Highly Porous Metal-Free Graphitic Carbon ... - ACS Publications

Mar 16, 2018 - Wenping Lv,. †. Xingyun Zhao,. †,‡ and Ren'an Wu*,†. †. Laboratory of High-Resolution Mass Spectrometry Technologies, CAS Key...
0 downloads 0 Views 2MB Size
Subscriber access provided by UNIV OF SCIENCES PHILADELPHIA

Functional Nanostructured Materials (including low-D carbon)

Highly porous metal-free graphitic carbon derived from metal-organic framework for profiling of N-linked glycans Xin Li, Guiju Xu, Jiaxi Peng, Shengju Liu, Hongyan Zhang, Jiawei Mao, Huan Niu, Wenping Lv, Xingyun Zhao, and Ren'an Wu ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.8b02423 • Publication Date (Web): 16 Mar 2018 Downloaded from http://pubs.acs.org on March 18, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Highly porous metal-free graphitic carbon derived from metal-organic framework for profiling of N-linked glycans Xin Li a,b, Guiju Xu a,b, Jiaxi Peng a,b, Shengju Liu a,b, Hongyan Zhang a,b, Jiawei Maoa,b, Huan Niu a,b, Wenping Lv a, Xingyun Zhao a,b, Ren’an Wua,*

a

Laboratory of High-Resolution Mass Spectrometry Technologies, CAS Key Laboratory

of Separation Science for Analytical Chemistry, Dalian Institute of Chemical Physics, Chinese Academy of Sciences (CAS), Dalian 116023, China b

The University of Chinese Academy of Sciences, Beijing 100049, China

Email: [email protected]

KEYWORDS: metal-free, porous graphitic carbon, metal organic framework, glycan, profiling, enrichment, size-exclusion

1 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ABSTRACT In this work, a highly efficient profiling of N-linked glycans was achieved by a facile and eco-friendly synthesized highly porous metal-free carbon material. The metal-free carbon was derived from a well-defined nanorod zinc metal-organic framework via the metal removal under a high-temperature carbonization, which exhibited a highly specific surface area of 1700 m2/g. After a further oxidation, the oxidized metal-free carbon was applied to the selective isolation of N-linked glycans from complex biological samples due to the strong interaction between carbon and glycan as well as the size-exclusion mechanism. 26 N-linked glycans could be identified from the digest of a standard glycoprotein ovalbumin at a concentration of 0.01 µg/µL, and the detection limit of glycans could be down to 1 ng/µL with 21 N-linked glycans identified. When the mass ratio of the interfering protein bovine serum albumin vs. a standard ovalbumin digest up to 500:1, there were 24 N-glycans confidentially identified. From a real complex sample of a healthy human serum, there were 43 N-linked glycans identified after the enrichment of oxidized metal-free carbon. In a word, the metal-free carbon is opening up new prospect for the high through-put identification of glycan.

1. INTRODUCTION Glycosylation, as one of the most important but complex post-translational modifications of proteins, plays the critical role in cell biological functions including cell growth, cell adhesion, cell-cell recognition, etc.1,2 Generally, protein glycosylation can be classified into two main types: the N-glycosylation and O-glycosylation.3 The aberrant alteration of the N-glycosylation has been found to be associated with many human diseases.4 The linked glycans on proteins have been highlighted in cancer research, for serving as important biomarkers and providing a set of specific targets for therapeutic intervention.5 Therefore, the identification of glycans is significantly important not only for the comprehensive understanding of protein glycosylations but also for regulating the development and progression of cancers. Yet, the confident identification of glycans from 2 / 27

ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

the complex biological system is quite difficult, which has pushed the development of analytical methods to separate, characterize, and quantify glycans from various biological samples (such as blood and tissue etc.).6-10 As the high-throughput and ultrahigh sensitivity, mass spectrometry (MS) has undertaken the critical task to identify glycans as well as the in-depth glycosylation analysis.11,12 However, the extremely low abundance of inherent glycans in real samples and the low ionization efficiency of glycans in mass spectrometry analysis both lead to the difficulty of glycans identification.13 In addition, the severe interferences especially from matrix proteins or/and other components in a complex biological sample made it more challengeable for the global analysis of glycans.14,15 Therefore, developing efficient glycan-isolated methods with strong anti-fouling ability before mass spectrometry analysis is significantly important for the glycan profiling. Up

to

now,

approaches

of

organic

solvent

precipitation,16

lectin-affinity17/boronate-affinity18,19 chromatography, and carbonic material isolation20-22 have been developed to separate and/or enrich glycans. Among of them, the carbon materials have been paid great attention in glycan isolation as the strong interaction between carbon and glycans.23-25 The carbon materials with high surface areas and suitable pore structures are important for the highly efficient isolation of glycans from complex biological samples.26,27 Diverse approaches in fabricating porous carbon materials including activation, thermal annealing, and templating have been developed.23,28 The hard-templating and the soft-templating approaches were the most frequently applied methods to prepare carbon materials in glycan enrichment.24,26 While, the tedious multi-steps for templates removal and the usage of strong corrosive chemical agents in subsequent activation procedures are the shortages of the templating approaches.29,30 Hence, a facile and eco-friendly preparation of clean and porous carbon materials is greatly in demand. Metal–organic frameworks (MOFs) are a class of porous and well-ordered 3 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 28

crystalline materials synthesized via the coordination of metal clusters or ions with organic ligands.31,32 Carbonizing MOF materials is receiving immense interests to construct porous carbon materials with the structural diversities, large specific surface areas and the tailored pores.33,34 Efforts on carbonization of MOFs have been made in multi-research fields not only sample separations in analytical chemistry but also drug delivery, thermal therapy, catalysis as well as the electrical energy storage.35-37 However, most of currently MOFs are prepared via solvothermal method operated at high temperature/pressure and the preparation process typically requires several hours to days.38 In addition, well-controlled the size and morphology of MOFs is difficult, especially some nanosized MOFs need special additives such as surfactants or moderators.39,40 As a result, rapid and facile synthesized MOFs route is very necessary. On the other side, as MOFs are consisting of metal moieties and organic ligands, the metal residuals (metal or metal oxide) in MOFs-derived carbon materials seemed very common during the carbonization process.41 If metal residuals are not the desired part of synthesized carbon materials such as the active moieties for catalysis, which otherwise are not favoring components to their further applications because of the possible adverse effects of the exposure and/or leakage of metal.42,43 Similarly, the residual metal of MOF-derived carbon materials is not desired for glycan enrichment not only the obvious decrease of the material specific surface area but also the possible nonspecific interfering arising from interaction between metal ion and other non-target biomolecules.44,45 Therefore,

employing

MOFs

to

derive

high-porosity

carbon

materials

with

metal-eliminated is of great expecting, not only for analytical chemistry requirements but also for their further applications in other areas. In this work, we came up with an efficient approach to prepare metal-free graphitic carbon via a rapid room-temperature synthesized zinc MOFs, and for the first time endowed its further application in high-throughput N-glycan profiling. The zinc MOFs were well-defined to nanosize and uniform rod-like morphology without use of any 4 / 27

ACS Paragon Plus Environment

Page 5 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

surfactants or modulators. After the carbonization of the zinc MOFs, a highly porous metal-free carbon rod was obtained and the carbon structure was investigated by Raman, X-ray photoelectron spectroscopy, as well as high-resolution transmission electron microscopy. It was found that the metal-free carbon demonstrated the strong interaction to glycans and the size-exclusion mechanism against proteins. 2. EXPERIMENTAL SECTION 2.1 Materials. Ovalbumin (OVA, chicken egg white), Bovine serum albumin (BSA), 2, 5-dihydroxy-benzoic acid (DHB), sinapinic acid (SA), trifluoroacetic acid (TFA), 1,3,5-benzenetricarboxylate (BTC) were obtained from Sigma-Aldrich. Acetonitrile (ACN) was chromatographic grade from Merck (Darmstadt, Germany). Concentrated nitric acid (HNO3, 65.0 wt %), NH4HCO3, ethyl alcohol, and Zn(OAc)2·2H2O were purchased from Sinopharm Chemical Reagent, Co., Ltd (Shanghai, China). PNGase F (Genetimes Technology) was acquired from New England Biolabs (Ipswich, MA, U.S.A.). The deionized water (18.2 MΩ cm-1) were purified with a Milli-Q water system (Millipore, USA). All the other chemicals are of analytical grade. Human serum from healthy volunteers were provided by Dalian Medical University and stored at -80 °C before analysis. 2.2 Preparation of Zinc Metal-Organic Framework Nanorods. The zinc metal-organic framework (Zn-MOF) nanorods were prepared by the facile coordination of zinc (II) acetate dehydrate and 1,3,5-benzenetricarboxylate at room temperature. 0.55 g zinc (II) acetate dehydrate (2.5 mmol) were dispersed in aqueous solution to form solution A, and 0.525 g 1,3,5-benzenetricarboxylate (BTC, equal to 2.5 mmol) were added into ethyl alcohol to form solution B. Then, solutions of A and B were mixed and carried into an ultrasonic bath (SK2210HP, KUDOS) at frequency of 53 KHz for 30 min at room temperature. The generated white solid was immediately isolated by centrifugation (5000 rpm, 2.0 min), washed with water and ethanol respectively for three 5 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

times, and dried under vacuum at 60 °C overnight. Varied Zn-MOFs were obtained by changing the volume ratios of water/ethanol solvents. 2.3 Preparation of Porous Carbon and other Materials. The porous carbon nanorods were prepared by carbonizing the uniform nanorod Zn-MOF that was synthesized with a well-defined VH2O/Vethanol ratio of 1:7. Specifically, the Zn-MOF nanorods were carbonized with a ramping rate of 5 °C/min in a nitrogen flow and maintained at thermostatic temperature (600 °C, 700 °C, 800 °C, 900 °C, or 1000 °C) for 5 h, respectively. The porous carbon obtained at 900 °C was oxidized with diluted HNO3. Typically, 0.1 g of as-synthesized was refluxed with 50 mL of HNO3 (26.0 wt %) at 90 °C for 1 h, and centrifuged, washed to neutral, dried under vacuum at 60 °C overnight. The ZnO/C material was prepared by carbonizing the as-synthesized Zn-MOF in a nitrogen flow with a ramping rate of 5 °C/min and maintained at 800 °C for 5 h. And, the ZnO material was obtained by heating the as-synthesized Zn-MOF at 600 °C for 1 h in air flow. 2.4 Measurements and Characterizations. Transmission electron microscopy measurements were carried out on a Model JEM-2000 EX (JEOL, Japan) microscope operated at 120 kV. Scanning electron microscopy was performed on a JSM-7800F (Japan) operated at 3.0 KV. Raman spectra were recorded on an in Via spectrometer (Renishaw, Hoffman Estates, IL) equipped with an argon-ion laser at an excitation wavelength of 514 nm. The nitrogen sorption/desorption isotherms were measured at 77 K using an Autosorb iQ2 adsorptometer (Quantachrome Instruments). Powder X-ray diffraction patterns were recorded on an Empyrean XRD system (PANalytical, Almelo) with Cu−Kα radiation (λ = 1.54056 Å) over a 2θ range of 5−80°. Thermo gravimetric analysis was carried out on a simultaneous thermal analyzer (STA 449 F3, NETZSCH) under flowing artificial air or nitrogen air. X-ray photoelectron spectroscopy was recorded on a Thermo Scientific ESCALAB 250Xi system equipped with a dual X-ray source using Al−Kα. 6 / 27

ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

MALDI-TOF-MS analysis results were obtained using a MALDITOF/TOF 5800 System (AB SCIEX, Foster City, CA) equipped with a 1 kHz OptiBeam on-axis laser. 2.5 Releasing of N-Linked Glycans from Proteins. The release of N-linked glycans from proteins was carried out as follows: standard glycoprotein OVA was dissolved in 50 mM NH4HCO3 solution (pH 7.5) and boiled for 5 min. After cooling to room temperature, PNGase F enzyme was added and incubated for 24 h at 37 °C for N-glycan releasing. For mixtures of OVA digest and BSA, the OVA digest was directly mixed with BSA at mass ratios of 1:10 or 1:500. To release N-linked glycans from human serum, the serum samples from healthy volunteers were thawed at 4 °C and centrifuged at 12 000g for 10 min. The collected supernatants were 10 times diluted with 50 mM NH4HCO3 (pH 7.5) solution, and boiled for 5 min. After cooling to room temperature, the samples were treated by an ultrafiltration membrane with MWCO of 10 kDa. The collected proteins on membrane were rinsed by NH4HCO3 buffer for three times and re-dissolved in 50 mM NH4HCO3 solution (10 mM, pH 7.5). The procedures of releasing glycans from collected human serum proteins were the same as for the above standard proteins. 2.6 Enrichment of N-Linked Glycans. A 10 µL of protein digest was diluted with 80 µL of deionized water and mixed with 10 µL of nanomaterial (10 mg/mL) suspension. The mixture was incubated at room temperature for 0.5 h and followed by the removal of supernatant by centrifuging at 20 000 g for 1 min. The precipitate was washed for three times with 100 µL of deionized water. Finally, the adsorbed glycans on material were eluted with 10 µL of 50 % ACN solution. The same procedure was carried out by all contrasted materials. 2.7 MALDI-TOF MS Analysis. DHB (10 mg/mL, in 50% ACN water solution with 10 mM of NaCl) was the MALDI matrix for the analysis of N-linked glycans. Sample aliquots (0.5 µL) were dropped on the 384 spots MALDI plate. After dried with 0.5 µL DHB matrix added, the 7 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 28

sample spots were dried under vacuum prior to TOF-MS analysis. Residual proteins were mixed with saturated SA with the ratio of 1:1. And, a 0.5µL solution was dipped onto the plate and dried at room temperature before the MS analysis. All MALDI-TOF mass spectrometric experiments were carried out using a MALDI TOF/TOF 5800 System (AB SCIEX) equipped with a neodymium: yttrium aluminum garnet (Nd:YAG) laser (355 nm wavelength). The mass spectrometric analyses of glycans were in positive reflection mode and were in positive linear mode for proteins. Mass spectrometric data interpretation and glycoform analysis was carried out by GlycoWorkbench software from Database GlycomeDB. 3. RESULTS AND DISCUSSION The synthesis of porous metal-free carbon (PMFC) was illustrated in Scheme 1a. We first synthesized one-dimensional (1D) Zn-MOF nanorods via a rapid approach at room-temperature

with

the

coordination

of

zinc

acetate

dehydrate

and

1,3,5-benzenetricarboxylate in water/ethanol solvents. Different surfactants have been explored to tune the structure and morphology of Zn-based MOF,46 and Jiang has pioneered to synthesize 1D Zn-BTC rods crystal (Zn3(BTC)2•12H2O) by the assistance of ultrasonic irradiation at room temperature, with diameters of 100–900 nm and lengths of more than 100 µm without use of surfactants.47 Recently, zinc MOF-74 nanorods with smaller diameters of 30–60 nm and shorter lengths of 200–500 nm were obtained by use of salicylic acid as a modulator.48 Herein, avoiding use of surfactants or modulators, a simple method by adjusting the reaction regents ratios was developed to tune the size and morphology of Zn-MOF nanorods.

8 / 27

ACS Paragon Plus Environment

Page 9 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Scheme 1. (a) the preparation procedure and (b) the glycan enrichment of PMFC.

As shown in Figure S1, diverse nanorod-liked micelles were synthesized at room temperature in 30 min via changing the volume ratios of water/ethanol. Interestingly, by the well-controlled volume ratio of water/ethanol at 1:7, the quite uniform rod-like particle was formed with well-defined nanosize of diameter at ca. 30 nm and length/diameter ratio at ca. 10 (Figure S1a). The powder X-ray diffraction (PXRD) analysis was carried out (Figure 1a). The sharp Bragg diffraction peaks were existed at the range of 5-70°, and all the diffraction peaks could be well-matched to the simulated structure of [Zn6(OH)3(BTC)3(H2O)3]·7H2O that was synthesized by a traditional hydrothermal method at 170 °C for several days.49 The strong and sharp Bragg diffraction peaks suggested the as-synthesized Zn-MOF nanorods with high crystallinity and good uniformity.31 The structural stability of the Zn-BTC was examined by immersing the material with different solvents. The PXRD patterns of the treated Zn-BTC were kept as same as before treatment, with the strong diffraction peaks remained regardless of the solvent treatment or not (Figure S2). This indicated the good solvent tolerance and structural stability of the as-synthesized Zn-MOF nanorods. Moreover, the thermal stability of the Zn-BTC was investigated by thermo gravimetric analysis (TGA) in air. Two loss ladders were appeared below 200 °C and 200-300 °C, respectively, which might attribute to free small 9 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

solvent molecules and combined water (Figure S3).50 The followed plateau and none-loss of mass up to 450 °C hinted the highly thermal robustness of this material.51 Combining with above observations, the robust MOF has been synthesized by a quite facile and high-efficiency procedure. (a)

(b) 1200

0

T(t)

(4) metal-free carbon

1000

Intensity (a.u.)

-20

TG (%)

800

(3) ZnO/C

-40

600 400

No mass loss

-60

(2) Synthesized Zn-MOF

Temperature (°C)

TG(t)

200

(1) Simulated Zn-MOF 10

20

30

40

50

60

-80

70

0 0

100

200

2 Theta (degree)

(d)

1500

1000 °C 900 °C 800 °C 700 °C 600 °C

1000

300

400

Time (min)

0.5

1.4 nm

1000 °C 900 °C 800 °C 700 °C 600 °C

0.4

dV/dD(cm3/g/nm)

(c)

Volume@STP (cc/g)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 28

500

0

0.3

0.2

0.1

0.0

0.0

0.2

0.4

0.6

0.8

1.0

0

2

4

6

Relative Pressure (P/P0)

8

10

12

14

16

18

20

Pore size (nm)

Figure 1 (a) PXRD patterns of (1) simulated Zn-BTC MOF, (2) synthesized Zn-MOF nanorods, (3) ZnO/C composite from 800 °C, and (4) metal-free carbon from 900 °C carbonization of Zn-MOF nanorods for 5 h. (b) TGA of Zn-MOF nanorods in nitrogen flow of maintaining at 900 °C for 5 h and followed by raising to 1000 °C. The black and red line were thermo gravimetric percent-time (TG-t) curve and temperature-time (T-t) curve, respectively. (c) Nitrogen adsorption-desorption and (d) pore size distribution curves of carbon materials derived from Zn-MOF nanorods at carbonization of 600 °C, 700 °C, 800 °C, 900 °C, and 1000 °C, respectively.

The thermal treatment of Zn-MOF nanorods was carried out in nitrogen flow to derive 1D carbon material. The carbonization temperature procedure was with a heating rate of 10 / 27

ACS Paragon Plus Environment

Page 11 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

5 °C/min to the terminal point temperature and cooling to the room temperature followed with a thermostatic period of 5 h. The products at different temperatures were monitored by PXRD analyses (Figure 1a,S4). At terminal point temperatures from 600 °C to 800 °C, the resulting carbon materials were with same PXRD patterns and all of their sharp Bragg reflection peaks were well-indexed to the hexagonal phase of zinc oxide (JCPDS card no. 36-1451).52 Thus Zn-MOF nanorods were initially decomposed to ZnO/C composite.53 Base on the ZnO residual mass in air (40.9 wt.%) in Figure S3,5, the zinc mass percentage in the Zn-MOF nanorods was estimated ca. 32.8 wt.% (Calc. 31.3 wt.%). Checking the PXRD pattern of the resulting carbon material at 900 °C, the spectrum was flat and no peaks of metal Zn-related was observed (Figure 1a). Only a weak and broad diffraction peak existed at around 43° which corresponding to the (101) diffraction peak of graphitic carbon (Figure S4).54 The results revealed that a metal-free carbon material was obtained at carbonization of 900 °C for 5 h. A simulated carbonization procedure for deriving Zn-MOF nanorods to carbon at 900 °C was monitored by the TGA (Figure 1b). At the thermostatic period of 900 °C, the mass loss rate was rapid at the first 40 min and the mass loss was ended at ~ 4 h, yet the mass loss could not finished at a contrasted of 800 °C (Figure S6,7). After the 900 °C process finished, a further investigation was performed by raising temperature to 1000 °C and maintained for a period. The mass was not changed at the whole process (Figure 1b). In addition, the surface area and porosity property of the derived carbon materials were characterized by N2 adsorption-desorption measurements. The N2 uptake capacity of carbon materials was increased along with raising the calcination temperature (Figure 1c), with a high Brunauer−Emmett−Teller (BET) surface areas (417-1700 m2 g−1, summarized in Table S1). All the carbon materials showed a narrow pore size mainly centered at 1.4 nm with slight mesopores (Figure 1d). The metal-free carbon obtained at 900 °C and 1000 °C exhibited approximate BET surface areas of 1700 m2 g−1 and 1673 m2 g−1 , respectively. Which surface areas were significantly improved than ZnO/C composite (at 800 °C) of 653 m2 g−1. Considering the experiment operability and friendly to calcinator equipment in lab, the 11 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

application of porous metal-free carbon (PMFC) obtained at 900 °C was further explored.

Figure 2. .TEM images of as-synthesized (a) Zn-BTC, (b) PMFC, and (c) oPMFC. SEM images of PMFC (d), and oPMFC (g). High resolution TEM images of (e, f) PMFC and (h, i) oPMFC.

To minimize the possible loss during subsequent sample treatment, as carbon materials keen on adhering EP tubes, a slight oxidation of PMFC was operated,23 and the resulting material was endowed as oPMFC. The surface area and pore structure of oPMFC were examined by N2 adsorption-desorption analysis (Figure S8,9). The results indicated the oPMFC was with a high specific surface area of 1670 m2/g and a concentrated pore size at 1.4 nm in consistent with PMFC. The morphology of oPMFC was investigated and contrasted with Zn-BTC as well as the PMFC by applying the transmission electron microscopy (TEM) and scanning electron microscopy (SEM) measurements. The PMFC (Figure 2b, d) and the oPMFC (Figure 2c, g) were all with uniformly rod-like morphology and maintained the inherent skeleton structure of the as-synthesized Zn-BTC (Figure 2a). But a slight decrease of the rod size of the PMFC compared with Zn-BTC could be observed, which due to the commonly shrinkage of the MOFs precursor during carbonization process.55 Furthermore, 12 / 27

ACS Paragon Plus Environment

Page 12 of 28

Page 13 of 28

comparing both TEM and SEM images of oPMFC and PMFC, no distinct shapes or surface structure changes were observed before and after oxidation. As already be certified by the above PXRD studies, the PMFC was with obvious graphitic carbon component. Also could be observed from the high-resolution TEM (HRTEM) images, the graphitic layers were dispersed randomly in the PMFC (Figure 2e, f) and oPMFC (Figure 2h, i). This was because the carbon derived from the organic ligands could be further oriented to the graphitic structure at a period of high temperature. In addition, the carbon structure of the PMFC and oPMFC were confirmed by Raman Spectroscopy (Figure 3a), the strong G-band peak along with a weak and broad 2D-band of graphitic carbon located at ~1600 cm-1 and ~2900 cm-1, respectively.56 Another distinct broad peak of ~1360 cm-1 assigned to the D-band of amorphous carbon.57 Compared to the single-layer graphene (2D band at ~2700 cm-1), the 2D-band of (o)PMFC was broad and exhibited a shifted location, which might attributed to the stacked effect of multilayered graphitic carbon and the inter-dispersed porosity structure.56,58 The results were consistent with the previous HRTEM and PXRD analysis. The band intensity ratio of ID/IG is widely used to evaluate the quality of carbon materials. As observed, the measured value of ID/IG for PMFC was approximate 0.83 which suggested that the PMFC derived from the Zn-BTC was a highly graphitized carbon. As the structure stability of carbon, the ID/IG value of the oPMFC was cal. 0.82 which was consistent with the PMFC. 5

(b)

C 1s

G

4

D 3

2D

2

curve 2

Intensity (a.u.)

(a) Intensity (*1000a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

O 1s

curve 1

1

curve 2

curve 1 0 500

1000

1500

2000

2500

3000

3500

0

200

-1 Raman shift (cm )

400

600

Binding Energy (eV)

13 / 27

ACS Paragon Plus Environment

800

ACS Applied Materials & Interfaces

(c)

(d)

2

PMFC

C-C sp

oPMFC

C-C sp

3

C-C sp

C-O C=O -COO

280

282

284

286

288

satallites

290

292

2

C-C sp

Intensity (a.u.)

Intensity (a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 28

3

C-O C=O -COO

280

282

Binding Energy (eV)

284

286

288

satallites

290

292

Binding Energy (eV)

Figure 3. .(a) Raman spectra of (curve 1) PMFC and (curve 2) oPMFC. (b) Typical survey scanned XPS spectrum of (curve 1) PMFC and (curve 2) oPMFC. The deconvoluted C1s spectrum of (c) PMFC and (d) oPMFC.

Besides, X-ray photoelectron spectroscopy (XPS) analysis was employed for more information about the features of the as-synthesized carbon materials. No metal elements were observed in the spectra of both PMFC and oPMFC (Figure 3b), and the two carbon materials both featured prominent C 1s peak at ∼284 eV and O 1s peak at ∼532 eV.59 Observing an obvious intensity change of O 1s peak between PMFC (Figure 3b, curve 1) and oPMFC (Figure 3b, curve 2), the oxygen atomic ratio was increased from 5.22% (PMFC, curve 1) to 15.78% (oPMFC, curve 2), which revealed that several oxygenic groups were introduced on the oPMFC surface. The high-resolution XPS of the deconvolution C 1s signals were investigated. Figure 3c and Figure 3d were for the PMFC and the oPMFC, respectively. As documented, two peaks at 284.6 and 285.6 eV, respectively, assigned to graphite-like carbon (C–C sp2) and diamond-like carbon (C–C sp3) dominated the spectra.59 According to the area ratio estimated, the graphitic carbon content of the materials was as high as ~ 70%. Correspondingly, the peaks centered at 286.3 eV, 287.4 eV and 289.0 eV were attributed to surface oxygen groups (designated as C–O, C=O and O=C–O, respectively). And the appearance of p–p* shake up satellites of sp2 graphite-like carbon above 290 eV indicated the further carbonization and aromatization during the prolonged carbonization time of 900 °C.48 The above results indicated that the Zn-BTC derived PMFC was born with high graphitic 14 / 27

ACS Paragon Plus Environment

Page 15 of 28

carbon content, which would be a potential seed for strong interaction with glycan molecules. In addition, by a facile oxidation of the PMFC, hydrophilic groups were demonstrated to be increased on the oPMFC material surface, and also can be seen in Figure S10, the water-dispersibility of oPMFC were significantly improved than PMFC. Which improvement was great importance for sample preparation and could minimize the possible sample loss during enrichment. As the strong interaction between carbon and glycan, and the relatively narrow nanopore size distribution might provide the size exclusion effect against large-proteins. Therefore, the high specific surface area and suitable nanopore size endowed the oPMFC to be a promising material for extracting glycans from complex biological samples.

1000

(a) before enrichment

15000

(b) ZnO/C

800 10000

600 400

3

16 20 91013 1819 23 14 12 22 24 25

6

200 Intensity (a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

0 15000

1000

1500

(c) PMFC

2000

5000

3 4

1 2

2500

0

16 6 7 9 19 13 10 5 8 14 1820 17 212324 25 26 1112 15 22

1000

1500

15000

m/z

2000 m/z

3

2500

(d) oPMFC

16

10000

10000 1

6 4

7 20 10 13 14 9 18 19 23 8 24 1112 25 21 15 17 26 22

3 16

5000 1 2

0

1000

5000 4 6 7 20 13 910 14 18 5 19 23 8 12 11 15 17 212224 25 26 1500

2000

2500

0

5 2

1000

1500

m/z

2000

2500

m/z

Figure 4. .MALDI-TOF MS spectra of N-glycans released from 1 µg µL-1 ovalbumin digestion before enrichment (a) and after enrichment (b) with ZnO/C-800, (c) with PMFC, and (d) with oPMFC. The captured N-glycans were marked with digital labels.

A standard glycoprotein digest (chicken ovalbumin, OVA, treated with PNGase F) 15 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

was employed to evaluate the N-linked glycans enrichment ability of the as-synthesized carbon materials. Before enrichment, there was only 15 N-linked glycans detected and with very low signal to noise (S/N) ratios (Figure 4a), this is due to the residual proteins interfering the MS analysis. However, after the oPMFC enrichment, 26 N-linked glycans signals were clearly observed with high S/N ratio and the significant signal intensity was about 35 times enhanced comparing with direct analysis (Figure 4d). The glycoforms of the identified N-linked glycans from OVA digest were listed in the Table S2. To grade the glycan enrichment abilities, we contrasted the oPMFC with other materials. Firstly, we investigated the ability of the ZnO/C composite (Figure 4b) and the PMFC (Figure 4c), both 26 N-linked glycans could be detected after enriched, but it is clearly that PMFC outperformed ZnO/C both the S/N ratios and signal intensities. As another investigation, none N-linked glycans were captured by ZnO material (Figure S11). We speculated that, after removing the ZnO component, both the larger specific surface area and the higher carbon content contributed the higher enrichment efficiency of the PMFC. Additionally, as the PMFC inferior to the oPMFC, which was because of the improved water-dispersibility of oPMFC made the sample handling easily and a slight oxidation surface might also increase polar interaction with glycan molecules. It was mentioning that, the oPMFC exhibited absolutely superiority to commercial active carbon (Figure S12). Therefore, the oPMFC was demonstrated to be a promising enrichment material for its strong interaction toward glycan molecules. Enrichment efficiency and sensitivity were important factors to evaluate the extraction substrate. Therefore, the limit of detection for oPMFC enrichment was further examined by declining the OVA digest concentrations. At a low concentration of 10 ng µL-1 OVA digest (Figure 5a), 26 N-linked glycans were detected with significantly enhanced S/N ratios and intensities after enrichment by the oPMFC. As the OVA digest decreased to 5 ng µL-1 (Figure 5b), the intensities of enriched glycans declined slightly but still with 24 glycans observed clearly. When the OVA digest concentration was as low 16 / 27

ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

as 1 ng µL-1 (Figure 5c), only 1 glycan was observed by direct MALDI-TOF MS analysis. Nevertheless, after enriched by using of oPMFC, 21 N-linked glycan signals were clearly detected (with S/N >10). Therefore, the oPMFC exhibited its high efficiency and sensitivity of N-glycan enrichment. 8000 6000

(a) 10 ng µL-1

Before enrichment

Intensity (a.u.)

4000 2000

3

4

6

10

7

13

16 14

181920

23

24

25

0 1000

8000

1500

2000

6000

10

7

13

4

4000 1

8

5

2

24 19 20 18 17 212223

14

6

3

2500

After enrichment

16

2000

9

11 12

15

25 26

0 1000

1500

2000

2500

m/z 3000

(b) 5 ng µL-1

Before enrichment

Intensity (a.u.)

2000 1000

4

0 3000 2000

10

7

1000

16 13

1500

3

19 18 20

9

11 12

8

25

2500

After enrichment

13 14

6

2

24

2000

16

10

7

4

1

1000

24

19 18 20 21 23 22

17

25 26

0 1000

1500

2000

2500

m/z 3000

(c) 1 ng µL-1

Before enrichment

2000

Intensity (a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

3

1000 0 1000

3000

1500

3

m/z

2000

2500

After enrichment

1

2000 4

1000

56

7

9

16 10 13 11 12 14

181920

24 22 23

25

26

0 1000

1500

2000

2500

m/z

Figure 5. . MALDI-TOF MS for N-Glycan before and after enriched by using oPMFC from 17 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ovalbumin digest with different concentrations of (a) 10 ng µL-1, (b) 5 ng µL-1, and (c) 1 ng µL-1.

Figure 6. .MALDI-TOF MS spectra for glycans and proteins from the mixture sample of the OVA digest and BSA protein at the mass ratio of 1:10 (a) direct analysis and (b) after enrichment with the oPMFC. The captured N-glycan was marked with a * label.

As the unique nanopore structure of the oPMFC, the size-exclusion effect of the material was demonstrated by a complex sample. Specifically, we investigated the selective ability in capturing N-linked glycans with interfering-proteins existed. Firstly, a mixture of OVA digest mingled with BSA protein at a mass ratio of 1:10 was tested. Before enrichment, no peaks of glycans (Figure 6a) but only protein signals of the BSA and the residual OVA (the inset of Figure 6a) were recognized. Yet, after enriched by the oPMFC (Figure 6b), 25 peaks of N-glycan were clearly detected with significant signal intensities and high S/N ratios. In the elution of the oPMFC enrichment, no proteins were detected (in the inset of Figure 6b); and in the supernatant after enrichment, only size-excluded proteins (BSA and residual OVA) were detected (Figure S13). That is to say, the oPMFC selectively enriched the N-linked glycans from the complex sample but with large-sized proteins excluded. Then, the selective enrichment efficiency of the oPMFC was further contrasted with commercial active carbon. A more complex mixture with higher abundant proteins (the mass ratio of OVA digest to BSA was 1: 500) were employed as test model. By using the oPMFC profiling (Figure 7a), 24 N-linked glycans from OVA digest were undoubtedly identified with strong S/N ratios. As for the active 18 / 27

ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28

carbon (Figure 7b), only 12 N-linked glycans were detected with weak signal intensities and lower S/N ratios. Again, the oPMFC exhibited its advantage in selective enrichment of glycans. Based on above observation, the as-synthesized oPMFC demonstrated its strong interaction with glycans as well as excellent anti-proteins interfering ability, which endowed it tremendous potential for the application of real samples analysis.

1200

*

(a) oPMFC

* 400

* 0 1000

*

** * ** * * * * * * * * * 1500

(b) active carbon

1200

*

* **

800

1600

*

*

*

Intensity (a.u.)

1600

Intensity (a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

2000

2500

800

*

400

** * * **** * * 0 1000

m/z

1500

2000

2500

m/z

Figure 7. .MALDI-TOF MS analysis of N-glycans released from a mixture of OVA digest with BSA at mass ratio of 1:500 after enrichment with (a) the oPMFC and (b) the active carbon. Peaks marked with a “*” symbol are the signals of N-linked glycans.

As the significantly physiological importance of glycan analysis from human serum which are closely related with many diseases.60 However, the isolation and identification of human serum glycans are challenging due to the complexity of the system. Encouraged by the excellent performance of the oPMFC for N-glycan enrichment, we utilized the oPMFC to trap N-glycans released from human serum. None signals of the N-linked glycans were observed by direct analysis of the PNGase F treated human serum sample (Figure S14). After the oPMFC enrichment, 43 N-linked glycans could be clearly identified with significant S/N ratios for a 0.2 µL healthy human serum sample (Figure 8), which attributed to the strong adsorption capacity of the oPMFC to glycan but also the 19 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

size-exclusion effect of oPMFC against large-sized proteins. The detailed information of these N-linked glycans was displayed in Table S3. In a word, the oPMFC derived from the as-synthesized Zn-BTC was an excellent candidate for the highly sensitive and selective enrichment of glycan, which is broadening the way of high-throughput glycans analysis from complex biological samples.

Figure 8. .MALDI-TOF MS analysis of 43 N-linked glycans released from human serum proteins after enrichment by the oPMFC.

4. CONCLUSIONS In summary, it was demonstrated that the rapid room-temperature synthesis of metal-organic framework provided an efficient approach to fabricate highly porous carbon material. For the first time, the well-defined zinc MOFs with nanosized and rod-shaped morphology was directly prepared without use of any surfactants or modulators and the entire process was accomplished in 30 minutes. Additionally, the route was facile and low cost. The porous metal-free carbon (PMFC) was derived from the synthesized zinc MOFs via one-step carbonization procedure. At the heating treatment of high temperature, zinc MOFs were self-sacrificed with organic ligand pyrolysis and metal-eliminating, thus oriented to a 3D porous and graphitic carbon 20 / 27

ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

structure but remained the rod-liked morphology. The resultant PMFC exhibited a large surface area and demonstrated to have strong interaction with glycan molecules. Furthermore, the uniform nanopore structure of PMFC endowed it selectively captured glycan molecules but large proteins excluded. The PMFC demonstrated its high sensitivity and strong anti-fouling ability for glycan profiling in biosamples. As a consequence, the simple method for fabrication of MOFs and porous carbon would give new insight into material research as well as biosamples analysis. ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website. The images of TEM (Figure S1), PXRD (Figure S2), and TGA (Figure S3,6,7) for the as-synthesized Zn-BTC nanorods. The PXRD of carbon materials obtained at different conditions (Figure S4,5). Nitrogen adsorption-desorption and pore size distribution of the oPMFC (Figure S8,9). The aqueous dispersion of PMFC and oPMFC (Figure S10). The glycan enrichment of ZnO ( Figure S11), active carbon ( Figure S12), and oPMFC ( Figure S13). Figure S14 was for the MALDI TOF-MS spectrum of direct analysis of human serum glycan. Summary of BET surface areas for carbon materials of different temperatures (Table S1). Table S2 and Table S3 were for the glycan information of enriched by oPMFC from OVA digest and human serum, respectively. AUTHOR INFORMATION Corresponding Author ∗

Ren’an Wu, Email:[email protected]

Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes The authors declare no competing financial interest. 21 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 28

ACKNOWLEDGMENTS The financial supports from the National Natural Science Foundation of China (Nos. 21175134, 21375125, 21505134, and 21675156), the Instrument Developing Project of the Chinese Academy of Sciences (YZ201503), the CAS Key Laboratory Foundation of Separation Science for Analytical Chemistry and the Innovation Program (DICP TMSR201601) of Science and Research from the Dalian Institute of Chemical Physics are greatly acknowledged. REFERENCES (1) Dennis, J. W.; Nabi, I. R.; Demetriou, M. Metabolism, Cell Surface Organization, and Disease. Cell 2009, 139, 1229-1241. (2) Pinho, S. S.; Reis, C. A. Glycosylation in cancer: mechanisms and clinical implications. Nat. Rev. Cancer 2015, 15, 540-555. (3) Helenius, A.; Aebi, M. Intracellular functions of N-linked glycans. Science 2001, 291, 2364-2369. (4) Sethi, M. K.; Hancock, W. S.; Fanayan, S. Identifying N-Glycan Biomarkers in Colorectal Cancer by Mass Spectrometry. Accounts. Chem. Res. 2016, 49, 2099-2106. (5) Wang, J. R.; Gao, W. N.; Grimm, R.; Jiang, S. B.; Liang, Y.; Ye, H.; Li, Z. G.; Yau, L. F.; Huang, H.; Liu, J.; Jiang, M.; Meng, Q.; Tong, T. T.; Huang, H. H.; Lee, S.; Zeng, X.; Liu, L.; Jiang, Z. H. A method to identify trace sulfated IgG N-glycans as biomarkers for rheumatoid arthritis. Nat. Commun. 2017, 8, 1-14. (6) Liu, T.; Qian, W. J.; Gritsenko, M. A.; Camp, D. G.; Monroe, M. E.; Moore, R. J.; Smith, R. D. Human plasma N-glycoproteome analysis by immunoaffinity subtraction, hydrazide chemistry, and mass spectrometry. J. Proteome Res. 2005, 4, 2070-2080. (7) Tousi, F.; Hancock, W. S.; Hincapie, M. Technologies and strategies for glycoproteomics and glycomics and their application to clinical biomarker research. Anal. Methods 2011, 3, 20-32. (8) Alley, W. R.; Mann, B. F.; Novotny, M. V. High-sensitivity Analytical Approaches for the Structural Characterization of Glycoproteins. Chem. Rev. 2013, 113, 2668-2732. (9) Jayo, R. G.; Thaysen-Andersen, M.; Lindenburg, P. W.; Haselberg, R.; Hankemeier, T.; Ramautar, R.; Chen, D. D. Y. Simple Capillary Electrophoresis-Mass Spectrometry Method for Complex Glycan Analysis Using a Flow-Through Microvial Interface. Anal. Chem. 2014, 86, 6479-6486. (10) Krenkova, J.; Lacher, N. A.; Svec, F. Multidimensional system enabling deglycosylation of proteins using a capillary reactor with peptide-N-glycosidase F immobilized on a porous polymer monolith and hydrophilic interaction liquid chromatography-mass spectrometry of glycans. J. Chromatogr. A 2009, 1216, 3252-3259. (11) Shubhakar, A.; Kozak, R. P.; Reiding, K. R.; Royle, L.; Spencer, D. I. R.; Fernandes, D. L.; Wuhrer, M. Automated

High-Throughput

Permethylation

for

Glycosylation

Analysis

of

Biologics

Using

MALDI-TOF-MS. Anal. Chem. 2016, 88, 8562-8569. (12) Liu, M. Q.; Zeng, W. F.; Fang, P.; Cao, W. Q.; Liu, C.; Yan, G. Q.; Zhang, Y.; Peng, C.; Wu, J. Q.; Zhang, X. J.; Tu, H. J.; Chi, H.; Sun, R. X.; Cao, Y.; Dong, M. Q.; Jiang, B. Y.; Huang, J. M.; Shen, H. L.; 22 / 27

ACS Paragon Plus Environment

Page 23 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Wong, C. C. L.; He, S. M.; Yang, P. Y. pGlyco 2.0 enables precision N-glycoproteomics with comprehensive quality control and one-step mass spectrometry for intact glycopeptide identification. Nat. Commun. 2017, 8, 1-14. (13) Aizpurua-Olaizola, O.; Sastre Toraño, J.; Falcon-Perez, J. M.; Williams, C.; Reichardt, N.; Boons, G. J. Mass spectrometry for glycan biomarker discovery. TrAC, Trends Anal. Chem. 2018, 100, 7-14. (14) Yang, Y.; Franc, V.; Heck, A. J. R. Glycoproteomics: A Balance between High-Throughput and In-Depth Analysis. Trends Biotechnol. 2017, 35, 598-609. (15) Diop, N. K.; Hrycyna, C. A. N-linked glycosylation of the human ABC transporter ABCG2 on asparagine 596 is not essential for expression, transport activity, or trafficking to the plasma membrane. Biochemistry. 2005, 44, 5420-5429. (16) An, H. J.; Peavy, T. R.; Hedrick, J. L.; Lebrilla, C. B. Determination of N-glycosylation sites and site heterogeneity in glycoproteins. Anal. Chem. 2003, 75, 5628-5637. (17) Chen, S. M.; LaRoche, T.; Hamelinck, D.; Bergsma, D.; Brenner, D.; Simeone, D.; Brand, R. E.; Haab, B. B. Multiplexed analysis of glycan variation on native proteins captured by antibody microarrays. Nat. Methods 2007, 4, 437-444. (18) Liu, Z.; He, H. Synthesis and Applications of Boronate Affinity Materials: From Class Selectivity to Biomimetic Specificity. Accounts. Chem. Res. 2017, 50, 2185-2193. (19) You, M.; Yang, S.; Tang, W. X.; Zhang, F.; He, P. G. Ultrasensitive Electrochemical Detection of Glycoprotein Based on Boronate Affinity Sandwich Assay and Signal Amplification with Functionalized SiO2@Au Nanocomposites. ACS Appl. Mater. Interfaces 2017, 9, 13855-13864. (20) Fan, J. Q.; Kondo, A.; Kato, I.; Lee, Y. C. High-Performance Liquid-Chromatography of Glycopeptides and Oligosaccharides on Graphitized Carbon Columns. Anal. Biochem. 1994, 219, 224-229. (21) Sun, N. R.; Yao, J. Z.; Deng, C. H. Designed synthesis of carbon-functional magnetic graphene mesoporous silica materials using polydopamine as carbon precursor for the selective enrichment of N-linked glycan. Talanta 2016, 148, 439-443. (22) Ge, S. G.; Lan, F. F.; Lang, L. L.; Ren, N.; Li, L.; Liu, H. Y.; Yan, M.; Yu, J. H. Ultrasensitive Photoelectrochemical Biosensing of Cell Surface N-Glycan Expression Based on the Enhancement of Nanogold-Assembled Mesoporous Silica Amplified by Graphene Quantum Dots and Hybridization Chain Reaction. ACS Appl. Mater. Interfaces 2017, 9, 6670-6678. (23) Qin, H. Q.; Zhao, L.; Li, R. B.; Wu, R. A.; Zou, H. F. Size-Selective Enrichment of N-Linked Glycans Using Highly Ordered Mesoporous Carbon Material and Detection by MALDI-TOF MS. Anal. Chem. 2011, 83, 7721-7728. (24) Sun, N.; Deng, C.; Li, Y.; Zhang, X. Highly selective enrichment of N-linked glycan by carbon-functionalized ordered graphene/mesoporous silica composites. Anal. Chem. 2014, 86, 2246-2250. (25) Zhang, Q. Q.; Huang, Y. Y.; Sun, H. F.; Gan, Y. Y.; Shen, H. L.; Jia, B.; Yang, P. Y. Facile preparation of 3-D floor-like ordered mesoporous carbon functionalized graphene composites and its application for selective enrichment of N-glycans from human serum. Talanta 2017, 174, 689-695. (26) Qin, H.; Hu, Z.; Wang, F.; Zhang, Y.; Zhao, L.; Xu, G.; Wu, R.; Zou, H. Facile preparation of ordered mesoporous silica-carbon composite nanoparticles for glycan enrichment. Chem. Commun. 2013, 49, 5162-5164. (27) Xu, L. N.; Qi, X. Y.; Li, X. J.; Bai, Y.; Liu, H. W. Recent advances in applications of nanomaterials for 23 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

sample preparation. Talanta 2016, 146, 714-726. (28) Lee, J.; Kim, J.; Hyeon, T. Recent progress in the synthesis of porous carbon materials. Adv. Mater. 2006, 18, 2073-2094. (29) Liang, C. D.; Li, Z. J.; Dai, S. Mesoporous carbon materials: Synthesis and modification. Angew. Chem., Int. Ed. 2008, 47, 3696-3717. (30) Ryoo, R.; Joo, S. H.; Kruk, M.; Jaroniec, M. Ordered mesoporous carbons. Adv. Mater. 2001, 13, 677-681. (31) Li, H.; Eddaoudi, M.; O'Keeffe, M.; Yaghi, O. M. Design and synthesis of an exceptionally stable and highly porous metal-organic framework. Nature 1999, 402, 276-279. (32) Furukawa, H.; Cordova, K. E.; O'Keeffe, M.; Yaghi, O. M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 974-987. (33) Jiang, H. L.; Liu, B.; Lan, Y. Q.; Kuratani, K.; Akita, T.; Shioyama, H.; Zong, F. Q.; Xu, Q. From Metal-Organic Framework to Nanoporous Carbon: Toward a Very High Surface Area and Hydrogen Uptake. J. Am. Chem. Soc. 2011, 133, 11854-11857. (34) Gao, J. K.; Miao, J. W.; Li, P. Z.; Teng, W. Y.; Yang, L.; Zhao, Y. L.; Liu, B.; Zhang, Q. C. A p-type Ti( IV)-based metal-organic framework with visible-light photo-response. Chem. Commun. 2014, 50, 3786-3788. (35) Chaikittisilp, W.; Ariga, K.; Yamauchi, Y. A new family of carbon materials: synthesis of MOF-derived nanoporous carbons and their promising applications. J. Mater. Chem. 2013, 1, 14-19. (36) Cao, X. H.; Tan, C. L.; Sindoro, M.; Zhang, H. Hybrid micro-/nano-structures derived from metal-organic frameworks: preparation and applications in energy storage and conversion. Chem. Soc. Rev. 2017, 46, 2660-2677. (37) Wang, J. P.; Cong, J. K.; Xu, H.; Wang, J. M.; Liu, H.; Liang, M.; Gao, J. K.; Ni, Q. Q.; Yao, J. M. Facile Gel-Based Morphological Control of Ag/g-C3N4 Porous Nanofibers for Photocatalytic Hydrogen Generation. ACS Sustainable Chem. Eng. 2017, 5, 10633-10639. (38) Rubio-Martinez, M.; Avci-Camur, C.; Thornton, A. W.; Imaz, I.; Maspoch, D.; Hill, M. R. New synthetic routes towards MOF production at scale. Chem. Soc. Rev. 2017, 46, 3453-3480. (39) Kaneti, Y. V.; Tang, J.; Salunkhe, R. R.; Jiang, X. C.; Yu, A. B.; Wu, K. C. W.; Yamauchi, Y. Nanoarchitectured Design of Porous Materials and Nanocomposites from Metal-Organic Frameworks. Adv. Mater. 2017, 29, 1604898-1604938. (40) Guo, C. Y.; Zhang, Y. H.; Guo, Y.; Zhang, L. G.; Zhang, Y.; Wang, J. D. A general and efficient approach for tuning the crystal morphology of classical MOFs. Chem. Commun. 2018, 54, 252-255. (41) Wang, Y. A.; Wang, J. X.; Gao, M. X.; Zhang, X. M. A novel carbon material with nanopores prepared using a metal-organic framework as precursor for highly selective enrichment of N-linked glycans. Anal. Bioanal. Chem. 2017, 409, 431-438. (42) Pandiaraj, S.; Aiyappa, H. B.; Banerjee, R.; Kurungot, S. Post modification of MOF derived carbon via g-C3N4 entrapment for an efficient metal-free oxygen reduction reaction. Chem. Commun. 2014, 50, 3363-3366. (43) Zhu, C. Z.; Li, H.; Fu, S. F.; Du, D.; Lin, Y. H. Highly efficient nonprecious metal catalysts towards oxygen reduction reaction based on three-dimensional porous carbon nanostructures. Chem. Soc. Rev. 2016, 45, 517-531. 24 / 27

ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(44) Yip, T. T.; Nakagawa, Y.; Porath, J. Evaluation of the Interaction of Peptides with Cu(Ii), Ni(Ii), and Zn(Ii) by High-Performance Immobilized Metal-Ion Affinity-Chromatography. Anal. Biochem. 1989, 183, 159-171. (45) Darnault, C.; Volbeda, A.; Kim, E. J.; Legrand, P.; Vernede, X.; Lindahl, P. A.; Fontecilla-Camps, J. C. Ni-Zn-[Fe-4-S-4] and Ni-Ni-[Fe-4-S-4] clusters in closed and open subunits of acetyl-CoA synthase/carbon monoxide dehydrogenase. Nat. Struct. Biol. 2003, 10, 271-279. (46) Gao, J. K.; Ye, K. Q.; Yang, L.; Xiong, W. W.; Ye, L.; Wang, Y.; Zhang, Q. C. Growing Crystalline Zinc-1,3,5-benzenetricarboxylate Metal-Organic Frameworks in Different Surfactants. Inorg. Chem. 2014, 53, 691-693. (47) Qiu, L. G.; Li, Z. Q.; Wu, Y.; Wang, W.; Xu, T.; Jiang, X. Facile synthesis of nanocrystals of a microporous metal-organic framework by an ultrasonic method and selective sensing of organoamines. Chem. Commun. 2008, 3642-3644. (48) Pachfule, P.; Shinde, D.; Majumder, M.; Xu, Q. Fabrication of carbon nanorods and graphene nanoribbons from a metal-organic framework. Nat. Chem. 2016, 8, 718-724. (49) Fu, Y.; Li, G. B.; Liao, F. H.; Xiong, M.; Lin, J. H. Two novel transition metal-organic frameworks based on 1,3,5-benzenetricarboxylate ligand: Syntheses, structures and thermal properties. J. Mol. Struct. 2011, 1004, 252-256. (50) Tamames-Tabar, C.; Imbuluzqueta, E.; Guillou, N.; Serre, C.; Miller, S. R.; Elkaim, E.; Horcajada, P.; Blanco-Prieto, M. J. A Zn azelate MOF: combining antibacterial effect. Crystengcomm 2015, 17, 456-462. (51) Gu, Z. Y.; Wang, G.; Yan, X. P. MOF-5 Metal-Organic Framework as Sorbent for In-Field Sampling and Preconcentration in Combination with Thermal Desorption GC/MS for Determination of Atmospheric Formaldehyde. Anal. Chem. 2010, 82, 1365-1370. (52) Liu, X.; Wu, X. H.; Cao, H.; Chang, R. P. H. Growth mechanism and properties of ZnO nanorods synthesized by plasma-enhanced chemical vapor deposition. J. Appl. Phys. 2004, 95, 3141-3147. (53) Liu, B.; Shioyama, H.; Akita, T.; Xu, Q. Metal-organic framework as a template for porous carbon synthesis. J. Am. Chem. Soc. 2008, 130, 5390-5391. (54) Kim, T. W.; Park, I. S.; Ryoo, R. A synthetic route to ordered mesoporous carbon materials with graphitic pore walls. Angew. Chem., Int. Ed. 2003, 42, 4375-4379. (55) Shih, Y. H.; Fu, C. P.; Liu, W. L.; Lin, C. H.; Huang, H. Y.; Ma, S. Nanoporous Carbons Derived from Metal-Organic Frameworks as Novel Matrices for Surface-Assisted Laser Desorption/Ionization Mass Spectrometry. Small 2016, 12, 2057-2066. (56) Xiang, Z. H.; Xue, Y. H.; Cao, D. P.; Huang, L.; Chen, J. F.; Dai, L. M. Highly Efficient Electrocatalysts for Oxygen Reduction Based on 2D Covalent Organic Polymers Complexed with Non-precious Metals. Angew. Chem., Int. Ed. 2014, 53, 2433-2437. (57) Li, X.; Xu, G. J.; Zhang, H. Y.; Liu, S. J.; Niu, H.; Peng, J. X.; Wu, J.; Wu, R. A. A homogeneous carbon nanosphere film-spot: For highly efficient laser desorption/ionization of small biomolecules. Carbon 2017, 121, 343-352. (58) Zhong, S.; Zhan, C. X.; Cao, D. P. Zeolitic imidazolate framework-derived nitrogen-doped porous carbons as high performance supercapacitor electrode materials. Carbon 2015, 85, 51-59. (59) Kosynkin, D. V.; Higginbotham, A. L.; Sinitskii, A.; Lomeda, J. R.; Dimiev, A.; Price, B. K.; Tour, J. M. Longitudinal unzipping of carbon nanotubes to form graphene nanoribbons. Nature 2009, 458, 872-876. 25 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(60) Drabik, A.; Bodzon-Kulakowska, A.; Suder, P.; Silberring, J.; Kulig, J.; Sierzega, M. Glycosylation Changes in Serum Proteins Identify Patients with Pancreatic Cancer. J. Proteome Res. 2017, 16, 1436-1444.

26 / 27

ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Highly porous metal-free graphitic carbon derived from metal-organic framework for profiling of N-linked glycans Xin Li a,b, Guiju Xu a,b, Jiaxi Peng a,b, Shengju Liu a,b, Hongyan Zhang a,b, Jiawei Maoa,b, Huan Niu a,b, Wenping Lv a, Xingyun Zhao a,b, Ren’an Wua,*

a

Laboratory of High-Resolution Mass Spectrometry Technologies, CAS Key Laboratory

of Separation Science for Analytical Chemistry, Dalian Institute of Chemical Physics, Chinese Academy of Sciences (CAS), Dalian 116023, China b

The University of Chinese Academy of Sciences, Beijing 100049, China

Table of Contents Graphic

27 / 27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Highly porous metal-free graphitic carbon derived from metal-organic framework for profiling of N-linked glycans 193x114mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 28 of 28