PE to gasoline on LSP zeolites abstraction; β-scis, β-scission; Isom, isomerization; Htf, hydride transfer.
external H2 or co-reactants in the reaction system is highly attractive β-scission and isomerization, resulting in commercial-grade gasoline
to industrial applications. Without the participation of external hydro- (alkenes <2 wt%) with a high research octane number of 88.0 (Fig. 1b).
gen source, only long-chain aromatics24 are formed over noble metal The inexpensive LSP-Z100 zeolites show excellent catalytic stability,
catalysts, and strategies using zeolites suffer from low polyethylene consolidating its great industrial potential.
(PE) conversion25 and generate mixtures of volatile hydrocarbons26–28.
Thus, powerful drivers exist to develop efficient, robust and economic Results and discussion
processes for the conversion of polyolefins to transportation fuels. Synthesis and characterization
In this Article, we report a self-supplied hydrogen (SSH) strat- A facile one-step hydrothermal reaction was conducted to syn-
egy to convert PE directly into gasoline with a selectivity of 99% and thesise LSP-Z100 using tetra(n-butyl)ammonium hydroxide as
yield of 81% over a unique layered self-pillared zeolite (LSP-Z100) at structure-directing agent (Methods, LSP-ZX refers to LSP zeolite
240 °C without using noble metals or any external hydrogen source with Si/Al ratio of X). It adopts MFI/MEL intergrowth structure with
(Fig. 1b). The layered structure endows LSP zeolites with extensive open self-pillared layers, which feature a large external surface area and a
framework tri-coordinated Al sites (oFTAl) as strong Lewis acid sites, series of mesopores (Fig. 2a–c, Supplementary Figs. 1 and 2 and Sup-
resulting in superior activity to activate the inert C–H bonds of PE to plementary Table 2). A similar structure was found with LSP-Z75 but
supply hydrogen internally. Time-resolved analysis, solid-state nuclear with reduced surface area and mesoporous volume (Supplementary
magnetic resonance (NMR), an isotope-labelling technique, X-ray Figs. 1 and 2 and Supplementary Table 2). A further decrease in the Si/
absorption spectroscopy and in situ inelastic neutron spectroscopy Al ratio prevents the formation of LSP structures. Previous literature
(INS) have revealed the SSH mechanism, where hydrogen is transferred shows that strong acid sites and shape selectivity of zeolites are impor-
from PE (or oligomers) to iso-alkenes formed by hydride abstraction, tant features for conversion of PE29, but conventional zeolites such as
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
a
LSP-Z100 LSP-Z100 LSP-Z100
50 nm 20 nm 5 nm
200 nm 20 nm 5 nm
b c 800 d
LSP-Z100 LSP-Z100 Experimental LSP-Z100
Adsorbed volume (STP) (mL/g)
HZSM-5 HZSM-5 Experimental HZSM-5
Cumulative fitted peak
600 Fited peak –102
Intensity (a.u.)
Intensity (a.u.)
400 –111 –92
–112 –104
200
–119
–93
0
5 10 15 20 25 30 35 0 0.2 0.4 0.6 0.8 1.0 −140 −130 −120 −110 −100 −90 −80
e f g
LSP-Z100 LSP-Z100 LSP-Z100
HZSM-5 HZSM-5 1,455 DTBPy@LSP-Z100
HZSM-5
1,545 DTBPy@HZSM-5 1,530
Intensity (a.u.)
Intensity (a.u.)
Intensity (a.u.)
150 °C 1,616
200 °C
3,610
250 °C 3,370
350 °C
450 °C 3,740
120 100 80 60 40 20 0 −20 −40 −60 −80 1,580 1,540 1,500 1,460 1,420 3,600 3,200 2,800 2,400 2,000 1,600
27
Al chemical shift (ppm) Wavenumber (cm–1) Wavenumber (cm–1)
27
Fig. 2 | Characterization of the catalysts. a, High-resolution transmission shapes. e, Solid-state Al NMR spectra of LSP-Z100 and HZSM-5. f, IR spectra
electron microscopy images of LSP-Z100 and HZSM-5, showing layered self- before and after adsorption of pyridine at variable temperatures on LSP-Z100
pillared structure of LSP-Z100. Intermittent lattice fringes (white arrowheads) and HZSM-5. The dashed lines indicate vibrational peaks of pyridine molecules
suggest that mesopores exist throughout LSP-Z100. Note that three sets of adsorbed on Brönsted acid sites (1545 cm–1) and Lewis acid sites (1455 cm–1).
transmission electron microscopy images were taken at different magnifications. g, IR spectra before and after adsorption of DTBPy at 150 °C on LSP-Z100 and
b, X-ray diffraction patterns of LSP-Z100 and HZSM-5. c, N2 adsorption/ HZSM-5. The dashed lines indicate vibrational peaks of zeolite silanol group
desorption isotherms of LSP-Z100 and HZSM-5. d, 1H–29Si cross-polarization/ (3740 cm–1), Brönsted acid site (3160 cm–1) and adsorbed DTBPy molecule
Magic Angle Spinning Nuclear Magnetic Resonance (MAS NMR) spectra of (3370, 1616 and 1530 cm−1). a.u., arbitrary units.
LSP-Z100 and HZSM-5, and de-convolution of 29Si-NMR spectra by Gaussian line
HZSM-5 and HY alone showed little activity due to the narrow pores a similar amount of acid sites to HZSM-5 (Supplementary Fig. 3 and
and limited acid sites on the external surface. Mesoporous materi- Supplementary Tables 4 and 5), but shows higher amounts of strong
als, such as MCM-41 and SAB-15, have only weak acid sites, which are Lewis acid sites (Fig. 2f, Supplementary Fig. 4 and Supplementary
incapable of cleaving the C(sp3)–C(sp3) bonds under mild conditions. Table 5). Moreover, the acid sites of LSP-Z100 are significantly more
Two-dimensional structured materials are emerging catalysts showing accessible to bulky molecules as confirmed by 2,6-di-tert-butylpyridine
unique advantages, such as high external surface area, but have not (DTBPy) infra-red (IR) spectroscopy (Fig. 2g, Supplementary Fig. 5
been used for PE upcycling. Conventional two-dimensional materials and Supplementary Discussion 1). This is in contrast to the conven-
can hardly afford shape selectivity due to unrestricted surface. LSP tional porous materials (for example, HZSM-5, HY, MCM-41 and
zeolites have uniquely pillared structures and accessible acid sites for SBA-15), which have either limited accessibilities of acid sites or no
bulky molecules, distinctly different from conventional zeolites. Com- strong acid sites.
pared with the commercial HZSM-5, LSP-Z100 showed much higher N2
adsorption (Fig. 2c) and higher content of Q3 [Si(OSi)3(OH)] (−102 ppm) Catalytic performance
and Q2 [Si(OSi)2(OH)2] (−92 ppm) Si species (Fig. 2d and Supplementary For a typical reaction, high-density PE (HDPE) was mixed with the
Table 3), due to its distinct LSP network and the presence of mesopores catalyst in a mass ratio of 5:1 and the mixture was loaded into an auto-
(Fig. 2a). LSP-Z100 exhibits trace extra-framework Al sites (Fig. 2e) and clave. After purging with N2, the autoclave was heated to 240 °C for 4 h.
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
Table 1 | Summary of the HDPE conversion and product yields over various catalystsa
Entry Catalysts Conversion (%) Yield (%) Components of gasoline range products (%)
C1–C3 C4–C12 >C 12
b
n-Alkanes i-Alkanes Alkene Cycloalkanes Aromatics
(gasoline range)
1 HZSM-5 35.1 0.6 34.5 0.0 16.0 43.0 27.8 5.7 7.5
2 HY 3.6 0.1 3.5 0.0 3.3 78.3 3.9 4.1 10.4
3 USY 8.6 0.1 8.5 0.0 3.8 77.2 1.0 6.1 11.9
4 Meso-HY 19.7 0.4 19.3 0.0 5.2 79.5 4.7 5.3 5.3
5 MCM-41 9.0 0.3 8.7 0.0 8.1 78.5 5.2 3.1 5.1
6 SBA-15 <1 – – – – – – – –
7 LSP-Z75 78.3 1.8 76.5 0.0 10.1 71.7 3.2 4.5 10.5
8 LSP-Z100 81.8 0.6 81.2 0.0 11.7 72.5 1.4 4.3 10.1
9 LSP-Z100 c
90.2 3.0 87.2 0.0 14.6 68.7 0.8 4.2 11.7
10 Commercial gasolined – – – – 9.8 44.1 7.4 6.2 20.3
11 Fresh FCC catalyst 5.8 0.2 5.6 0.0 8.5 86.7 4.4 0.4 0.0
12 Spent FCC catalyst <1 – – – – – – – –
13 Short b-axis ZSM-5 12.0 0.2 11.8 0.0 10.1 21.4 53.4 6.7 8.5
14 [C4Py]Cl–AlCl3 <1 – – – – – – – –
15 Ru/C <1 – – – – – – – –
16 Pt/γ–Al2O3 <1 – – – – – – – –
17 Pt/WO3/ZrO2 + HY(30) 12.5 0.3 12.2 0.0 7.9 81.2 1.2 0.4 9.3
a
Reaction conditions, catalyst, 0.09 g; HDPE, 0.45 g; temperature, 240 °C; reaction time, 4 h; N2 atmosphere, 0.1 MPa. bC12+ compounds in liquid products were not detected by GC. cReaction
time is 24.5 h. dCommercial gasoline with 12.2% ethanol as an additive.
The reaction was then ceased by cooling to room temperature, and SSH pathway
the products collected for analysis. Among all the tested microporous The time course of HDPE depolymerization over LSP-Z100 at 240 °C
(HZSM-5, HY and USY) and mesoporous (LSP, meso-HY, MCM-41 and was studied by analysing the products using gas chromatography (GC),
SBA-15) catalysts, only LSP zeolites exhibited substantial catalytic activi- GC mass spectroscopy (MS) and by analysing the solid residues using
ties (Table 1, entries 7–9, and Supplementary Table 6). Importantly, elemental analysis, 13C MAS NMR, 1H NMR and glow discharge electro-
LSP-Z100 gave a high conversion of HDPE of 81.8% (Table 1, entry 8) spray ionization (GD ESI) mass spectrometry. At 0.2 h (Fig. 3a), alkanes
and an unprecedented selectivity of >99% to C4–C12 compounds (gaso- and alkenes were observed, indicating that LSP-Z100 is extremely active
line range) with negligible C1–C3 compounds (<1%) (Supplementary to crack PE. Between 0.2–4 h, the yield of alkenes experienced a slight
Tables 7 and 8). This is attributed to the advantage that the formation increase and then decreased to <2%, whereas rapid formation of alkanes
of C4+ compounds, particularly branched C4+ compounds through continued until the completion of reaction when the yield of alkanes
A-type (tertiary–tertiary) or B-type (secondary–tertiary) β-scission, reached 75%. From 4 h to 24.5 h, secondary cracking of the C6–C9 prod-
is energetically more favourable30,31 and can rapidly take place at mild ucts took place (Extended Data Fig. 2), as well as cyclization and aroma-
reaction temperatures (Supplementary Discussion 2). With a prolonged tization to produce only a small fraction of aromatics and cycloalkanes
reaction time of 24.5 h, the yield of gasoline increased to 87.2% (Table 1, (Fig. 3a). The 13C NMR spectrum of the solid residue from the reaction
entry 9). Compared with commercial gasoline, the liquid product showed a clear signal of methyl group (18 ppm), branched (21 ppm)
from this reaction is of premium quality in terms of higher content of and unsaturated species (120–150 ppm) (Fig. 3b, Supplementary Fig. 6,
branched alkanes (44.1% versus 72.5%), comparable research octane Supplementary Tables 11 and 12 and Supplementary Discussion 3). This
number (86.6 versus 88.0, Supplementary Table 9) and notably lower indicates that part of PE undergoes partial cracking, isomerization and
contents of contaminating aromatics (20.3% versus 10.1%) and olefins hydride transfer to form unsaturated oligomers (H/C <2) as internal
(7.4% versus 1.4%) (Table 1, entries 8 and 10, and Extended Data Fig. 1). hydrogen source. The SSH pathway is validated by elemental analysis
The reduced olefin content is probably related to the shortened diffu- of products and reaction residues. As shown in Fig. 3c, C and H content
sion path in LSP-Z100. This suggests that intermediate alkenes diffuse transfer from the solid residue to the product with the increase of reac-
more rapidly within the catalyst, facilitating adequate interaction with tion time. The H/C ratio of products raises from initially 2.03 to ~2.2 as
active sites in a confined system. Consequently, this promotes the the reaction starts, while the H/C ratio of residue decreases during the
transformation of intermediate alkenes into alkanes (that is, the main reaction, confirming the internal hydrogen transfer (Fig. 3d). The H/C
component of gasoline). For instance, reducing the particle size of ratio of total hydrocarbons (products and solid residues) maintained
HZSM-5 from 4 μm to 200 nm resulted in a decrease in selectivity for at ~2.03 throughout the reaction (Fig. 3c and Supplementary Table 13)
alkenes (Supplementary Table 10, entries 1 and 2). A series of reported and total mass balances for all hydrocarbons typically closed to within
best-behaving catalysts (for example, fresh fluid catalytic cracking 1.8% (Supplementary Table 14). The SSH strategy inevitably generates
(FCC) catalysts27, spent FCC catalysts27, short b-axis ZSM-5 (ref. 28), a small amount (18.2%) of solid residues (H/C <2). The solid residues
[C4Py]Cl–AlCl3(ref. 23), Ru/C (ref. 16), Pt/γ–Al2O3 (ref. 24) and Pt/WO3/ were treated with hydrofluoric acid (HF) to remove the zeolite, and the
ZrO2 + HY(30) (ref. 21)) have also been tested in our system (Table 1, remaining hydrocarbons were separated through dichloromethane
entries 11–17), but they show regrettable activity due to the lack of (DCM) extraction (Supplementary Table 15) for further characteriza-
external hydrogen source or additives. tion by 1H NMR, 13C NMR and GD ESI mass spectrometry. The fraction
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
a 100 100 b c
60 C Products
2.6
Residue
H
Content of C or H (mmol)
80 80 In total
40 2.4
Simulated
Conversion (%)
Intensity (a.u.)
20
60 60 24.5 h 2.2
H/C ratio
Yield (%)
Products
Isoalkane 0
7.5 h Residue 2.0
40 Alkene 40
n-Alkane 20
4h 1.8
Aromatics
20 Cycloalkane 20 1h 40
1.6
HDPE 60
0 0 1.4
0 5 10 15 20 25 160 140 120 100 80 60 40 20 0 0 0.1 2.5 4 7.5 24.5
Time (h) δ (ppm) Time (h)
d PE e f
Habs Habs 10,000.00 3.2
β-scis β-scis LSP-Z100 + HDPE
1,000.00 0.5 h Dm(LSP-Z100) = 2.69
2h
Unsaturated 2.8
Fractal dimensions
Htf
Olefins n-Alkanes 100.00 4h
Intensity (cm–1)
oligomers
LSP-Z100 2.68
Isom Isom 10.00 2.4 2.55
2.0
0.10
Dm(LSP-Z100 + HDPE) = 2.03
SSH [H] 0.01
1.6
0.01 0.10 1.00 0 1 2 3 4 5
Q (Å–1) Reaction time (h)
g h i
Virgin PE Product Residue Products 100
100 C
60 Residue
81.8 77.9 72.5 70.5 2.4
Content of C or H (mmol)
H In total 80
20 40 20
1.8
18.2 18.2 22.1 22.1 27.5 27.5 29.5 60
0 0
R1,0 R1,t R2,0 R2,t R3,0 R3,t R4,0 R4,t PE 1 2 3 4 5 1 2 3 4 5
Cycles Cycles
Fig. 3 | Time course of conversion of HDPE over LSP-Z100 and stability of fractal dimensions of LSP-Z100, mixture of LSP-Z100 and HDPE, and reaction
LSP zeolites. a, The trends of conversion of PE and yield of products. Reaction of mixture at different reaction time obtained from SANS spectra. Simplified
conditions: catalyst, 0.09 g; HDPE powder, 0.45 g; reaction temperature, models of LSP-Z100 and HDPE are shown based on SANS results (LSP-Z100,
240 °C; N2, 0.1 MPa. b, Solid-state 13C NMR spectra of the reaction residue, orange; HDPE, grey; argon, cyan). g, A comparison of reactions 1–4 over LSP
showing the presence of unsaturated species and methylene group in residue zeolites at 240 °C for 4 h. After each reaction, the catalyst was used directly after
(32 ppm, Supplementary Fig. 6 and Supplementary Table 12). a.u., arbitrary drying. Fresh PE was added to maintain the same mass of hydrocarbons in Rx,0 for
units. c, Variation of C and H contents and H/C ratio of short chain products each reaction. Rx,0, before reaction; Rx,t, after reaction, where x = 1,2,3,4. Atomic
and solid residues as a function of reaction time. Bars with dots are attributed utilization = (sum of products/total added virgin PE) × 100%. h, The variation of
to unreacted PE. d, Scheme of PE depolymerization pathway and SSH pathway. C and H content and H/C ratio of products and solid residue in reactions 1–4 over
Habs, hydride abstraction; β-scis, β-scission; Isom, isomerization; Htf, hydride LSP zeolites at 240 °C for 4 h. i, Comparison of gasoline yield over five cycles
transfer. e, SANS spectra for HDPE and reacted HDPE at 190 °C. All SANS spectra of reactions over LSP zeolite at 240 °C for 4 h. After each cycle, the catalyst is
shown are after the subtraction of SANS spectrum of the empty cell. f, Mass calcined at 550 °C under air.
insoluble in DCM comprises unreacted PE (Supplementary Fig. 7). The the reaction (Fig. 3e). The decrease on intensity of neutron scattering in
DCM-soluble portion is predominantly composed of long-chain alkylar- the high Q region (>0.2 Å−1) is due to lower incoherent scattering caused
omatics and long-chain alkylcyclic compounds (unsaturated oligomers), by hydrogen atom, suggesting the cracking of HDPE and evaporation
with a molecular weight range between 100 and 1,000 (Supplementary of generated light hydrocarbons (Fig. 3e). In addition, the change of
Figs. 8–11 and Supplementary Tables 16 and 17). The observation of these mass fractal dimension in the Q region of 0.02–0.2 Å−1 reflects that the
unsaturated oligomers further corroborates the proposed SSH mode. components of reaction mixture had altered during reaction (Fig. 3f
Thermogravimetric analysis (TGA) shows that the yield of coke is only and Supplementary Fig. 14). Before 4 h, the mass fractal dimension
0.58% (Supplementary Figs. 12 and 13 and Supplementary Discussion 4) increased gradually, indicating the surface-assisted cracking of PE
and thus the residual comprises primarily unsaturated oligomers and chains around active sites. In this period, the viscosity of the system
uncracked PE, probably restricted by hindered diffusion. decreases because the cracking of long-chain PE into shorter chain
In situ small-angle neutron scattering (SANS) was applied to oligomers32. At 4 h, the mass fractal reached 2.68 (close to 2.69, the
investigate the diffusion of PE and intermediates to the catalysts during mass fractal dimension of LSP-Z100), indicating the approximate
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
depletion of coil chain of PE upon reaction. This reveals the hindered 64 and 68 ppm) exhibit slight shifts compared with that of HZSM-5
diffusion of residual oligomers to the active sites, consistent with the (63, 65 and 69 ppm), since the Al sites on layered structure of LSP-Z100
catalysis and TGA results. are extensively open and have different chemical environment, con-
To promote the diffusion and atomic economy, fresh PE (81.8%) sistent with the DTBPy IR spectroscopy analysis (Fig. 2g). Thus, the
is mixed with the reaction residue (18.2%) after 4 h (R1) and undergoes super-strong Lewis acid sites (strongly bound with pyridine even at
a new cycle of reaction (R2) (Fig. 3g). The newly added PE acts as sol- 450 °C as shown in Fig. 2f) of LSP-Z100 originate from oFTAl. Moreover,
vent to facilitate mass transport of residual oligomers as well as reac- the hydride transfer between 2-methylbutane (i-C5H12) and deuterated
tion substrate in the new reaction. Importantly, the result of the new hexane (C6D14) has been studied at 240 °C to demonstrate the activity
cycle is comparable with the first reaction (Fig. 3g and Supplementary of oFTAl. On LSP-Z100, i-C5H12-xDx (x = 1–10), C6D14–yHy (y = 1–3) were
Table 18), suggesting that the previous reaction residues can be con- observed (Fig. 4e–h, Extended Data Figs. 3 and 4, Supplementary
verted to products with fresh PE, although another solid residue (22.1%) Figs. 20 and 21, Supplementary Tables 22 and 23 and Supplementary
was formed from the new cycle. This reveals the key role of promoted Discussion 5), while HZSM-5 shows little activity, indicating that the
diffusion of residues, thus preserving the activity of LSP-Z100. After hydride abstraction and hydride transfer occur primarily on oFTAl. By
four cycles of reaction, the yield of gasoline is maintained at >70% extending the reaction time for HZSM-5 to 8 h, a similar deuteration
and the total atomic utilization is as high as 91% (Fig. 3g). The H/C level of 2-methylbutane and selectivities of products were observed,
contents are balanced during the four reactions (Fig. 3h and Supple- compared with that of LZP-Z100 achieved in 2 h (Supplementary
mentary Table 19) and little coke was formed (~0.5%, Supplementary Figs. 22–25), thus confirming the higher efficiency of LSP-Z100 in
Table 20 and Supplementary Fig. 15) owing to the widely opened pores hydride abstraction and transfer. Controlled experiments on LSP-Z100
of LSP zeolites to prevent coke formation. The excellent stability of with decreased accessible acid sites (poisoned by DTBPy) or oFTAl sites
LSP-Z100 was further demonstrated by cycling testing (the residues (partially removed by ammonium hexafluorosilicate) demonstrated a
were removed by calcination before next cycle), in which the yield of notable drop in PE conversion (Supplementary Fig. 26, Supplementary
gasoline remained at 80% over five cycles (Fig. 3i). The characteriza- Tables 10 and 24 and Supplementary Discussion 6). Also, the contribu-
tion of used catalysts confirms little change to the structure and acidic tion of silanol groups with weak Lewis acidity to the conversion of PE is
sites, demonstrating the excellent structural stability of LSP-Z100 limited (Supplementary Fig. 27 and Supplementary Table 10, entries
(Supplementary Figs. 16–18). Thus, LSP-Z100 enables the success of 5–9). These results demonstrate that oFTAl and accessible Brönsted
SSH mode for conversion of PE to gasoline with coke resistance and acid sites play the primary role in the PE conversion.
high atomic utilization.
Study of reaction mechanism
Identification of active sites To further understand the reaction mechanism at an atomic level,
The local environment of the acid sites Si–O–Al was interrogated in situ INS (Supplementary Fig. 28 and Supplementary Discussion
through near-edge X-ray absorption fine structure (NEXAFS) analysis 7), combined with density functional theory (DFT) calculations, was
at the oxygen K edge. The absorption edges of LSP-Z100 were consist- employed to investigate the conversion of HDPE on LSP-Z100. The
ent with those of SiO2 and Al2O3 (Fig. 4a). The peak at 534 eV is due to INS spectrum of activated LSP-Z100 gave a clean background with
adsorbed water in the materials33. The region between 538 eV and 543 eV no prominent features at 0–1,600 cm−1 (Supplementary Figs. 29 and
(white line) in the O K-edge NEXAFS spectrum is attributed to transi- 30). The mixture of HDPE with LSP-Z100 showed a similar spectrum to
tions from O 1s to O 2p anti-bonding states hybridized with Si 3s, Si 3p, that of the bare HDPE, indicating little interaction between HDPE and
Al 3s or Al 3p character34–36. To understand the interaction between acid LSP-Z100 upon mixing at room temperature (Supplementary Fig. 30).
sites and guest molecules, the conversion of HDPE over LSP-Z100 was HDPE on LSP-Z100 underwent the first catalytic conversion at 190 °C
studied at different reaction stages. At stage 1 of reaction (190 °C for for 2 h (stage 1 of reaction, Fig.5a and Supplementary Fig. 31). In com-
2 h), LSP-Z100 showed decrease in the white line intensity (538–543 eV) parison with the spectrum of HDPE, the spectrum of stage 1 of the reac-
relative to bare LSP-Z100 (Fig. 4b and Supplementary Fig. 19), suggest- tion showed a decrease in intensity at 130, 201 and 726 cm−1, which are
ing a partial filling of the O 2p anti-bonding orbital. This confirms that assigned to the in-plane and out-of-plane skeletal stretching of HDPE,
the generated unsaturated species (for example, alkenes) during the and to the methylene (–CH2–) rocking, respectively. This suggests the
reaction were bound to the Si–O–Al sites by donating electrons to the scission of C–C bonds and depolymerization of PE chains. Meanwhile,
O 2p anti-bonding orbital. In addition, the increase in intensity in the a broad feature at low energy (below 200 cm−1) appeared, suggesting
region of 545–555 eV is due to scattering from the adsorbed unsaturated that the intermediate species were disordered over the catalyst surface
species on Si–O–Al sites. With the reaction ongoing (stage 2 of reaction, showing restricted translational and rotational dynamics. In addi-
230 °C for 1 h, Fig. 4b), the white line intensity increased, indicating the tion, new peaks at 212, 235, 255, 266, 336, 440, 913 and 971 cm−1 were
adsorbed alkenes are saturated because of hydride transfer reaction. observed. The appearance of peaks at 212–266 cm−1 (assigned to methyl
Importantly, at the stage 3 of reaction (240 °C for 2 h), the intensity torsion, Supplementary Table 25) further confirms the cleavage of PE
was closed to bare LSP-Z100, indicating the complete regeneration of chains. Notably, the overall spectrum profile of stage 1 of reaction is
Si–O–Al sites upon conversion of alkenes to alkanes that could desorb consistent with that of adsorbed unsaturated oligomers (Fig. 5b and
readily from the catalysts. Supplementary Discussion 8), which is in excellent agreement with
The structures of active sites were further investigated by 31P NMR the formation of oligomers upon cracking of PE as confirmed by 13C
spectroscopy of trimethylphosphine oxide (TMPO) as a probe mol- NMR in the time course study. Specifically, shoulder peaks at 212, 255
ecule37. LSP-Z100 possesses a large amount of FTAl sites, which act as and 266 cm−1 were identified as methyl torsion of 2-methylpentane
Lewis acid sites to coordinate with lone pair electrons of basic probe and 3-methylpentane (Fig. 5c, Supplementary Fig. 32 and Supple-
molecule (64 ppm, TMPO ∙ ∙∙Al(OSi)3; 62 ppm, TMPO ∙ ∙∙Al(OSi)2(OH); mentary Table 25) and the peaks at 336 and 440 cm−1 as the skeletal
68 ppm, TMPO ∙ ∙∙Al(OSi)(OH)2) (Fig. 4c,d and Supplementary Table 21). stretching modes of short alkanes (Fig. 5b, Supplementary Fig. 32 and
In contrast, the spectrum of TMPO adsorbed on HZSM-5 shows higher Supplementary Table 25). This result indicates that PE was activated
intensity of TMPOH+ ions protonated by Brönsted acid sites (76 ppm) via hydride abstraction on oFTAl and underwent β-scission/isomeri-
and less prominent signal of Lewis acid coordinated molecules (63– zation/hydride transfer to generate iso-alkanes. Although oFTAl are
69 ppm). This result is consistent with the Py-IR experiment that shows able to activate C(sp3)–H of PE/alkanes, the alkanes produced can
higher L/B ratio of LSP-Z100 than HZSM-5 (Fig. 2f and Supplementary readily desorb as supported by the INS spectra of adsorbed alkanes
Table 5). In addition, the peaks related to FTAl sites of LSP-Z100 (62, on the LSP-Z100 (Extended Data Fig. 5a,b). In contrast, alkenes bound
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
a 4 b 1.8 c
LSP-Z100 Stage 3 Experimental LSP-Z100
1.6 Experimental HZSM-5
SiO2 Stage 2
Intensity (a.u.)
1.2
74 68 62
1.0
2
0.8 89
0.6
76 65
1
0.4 63
83 69
0.2
0 0
530 535 540 545 550 555 560 565 570 530 535 540 545 550 555 560 565 570 100 90 80 70 60 50 40
31
Energy (eV) Energy (eV) P chemical shift (ppm)
d e h
Htf reaction on LSP-Z100
Htf reaction on HZSM-5
II
I
Intensity (a.u.)
III IV V
f g
I@LSP-Z100: i-C5H12+i-C5H12-XDX II@LSP-Z100: C6D14+C6D14-yHy
I@HZSM-5: i-C5H12+i-C5H11D II@HZSM-5: C6D14+C6D13H
Normalized intensity (a.u.)
Normalized intensity (a.u.)
Fig. 4 | Identification of active sites of LSP-Z100 by NEXAFS, NMR and of i-C5H12/C6D14 hydride transfer reactions by LSP-Z100 and HZSM-5. I–V are
isotope-labelling technique. a, O K-edge NEXAFS for LSP-Z100, SiO2 and Al2O3. mainly composed of i-C5H11D, C6D13H, i-C5D11H, i-C6D13H and C6D12H2, respectively
The absorption edges of LSP-Z100 were consistent with those of SiO2 and Al2O3. (Extended Data Figs. 3 and 4). f, Mass spectra of 2-methylbutane after reaction on
b, O K-edge NEXAFS spectra for LSP-Z100 before reaction and LSP-Z100 during HZSM-5 and LSP-Z100. g, Mass spectra of hexane after reaction on HZSM-5 and
the reaction. c, 31P MAS NMR spectra of LSP-Z100 and HZSM-5 loaded with LSP-Z100. h, Scheme of i-C5H12/C6D14 hydride transfer reactions. Habs, hydride
TMPO. d, View of structure of Al sites with adsorption of TMPO. e, GC traces abstraction; a.u., arbitrary units.
strongly to active sites to promote their further conversion with SSH A full catalytic circle can be established (Fig. 5f). Upon adsorp-
(Extended Data Fig. 5c). tion on the LSP zeolite, PE is activated via hydride abstraction38,39
Then, the stage 2 of reaction was studied (230 °C for 1 h, Fig. 5a). on oFTAl or via protonation on the Brönsted acid sites40 to give
The very strong peak at 235 cm−1, corresponding to the torsional mode carbenium ions or carbonium ions as the initiation step. The carbo-
of oligomers, decreased notably in intensity. Meanwhile, the peaks at nium ions formed on the Brönsted acid sites are unstable, rapidly
212 and 255 cm−1, corresponding to the methyl torsion of iso-alkanes, collapsing to afford carbenium ions along with H2 or alkanes. The
grew in intensity. This result suggests the rapid conversion of activated isotope-labelling reactions (Fig. 4e–h) suggest that the activation via
HDPE or oligomers to iso-alkanes on the zeolite via β-scission/isomeri- hydride abstraction on oFTAl dominates at 240 °C. Then the activated
zation/hydride transfer reactions. However, the peaks of HDPE (130, PE (that is, the carbenium ion originated from hydride abstraction
201 and 726 cm−1) were still observed. Finally, the stage 3 of reaction or carbonium ion collapse) binds strongly to the Brönsted acid sites
was also investigated (240 °C for 2 h, Fig. 5a). The peaks at 235 and and undergoes rapid β-scission and skeletal isomerization via a
726 cm−1 disappeared and the peaks at 212, 255 and 266 cm−1 increased three-membered ring structure41–43, followed by intramolecular
dramatically in intensity. The spectrum of stage 3 of reaction showed hydrogen transfer to generate stable tertiary carbenium ions. Mean-
notable similarity with iso-hexanes (Fig. 5d,e), confirming the efficient while, oFTAl abstract hydrogen from PE or oligomers. Subsequen-
transformation of HDPE into iso-alkanes. tially, tertiary carbenium ions react with the SSH through hydride
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
a Neutron energy loss (meV) b Neutron energy loss (meV) c Neutron energy loss (meV)
0 20 40 60 80 100 120 140 160 180 200 0 20 40 60 80 100 120 140 160 180 200 0 10 20 30 40 50
S (Q, ω) (a.u.)
S (Q, ω) (a.u.)
255
212 266
235
0 200 400 600 800 1,000 1,200 1,400 1,600 0 200 400 600 800 1,000 1,200 1,400 1,600 0 50 100 150 200 250 300 350 400
Neutron energy loss (cm–1) Neutron energy loss (cm–1) Neutron energy loss (cm–1)
Adsorbed 2-MP
Adsorbed 3-MP
Stage 3
Olig, 726 cm–1 3-MP, 212, 266 cm–1 3-MP, 384 cm–1 2-MP, 255 cm–1
0 200 400 600 800 1,000 1,200 1,400 1,600
f n
n
n n
Fig. 5 | INS spectra for LSP-Z100 zeolite on the catalytic conversion of adsorbed 2-MP, adsorbed 3-MP and reacted HDPE (stage 3 of the reaction) over
HDPE and proposed reaction mechanism. a, A comparison of INS spectra for LSP-Z100. All INS spectra shown here are after the subtraction of INS spectrum
solid HDPE and reacted HDPE over LSP-Z100. b, A comparison of INS spectra of the empty cell and zeolite. e, Selected vibrational modes of PE, oligomers,
of unsaturated oligomers, adsorbed 2-methylpentane (2-MP), adsorbed 3-MP and 2-MP (C, grey; H, white) observed by experiments. f, The proposed
3-methylpentane (3-MP) and reacted HDPE (stage 1 of the reaction) over main reaction pathway for the conversion of PE to iso-alkane over LSP-Z100,
LSP-Z100. The INS spectrum of adsorbed unsaturated oligomer was generated by including initiation, cracking/isomerization and hydride transfer. Habs, hydride
1-butene adsorbed on LSP-Z100. The dosed 1-butene oligomerizes over LSP-Z100 abstraction; Prot, protonation; β-scis, β-scission; Isom, isomerization; Htf,
and unsaturated bonds of oligomer bind to Brönsted acid site to form carbenium hydrogen transfer.
ions. c, Enlarged spectra of b at 0–400 cm−1. d, A comparison of INS spectra of
transfer to yield iso-alkanes for facile desorption and the acid sites PE molecules through hydride abstraction on oFTAl sites, β-scission/
regenerated. isomerization on Brönsted acid sites and hydride transfer on oFTAl
sites, collectively resulting in the production of gasoline with both high
Conclusion selectivity and yield. Moreover, LSP zeolites show excellent catalytic
The LSP structures endow the materials with oFTAl sites and accessible performance for the production of commercial-grade gasoline from
Brönsted acid sites, which facilitate the SSH mode to activate bulk both low-density and high-density PE waste (Supplementary Fig. 33),
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
which account for about 25% of the plastic waste today. The potential 19. Chen, L. et al. Disordered, sub-nanometer Ru structures on CeO2
profit of this route makes it economically attractive (Supplementary are highly efficient and selective catalysts in polymer upcycling
Discussion 9 and Supplementary Tables 26–28, the analysis is subject to by hydrogenolysis. ACS Catal. 12, 4618–4627 (2022).
uncertainties as discussed in Supplementary Information). Moreover, 20. Tennakoon, A. et al. Catalytic upcycling of high-density
this PE-based route to produce gasoline has a reduced carbon emis- polyethylene via a processive mechanism. Nat. Catal. 3,
sion compared with the conventional oil-based route (Supplementary 893–901 (2020).
Discussion 10, Supplementary Figs. 34–37, Supplementary Table 29). 21. Liu, S., Kots, P. A., Vance, B. C., Danielson, A. & Vlachos, D. G.
The integration of low energy input, inexpensive and noble metal-free Plastic waste to fuels by hydrocracking at mild conditions. Sci.
and highly stable catalysts, removal of external hydrogen source and Adv. 7, eabf8283 (2021).
products for direct use as transportation fuels with minimized envi- 22. Jia, X., Qin, C., Friedberger, T., Guan, Z. & Huang, Z. Efficient and
ronmental impact affords a promising solution to the abatement of selective degradation of polyethylenes into liquid fuels and waxes
future plastic pollution via the ‘waste-to-chemical’ strategy. under mild conditions. Sci. Adv. 2, e1501591 (2016).
23. Zhang, W. et al. Low-temperature upcycling of polyolefins into
Online content liquid alkanes via tandem cracking–alkylation. Science 379,
Any methods, additional references, Nature Portfolio reporting sum- 807–811 (2023).
maries, source data, extended data, supplementary information, 24. Zhang, F. et al. Polyethylene upcycling to long-chain
acknowledgements, peer review information; details of author contri- alkylaromatics by tandem hydrogenolysis/aromatization. Science
butions and competing interests; and statements of data and code avail- 370, 437–441 (2020).
ability are available at https://doi.org/10.1038/s41557-024-01506-z. 25. Tan, J. Z., Hullfish, C. W., Zheng, Y., Koel, B. E. & Sarazen, M. L.
Conversion of polyethylene waste to short chain hydrocarbons
References under mild temperature and hydrogen pressure with metal-free
1. Geyer, R., Jambeck, J. R. & Law, K. L. Production, use, and fate of and metal-loaded MFI zeolites. Appl. Catal. B 338, 123028
all plastics ever made. Sci. Adv. 3, e1700782 (2017). (2023).
2. Borrelle, S. B. et al. Predicted growth in plastic waste exceeds 26. Zhang, Z. et al. Recovering waste plastics using shape-
efforts to mitigate plastic pollution. Science 369, 1515–1518 selective nano-scale reactors as catalysts. Nat. Sustain. 2,
(2020). 39–42 (2019).
3. MacLeod, M., Arp, H. P. H., Tekman, M. B. & Jahnke, A. The global 27. Vollmer, I., Jenks, M. J. F., Mayorga González, R., Meirer, F. &
threat from plastic pollution. Science 373, 61–65 (2021). Weckhuysen, B. M. Plastic waste conversion over a refinery waste
4. Rochman, C. M. & Hoellein, T. The global odyssey of plastic catalyst. Angew. Chem. Int. Ed. 60, 16101–16108 (2021).
pollution. Science 368, 1184–1185 (2020). 28. Duan, J. et al. Coking-resistant polyethylene upcycling modulated
5. Ragusa, A. et al. Plasticenta: first evidence of microplastics in by zeolite micropore diffusion. J. Am. Chem. Soc. 144, 14269–
human placenta. Environ. Int. 146, 106274 (2021). 14277 (2022).
6. Rahimi, A. & García, J. M. Chemical recycling of waste plastics for 29. Manos, G., Garforth, A. & Dwyer, J. Catalytic degradation of
new materials production. Nat. Rev. Chem. 1, 0046 (2017). high-density polyethylene over different zeolitic structures. Ind.
7. Abel, B. A., Snyder, R. L. & Coates, G. W. Chemically recyclable Eng. Chem. Res. 39, 1198–1202 (2000).
thermoplastics from reversible-deactivation polymerization of 30. Weitkamp, J., Jacobs, P. A. & Martens, J. A. Isomerization and
cyclic acetals. Science 373, 783–789 (2021). hydrocracking of C9 through C16 n-alkanes on Pt/HZSM-5 zeolite.
8. Korley, L. T. J., Epps, T. H., Helms, B. A. & Ryan, A. J. Toward Appl. Catal. 8, 123–141 (1983).
polymer upcycling—adding value and tackling circularity. 31. Lin, L. F. et al. Acid strength controlled reaction pathways for the
Science 373, 66–69 (2021). catalytic cracking of 1-pentene to propene over ZSM-5. ACS Catal.
9. Jehanno, C. et al. Critical advances and future opportunities in 5, 4048–4059 (2015).
upcycling commodity polymers. Nature 603, 803–814 (2022). 32. Makrodimitri, Z. et al. Viscosity of heavy n-alkanes and diffusion
10. Du, J. et al. Efficient solvent- and hydrogen-free upcycling of of gases therein based on molecular dynamics simulations and
high-density polyethylene into separable cyclic hydrocarbons. empirical correlations. J. Chem. Thermodynamics 91, 101–107
Nat. Nanotechnol. 18, 772–779 (2023). (2015).
11. Conk, R. J. et al. Catalytic deconstruction of waste polyethylene 33. Parent, P. H., Laffon, C., Mangeney, C., Bournel, F. & Tronc, M.
with ethylene to form propylene. Science 377, 1561–1566 (2022). Structure of the water ice surface studied by x-ray absorption
12. Dong, Q. et al. Depolymerization of plastics by means of spectroscopy at the O K-edge. J. Chem. Phys. 117, 10842–10851
electrified spatiotemporal heating. Nature 616, 488–494 (2023). (2002).
13. Sullivan, K. P. et al. Mixed plastics waste valorization through 34. Wu, Z. Y., Jollet, F. & Seifert, F. Electronic structure analysis of
tandem chemical oxidation and biological funneling. Science via x-ray absorption near-edge structure at the Si K, L2,3 and O K
378, 207–211 (2022). edges. J. Phys. Condens. Matter 10, 8083–8092 (1998).
14. García, J. M. Catalyst: design challenges for the future of plastics 35. Gautier, M. et al. Alpha-Al2O3 (0001) surfaces: atomic
recycling. Chem 1, 813–815 (2016). and electronic structure. J. Am. Ceram. Soc. 77, 323–334
15. Rorrer, J. E., Beckham, G. T. & Román-Leshkov, Y. Conversion of (1994).
polyolefin waste to liquid alkanes with Ru-based catalysts under 36. Uchino, T., Sakka, T., Ogata, Y. & Iwasaki, M. Local structure of
mild conditions. JACS Au 1, 8–12 (2021). sodium aluminosilicate glass: an ab initio molecular orbital study.
16. Rorrer, J. E., Troyano-Valls, C., Beckham, G. T. & Román-Leshkov, J. Phys. Chem. 97, 9642–9649 (1993).
Y. Hydrogenolysis of polypropylene and mixed polyolefin plastic 37. Hu, M. et al. Unravelling the reactivity of framework Lewis acid
waste over Ru/C to produce liquid alkanes. ACS Sustain. Chem. sites towards methanol activation on H‐ZSM‐5 zeolite with solid‐
Eng. 9, 11661–11666 (2021). state NMR spectroscopy. Angew. Chem. Int. Ed. 61, e202207400
17. Wang, C. et al. Polyethylene hydrogenolysis at mild conditions over (2022).
ruthenium on tungstated zirconia. JACS Au 1, 1422–1434 (2021). 38. Dapsens, P. Y., Mondelli, C. & Pérez-Ramírez, J. Design of
18. Kots, P. A. et al. Polypropylene plastic waste conversion to Lewis-acid centres in zeolitic matrices for the conversion of
lubricants over Ru/TiO2 catalysts. ACS Catal. 11, 8104–8115 (2021). renewables. Chem. Soc. Rev. 44, 7025–7043 (2015).
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
39. Primo, A. & Garcia, H. Zeolites as catalysts in oil refining. Chem. Publisher’s note Springer Nature remains neutral with regard to
Soc. Rev. 43, 7548–7561 (2014). jurisdictional claims in published maps and institutional affiliations.
40. Kotrel, S., Knözinger, H. & Gates, B. C. The Haag–Dessau
mechanism of protolytic cracking of alkanes. Microporous Open Access This article is licensed under a Creative Commons
Mesoporous Mater. 35–36, 11–20 (2000). Attribution 4.0 International License, which permits use, sharing,
41. Mooiweer, H. H., de Jong, K. P., Kraushaar-Czarnetzki, B., adaptation, distribution and reproduction in any medium or format,
Stork, W. H. J. & Krutzen, B. C. H. Skeletal isomerization of olefins as long as you give appropriate credit to the original author(s) and the
with the zeolite ferrierite as catalyst. Stud. Surf. Sci. Catal. 84, source, provide a link to the Creative Commons licence, and indicate
2327–2334 (1994). if changes were made. The images or other third party material in this
42. Trombetta, M. et al. FT-IR studies on light olefin skeletal article are included in the article’s Creative Commons licence, unless
isomerization catalysis: III. surface acidity and activity of indicated otherwise in a credit line to the material. If material is not
amorphous and crystalline catalysts belonging to the SiO2–Al2O3 included in the article’s Creative Commons licence and your intended
system. J. Catal. 179, 581–596 (1998). use is not permitted by statutory regulation or exceeds the permitted
43. Rey, J., Gomez, A., Raybaud, P., Chizallet, C. & Bučko, T. use, you will need to obtain permission directly from the copyright
On the origin of the difference between type A and type B holder. To view a copy of this licence, visit http://creativecommons.
skeletal isomerization of alkenes catalyzed by zeolites: the org/licenses/by/4.0/.
crucial input of ab initio molecular dynamics. J. Catal. 373,
361–373 (2019). © The Author(s) 2024
1
Beijing National Laboratory for Molecular Sciences, CAS Laboratory of Colloid and Interface and Thermodynamics, CAS Research/Education Center for
Excellence in Molecular Sciences, Center for Carbon Neutral Chemistry, Institute of Chemistry, Chinese Academy of Sciences, Beijing, China.
2
School of Chemical Sciences, University of Chinese Academy of Sciences, Beijing, China. 3College of Chemistry, Beijing Normal University, Beijing,
China. 4College of Chemistry and Molecular Engineering, Beijing National Laboratory for Molecular Sciences, Peking University, Beijing, China.
5
Department of Chemistry, University of Manchester, Manchester, UK. 6Shanghai Key Laboratory of Green Chemistry and Chemical Processes, State Key
Laboratory of Petroleum Molecular and Process Engineering, School of Chemistry and Molecular Engineering, East China Normal University, Shanghai,
China. 7China Spallation Neutron Source, Institute of High Energy Physics, Dongguan, China. 8Institute of High Energy Physics, Chinese Academy of
Sciences, Beijing, China. 9Center for Physicochemical Analysis Measurements, Institute of Chemistry, Chinese Academy of Sciences, Beijing, China.
10
SINOPEC Research Institute of Petroleum Processing, Beijing, China. 11Neutron Scattering Division, Neutron Sciences Directorate, Oak Ridge National
Laboratory, Oak Ridge, TN, USA. 12Diamond Light Source, Harwell Science and Innovation Campus, Didcot, UK. 13Institute of Eco-Chongming,
Shanghai, China. e-mail: xue.han@bnu.edu.cn; linlongfei@iccas.ac.cn; sihai.yang@pku.edu.cn; hanbx@iccas.ac.cn
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
platform based on DFT calculation46. On this platform, molecular con- Data Environment for Science. We thank Z. Wang for the assistance
struction and visualization modules are built using Ketcher, JSmol and during the acquisition of GD ESI mass spectra.
molview, three-dimensional conformation of two-dimensional molecu-
lar structures and the optimization of molecular structures based on Author contributions
Merck Molecular Force Field are performed using openbabel 2.3.1, and Z.C. carried out syntheses, characterization and catalytic tests of the
optimization of molecular structure based on DFT is performed using zeolite samples. W.H., W.W., W.Y., M.D., Z.M., F.L. and Y.L. participated in
Gaussian 09 at the M06-2X/6-31G(d) level. sample characterization. Z.C., Q.L., N.W. and J.X. carried out solid-state
NMR experiments. Z.C., W.W., W.Y., L.L., S.L., J.L., J.D., J.Z., P.F., D.C.G. and
Data availability G.H. collected and analysed NEXAFS data. Y.K., Z.C. and L.L. collected
All data are available in the main text or Supplementary Information. and analysed SANS data. S.Y., L.L., Z.C., Y.C., L.L.D. and A.J.R.-C.
Source data are provided with this paper. collected and analysed INS and DFT data. H.W., L.C. and Y.H. calculated
CO2 emission. X. H., S.Y., L.L. and B.H. directed and supervised the
References project. Z.C., X. H., S.Y., L.L. and B.H. prepared the manuscript.
44. Zhang, X. et al. Synthesis of self-pillared zeolite nanosheets by
repetitive branching. Science 336, 1684–1687 (2012). Competing interests
45. Ramirez-Cuesta, A. J. aCLIMAX 4.0.1, the new version of the The characterization of catalysts, catalytic method and catalytic
software for analyzing and interpreting INS spectra. Comput. results are described in a patent invented by L.L., Z.C. and B.H.
Phys. Commun. 157, 226–238 (2004). (Chinese patent, Institute of Chemistry, Chinese Academy of Sciences,
46. Li, Q., Tang, Y. & Xiang, J. An on-line NMR chemical shift prediction Longfei Lin, Ziyu Cen, Buxing Han, 202310750469.6, patent pending).
platform based on density functional theory. Chinese J. Magn. The other authors declare no competing interests.
Reson. 38, 22–31 (2021).
Additional information
Acknowledgements Extended data is available for this paper at
This work was supported by the National Natural Science Foundation https://doi.org/10.1038/s41557-024-01506-z.
of China (grant no. 22293012 to L.L., grant no. 22293015 to B.H.,
grant no. 22121002 to B.H.), Engineering and Physical Sciences Supplementary information The online version
Research Council (EPSRC) (grant no. EP/V056409/1 to S.Y.), the contains supplementary material available at
University of Manchester, Beijing National Laboratory for Molecular https://doi.org/10.1038/s41557-024-01506-z.
Sciences (BNLMS) and Peking University. We thank Beijing
Synchrotron Radiation Facility, Diamond Light Source and China Correspondence and requests for materials should be addressed to
Spallation Neutron Source for access to the beamlines 4B7B, B07-B Xue Han, Longfei Lin, Sihai Yang or Buxing Han.
(SI33962) and Small Angle Neutron Diffractometer, respectively.
A portion of this research used resources at the Spallation Neutron Peer review information Nature Chemistry thanks Jeff Armstrong and
Source, a Department of Energy (DOE) Office of Science User the other, anonymous, reviewer(s) for their contribution to the peer
Facility operated by Oak Ridge National Laboratory. The computing review of this work.
resources at Oak Ridge National Laboratory were made available
through the VirtuES and the ICE-MAN projects, funded by Laboratory Reprints and permissions information is available at
Directed Research and Development program and Compute and www.nature.com/reprints.
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
Extended Data Fig. 1 | GC trace for commercial gasoline and gasoline obtained from conversion of HDPE. Reaction conditions: LSP-Z100, 0.09 g; HDPE, 0.45 g;
temperature, 240 °C; reaction time, 4 h; N2 atmosphere, 0.1 MPa. The ethanol in commercial gasoline is an additive.
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
Extended Data Fig. 2 | Trends of product yield in carbon numbers in the time course study. The profiles in (a) have been moved along Y axis, which only show trend
rather than absolute value. The profiles in (b) are plotted from original data and show the actual yields of species with certain carbon numbers.
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
Extended Data Fig. 3 | Standard mass spectrum of 2-methylbutane (a) of III@LSP-Z100 is because of dehydrogenation, oligomerisation of C6D14
and mass spectra of 2-methylbutane after the reaction on HZSM-5 (b) and and cleavage of the formed oligomers. Reaction conditions: catalyst, 0.075 g;
LSP-Z100 (c, d). The Mass spectra of GC peaks I@LSP-Z100 and III@LSP-Z100 in 2-methylbutane, 0.075 g; n-hexane (d-14), 0.15 g; temperature, 240 °C; reaction
GC traces show that I@LSP-Z100 is composed of i-C5H12 and i-C5H11D, and time, 1 hour; N2 atmosphere, 0.1 MPa.
III@LSP-Z100 is composed of i-C5D12, i-C5D11H and i-C5D10H2. The generation
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
Extended Data Fig. 4 | Standard mass spectrum of C6D14 (a) and mass spectra at m/z = 48 demonstrates the deuterated hexane has isomerized into deuterated
of deuterated hexane after reaction on HZSM-5 (b) and LSP-Z100 (c-e). 2-methylbutane. Reaction conditions: catalyst, 0.075 g; 2-methylbutane,
The Mass spectra show that II@LSP-Z100 is composed of C6D13H and C6D14, 0.075 g; n-hexane (d-14), 0.15 g; temperature, 240 °C; reaction time, 1 hour;
and IV@LSP-Z100 is composed of i-C6D13H and i-C6D12H2, and V@LSP-Z100 is N2 atmosphere, 0.1 MPa.
composed of C6D12H2 and C6D11H3. The base peak of IV@LSP-Z100 mass spectra
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41557-024-01506-z
Extended Data Fig. 5 | Comparison of INS spectra of adsorbed 2-methylpentane (2-MP), 3-methylpentane (3-MP) and 1-butene (unsaturated oligomer) with
that of solid 2-MP, 3-MP and 1-butene. a, INS spectra of adsorbed 2-MP and solid 2-MP. b, INS spectra of adsorbed 3-MP and solid 3-MP. c, INS spectra of adsorbed
1-butene and solid 1-butene.
Nature Chemistry
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
1. use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
2. use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com