这是用户在 2024-10-16 22:59 为 https://pubs.acs.org/doi/10.1021/acscatal.3c02736?fig=sch1&ref=pdf 保存的双语快照页面,由 沉浸式翻译 提供双语支持。了解如何保存?

Pair your accounts. 配对您的帐户。

Export articles to Mendeley
导出文章至 Mendeley

Get article recommendations from ACS based on references in your Mendeley library.
根据 Mendeley 文库中的参考文献从 ACS 获取文章推荐。

Pair your accounts. 配对您的帐户。

Export articles to Mendeley
导出文章至 Mendeley

Get article recommendations from ACS based on references in your Mendeley library.
根据 Mendeley 文库中的参考文献从 ACS 获取文章推荐。

You’ve supercharged your research process with ACS and Mendeley!
您已经用 ACS 和 Mendeley 为您的研究过程注入了活力!

STEP 1: 步骤1:
Click to create an ACS ID
单击以创建 ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.
请注意:如果您切换到其他设备,系统可能会要求您仅使用 ACS ID 再次登录。

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.
请注意:如果您切换到其他设备,系统可能会要求您仅使用 ACS ID 再次登录。

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.
请注意:如果您切换到其他设备,系统可能会要求您仅使用 ACS ID 再次登录。

MENDELEY PAIRING EXPIRED MENDELEY 配对已过期
Your Mendeley pairing has expired. Please reconnect
您的 Mendeley 配对已过期。请重新连接
ACS Publications. Most Trusted. Most Cited. Most Read
Catalytic Activation of Molecular Oxygen Toward Producing Hydroxyl Radicals Controllably for Highly Selective Oxidation of Hydroxyl Compounds under Mild Conditions
My Activity
  • Subscribed
Research Article
研究论文

Catalytic Activation of Molecular Oxygen Toward Producing Hydroxyl Radicals Controllably for Highly Selective Oxidation of Hydroxyl Compounds under Mild Conditions
在温和条件下催化氧以产生羟基自由基以产生羟基自由基,用于羟基化合物的高选择性氧化
Click to copy article link
点击复制文章链接
Article link copied!

  • Xin Liu 刘欣
    Xin Liu
    State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. China
    University of Chinese Academy of Sciences, Beijing 100049, P. R. China
    More by Xin Liu
  • Hong Ma* 马宏*
    Hong Ma
    State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. China
    *Email: mahong@dicp.ac.cn
    More by Hong Ma
  • Meiyun Zhang 张美云
    Meiyun Zhang
    State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. China
    University of Chinese Academy of Sciences, Beijing 100049, P. R. China
    More by Meiyun Zhang
  • Penghua Che 车鹏华
    Penghua Che
    State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. China
    More by Penghua Che
  • Yang Luo 罗阳
    Yang Luo
    State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. China
    More by Yang Luo
  • Shujing Zhang 张淑静
    Shujing Zhang
    State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. China
  • Jie Xu* 徐杰*
    Jie Xu
    State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. China
    *Email: xujie@dicp.ac.cn
    More by Jie Xu
Open PDF 打开 PDFSupporting Information (1)
支持信息 (1)
OpenURL XIAMEN UNIV

ACS Catalysis

Cite this: ACS Catal. 2023, 13, 16, 11104–11116
Click to copy citationCitation copied!
https://doi.org/10.1021/acscatal.3c02736
Published August 7, 2023
Copyright © 2023 American Chemical Society

Abstract 抽象

Click to copy section link
单击以复制部分链接
Section link copied!

Hydroxyl radicals (OH), as one of the most highly oxidizing oxidants, are difficult to utilize in chemical synthesis since they usually cause unavoidable oxidative damage with CO2 released. Here, a Co–N–C nanomaterial composed of CoN4 sites on graphitic carbon nitride was constructed, which fulfills the self-cascade activation of O2 in water, wherein OH is guided to generate in a controlled level and exerts a highly selective oxidation role for efficient synthesis of a variety of organic acids by oxidation of furan alcohols, aromatic alcohols, and fatty alcohols under mild conditions. 5-Hydroxymethylfurfural (HMF) reached >99.9% conversion with 94.2% yield of 2,5-furandicarboxylic acid (FDCA) at 25 °C under atmospheric O2 pressure. This study allows a promising utility of O2 in a low energy-consuming way independent of light or electricity.
羟基自由基 (OH) 作为最具高度氧化性的氧化剂之一,很难用于化学合成,因为它们通常会在释放 CO2 时造成不可避免的氧化损伤。在这里,构建了一种由石墨氮化碳上的 CoN4 位点组成的 Co-N-C 纳米材料,它实现了 O2 在水中的自级联活化,其中 OH 被引导在受控水平生成并发挥高度选择性的氧化作用,通过氧化呋喃醇高效合成多种有机酸, 芳香醇和温和条件下的脂肪醇。在 25 °C 的大气 O2 压力下,5-羟甲基糠醛 (HMF) 的转化率达到 >99.9%,2,5-呋喃二羧酸 (FDCA) 的产率为 94.2%。这项研究允许 O2 以低能耗的方式独立于光或电而具有广阔的用途。

This publication is licensed under the terms of your institutional subscription. Request reuse permissions.
本出版物根据您的机构订阅条款获得许可。请求重用权限。

Copyright © 2023 American Chemical Society
版权所有 © 2023 美国化学会

1. Introduction 1. 引言

Click to copy section link
单击以复制部分链接
Section link copied!

Molecular oxygen (O2) is well acknowledged as an ideal oxidant for high atomic economic production of oxygen-containing chemicals, and facilitating its utility at a low temperature (25–60 °C) close to that of natural enzymes has always been one of the most difficult pursued goals. (1,2) An attractive strategy is to activate O2 into reactive oxygen species (ROS) such as hydroxyl radicals (OH), superoxide radicals (O2•–), hydrogen peroxide (H2O2), and singlet oxygen (1O2), (3) so as to break the limit of spin forbidden between thermodynamically inert O2 in the triplet state and organic substrates in the singlet state. (4) OH, as one of the most highly oxidizing oxidants and typical electrophilic species from O2, has high positive redox potential (E0(OH/OH) = 1.90 V, E0(OH + H+/H2O) = 2.73 V). (5) Such superhigh oxidizing ability is of particular interest as critical reaction conditions are not required to overcome the activation energy barrier but generally leads to ability in breaking chemical bonds, destructing organic structures, sterilizing and damaging cells, etc. (5,6) Typically, OH provided by an external source of H2O2 (e.g., Fenton’s reagent) is adopted in advanced oxidation processes (AOPs) for wastewater treatment, completely degrading water pollutants into CO2 without leaving residues. (7) However, the excessively generated OH due to the rapid reduction by metallic ions such as Fe2+ (8) is difficult to exert a highly selective oxidation role for synthesis application purposes. Although there are some examples of using photo- or electrocatalysts (9,10) as well as noble metal catalysts such as Au NPs, (11,12) to achieve in situ formation of OH, it is still a challenge for OH to be controllably formed from O2 under mild conditions if only thermocatalysis is to be relied on. For example, Iglesia et al. revealed that catalytic generation of OH on the surface of Mn/Na2WO4/SiO2 could enhance the rate and selectivity for oxidative coupling of methane, but operation at a high thermal induction temperature of 1073 K makes this method merely applicable for gas-phase oxidation. (13) To the best of our knowledge, the utilization of OH for thermocatalytic selective oxidation over heterogeneous non-noble metal catalysts at room temperature is very difficult and has not been reported.
分子氧 (O2) 是公认的高原子经济生产含氧化学品的理想氧化剂,在接近天然酶的低温 (25-60 °C) 下促进其效用一直是最困难追求的目标之一。(1,2) 一个有吸引力的策略是将 O2 活化成活性氧 (ROS),例如羟基自由基 (OH)、超氧自由基 (O2•–)、过氧化氢 (H2O2) 和单重态氧 (1O2),(3) 以打破三重态的热力学惰性 O2 和单重态的有机底物之间的自旋限制。(4) OH作为氧化性最强的氧化剂之一,也是O2的典型亲电物质,具有高正氧化还原电位(E0(OH/OH) = 1.90 V,E0(OH + H+/H2O) = 2.73 V)。(5) 这种超高的氧化能力特别值得关注,因为不需要关键反应条件来克服活化能屏障,但通常会导致破坏化学键、破坏有机结构、消毒和破坏细胞等的能力。(5,6) 通常,OH 由 H2O2 的外部来源提供(例如 Fenton 试剂)用于废水处理的高级氧化过程 (AOP),将水污染物完全降解为 CO2,而不会留下残留物。 (7) 然而,由于金属离子如 Fe2+(8) 的快速还原而产生的过量 OH 难以发挥用于合成应用目的的高选择性氧化作用。尽管有一些使用光或电催化剂 (9,10) 以及贵金属催化剂(如 Au NPs) (11,12) 来实现 OH的原位形成的例子,但如果仅依靠热催化,在温和的条件下由 O2 可控地形成 OH 仍然是一个挑战。例如,Iglesia 等人揭示了在 Mn/Na2WO4/SiO2 表面催化生成 OH 可以提高甲烷氧化偶联的速率和选择性,但在 1073 K 的高热感应温度下运行使该方法仅适用于气相氧化。(13) 据我们所知,在室温下利用 OH 对非均相非贵金属催化剂进行热催化选择性氧化非常困难,尚未报道。
To bridge the high oxidizing ability of OH and the synthetic application, the key point is producing OH in a controlled manner under mild conditions and avoid its excessive generation during the oxidation process. Recently, the great advancement achieved with certain nanomaterial-based artificial enzymes has encouraged the development in biomimetic activation of O2 into various ROS. (14,15) It is noteworthy that inspired by the metal coordination structure of natural enzymes, some biomimetic single-atom nanomaterials such as the ZIF-8-derived Zn–N–C can mimic the activity of peroxidase (POD) to facilely yield OH from H2O2 decomposition at 25 °C, overcoming the disadvantages of native enzymes such as structural vulnerability to damage, narrow pH value scale, etc. (16) By density functional theory (DFT) calculation and activity analysis experiments, the oxidase (OXD)-like activity has been demonstrated to depend largely on the local N coordination environment around the Co sites such as Co–N4. (17) Even more interestingly, when some peroxidase-like nanomaterials combine with natural oxidases, a cascade catalysis may be created to enable the generation of OH via promoting the interconversion among ROS. For example, H2O2 could be initially produced from O2 by using glucose oxidase, and subsequently, H2O2 is decomposed into OH catalyzed with the CuO-based artificial enzyme. (18) These findings bring new opportunities for the selective generation of OH via biomimetic cascade catalysis. Unfortunately, such prominent ability is exploited primarily for disease therapy, antibacterial application, biosensing, fluorescence analysis, degradation of environmental pollutants, etc. (14,15) So far, there is an urgent need to guide the activation of O2 for the generation of reactivity-controllable OH under the inspiration of cascade catalysis and realize the production of important organic chemicals, especially acids, aldehydes, ketones, etc., which are in large demand and with wide applications in modern chemistry. (19−25)
为了平衡 OH 的高氧化能力和合成应用,关键是在温和的条件下以受控方式生产 OH 并避免在氧化过程中过度产生。最近,某些基于纳米材料的人工酶取得的巨大进步促进了 O2 仿生激活成各种 ROS 的发展。(14,15) 值得注意的是,受天然酶的金属配位结构的启发,一些仿生单原子纳米材料,如 ZIF-8 衍生的 Zn-N-C,可以模拟过氧化物酶 (POD) 的活性,在 25 °C 下从 H2O2 分解中轻松产生 OH,克服了天然酶的缺点,例如结构易受损伤, 窄 pH 值尺度等。(16) 通过密度泛函理论 (DFT) 计算和活性分析实验,已证明氧化酶 (OXD) 样活性在很大程度上取决于 Co 位点周围的局部 N 配位环境,例如 Co-N4(17) 更有趣的是,当一些类似过氧化物酶的纳米材料与天然氧化酶结合时,可能会产生级联催化,通过促进 ROS 之间的相互转化来产生 OH 。例如,H2O2 最初可以通过使用葡萄糖氧化酶从 O2 产生,随后,H2O2 在基于 CuO 的人工酶的催化下分解成 OH(18) 这些发现为通过仿生级联催化选择性生成 OH 带来了新的机会。 不幸的是,这种突出的能力主要用于疾病治疗、抗菌应用、生物传感、荧光分析、环境污染物降解等。(14,15) 到目前为止,迫切需要在级联催化的启发下指导 O2 的活化以产生反应性可控的 OH 并实现重要有机化学品的生产,尤其是酸, 醛、酮等,在现代化学中需求量很大,应用广泛。(19−25)
Our laboratory has been conducting a series of studies on catalytic selective oxidations of biomass feedstocks to synthesize chemicals with O2. (26−30) We proposed a strategy by using hydrogen binding-initiated activation of O–H bonds to realize the highly efficient oxidation of 5-hydroxymethylfurfural (HMF) to 2,5-furandicarboxylic acid (FDCA). (27) Although the reaction kinetics was clearly promoted, the dependence of the reaction on temperature still persisted. The activity declined seriously once the temperature is lower than 80 °C, which is consistent with the mostly adopted temperature range of 80–130 °C for non-noble metal catalysts. (31) Besides O–H bonds, the H-elimination of C–H bonds at the hydroxymethyl group is another key factor that influence the oxidation. We noticed in our previously designed biomimetic catalytic system, NHPI/anthraquinone/HY zeolite, that the strong electrophilic species PINO showed remarkable ability in facilitating the cleavage of inert C–H bonds of ethylbenzene by single electron transformation. (32) Considering that OH is also a highly electrophilic species and formed from O2 itself, we speculated that it may act as the ideal ROS toward biomimetic oxidative synthesis. Herein, a single-atom Co–N–C catalyst has been designed to activate O2 by self-cascade catalysis, wherein OH plays the selective oxidation role at a regulated level in water, so as to ensure the biomass-derived platform HMF oxidized with a high yield at 25 °C and atmospheric oxygen pressure. It combines the advantages of enzymatic and thermal catalysis, independent of stoichiometric co-reductants compared to natural enzymes.
我们的实验室一直在进行一系列关于生物质原料催化选择性氧化以合成具有 O2 的化学品的研究。(26−30) 我们提出了一种策略,通过使用氢结合引发的 O-H 键激活来实现 5-羟甲基糠醛 (HMF) 高效氧化为 2,5-呋喃二羧酸 (FDCA)。(27) 尽管反应动力学明显得到促进,但反应对温度的依赖性仍然存在。当温度低于 80 °C 时,活性严重下降,这与非贵金属催化剂主要采用的 80–130 °C 温度范围一致。(31) 除了 O-H 键外,羟甲基 C-H 键的 H 消除是影响氧化的另一个关键因素。我们注意到在我们之前设计的仿生催化系统 NHPI/蒽醌/HY 沸石中,强亲电物种 PINO 在促进乙苯惰性 C-H 键裂解方面表现出显着的能力通过单电子转化。(32) 考虑到 OH 也是一种高度亲电的物质,并且由 O2 本身形成,我们推测它可能作为仿生氧化合成的理想 ROS。在此,设计了一种单原子 Co-N-C 催化剂,通过自级联催化活化 O2,其中 OH 在水中发挥调节水平的选择性氧化作用,从而保证生物质衍生平台 HMF 在 25 °C 和大气氧压下以高产率被氧化。它结合了酶和热催化的优点,与天然酶相比,它独立于化学计量共还原剂。

2. Results and Discussion
2. 结果与讨论

Click to copy section link
单击以复制部分链接
Section link copied!

2.1. Design, Characterization, and Biomimetic Catalytic Performance of the Co–N–C Catalyst
2.1. Co-N-C 催化剂的设计、表征和仿生催化性能

We proposed a strategy of activating O2 into OH at a controllable level by mimicking the cascaded catalysis of enzymes, so as to enable Co–N–C to accomplish the biomimetic oxidation synthesis. Scheme 1a illustrates that in the biomimetic catalytic system that uses Co–N–C with O2 as the oxidant, the concentration of OH is reduced to a sufficiently low level and in control throughout the oxidation process, and thus, its oxidative capacity is reduced to a suitable degree since the oxidative capacity is generally positively correlated with the concentration in the oxidized state. On the contrary, using H2O2 as an OH source will lead to an uncontrollable production of OH at a relatively high level, which causes a high oxidative damage effect. By design of a biomimetic Co–N–C system, in situ and continuous generation of OH can be fulfilled (Scheme 1b). The Co–N–C catalyst containing CoN4 sites has exhibited the ability to mimick enzymically cascaded catalysis of oxidase (OXD), superoxide dismutase (SOD), and peroxidase (POD), accomplishing the interconversions among ROS. Three ROS including O2•–, H2O2, and OH, with different oxidative capacities, were generated via cascading steps in sequence. Even at 25 °C, O2 can be initially activated into O2•– in situ and subsequently into H2O2, which is further converted into OH at a sufficiently low level by adjusting the pH value of the aqueous solution, guaranteeing that OH plays a selective oxidation role in the oxidation of hydroxyl compounds, instead of the oxidative degradation effect (Scheme 1c). Considering that HMF is prone to staying in various oxidation stages and forming intermediates with different reactivities, which can be used to evaluate the oxidation ability of ROS in situ generated, the selective oxidation of the biomass platform HMF to FDCA was selected as a model reaction. In addition, the catalytic performance of Co–N–C was evaluated at room temperature (25 °C) and atmospheric pressure with O2 bubbling to test whether it can achieve excellent performance comparable to that of natural enzymes under such mild conditions.
我们提出了一种通过模拟酶的级联催化在可控水平上将 O2 激活成 OH 的策略,从而使 Co-N-C 能够完成仿生氧化合成。方案 1a 表明,在使用带有 O2 的 Co-N-C 作为氧化剂的仿生催化系统中,OH 的浓度降低到足够低的水平并在整个氧化过程中处于受控状态,因此,其氧化能力降低到适当的程度,因为氧化能力通常与氧化状态下的浓度呈正相关。相反,使用 H2O2 作为 OH 源会导致 OH 在相对较高的水平上无法控制地产生,从而产生很高的氧化损伤效应。通过设计仿生 Co-N-C 系统,可以实现 OH 的原位和连续生成(方案 1b)。含有 CoN4 位点的 Co-N-C 催化剂表现出模拟氧化酶 (OXD)、超氧化物歧化酶 (SOD) 和过氧化物酶 (POD) 的酶促级联催化的能力,从而完成 ROS 之间的相互转化。通过级联步骤依次生成具有不同氧化能力的三种 ROS,包括 O2•–、H2、O2 和 OH。 即使在 25 °C 下,O2 也可以最初原位活化为 O2•–,随后转化为 H2O2,通过调节水溶液的 pH 值,在足够低的水平上进一步转化为 OH,保证 OH 在羟基化合物的氧化中起选择性氧化作用, 而不是氧化降解效应(方案 1c)。考虑到 HMF 易停留在各种氧化阶段并形成具有不同反应性的中间体,可用于评价原位生成的 ROS 的氧化能力,因此选择生物质平台 HMF 选择性氧化为 FDCA 作为模型反应。此外,在室温 (25 °C) 和 O2 鼓泡的大气压下评估了 Co-N-C 的催化性能,以测试它在如此温和的条件下是否能达到与天然酶相当的优异性能。

Scheme 1 方案 1

Scheme 1. Comparison of OH Generation in Different Systems (a); Design of the Biomimetic System for In Situ and Continuous Generation of OH (b); Conversion of RCH2OH in Different OH Generation Systems: Oxidative Degradation of RCH2OH with Uncontrollable Production of OH (c, Left), Selective Oxidation of RCH2OH with Controllable Production of OH (c, Right)
方案 1.不同系统中 OH 生成的比较 (a);原位和连续生成 OH 的仿生系统设计 (b);RCH2OH 在不同 OH 生成系统中的转化:RCH2OH 的氧化降解与无法控制的 OH产生 (c,左),RCH2OH 的选择性氧化与 OH 的可控产生 (c,右)
To our delight, the Co–N–C single-atom catalyst (SAC) showed outstanding activity at 25 °C, affording 94.2% yield of FDCA at >99.9% conversion of HMF for 10 h in an aqueous solution when the pH value varied to 12.79 (Table 1, entry 1). It is worth noting that such mild conditions are very intriguing for non-noble metal catalysts achieving high activity proximity to natural oxidase, instead of deep oxidative degradation with C–C and C–O bond breaking. The oxidation time in the Co–N–C system could be shortened by 2–6 times compared with the natural enzymes such as the HMFO enzyme (25 °C, 24 h) (33) and GOase + HLADH enzyme (30 °C, 60 h). (34) In contrast, typical non-SAC cobalt-based catalysts for comparison showed poor activities, including other N-doped carbon-supported cobalt nanocatalysts such as Co/N–C-1, Co/N–C-2, CoOx@N–C, and CoOx-N/C, homogeneous biomimetic cobalt complexes (Co(Phen)2 and CoTPP) containing similar CoN4 coordination structures, and Co/AC without N-doped carbon structures (Table 1, entries 2–8). To the best of our knowledge, there is no example that can be accomplished at 25 °C by using non-noble metal catalysts for this model reaction, which generally requires at least 80 °C and above and mostly in the pressured O2 (i.e., 0.3–2.0 MPa) (Figure S1). (31) Moreover, Co–N–C provides a broad range of reaction temperatures (25–60 °C) that can maintain high activity and selectivity simultaneously. By reacting at optimized 60 °C, it could further improve the catalytic efficiency and shorten oxidation time while achieving a high yield of FDCA (98.0%, 5 h). However, when the temperature continued to rise to 120 °C, the FDCA yield decreased to 6.7% with a serious carbon imbalance due to the side reactions. In the subsequent investigations, 60 °C was chosen as the experimental temperature to ensure the accurate progress of the exploration as efficiently as possible.
令我们高兴的是,Co-N-C 单原子催化剂 (SAC) 在 25 °C 下表现出出色的活性,当 pH 值变化至 12.79 时,在水溶液中 10 小时内 HMF 的 >99.9% 转化率下,FDCA 收率为 94.2%(表 1,条目 1)。值得注意的是,这种温和的条件对于非贵金属催化剂来说非常有趣,因为它在天然氧化酶附近实现了高活性,而不是通过 C-C 和 C-O 键断裂进行深度氧化降解。与天然酶(如 HMFO 酶(25 °C,24 小时)(33)和 GOase + HLADH 酶(30 °C,60 小时))相比,Co-N-C 系统中的氧化时间可以缩短 2-6 倍。(34) 相比之下,用于比较的典型非 SAC 钴基催化剂表现出较差的活性,包括其他 N 掺杂碳负载钴纳米催化剂,如 Co/N-C-1、Co/N-C-2、CoOx@N-C 和 CoOx-N/C,含有相似 CoN4 配位结构的均相仿生钴配合物(Co(Phen)2 和 CoTPP),以及不含 N 掺杂碳结构的 Co/AC(表 1,条目 2-8)。据我们所知,在25 °C下使用非贵金属催化剂进行该模型反应尚无示例,该模型反应通常需要至少80 °C及以上的温度,并且主要在加压的O2(即0.3-2.0 MPa)中(图S1)。(31) 此外,Co-N-C 提供广泛的反应温度范围 (25–60 °C),可以同时保持高活性和选择性。通过在优化的 60 °C 下反应,它可以进一步提高催化效率并缩短氧化时间,同时实现高 FDCA 收率(98.0%,5 h)。 然而,当温度继续升高到 120 °C 时,FDCA 收率下降到 6.7%,副反应导致严重的碳失衡。在随后的调查中,选择 60 °C 作为实验温度,以尽可能高效地保证勘探的准确进行。
Table 1. Comparison of the Catalytic Performance of Cobalt-Based Catalysts for HMF Oxidation at 25 °Ca
表 1.钴基催化剂在 25 °Ca 下催化 HMF 氧化的性能比较
entry 进入catalyst 催化剂HMF conv. (mol %) HMF 转化率 (mol %)FDCA yield (mol %) FDCA 产量 (mol %)
1Co–N–C 钴 N- C>99.994.2
2Co/N–C-1 钴/N-C-1>99.943.2
3Co/N–C-2 钴/N-C-2>99.915.8
4CoOx@N–C>99.945.9
5CoOx-N/C
CoOx-N/C
>99.922.7
6Co(Phen)2 钴(Phen)249.10.6
7CoTPP CoTPP 协议41.60.1
8Co/AC Co/AC 公司51.00.1
a

Reaction conditions: 0.5 mmol HMF, 1.25 mmol NaOH, 0.032 mmol Co in the catalyst, 5 mL of H2O, 40 mL/min O2, 25 °C, 10 h. Phen: 1,10-phenanthroline, and TPP: meso-tetraphenylporphyrin.


a

反应条件:催化剂中加入 0.5 mmol HMF、1.25 mmol NaOH、0.032 mmol Co、5 mL H2O、40 mL/min O2、25 °C、10 h。Phen:1,10-菲咯啉和 TPP:内消旋四苯卟啉。

Afterward, the Co–N–C catalyst has been characterized in detail and demonstrated to have atomically dispersed cobalt-coordinated macrocycles along with the nitrogen doping graphitization structure. In the presented transmission electron microscopy (TEM) image (Figure 1a) and aberration-corrected high-angular annular dark-field scanning transmission electron microscopy (AC HAADF-STEM) image (Figure 1i), no metallic cobalt or cobalt oxide nanoparticles were detected on the graphitic carbon sheet structure of Co–N–C, which was consistent with the X-ray diffraction (XRD) pattern (Figure 1m). The diffraction peaks at 25 and 43 degrees corresponded to the (002) and (004) planes of carbon, respectively, indicating a higher degree of graphitization. (27) The energy dispersive X-ray spectroscopy (EDX) analysis revealed that Co, N, and O were homogeneously dispersed on the carbon matrix (Figure 1b–d). As revealed by the X-ray absorption near-edge structure (XANES) analysis, the Co–N–C had almost the same first inflection point (E0 value) as the CoO sample (Figure 1e), and the valence state of cobalt in Co–N–C was +2. (35) Both the extended X-ray absorption fine structure (EXAFS) spectrum and the wavelet transforms for the EXAFS spectrum showed a Co–N-characteristic signal and no obvious Co–Co and Co–O signals (Figure 1f–g). (35) The EXAFS fitting revealed that one Co atom coordinated with four N atoms, i.e., CoN4 moiety was formed (Figure S2). The nitrogen content of Co–N–C was as high as 10.7 wt %, and the doped nitrogen species existed in the forms of pyridinic N/pyridinic N–Co, pyrrolic N, graphitic N, and pyridine N-oxide by X-ray photoelectron spectroscopy (XPS) analysis (Figure 1h). (36) The density of Co atoms in Co–N–C was approximately 0.72 Co atoms per nm2, which was calculated based on the Brunauer–Emmett–Teller (BET) surface area (641 m2/g) and the cobalt loading (4.3 wt %) of the catalyst. Such high surface area and highly dispersed cobalt atoms were desirable for catalytic applications.
之后,对 Co-N-C 催化剂进行了详细表征,并证明具有原子分散的钴配位大环以及氮掺杂石墨化结构。在所呈现的透射电子显微镜 (TEM) 图像(图 1a)和像差校正的高角度环形暗场扫描透射电子显微镜 (AC HAADF-STEM) 图像(图 1i)中,在 Co-N-C 的石墨碳片结构上未检测到金属钴或钴氧化物纳米颗粒,这与 X 射线衍射 (XRD) 图样一致(图 1m). 25 度和 43 度处的衍射峰分别对应于碳的 (002) 和 (004) 平面,表明石墨化程度较高。(27) 能量色散 X 射线光谱 (EDX) 分析显示,Co、N 和 O 均匀分散在碳基体上(图 1b-d)。正如 X 射线吸收近边缘结构 (XANES) 分析所揭示的那样,Co-N-C 与 CoO 样品具有几乎相同的第一拐点(E0 值)(图 1e),钴在 Co-N-C 中的价态为 +2。(35) 扩展 X 射线吸收精细结构 (EXAFS) 光谱和 EXAFS 光谱的小波变换都显示 Co-N 特征信号,没有明显的 Co-Co 和 Co-O 信号(图 1f-g)。(35) EXAFS 拟合显示一个 Co 原子与四个 N 原子配位,即形成 CoN4 部分(图 S2)。Co–N–C 的氮含量高达 10。7 wt %,掺杂的氮种类以吡啶 N/吡啶 N-Co、吡咯 N、石墨 N 和吡啶 N-氧化物的形式存在通过 X 射线光电子能谱 (XPS) 分析(图 1h)。(36) Co-N-C 中 Co 原子的密度约为 0.72 个 Co 原子/nm2,这是根据 Brunauer-Emmett-Teller (BET) 表面积 (641 m2/g) 和催化剂的钴负载量 (4.3 wt %) 计算得出的。如此高的表面积和高度分散的钴原子是催化应用的理想选择。

Figure 1 图 1

Figure 1. TEM image (a), HAADF STEM image (b), and corresponding EDX maps for the overlapping Co, N, C, and O (c, d) of Co–N–C; normalized XANES spectra (e), EXAFS spectra (f), and wavelet transforms for the EXAFS spectra (g) of Co–N–C and Co foil; N 1s XPS spectrum (h), AC HAADF-STEM image (i) of Co–N–C; TEM images of Co/N–C-1 (j), Co/N–C-2 (k), and Co/AC (l); XRD patterns (m), ratio of the intensity of DMPO-OH adducts in the electron spin resonance (ESR) spectra at the reaction times of 20 min (n), and the catalytic performance of HMF oxidation (o) of different Co-based catalysts. Reaction conditions: 0.5 mmol HMF, 1.25 mmol NaOH, 0.032 mmol Co in the catalyst, 5 mL of H2O, 40 mL/min O2, 60 °C, and 5 h.
图 1.TEM 图像 (a)、HAADF STEM 图像 (b) 以及 Co-N-C 重叠的 Co、N、C 和 O (c, d) 的相应 EDX 图;Co-N-C 和 Co 箔的 EXAFS 光谱 (g) 的归一化 XANES 光谱 (e)、EXAFS 光谱 (f) 和小波变换;N 1s XPS 谱图 (h),Co–N–C 的 AC HAADF-STEM 图像 (i);Co/N-C-1 (j)、Co/N-C-2 (k) 和 Co/AC (l) 的 TEM 图像;XRD 图谱 (m),反应时间为 20 min 时电子自旋共振 (ESR) 光谱中 DMPO-OH 加合物强度的比值 (n),以及不同钴基催化剂的 HMF 氧化催化性能 (o)。反应条件:催化剂中加入 0.5 mmol HMF、1.25 mmol NaOH、0.032 mmol Co、5 mL H2O、40 mL/min O2、60 °C 和 5 h。

Next, the significant effect of the Co–N–C SAC structure on the formation of OH has been evidenced by electron spin resonance spectroscopy (ESR) using the trapping agent 5,5-dimethyl-1-pyrroline N-oxide (DMPO). The relative ability in generating OH was determined in terms of comparing the ratio of I and I0 (I/I0), where I represents the intensity of the DMPO-OH signal and I0 represents that in the blank experiment (the aqueous solution at a pH value of 12.79), which afforded trace OH with a weak signal of DMPO-OH without catalysts (Figure S4). Compared to the atomically dispersed CoN4 sites in Co–N–C, increasing the average size of Co composites in Co/N–C-1 (3.21 nm) and Co/N–C-2 (13.78 nm) led to an obvious decrease in I/I0 as well as the selectivity of FDCA (Figure 1n,o). In addition to the CoN4 moiety, the Co/N–C-1 and Co/N–C-2 catalysts also contained Co species in the forms of Co (0) and CoOx as evidenced by the nanoparticle morphology shown in the TEM images (Figure 1j,k) and the XRD patterns (Figure 1m). For the Co/AC catalyst with large aggregated Co composite nanoparticles (average size of around 100.06 nm) and the N–C catalyst without any Co loaded, the I/I0 values of DMPO-OH were close to 1. There was no CoN4 moiety existing in Co/AC and N–C catalysts, and both consequently exhibited poor activity for selective oxidation of HMF (Table S1). This obvious distinction in I/I0 and activities indicated that the OH generation was closely related to the atomically dispersed CoN4 sites. Toxicity experiments on CoN4 sites were conducted by employing KSCN as the toxic reagent to further verify the facilitation of CoN4 sites. A significant reduction in the catalytic activity for HMF oxidation over Co–N–C pretreated with KSCN indicated that the highly dispersed CoN4 sites in Co–N–C were the core of OH production and selective oxidation of HMF and those unevenly distributed cobalt nanoparticles do not play a major role.
接下来,Co-N-C SAC 结构对 OH 形成的显着影响已通过使用捕获剂 5,5-二甲基-1-吡咯啉 N-氧化物 (DMPO) 的电子自旋共振光谱 (ESR) 得到证明。通过比较 II0I/I0) 的比率来确定产生 OH 的相对能力,其中 I 代表 DMPO-OH 信号的强度,I0 代表在空白实验中(pH 值为 12.79 的水溶液)中的强度,这提供了痕量 OH 具有 DMPO-OH 弱信号• 没有催化剂(图 S4).与 Co-N-C 中原子分散的 CoN4 位点相比,Co/N-C-1 (3.21 nm) 和 Co/N-C-2 (13.78 nm) 中 Co 复合材料的平均尺寸增加导致 I/I0 以及 FDCA 选择性的明显降低(图 1n,o)。除了 CoN4 部分外,Co/N-C-1 和 Co/N-C-2 催化剂还含有 Co (0) 和 CoOx 形式的 Co 物质,TEM 图像(图 1j,k)和 XRD 图谱(图 1m)中显示的纳米颗粒形态证明了这一点。对于具有大聚集体 Co 复合纳米颗粒(平均尺寸约为 100.06 nm)的 Co/AC 催化剂和没有任何 Co 负载的 N-C 催化剂,DMPO-OHI/I0 值接近 1。Co/AC 和 N-C 催化剂中不存在 CoN4 部分,因此两者都表现出较差的 HMF 选择性氧化活性(表 S1)。 I/I0 和活性的这种明显差异表明 OH 生成与原子分散的 CoN4 位点密切相关。采用 KSCN 作为毒性试剂对 CoN4 位点进行毒性实验,以进一步验证 CoN4 位点的便利性。与用 KSCN 预处理的 Co-N-C 相比,HMF 氧化的催化活性显着降低,这表明 Co-N-C 中高度分散的 CoN4 位点是 HMF 产生 OH• 和选择性氧化的核心,那些分布不均匀的钴纳米颗粒不起主要作用。
For comparison, we continued to check the OH level with the classical Fenton system (H2O2 + Fe2+) and alkaline hydrogen peroxide system (H2O2 + OH), which are frequently used for oxidative degradation at room temperature. As expected, a rather stronger signal of the DMPO-OH adduct has been detected by the ESR analysis, indicating the elevated OH concentration. (37) It evidenced OH produced from H2O2 decomposition in an uncontrollable way, leading to fully oxidative degradation of HMF into CO2. This observation agreed well with previous studies that the destructive capacity came from the high level of OH concentration, which was closely related to the concentration of H2O2. (38) Such differences in oxidation product types showed that Co–N–C could activate O2 and form OH in a controlled manner, with properties similar to those of enzymes. This characteristic has not been reported in previous studies, which only found that the CoN4 structure could facilitate the adsorption of O2; (39) the specific mechanism needs to be further clarified.
为了进行比较,我们继续检查经典芬顿系统 (H2O2 + Fe2+) 和碱性过氧化氢系统 (H2O2 + OH) 的 OH 水平,它们经常用于室温下的氧化降解。正如预期的那样,ESR 分析检测到 DMPO-OH 加合物的信号相当强,表明 OH 浓度升高。(37) 它证明 H2O2 以不可控的方式分解产生 OH,导致 HMF 完全氧化降解为 CO2。这一观察结果与以前的研究非常吻合,即破坏能力来自高水平的 OH 浓度,这与 H2O2 的浓度密切相关。(38) 氧化产物类型的这种差异表明,Co-N-C 可以以受控方式激活 O2 并形成 OH,其性质与酶相似。这个特性在以前的研究中没有报道,只发现 CoN4 结构可以促进 O2 的吸附;(39) 具体机制有待进一步明确。
In light of the unique property, Co–N–C could be considered a new nanomaterial-based artificial enzyme to produce OH in situ, via the activation of O2. By taking advantage of the OH with high oxidation ability, the −CH2OH will be activated under mild conditions and substrates may follow the selective oxidation pathway (Scheme 1c, right), instead of undergoing the oxidative degradation pathway for the cleavage of C–C and C–O bonds and releasing CO2 caused by the high concentration of OH (Scheme 1c, left). It is attractive for Co–N–C simulating enzyme-catalyzed O2 activation and requires in-depth exploration of specific generation, control, and mechanism.
鉴于其独特的特性,Co-N-C 可以被认为是一种新的基于纳米材料的人工酶,通过激活 O2 原位产生 OH。通过利用具有高氧化能力的 OH,-CH2OH 将在温和的条件下被激活,底物可以遵循选择性氧化途径(方案 1c,右),而不是经历氧化降解途径以裂解 C-C 和 C-O 键并释放由高浓度 OH 引起的 CO2方案 1c, left)。它对 Co-N-C 模拟酶催化的 O2 活化很有吸引力,需要深入探索特异性生成、控制和机制。

2.2. Self-Cascade Catalysis for the Generation of OH
2.2. 用于生成 OH 的自级联催化

To distinguish the transformation of ROS via mimicking enzymically cascaded catalysis that ultimately generates OH, experiments consisting of three steps were performed under specific conditions (Figure 2a). ROS including O2•–, 1O2, OH, and H2O2 were detected by free radical trapping–ESR analysis using specific radical quenchers, in combination with gas chromatography–mass spectrometry (GC–MS) and titration. In step 1, Co–N–C showed activity like OXD in activating O2 to O2•– as trapped by DMPO in toluene (Figure 2b), in which O2•– is stabilized enough for measurement purposes but undetectable in water due to the short lifetime. (40) When replacing O2 with N2, the six characteristic ESR peaks of the DMPO-O2•– adduct disappeared (Figure S6a). When the inhibitor p-benzoquinone (PBQ) was introduced into step 1, O2•– was no longer observed (Figure S6b). These results implied the electron transfer from Co–N–C to O2. It agreed well with previous reports that conjugated N-doped carbon structures were beneficial to the enrichment of electrons on the surface and thus profited the adsorption of O2 on atomically dispersed CoN4 sites. (39) The ζ-potential of Co–N–C was measured, which is less than 0, proving the existence of negative charges on the catalyst surface in the aqueous solution. The treatment of Co–N–C with the cationic reagent 3-chloro-2-hydroxypropyltrimethylammonium chloride could shield the surface negative charges, thus leading to a significant decrease in HMF conversion and FDCA selectivity, further indicating that the electron-rich surface of Co–N–C played a crucial role in transferring electrons to O2. The present biomimetic case was free of co-reductants, such as NADH required by natural oxidases. (41)
为了区分通过模拟最终产生 OH的酶级联催化来转化 ROS,在特定条件下进行了由三个步骤组成的实验(图 2a)。使用特异性自由基淬灭剂通过自由基捕获-ESR 分析结合气相色谱-质谱 (GC-MS) 和滴定法检测 ROS,包括 O2•–、1 O2OH和 H2O2。在第 1 步中,Co-N-C 显示出类似 OXD 的活性,即激活 O2 到 O2•– 被 DMPO 捕获在甲苯中(图 2b),其中 O2•– 足够稳定,可用于测量目的,但由于寿命短,在水中无法检测到。(40) 当用 N2 代替 O2 时,DMPO-O2•– 加合物的六个特征 ESR 峰消失了(图 S6a)。当将抑制剂苯醌 (PBQ) 引入步骤 1 时,不再观察到 O2•–图 S6b)。这些结果表明电子从 Co-N-C 转移到 O2。它与以前的报道非常吻合,即共轭 N 掺杂碳结构有利于表面电子的富集,从而有利于 O2 在原子分散的 CoN4 位点上的吸附。(39) 测量了 Co-N-C 的 ζ 电位,小于 0,证明水溶液中催化剂表面存在负电荷。 用阳离子试剂 3-氯-2-羟丙基三甲基氯化铵处理 Co-N-C 可以屏蔽表面负电荷,从而导致 HMF 转化率和 FDCA 选择性显着降低,进一步表明 Co-N-C 的富电子表面在将电子转移到 O2 中起着至关重要的作用。目前的仿生案例不含共还原剂,例如天然氧化酶所需的 NADH。(41)

Figure 2 图 2

Figure 2. Cascade process of OH generation by the activation of O2 and H2O over Co–N–C (a); ESR spectrum of the DMPO-O2•– in the toluene phase of step 1 (b); MS spectrum of triphenylphosphine oxide formed in the water phase of step 2 (c); ESR spectrum of the DMPO-OH in the water phase of step 3 (d); and effect of solvents and inhibitors on the generation of ROS (e, f). Reaction conditions: for the detection of H2O2 and OH in water, 0.032 mmol Co in Co–N–C, 1.25 mmol NaOH, 1 mmol scavenger, 2.5 mL of H2O, 40 mL/min O2, and 60 °C and for the detection of O2•– in toluene, 0.032 mmol Co in Co–N–C, 1.25 mmol NaOH, 1 mmol scavenger, 2.5 mL of toluene, 0.6 MPa O2, and 100 °C.
图 2.通过 Co-N-C (a) 上 O2 和 H2O 的活化生成 OH 的级联过程;DMPO-O 的 ESR 谱2•– 在步骤 1 (b) 的甲苯相中;在步骤 2 (c) 的水相中形成的三苯基氧化膦的 MS 谱图;步骤 3 (d) 中 DMPO-OH • 在水相中的 ESR 谱图;以及溶剂和抑制剂对 ROS 生成的影响 (e, f)。反应条件:用于检测水中的 H2O2 和 OH、0.032 mmol Co 的 Co-N-C 溶液、1.25 mmol NaOH、1 mmol 清除剂、2.5 mL 的 H2O、40 mL/min O2 和 60 °C,以及用于检测甲苯中的 O2•--0.032 mmol Co 和 Co-N-C 溶液, 1.25 mmol NaOH、1 mmol 清除剂、2.5 mL 甲苯、0.6 MPa O2 和 100 °C。

In step 2, specific biphasic experiments have been conducted and revealed that O2•– facilely transformed into H2O2 when encountering water. To the suspension of toluene in step 1 (containing O2•–) was added water to form a two-phase system, and O2•– remained in the upper toluene phase, while it was undetectable in the lower aqueous phase as being immediately converted to H2O2 after its formation, supported by the typical oxygenated product Ph3P═O arising from the oxidation of Ph3P by H2O2 (Figure 2c). The chromogenic reaction of KI also confirmed this. After adding the KI solution into the isolated water phase, the aqueous phase changed brown, suggesting that I was oxidized to I2 by H2O2 (Figure S8). If PBQ was added into the toluene phase, O2•– was inhibited, and the subsequent addition of water would not enable H2O2 generation (Figure S9). Such biomimetic activity of Co–N–C may be attributed to the CoN4 sites and graphitized structure, mimicking the electronic and structural characteristics of natural SOD. The disproportionation of O2•– to H2O2 could occur as metal–nitrogen sites in the SOD could adsorb O2•– and facilitate the electron transfer of O2•–. (42) Together, the Co–N–C possessed SOD-like activity to generate H2O2 from O2•– and H2O.
在第 2 步中,进行了特定的双相实验,结果表明 O2•– 遇到水时很容易转化为 H2O2。在步骤 1 中的甲苯悬浮液(含有 O2•–)中加入水以形成两相体系,O2•– 保留在上层甲苯相中,而它在下层水相中检测不到,因为它在形成后立即转化为 H2O2,由典型的含氧产物 Ph 3 P═O 支持,该产物 Ph3P 由H2O2 氧化产生(图 2KI 的显色反应也证实了这一点。将 KI 溶液加入分离的水相中后,水相变为棕色,表明 I 被 H2O2 氧化成 I2图 S8)。如果将 PBQ 添加到甲苯相中,O2•– 被抑制,随后加水不会使 H2O2 产生(图 S9)。Co-N-C 的这种仿生活性可能归因于 CoN4 位点和石墨化结构,模仿天然 SOD 的电子和结构特性。O2•– 到 H2O2 的歧化可能是因为 SOD 中的金属-氮位点可以吸附 O2•– 并促进 O2•–的电子转移。(42) Co-N-C 一起具有类似 SOD 的活性,可从 O2•– 和 H2O 生成 H2O2
Step 3 demonstrated the activity of Co–N–C like POD under the alkaline environment. Once the water phase in step 2 was adjusted from neutral (pH = 7.02) to alkaline (pH = 12.79) by adding NaOH, it could be found that in the aqueous phase in addition to H2O2 (slightly decomposed), the OH species also appeared as evidenced by the ESR signal of the DMPO-OH adduct (Figure 2d). If isopropyl alcohol (IPA), an inhibitor of OH, was introduced into the system, the OH disappeared, while H2O2 remained. Afterward, OH could also generate when H2O2 was introduced into the aqueous system (containing the Co–N–C catalyst) instead of O2 bubbling, which verified that OH originated from the decomposition of the in situ generated H2O2 (Figure S10). Such POD-like behavior in decomposing H2O2 into OH was also found in some nanomaterial-based artificial enzymes containing the CoN4 macrocyclic structures. (43) The present Co–N–C system exhibited strong alkaline resistance and was clearly distinguished from common enzymes.
步骤 3 展示了 Co-N-C 样 POD 在碱性环境下的活性。通过添加 NaOH 将步骤 2 中的水相从中性 (pH = 7.02) 调整为碱性 (pH = 12.79) 后,可以发现在水相中,除了 H2O2(略微分解)外,还出现了 OH 物质,DMPO-OH 加合物的 ESR 信号证明了这一点(图 2d)。如果将 OH抑制剂异丙醇 (IPA) 引入系统,OH 消失,而 H2O2 仍然存在。之后,当 H2O2 被引入水性系统(包含 Co-N-C 催化剂)中时,也会产生 OH 而不是 O2 鼓泡,这验证了 OH 起源于原位生成的 H2O2 的分解(图 S10)。在一些含有 CoN4 大环结构的基于纳米材料的人工酶中也发现了这种将 H2O2 分解成 OH 的类似 POD 的行为。(43) 目前的 Co-N-C 系统表现出很强的耐碱性,与普通酶明显不同。
In addition, when PBQ was added in advance in step 1, followed by the subsequent steps 2 and 3, H2O2 and OH that should have been generated subsequently could no longer be detected (Figure 2e,f). It suggested that the first step for activation of O2 to O2•– was indispensable for the cascade process. Taken together, the mimicking cascaded catalysis of the OXD, SOD, and POD of the Co–N–C catalyst started with the activation of O2 to O2•–, and O2•– subsequently converted into H2O2 once the water was present; then, H2O2 further converted to OH in a basic solution (i.e., pH = 12.79). The source of OH was also explored via the OH trapping experiment and liquid chromatography quadrupole time-of-flight mass spectrometry (LC–QTOF–MS) analysis by using isotope-labeled 18O2, H218O, and D2O. As shown in Figure 3, when 18O2 and H218O were employed, the 18O-labeled DMPO-18OH was detected as the main adduct in both cases. Replacing H2O with D2O, DMPO-OD was formed, demonstrating that the source of H in OH was from H2O. These results indicated that OH was produced from O2 and H2O through Co–N–C. Compared with the conventional enzyme cascade process of O2 activation via glucose oxidase and peroxidase, (18) this cascade process exhibited higher temperature and pH tolerance and higher efficiency.
此外,当在步骤 1 中预先添加 PBQ 时,随后是后续步骤 2 和 3,本应随后生成的 H2、O2 和 OH 无法再检测到(图 2e、f)。它表明 O2 到 O2•– 激活的第一步对于级联过程是必不可少的。综上所述,Co-N-C 催化剂的 OXD、SOD 和 POD 的模拟级联催化始于 O2 到 O2•–,而 O2•– 随后在水存在后转化为 H2O2;然后,H2O2 在碱性溶液(即 pH = 12.79)中进一步转化为 OH。还通过 OH 捕获实验和液相色谱四极杆飞行时间质谱 (LC-QTOF-MS) 分析,使用同位素标记的 18O2、H218O 和 D2O 来探索 OH 的来源。如图 3 所示,当使用 18O2 和 H218O 时,在两种情况下都检测到 18O 标记的 DMPO-18 OH 作为主要加合物。用 D2O 代替 H2O,形成 DMPO-OD ,表明 OH 中 H 的来源来自 H2O。这些结果表明,OH 是由 O2 和 H2O 通过 Co-N-C 产生的。与通过葡萄糖氧化酶和过氧化物酶激活 O2 的常规酶级联过程相比,(18) 这种级联过程表现出更高的温度和 pH 耐受性以及更高的效率。

Figure 3 图 3

Figure 3. Mass spectra of the OH trapping experiments by DMPO in isotope-labeled H2O and O2. Reaction conditions: 1.25 mmol NaOH, 0.032 mmol Co in Co–N–C, 5 mL of H2O, 60 °C, 40 mL/min O2, and 30 min.
图 3.DMPO 在同位素标记的 H2O 和 O2 中进行 OH 捕获实验的质谱。反应条件:1.25 mmol NaOH、0.032 mmol Co in Co–N-C、5 mL H2O、60 °C、40 mL/min O2 和 30 min。

2.3. Identifying the Role of ROS in Selectivity Oxidation of HMF
2.3. 确定 ROS 在 HMF 选择性氧化中的作用

Because O2•–, H2O2, and OH were involved in the self-cascade catalysis for O2 activation over Co–N–C, it is necessary to identify the oxidation ability of different ROS in the oxidation of HMF. The generation of different types of ROS was performed in different conditions by altering the reaction solvents or pH values. We examined the capacity of O2•– by conducting the oxidation of HMF in toluene because the role of O2•– is difficult to resolve owing to its instability and extremely short lifetime in water. The Co–N–C revealed no activity at 25 °C and could afford 49.0% yield of 2,5-furandicarboxaldehyde (DFF) as the major oxidation product at 100 °C (Table 2, entry 1). O2•– demonstrated that the only ROS involved is suggested to play a role in selectively oxidizing −CH2OH to −CHO, showing insufficient oxidative capacity to oxidize −CH2OH to −COOH even at 100 °C. Next, by adding IPA to scavenge OH in the Co–N–C + water + OH system that pregenerates H2O2 and OH, H2O2 can be maintained as a dominant ROS. Accordingly, FDCA was formed with a yield of 30.5%, accompanied by partially unconverted 5-hydroxymethyl-2-furancarboxylic acid (HMFCA, yield of 35.2%) with a carbon balance of only 66.0% (Table 2, entry 2). H2O2 exhibited weak oxidation ability to the oxidation of −CH2OH to −COOH and was failed to convert HMF in time, resulting in side reactions with carbon imbalance. If OH remained in the Co–N–C system, the intermediates DFF and HMFCA with relative higher stability are not detected. The high yield of 94.8% for the oxidative product FDCA confirmed the high oxidative capacity of OH (Table 2, entry 3).
由于 O2•–、H2、O2 和 OH 参与了 Co-N-C 上 O2 活化的自级联催化,因此有必要确定不同 ROS 在 HMF 氧化中的氧化能力。通过改变反应溶剂或 pH 值,在不同条件下生成不同类型的 ROS。我们通过传导 HMF 在甲苯中的氧化来检查 O2•– 的容量,因为 O2•– 的作用由于其不稳定性和在水中的寿命极短而难以解决。Co-N-C 在 25 °C 下没有活性,在 100 °C 下可以得到 49.0% 的 2,5-呋喃二甲醛 (DFF) 作为主要氧化产物(表 2,条目 1)。O2•– 证明唯一涉及的 ROS 被认为在选择性地将 -CH2OH 氧化为 -CHO 中发挥作用,即使在 100 °C 下也显示出不足的氧化能力,无法将 -CH2OH 氧化为 -COOH。 接下来,通过在预生成 H2O2 和 OH的 Co–N–C + 水 + OH– 系统中添加 IPA 来清除 OH,H2O2 可以保持为占主导地位的 ROS。因此,形成 FDCA,产率为 30.5%,伴有部分未转化的 5-羟甲基-2-呋喃羧酸(HMFCA,产率 35.2%),碳平衡仅为 66.0%(表 2,条目 2)。H2O2 对 -CH2OH 氧化为 -COOH 表现出弱的氧化能力,并且未能及时转化 HMF,导致与碳不平衡的副反应。如果 OH 保留在 Co-N-C 系统中,则不会检测到稳定性相对较高的中间体 DFF 和 HMFCA。 氧化产物 FDCA 的 94.8% 的高产量证实了 OH 的高氧化能力(表 2,条目 3)。
Table 2. Comparison of the Effect of Different ROS in HMF Oxidation over Co–N–C
表 2.不同 ROS 在 HMF 氧化对 Co-N-C 的影响比较
      yield (%) 产量 (%)
entry 进入solvent 溶剂base 基础additive 添加剂dominant ROS 优势 ROSconv. (%) 转化率 (%)FDCAHMFCAFFCADFF
1a 1toluene 甲苯  O2•–87.1ndndnd49.0
2b 2water NaOH 氢氧化钠IPAH2O2>99.930.535.20.3nd
3b 3water NaOH 氢氧化钠 OH>99.994.8ndndnd
a

Reaction conditions: 0.5 mmol HMF, 0.032 mmol Co in Co–N–C, 5 mL of toluene, 0.6 MPa O2, 100 °C, and 10 h.


a

反应条件:0.5 mmol HMF、0.032 mmol Co in Co–N-C、5 mL 甲苯、0.6 MPa O2、100 °C 和 10 h。

b

0.5 mmol HMF, 1.25 mmol NaOH, 0.032 mmol Co in Co–N–C, 1.0 mmol additive, 5 mL of H2O, 40 mL/min O2, 60 °C, and 4 h.


b

0.5 mmol HMF、1.25 mmol NaOH、0.032 mmol Co in Co–N–C、1.0 mmol 添加剂、5 mL H2O、40 mL/min O2、60 °C 和 4 h。

Subsequently, the pulse experiments were conducted to explore whether OH is the dominant ROS playing the selective oxidation role. With O2 bubbling, 0.05 mmol HMF was added to the Co–N–C + water + OH system at 14 min intervals. The intensity of OH was monitored by ESR analysis and exhibited a cyclic change dependent on the addition and consumption of HMF. The OH appeared to decrease first and then increase after each addition of HMF (Figure 4a above). By correlating the concentrations of HMF and FDCA with the intensity ratio of OH, one could find that the consumption of HMF was basically synchronous with the consumption of OH (Figure 4a below), indicating that the selective oxidation of HMF would constantly consume OH, resulting in a concentration close to the blank experiment (the intensity ratio consistently remained around 1) when HMF depleted for each addition.
随后,进行脉冲实验,探讨 OH 是否是发挥选择性氧化作用的优势 ROS。在 O2 鼓泡的情况下,以 14 分钟的间隔向 Co–N–C + 水 + OH 系统中加入 0.05 mmol HMF。通过 ESR 分析监测 OH 的强度,并表现出取决于 HMF 添加和消耗的循环变化。OH 在每次添加 HMF 后似乎先降低,然后增加(上面的图 4a)。通过将 HMF 和 FDCA 的浓度与 OH的强度比相关联,可以发现 HMF 的消耗量与 OH 的消耗量基本同步(下图 4a),表明 HMF 的选择性氧化会不断消耗 OH,导致浓度接近空白实验(强度比始终保持在 1 左右),当 HMF 每次添加耗尽时。

Figure 4 图 4

Figure 4. Time curve of the intensity of DMPO-OH adducts in the ESR spectra of the Co–N–C system with HMF addition (a, above), the concentrations of HMF and FDCA during the HMF oxidation process in the Co–N–C system (a, below); the high-performance liquid chromatography (HPLC) traces for HMF oxidation in the Co–N–C system (b); the time curve of the intensity of DMPO-OH adducts in the ESR spectra of the Co–N–C system without HMF addition (c); the time curve of the intensity of DMPO-OH adducts in the ESR spectra of the H2O2 + OH system with and without HMF addition (d). Reaction conditions: in the Co–N–C system, 1.25 mmol NaOH, 0.032 mmol Co in Co–N–C, 5 mL of H2O, 40 mL/min O2, and 60 °C and in the H2O2 + OH system, 1.25 mmol NaOH, 1 mL of 30 wt % H2O2, 4 mL of H2O, and 60 °C. In these two systems, 0.05 mmol HMF was added in the 11th min, 25th min, 39th min, and 53rd min.
图 4.添加 HMF 后 Co-N-C 系统 ESR 光谱中 DMPO-OH 加合物强度的时间曲线(a,上图),Co-N-C 系统中 HMF 氧化过程中 HMF 和 FDCA 的浓度(a,下图);Co-N-C 系统中 HMF 氧化的高效液相色谱 (HPLC) 迹线 (b);未添加 HMF 的 Co-N-C 系统的 ESR 光谱中 DMPO-OH 加合物强度的时间曲线 (c);H2O2 + OH– 添加和不添加 HMF 的 ESR 光谱中 DMPO-OH 加合物强度的时间曲线 (d)。反应条件:在 Co-N-C 系统中,1.25 mmol NaOH、0.032 mmol Co 的 Co-N-C 溶液、5 mL 的 H2O、40 mL/min O2 和 60 °C,以及在 H2O2 + OH 系统中,1.25 mmol NaOH,1 mL 30 wt % H2O2,4 mL H2O, 和 60 °C。 在这两个系统中,在第 11 min、第 25 min、第 39 min 和第 53 min 添加 0.05 mmol HMF。

Meanwhile, similar trends were observed for the peaks of HMF and intermediates in the HPLC spectra with each interval addition of HMF (Figure 4b). HMF appeared first and then gradually decreased to exhaustion. During the interval between each two additions of HMF, intermediates HMFCA as well as a small amount of 5-formyl-2-furancarboxylic acid (FFCA) also appeared first and then decreased to zero. FDCA, which accumulated with the addition of HMF, continued to be elevated in its concentration. The intermediates HMFCA and FFCA showed a tendency to rise first and then decreased due to further conversion. When HMF and intermediates were consumed, the OH resumed to form and accumulated. When HMF was added again, similar cycles continued to follow, and at the end of each cycle, the level of OH still returned close to the initial. If HMF was not added to the Co–N–C + water + OH system, the I/I0 values of DMPO-OH increased gradually and remained constant at 2.0 after 90 min, indicating that an equilibrium was reached after a certain time (Figure 4c), and the concentration of OH was no longer improved and maintained at a certain concentration. These results indicated that OH in the Co–N–C system was produced in situ and continuously, which ensured the highly selective oxidation of hydroxyl compounds to the corresponding carboxylic acids.
同时,在每次间隔添加 HMF 后,HPLC 谱图中 HMF 和中间体的峰也观察到类似的趋势(图 4b)。HMF 首先出现,然后逐渐下降到疲惫。在每添加两次 HMF 的间隔期间,中间体 HMFCA 以及少量 5-甲酰基-2-呋喃羧酸 (FFCA) 也首先出现,然后减少到零。随着 HMF 的添加而积累的 FDCA 浓度继续升高。中间体 HMFCA 和 FFCA 表现出先上升后下降的趋势,这是由于进一步的转化。当 HMF 和中间体被消耗时,OH 恢复形成并积累。当再次添加 HMF 时,类似的循环继续进行,在每个循环结束时,OH 的水平仍然恢复到接近初始水平。如果未将 HMF 添加到 Co–N–C + 水 + OH 系统中,DMPO-OHI/I0 值逐渐增加,并在 90 分钟后保持在 2.0 恒定,表明一定时间后达到平衡(图 4c),OH 的浓度不再改善并维持在一定浓度。这些结果表明,Co-N-C 系统中的 OH 是原位连续产生的,这确保了羟基化合物高度选择性地氧化为相应的羧酸。
When H2O2 was directly used instead of O2 as the oxidant, the intermittent addition of HMF into the H2O2 + OH system led to a quite different trend in the I/I0 values of DMPO-OH, which decreased during all the pulse experiments, indicating that OH only arose from the decomposition of H2O2 without the Co–N–C catalyst. If H2O2 was fed in one batch and no HMF was added in the H2O2 + OH system, a large amount of OH was immediately generated at high intensity. For example, its I/I0 value at 15 min was 8.4 times higher than that of the Co–N–C system. The I/I0 value of DMPO-OH gradually decreased with the decomposition of H2O2 until exhaustion (Figure 4d). As a result, in the case of intermittent addition of HMF into the H2O2 + OH system, OH at a high level led to the oxidative degradation of HMF into CO2 without any selective oxidation products (Figure S16). Such oxidative degradation also led to OH being quickly consumed within 66 min. This was consistent with the need of continuously replenishing H2O2 for the application of the H2O2 system in the oxidative degradation. (44) Therefore, a high level of OH in the H2O2 + OH system was not conducive to the selective oxidation.
当直接使用 H2O2 代替 O2 作为氧化剂时,HMF 间歇性添加到 H2O2 + OH 体系中导致 DMPO-OHI/I0 值出现完全不同的趋势,在所有脉冲实验中均有所下降,表明 OH 仅由 H2O2 分解产生,没有 Co-N-C 催化剂。如果 H2O2 分批进料,并且在 H2O2 + OH 系统中未添加 HMF,则立即以高强度产生大量 OH。例如,它在 15 min 时的 I/I0 值比 Co-N-C 系统的 8.4 倍高。DMPO-OHI/I0 值随着 H2O2 的分解而逐渐降低,直至耗尽(图 4d)。因此,在将 HMF 间歇性添加到 H2O2 + OH 系统的情况下,高水平的 OH 导致 HMF 氧化降解成 CO2,而没有任何选择性氧化产物(图 S16)。这种氧化降解也导致 OH 在 66 分钟内被迅速消耗。这与不断补充 H2O2 以将 H2O2 系统应用于氧化降解的需求是一致的。(44) 因此,H2O2 + OH 系统中高水平的 OH 不利于选择性氧化。
According to the time course of HMF oxidation catalyzed by the Co–N–C catalyst (Figure 5a), the reaction mainly proceeded along the route in which HMF was oxidized to the intermediate HMFCA and then converted to FDCA. The oxidation of −CH2OH in HMFCA to −COOH is the key step affecting the whole reaction rate. After adding IPA to inhibit the formation of OH, the generation rate of FDCA was reduced significantly and so was the conversion rate of HMFCA (Figure 5b), confirming the crucial role of OH in the oxidation of −CH2OH to −COOH. To further verify whether OH played a dominant role, the oxidation using HMFCA as a starting substrate was performed (Figure 5c). By addition of IPA, OH was inhibited and H2O2 remained, and the conversion of HMFCA was greatly reduced with the yield of FDCA decreased from 92.0 to 7.6% (Figure 5d), indicating that OH played a major role in the oxidation of −CH2OH to −COOH. If both H2O2 and OH were inhibited by PBQ via inhibiting the starting species O2•–, it could lead to a further decrease in HMFCA conversion (from >99.9 to 10.9%) with no FDCA yielded, suggesting that H2O2 made a weak contribution to the oxidation of −CH2OH in this key step. Taken together, three ROS with different oxidizing abilities were generated via O2 activation over the Co–N–C catalyst, and the OH produced by mimicking the cascaded catalysis process exhibited the highest oxidizing ability among the three ROS and played the key role in the oxidation of −CH2OH to −COOH. It was suggested that the OH promoted the key step of the conversion of −CH2OH to −COOH in HMFCA oxidation, so as to enhance the efficiency in production of FDCA with high selectivity.
根据 Co-N-C 催化剂催化的 HMF 氧化的时间进程(图 5a),反应主要沿着 HMF 被氧化成中间体 HMFCA,然后转化为 FDCA 的路线进行。HMFCA 中的 -CH2OH 氧化为 -COOH 是影响整个反应速率的关键步骤。添加 IPA 抑制 OH 的形成后,FDCA 的生成速率显著降低,HMFCA 的转化率也显著降低(图 5b),证实了 OH 在 −CH2OH 氧化为 −COOH 中的关键作用。为了进一步验证 OH 是否起主导作用,使用 HMFCA 作为起始底物进行了氧化(图 5c)。通过添加 IPA,OH 被抑制,H2O2 仍然存在,HMFCA 的转化率大大降低,FDCA 的产量从 92.0% 下降到 7.6%(图 5d),表明 OH 在 -CH2OH 氧化为 -COOH 中起主要作用。如果 H2O2 和 OH 都通过抑制起始物质 O2•– 被 PBQ 抑制,则可能导致 HMFCA 转化率进一步降低(从 >99.9% 降低到 10.9%),而没有产生 FDCA,这表明 H2O2 在这个关键步骤中对 -CH2OH 的氧化贡献微弱。综上所述,通过在 Co-N-C 催化剂上进行 O2 活化生成了三种具有不同氧化能力的 ROS,通过模拟级联催化过程产生的 OH 在三种 ROS 中表现出最高的氧化能力,并在 -CH2OH 氧化为 -COOH 中起关键作用。 有人提出 OH 促进了 HMFCA 氧化中 -CH2OH 转化为 -COOH 的关键步骤,从而提高了高选择性 FDCA 的生产效率。

Figure 5 图 5

Figure 5. Reaction time profile of HMF oxidation over Co–N–C (a) and inhibited by IPA (b) and the reaction time profile of HMFCA oxidation over Co–N–C (c) and inhibited by IPA (d). Reaction conditions: 0.5 mmol HMF or HMFCA, 0.032 mmol Co in Co–N–C, 1.25 mmol NaOH, 5 mL of H2O, 1.0 mmol IPA, 40 mL/min O2, and 60 °C.
图 5.HMF 在 Co-N-C (a) 上被 IPA 抑制的反应时间曲线,以及 HMFCA 氧化在 Co-N-C 上被 (c) 和被 IPA 抑制的反应时间曲线 (d)。反应条件:0.5 mmol HMF 或 HMFCA、0.032 mmol Co in Co–N–C、1.25 mmol NaOH、5 mL H2O、1.0 mmol IPA、40 mL/min O2 和 60 °C。

2.4. Comparison of the External and In Situ Formed H2O2
2.4. 外部形成和原位形成的 H2O2 的比较

The oxidation of HMF was monitored by UV–vis spectroscopy and is presented in Figure 6a,b. In the case of bubbling O2 in the Co–N–C + water + OH system, HMF quickly converted in the initial period and then oxidized into FDCA, as shown by the sharp decrease of absorbance intensity at 284 nm (HMF) and increase at 259 nm (acidic products) versus time. The variations of UV–vis spectra agreed well with the trend in FDCA concentration (Figure 6c) and excellent results (FDCA selectivity of 98.0% and carbon balance of nearly 100%). In the case of the H2O2 + OH system, HMF was also quickly consumed, but no peaks of acidic products were observed. As demonstrated above, the one-time addition of H2O2 immediately formed a large amount of OH and led to the complete degradation of HMF to CO2.
通过紫外-可见光谱法监测 HMF 的氧化,如图 6a、b 所示。在 Co–N–C + 水 + OH 系统中冒泡 O2 的情况下,HMF 在初始阶段迅速转化,然后氧化成 FDCA,如 284 nm 处吸光度强度 (HMF) 和 259 nm 处吸光度强度(酸性产物)随时间急剧增加所示。紫外-可见光谱的变化与 FDCA 浓度(图 6c)和优异结果(FDCA 选择性为 98.0%,碳平衡接近 100%)的趋势非常吻合。在 H2O2 + OH 系统的情况下,HMF 也被迅速消耗,但没有观察到酸性产物的峰值。如上所述,一次性添加 H2O2 立即形成大量 OH 并导致 HMF 完全降解为 CO2

Figure 6 图 6

Figure 6. UV–vis spectra of the HMF oxidation process in the Co–N–C system (a) and H2O2 + OH system (b); evolution of FDCA concentration in HMF oxidation in the Co–N–C system with O2 bubbling (A), with H2O2 addition 10 times (B), with H2O2 addition 1 time (C), and in the H2O2 + OH system (D) (c); ESR spectra of DMPO-OH adducts at the reaction time of 1 min in the Co–N–C system with O2 bubbling, with H2O2 addition 10 times and 1 time and in the H2O2 + OH system (d). Reaction conditions: 0.5 mmol HMF, 0.032 mmol Co in Co–N–C, 1.25 mmol NaOH, 5 mL of H2O, 60 °C, and 40 mL/min O2 (A); 0.1 mL of 30 wt % H2O2 was added 10 times at equal intervals (B); 1 mL of 30 wt % H2O2 was added 1 time (C); 1 mL of 30 wt % H2O2 was added 1 time without Co–N–C addition (D). Reaction conditions of UV–vis monitoring in (a, b) were consistent with those of (A) and (D) in (c), respectively.
图 6.Co-N-C 系统 (a) 和 H2O2 + OH 系统 (b) 中 HMF 氧化过程的紫外-可见光谱;在 Co-N-C 系统中 O2 鼓泡 (A)、H2O2 添加 10 倍 (B)、H2O2 添加 1 倍 (C) 和 H2O2 + OH 系统 (D) (c) 中 HMF 氧化中 FDCA 浓度的演变;DMPO-OH 加合物在 Co–N-C 系统中反应时间为 1 min 时的 ESR 光谱,O2 鼓泡,H2O2 添加 10 次和 1 次,以及在 H2O2 + OH 系统 (d) 中。反应条件:0.5 mmol HMF、0.032 mmol Co in Co–N–C、1.25 mmol NaOH、5 mL H2O、60 °C 和 40 mL/min O2 (A);以相等的间隔添加 0.1 mL 30 wt % H2O2 10 次 (B);添加 1 mL 30 wt % H2O2 1 次 (C);加入 1 mL 30 wt % H2O2 1 次,不添加 Co-N-C (D)。(a、b) 中紫外-可见分光光度计监测的反应条件分别与 (c) 中 (A) 和 (D) 的反应条件一致。

It makes us curious about what changes will occur when this H2O2 + OH system is added by the Co–N–C catalyst, that is, external H2O2 is used to provide the OH source for HMF oxidation. If the addition of H2O2 was changed once to several times to reduce its concentration, would the pathway over the Co–N–C catalyst be changed? Then, we compared the HMF oxidation experiments over Co–N–C with 1 and 10 additions of H2O2 for validation (Figure 6c). It was intriguing that when 1 mL of 30 wt % H2O2 (8.8 mmol) was added once to replace the O2 feeding, the Co–N–C + H2O2 + OH system undergoes selective oxidation in addition to oxidative degradation. As evidenced by ESR (Figure 6d), Co–N–C showed POD-like activity to convert H2O2 into OH, with the I/I0 of OH at 1.5 in the initial stage of the reaction (1 min). It was higher than that of the O2 feeding case (1.1), further confirming that the one-time addition of H2O2 would lead to the formation of a relatively higher concentration of OH, which led to 33.0% of HMF decomposition at 300 min. With the gradual consumption of H2O2, the OH concentration decreased accordingly, thus exhibiting an insufficient ability to oxidize HMF to FDCA, and FDCA was produced with lower selectivity (33.7%) when HMF was consumed completely. If 0.1 mL of 30 wt % H2O2 was added 10 times at equal intervals, the initial intensity of OH was lower, with no longer continuous decrease with supplementation of H2O2. However, the instantaneous concentration of OH was still higher than that of the O2 feeding case (Figure 6d). Although the FDCA selectivity was increased to 70.9% in this case, there was still 20.6% of the added HMF decomposed in the initial stage of the reaction. Moreover, the carbon balance was not satisfied and only reached 79.4%.
这让我们很好奇,当这个 H2O2 + OH 系统由 Co-N-C 催化剂添加时会发生什么变化,即使用外部 H2O2 为 HMF 氧化提供 OH 源。如果将 H2O2 的添加量改为多次以降低其浓度,Co-N-C 催化剂上的途径会改变吗?然后,我们将 Co-N-C 与 H2 O2 的 1 次和 10 次添加进行了 HMF 氧化实验以进行验证(图 6c)。有趣的是,当添加 1 mL 的 30 wt % H2O2 (8.8 mmol) 以代替 O2 进料时,Co-N-C + H2O2 + OH 系统除了氧化降解外,还会发生选择性氧化。正如 ESR 所证明的(图 6d),Co-N-C 显示出将 H2O2 转化为 OH的 POD 样活性,OHI/I0 在反应的初始阶段(1 分钟)为 1.5。它高于 O2 进料情况 (1.1),进一步证实了 H2O2 的一次性添加会导致形成相对更高浓度的 OH,这导致 300 min 时 HMF 分解 33.0%。随着 H2O2 的逐渐消耗,OH 浓度相应降低,从而表现出将 HMF 氧化为 FDCA 的能力不足,当 HMF 完全消耗时,FDCA 的选择性较低 (33.7%)。如果为 0.以相等的间隔添加 10 次 30 wt % H2O2 的 1 mL,OH 的初始强度较低,添加 H2O2 不再持续降低。然而,OH 的瞬时浓度仍然高于 O2 进料情况(图 6d)。虽然在这种情况下 FDCA 选择性提高到 70.9%,但在反应的初始阶段仍有 20.6% 的添加 HMF 被分解。此外,碳平衡并未得到满足,仅达到 79.4%。
In the case of adding H2O2 at one time, i.e., the H2O2 + OH system, the I/I0 of OH reached 4.8 in the initial stage of the reaction. It was more than 3 times higher than those of the above-mentioned three cases (Figure 6d, lines A, B, and C). Thus, HMF was degraded rapidly, and the mass spectrum of the gas product confirmed that the main degradation product was CO2 (Figure S16). These results further suggested that the transformation of ROS (O2•– and H2O2) in the Co–N–C system could ensure the in situ and continuous generation of OH with low concentration, and only in situ controllable production of OH was beneficial to the highly selective oxidation of hydroxyl compounds. Reducing the concentration of externally added H2O2 is beneficial for selective oxidation but not enough to reduce the oxidative degradation to a sufficiently low extent as in Co–N–C.
在一次加入 H2O2 的情况下,即 H2O2 + OH 系统,OHI/I0 在反应的初始阶段达到 4.8。它比上述三种情况高 3 倍以上(图 6d,A、B 和 C 行)。因此,HMF 迅速降解,气体产物的质谱证实主要降解产物是 CO2图 S16)。这些结果进一步表明,ROS(O2•– 和 H2O2)在 Co–N–C 系统中的转化可以保证低浓度 OH 的原位连续生成,并且只有原位可控地产生 OH 才有利于羟基化合物的高选择性氧化。降低外部添加的 H2O2 的浓度有利于选择性氧化,但不足以将氧化降解降低到足够低的程度,如 Co-N-C。

2.5. Control in OH Produces via pH Variation
2.5. OH中的控制• 通过改变 pH 值产生

Obviously, the concentration of OH formed in situ was sufficiently low compared to that of the externally added H2O2 to ensure selective oxidation rather than oxidative degradation. Further studies found that adjusting the pH value in the Co–N–C system could tune the intensity of the OH formed in situ. As shown in Figure 7a, no signal for OH appeared in a neutral aqueous solution (pH = 7.02), and HMF was converted slowly with extremely low selectivity for FDCA. When increasing to pH = 8.16 by adding NaHCO3, weak peaks of the OH signal could be observed. As the pH value increased to 13.32 by adding NaOH, the OH intensity continuously increased. The results showed that the OH formation activity of Co–N–C closely depended on the basic intensity of the aqueous solution. The yield of FDCA and pH value had a volcano curve relationship, in which pH = 12.79 was the best with the highest yield of 98.0% (Figure 7d). This is consistent with the reported studies that inorganic bases such as NaOH, Na2CO3, NaHCO3, etc. were necessary for the oxidation of HMF to FDCA when using non-noble metal catalysts as alkaline conditions could facilitate the dehydrogenation process and prevent the adsorption of FDCA on the catalyst surface. (31) While the highly selective oxidation of HMF under neutral or weakly basic environments could occur on noble metal catalysts. (24,25) NaOH in this Co–N–C system also played an additional role, that is, by adjusting to appropriate alkaline conditions, O2 was activated to OH with a suitable concentration, which could ensure the highly selective oxidation of HMF at room temperature. Clearly, excessive OH was not conducive to the selective oxidation due to causing side reactions. In contrast, HMF almost completely decomposed without FDCA formation by using the H2O2 + OH system with initial OH intensity 4 times higher than that of the Co–N–C system at the optimized pH value (12.79). These results verified that the high OH concentration was unfavorable for the enhancement of catalytic selectivity. Therefore, regulation of OH concentration to the appropriate level could be realized by adjusting the pH value in the Co–N–C system, thus exhibiting the optimal performance of catalytic selective oxidation of hydroxyl compounds.
显然,与外部添加的 H2O2 相比,原位形成的 OH 浓度足够低,以确保选择性氧化而不是氧化降解。进一步的研究发现,调整 Co-N-C 系统中的 pH 值可以调节原位形成的 OH 的强度。如图 7a 所示,在中性水溶液 (pH = 7.02) 中没有出现 OH 信号,并且 HMF 转化缓慢,对 FDCA 的选择性极低。当通过添加 NaHCO3 增加到 pH = 8.16 时,可以观察到 OH 信号的弱峰。随着 NaOH 的加入使 pH 值升高到 13.32,OH 强度持续增加。结果表明,Co-N-C 的 OH 形成活性与水溶液的基本强度密切相关。FDCA 的产量与 pH 值呈火山曲线关系,其中 pH = 12.79 最好,最高产量为 98.0%(图 7d)。这与报道的研究一致,即当使用非贵金属催化剂作为碱性条件时,HMF 氧化成 FDCA 所必需的无机碱如 NaOH、Na2、CO3、NaHCO3 等可以促进脱氢过程并防止 FDCA 在催化剂表面的吸附。(31) 而 HMF 在中性或弱碱性环境中的高选择性氧化可能发生在贵金属催化剂上。 (24,25) NaOH 在这个 Co-N-C 体系中还起了额外的作用,即通过调整到适当的碱性条件,O2 被活化为具有适当浓度的 OH ,这可以保证 HMF 在室温下的高度选择性氧化。显然,过量的 OH 由于引起副反应,不利于选择性氧化。相比之下,在最佳 pH 值 (4) 下,HMF 使用 H2O2 + OH 系统几乎完全分解,没有形成 FDCA,初始 OH 强度比 Co-N-C 系统高 12.79 倍。这些结果验证了高 OH 浓度不利于催化选择性的增强。因此,可以通过调节 Co-N-C 系统中的 pH 值来实现将 OH 浓度调节到适当的水平,从而表现出羟基化合物催化选择性氧化的最佳性能。

Figure 7 图 7

Figure 7. ESR spectra of DMPO-OH adducts at the reaction time of 30 min in the H2O2 + OH system at a pH value of 12.79 and Co–N–C system with different pH values (a); the intensity of the DMPO-OH adducts in the ESR spectra (b); Arrhenius plots of HMF oxidation with and without OH in the Co–N–C system (c); the product yield in HMF oxidation in the Co–N–C system with different pH values and in the H2O2 + OH system at a pH value of 12.79 (d); and the schematic illustration of the HMF oxidation with OH formation in the Co–N–C system (e). Reaction conditions: 0.5 mmol HMF, 0.032 mmol Co in Co–N–C, 5 mL of H2O, 60 °C, 40 mL/min O2, 5 h, at different pH values; a0.5 mmol HMF, 1 mL of 30 wt % H2O2, 4 mL of H2O, 60 °C, 5 h, and pH = 12.79.
图 7.DMPO-OH 加合物在 pH 值为 12.79 的 H2O2 + OH 系统和不同 pH 值的 Co-N-C 系统中反应时间为 30 分钟的 ESR 光谱 (a);ESR 光谱中 DMPO-OH 加合物的强度 (b);Co-N-C 系统中含和不含 OH 的 HMF 氧化的 Arrhenius 图 (c);在不同 pH 值的 Co-N-C 体系和 pH 值为 12.79 的 H2O2 + OH 体系中 HMF 氧化的产物收率 (d);以及 Co-N-C 系统中 HMF 氧化与 OH 形成的示意图 (e)。反应条件:0.5 mmol HMF,0.032 mmol Co in Co–N–C,5 mL H2O,60 °C,40 mL/min O2,5 h,不同 pH 值;0.5mmol HMF、1 mL 30 wt % H2O2、4 mL H2O、60 °C、5 h 和 pH = 12.79。

The calculated apparent activation energy (Ea) of HMF oxidation in the Co–N–C system with different pH values was also determined (Figure 7c). Compared with OH free in pH = 7.02, the Ea value in pH = 12.79 with OH formation was reduced from 58.4 to 25.7 kJ/mol, and lower than that reported for the noble metal catalyst (Pt/C, 29.0 kJ/mol). (45) The controllable production of OH with strong oxidizing ability by the activation of O2 and H2O over the Co–N–C catalyst contributed to the decrease of the apparent activation energy in the oxidation of hydroxyl compounds.
还确定了不同 pH 值下 Co-N-C 系统中 HMF 氧化的计算表观活化能 (Ea)(图 7c)。与 pH = 7.02 中的不含 OH 相比,OH 形成时 pH = 12.79 时的 Ea 值从 58.4 kJ/mol 降低到 25.7 kJ/mol,低于贵金属催化剂报道的值(Pt/C,29.0 kJ/mol)。(45) 通过在 Co-N-C 催化剂上活化 O2 和 H2O 可控地产生具有强氧化能力的 OH 有助于降低羟基化合物氧化过程中的表观活化能。

2.6. OH Accelerating the H-Abstraction of C–H Bonds
2.6. OH 加速 C-H 键的 H 抽象

The oxidation of −CH2OH groups to −C═O groups often required hydrogen abstraction, i.e., cleavage of two classes of chemical bonds, O–H and C–H. In the conventional OH generating systems, the strong electrophilicity of OH was employed in the H-atom abstraction of alkanes or aromatics, thus leading to the degradation of them. (5) To explore the relationship between OH and hydrogen abstraction in −CH2OH oxidation in this Co–N–C system, the kinetic isotope effects (KIE) were studied by using model substrates of benzyl alcohol deuterated at the methylene group (PhCD2OH) and at the hydroxyl group (PhCH2OD), respectively. As shown in Figure S19 and Table S2, the KIE value (k(PhCH2OH)/k(PhCH2OD)) for PhCH2OD was 1.27, while the PhCD2OH gave a higher KIE value (k(PhCH2OH)/k(PhCD2OH)) of 2.40. These results indicated that the hydrogen abstraction of the C–H bond in the −CH2OH group was involved in the rate-determining step of the PhCH2OH oxidation. Then, increasing the OH concentration by the enhancement of pH values in the reaction system led to all the values of k(PhCH2OH), k(PhCD2OH), and k(PhCH2OD) increasing. Meanwhile, the KIE value for PhCD2OH decreased from 2.40 to 1.82 with the increase of the OH concentration. In contrast, the KIE value for PhCH2OD was almost unchanged. Although the acidity of hydroxyl hydrogen is stronger than that of C–H in the −CH2OH group, the enhancement of alkalinity in the reaction system caused by the pH increase did not lead to the activation of O–H in the PhCH2OH oxidation process. Therefore, the promoting effect of the enhancement of alkalinity on the breaking of C–H bonds could be ignored. These results indicated that the hydrogen abstraction of C–H bonds could be favored by OH at appropriate concentrations during the oxidation of −CH2OH groups in hydroxyl compounds.
-CH2OH 基团氧化成 -C═O 基团通常需要提取氢,即裂解两类化学键,O-H 和 C-H。在传统的 OH 生成系统中,OH 的强亲电性被用于烷烃或芳烃的 H 原子提取,从而导致它们的降解。(5) 为了探索该 Co-N-C 系统中 -CH2OH 氧化中 OH 和氢提取之间的关系,分别使用在亚甲基 (PhCD2OH) 和羟基 (PhCH2OD) 处氘化的苯甲醇的模型底物研究了动力学同位素效应 (KIE)。如图 S19 和表 S2 所示,PhCH2OD 的 KIE 值 (k(PhCH2OH)/k(PhCH2OD)) 为 1.27,而 PhCD2OH 的 KIE 值 (k(PhCH2OH)/k(PhCD2OH)) 更高,为 2.40。这些结果表明,−CH2OH 基团中 C-H 键的氢提取参与了 PhCH2OH 氧化的速率确定步骤。然后,通过提高反应体系中的 pH 值来增加 OH 浓度,导致 k(PhCH2OH)、k(PhCD2OH) 和 k(PhCH2OD) 的所有值都增加。同时,随着 OH 浓度的增加,PhCD2OH 的 KIE 值从 2.40 降低到 1.82。相比之下,PhCH2OD 的 KIE 值几乎没有变化。 尽管在 −CH2OH 基团中羟基氢的酸度比 C-H 强,但 pH 值升高引起的反应系统中碱度的增强并不会导致 PhCH2OH 氧化过程中 O-H 的激活。因此,碱度增强对 C-H 键断裂的促进作用可以忽略不计。这些结果表明,在羟基化合物中 −CH2OH 基团的氧化过程中,适当浓度的 OH 可能有利于 C-H 键的氢提取。
The substrate scope for the biomimetic Co–N–C catalyst was extended to a variety of model hydroxyl compounds, including cellulose-derived furan alcohols, lignin-derived aromatic alcohols, and oil-based aromatic alcohols, with excellent yields (84.1–97.4%) of the corresponding acids obtained (Table 3). In comparison with the previous studies, (31,46−48) the thermal catalytic oxidation of these hydroxyl compounds was usually performed under harsh reaction conditions when employing non-noble metal catalysts for the purpose of acid products. For example, cobalt-based catalysts (Co/GS@C, Co-Co3O4@NC, Co3O4, etc.) generally required a temperature above 100 °C for the oxidation of typical lignin-derived model aromatic alcohols. In the case of 3,4,5-trimethoxybenzyl alcohol, Co3O4 only afforded 16.1% yield of 3,4,5-trimethoxybenzoic acid by catalytic oxidation at 140 °C under 4 MPa O2. (48) In contrast, the present non-noble Co–N–C catalyst achieved 95.6% yield of 3,4,5-trimethoxybenzoic acid under significantly milder conditions (at 60 °C under the atmospheric pressure of O2). Considering there were few reports on the non-noble metals catalyzed oxidation of specific fatty alcohols in water, the activity of Co–N–C was tested in the oxidation of 1-butanol. It achieved as high as 79.1% yield of butyric acid, which is not accessible by the reported heterogeneous non-noble catalysts. These results demonstrated the wide generalizability of Co–N–C in the hydroxyl compounds oxidation. The stability and reusability of Co–N–C were also explored. The Co–N–C catalyst could retain its activity for at least four recycling runs, with only a slight reduction of FDCA yield (Figure S20). It is noteworthy that the catalyst could be reused by simply washing with water and soaping in NaOH aqueous solution. While for the majority of the reported N-doped carbon-supported metal-based catalysts for catalytic oxidation, the reduction under H2 or calcination in inert gas at high temperatures was usually required to restore their activity. (30,49) This difference reflected the excellent stability and reusability of Co–N–C. As discussed above, the in situ and continuously generated OH by self-cascade catalysis in the Co–N–C system was conducive to the hydroxyl compounds oxidation via promoting the hydrogen abstraction of C–H bonds in the −CH2OH groups, thus realizing that the highly selective oxidation of multiple sources of hydroxyl compounds to the corresponding acids could be performed under mild conditions.
仿生 Co-N-C 催化剂的底物范围扩展到各种模型羟基化合物,包括纤维素衍生的呋喃醇、木质素衍生的芳香醇和油基芳香醇,获得的相应酸的产率极高 (84.1–97.4%)(表 3)。与以前的研究相比,(31,46−48) 当使用非贵金属催化剂生产酸产品时,这些羟基化合物的热催化氧化通常是在恶劣的反应条件下进行的。例如,钴基催化剂(Co/GS@C、Co-Co3O4@NC、Co3O4 等)通常需要 100 °C 以上的温度才能氧化典型的木质素衍生的模型芳香醇。在 3,4,5-三甲氧基苯甲醇的情况下,Co3O4 在 140 °C 下在 4 MPa O2 下催化氧化,仅得到 3,4,5-三甲氧基苯甲酸的 16.1% 产率。(48) 相比之下,目前的非贵金属 Co-N-C 催化剂在明显较温和的条件下(在 O2 的大气压下为 60 °C)实现了 3,4,5-三甲氧基苯甲酸的 95.6% 产率。考虑到关于非贵金属催化水中特定脂肪醇氧化的报道很少,因此在 1-丁醇的氧化中测试了 Co-N-C 的活性。它实现了高达 79.1% 的丁酸收率,这是已报道的非均相非贵金属催化剂无法获得的。这些结果表明 Co-N-C 在羟基化合物氧化中的广泛泛化性。还探讨了 Co-N-C 的稳定性和可再利用性。 Co-N-C 催化剂可以保持其活性至少四次回收运行,而 FDCA 收率仅略有降低(图 S20)。值得注意的是,只需用水清洗并在 NaOH 水溶液中皂洗即可重复使用该催化剂。而对于大多数已报道的用于催化氧化的 N 掺杂碳负载金属基催化剂,通常需要在 H2 下还原或在高温下在惰性气体中煅烧以恢复其活性。(30,49) 这种差异反映了 Co-N-C 出色的稳定性和可重用性。如上所述,在 Co-N-C 系统中通过自级联催化原位和连续生成的 OH 通过促进 -CH2OH 基团中 C-H 键的氢提取,有利于羟基化合物氧化,从而实现多种来源的羟基化合物高度选择性氧化为相应的酸在温和的条件下。
Table 3. Catalytic Oxidation of Various Hydroxyl Compounds over the Co–N–C Catalyst
表 3.各种羟基化合物在 Co-N-C 催化剂上的催化氧化
a

Reaction condition: 0.5 mmol substrate, 0.022 mmol Co in Co–N–C, 1.0 mmol NaOH, 5 mL of H2O, 40 mL/min O2, and 60 °C.


a

反应条件:0.5 mmol 底物、0.022 mmol Co in Co–N–C、1.0 mmol NaOH、5 mL H2O、40 mL/min O2 和 60 °C。

b

0.032 mmol Co in Co–N–C, 1.25 mmol NaOH, 60 °C.


b

0.032 mmol Co 的 Co–N–C 溶液,1.25 mmol NaOH,60 °C。

c

0.016 mmol Co in Co–N–C, 1.25 mmol NaOH, 50 °C.


c

0.016 mmol Co 的 Co–N–C 溶液,1.25 mmol NaOH,50 °C。

3. Conclusions 3. 结论

Click to copy section link
单击以复制部分链接
Section link copied!

In summary, the Co–N–C catalyst enabled the regulated activation of molecular oxygen into reactivity controllable OH via self-cascade catalysis for efficient selective oxidation of a variety of biomass and petroleum-based hydroxyl compounds to organic acids under mild conditions. For the HMF oxidation, the >99.9% conversion of HMF with 94.2% selectivity of FDCA could be obtained at 25 °C under atmospheric pressure, with a high FDCA productivity value (8.1 h–1). The OH with high oxidizing potential could be controllably and continuously generated in situ from O2 and H2O with simple pH variation. The high catalytic performance of the Co–N–C catalyst was closely related to its atomically dispersed CoN4 sites and large conjugated aromatic coordination structures. Moreover, a promotion effect by OH in the hydrogen abstraction of the C–H bond in the −CH2OH group resulted in obviously reduced activation energy. This study provides a promising strategy of regulated activation of molecular oxygen into designed reactive oxygen species for oxidative synthesis of chemicals.
总之,Co-N-C 催化剂能够通过自级联催化将分子氧调节活化为反应性可控的 OH ,从而在温和条件下将各种生物质和石油基羟基化合物高效选择性氧化为有机酸。对于 HMF 氧化,在 25 °C 大气压下可以获得 >99.9% 的 HMF 转化率和 94.2% 的 FDCA 选择性,具有较高的 FDCA 生产率值 (8.1 h–1)。具有高氧化电位的 OH 可以从 O2 和 H2O 原位受控地连续生成,pH 值变化简单。Co-N-C 催化剂的高催化性能与其原子分散的 CoN4 位点和大的共轭芳烃配位结构密切相关。此外,OH 在 −CH2OH 基团中 C-H 键的氢提取中的促进作用导致活化能明显降低。本研究提供了一种很有前途的策略,将分子氧调节活化成设计的活性氧,用于化学品的氧化合成。

Supporting Information 支持信息

Click to copy section link
单击以复制部分链接
Section link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acscatal.3c02736.
支持信息可在 https://pubs.acs.org/doi/10.1021/acscatal.3c02736 免费获取。

  • Details of experimental procedures and data processing, comparison of Co–N–C and reported non-noble catalysts for HMF oxidation, additional characterization data of Co-based catalysts, ESR spectra of DMPO-radical adducts, UV–vis spectra of KI oxidation, GC spectra of Ph3P oxidation, mass spectrum of CO2, reaction time profiles for HMF and HMFCA oxidations, recycling experiments of Co–N–C, kinetics analysis for benzyl alcohol oxidation, and supplementary catalytic evaluation (PDF)
    实验程序和数据处理的详细信息,用于 HMF 氧化的 Co-N-C 和已报道的非贵金属催化剂的比较,Co 基催化剂的其他表征数据,DMPO 自由基加合物的 ESR 光谱,KI 氧化的紫外-可见光谱,Ph3P 氧化的 GC 光谱,CO2 的质谱,HMF 和 HMFCA 氧化的反应时间曲线, Co-N-C 回收实验、苯甲醇氧化动力学分析和补充催化评估 (PDF

Terms & Conditions  条款和条件

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
大多数电子支持信息文件无需订阅 ACS Web 版本即可获得。此类文件可以按文章下载用于研究用途(如果有链接到相关文章的公共使用许可证,则该许可证可能允许其他用途)。可以通过 RightsLink 权限系统请求 ACS 以用于其他用途:http://pubs.acs.org/page/copyright/permissions.html

Author Information 作者信息

Click to copy section link
单击以复制部分链接
Section link copied!

  • Corresponding Authors 通讯作者
    • Hong Ma - State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. China Email: mahong@dicp.ac.cn
      Hong 马 - 中国科学院大连化学物理研究所催化国家重点实验室,中国116023连市中山路457号; 电子邮件: mahong@dicp.ac.cn
    • Jie Xu - State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. ChinaOrcidhttps://orcid.org/0000-0003-2535-094X Email: xujie@dicp.ac.cn
      徐杰 - 中国科学院大连化学物理研究所催化国家重点实验室,中国大连116023中山路457号; Orcid https://orcid.org/0000-0003-2535-094X 电子邮件: xujie@dicp.ac.cn
  • Authors 作者
    • Xin Liu - State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. ChinaUniversity of Chinese Academy of Sciences, Beijing 100049, P. R. China
      刘欣 - 中国科学院大连化学物理研究所催化国家重点实验室,中国大连116023中山路 457 号;中国科学院大学, 北京 100049
    • Meiyun Zhang - State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. ChinaUniversity of Chinese Academy of Sciences, Beijing 100049, P. R. China
      Meiyun Zhang - 中国科学院大连化学物理研究所催化国家重点实验室,中国大连116023中山路457号;中国科学院大学, 北京 100049
    • Penghua Che - State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. China
      Penghua Che - 中国科学院大连化学物理研究所催化国家重点实验室,中国大连116023中山路 457
    • Yang Luo - State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. China
      罗洋 - 中国科学院大连化学物理研究所催化国家重点实验室,中国116023连市中山路457
    • Shujing Zhang - State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, P. R. China
      张淑静 - 中国科学院大连化学物理研究所催化国家重点实验室,中国大连116023中山路457
  • Author Contributions 作者贡献

    J.X., H.M., and X.L. designed the research. X.L. performed the catalyst preparation, characterizations, and tests, and completed the paper. M.Z., P.C., Y.L., and S.Z. participated in beneficial discussions. All authors reviewed and commented on the article.
    J.X.、H.M. 和 X.L. 设计了这项研究。X.L. 进行了催化剂制备、表征和测试,并完成了论文。M.Z.、P.C.、Y.L. 和 S.Z. 参与了有益的讨论。所有作者都对文章进行了审阅和评论。

  • Notes 笔记
    The authors declare no competing financial interest.
    作者声明没有竞争性的经济利益。

Acknowledgments 确认

Click to copy section link
单击以复制部分链接
Section link copied!

This work was supported by the National Natural Science Foundation of China (21790331, 22272172, and 22072149) and the National Key R&D Program of China (2022YFA1504902).
这项工作得到了国家自然科学基金 (21790331、22272172 和 22072149) 和国家重点研发计划 (2022YFA1504902) 的支持。

References 引用

Click to copy section link
单击以复制部分链接
Section link copied!

This article references 49 other publications.
本文引用了其他 49 种出版物。

  1. 1
    Pereira, M. M.; Dias, L. D.; Calvete, M. J. F. Metalloporphyrins: Bioinspired Oxidation Catalysts. ACS Catal. 2018, 8, 1078410808,  DOI: 10.1021/acscatal.8b01871

    1
    佩雷拉,MM;迪亚斯,LD;卡尔维特,MJ F. 金属卟啉:生物启发式氧化催化剂ACS 加泰罗尼亚。 201881078410808 DOI: 10.1021/acscatal.8b01871
    内容
  2. 2
    Piera, J.; Bäckvall, J.-E. Catalytic Oxidation of Organic Substrates by Molecular Oxygen and Hydrogen Peroxide by Multistep Electron Transfer─a Biomimetic Approach. Angew. Chem., Int. Ed. 2008, 47, 35063523,  DOI: 10.1002/anie.200700604

    2
    皮埃拉,J.;Bäckvall, J.-E. 通过多步电子转移通过分子氧和过氧化氢催化氧化有机底物 - 一种仿生方法Angew. Chem., 国际教育 20084735063523 DOI: 10.1002/anie.200700604
    内容
  3. 3
    Li, Q.; Li, F.-t. Recent Advances in Molecular Oxygen Activation via Photocatalysis and its Application in Oxidation Reactions. Chem. Eng. J. 2021, 421, 129915  DOI: 10.1016/j.cej.2021.129915

    3
    李婉;李 F.-t. 光催化分子氧活化及其在氧化反应中的应用的最新进展化学工程学杂志 2021421129915  DOI: 10.1016/j.cej.2021.129915
    内容
  4. 4
    Metz, M.; Solomon, E. I. Dioxygen Binding to Deoxyhemocyanin: Electronic Structure and Mechanism of the Spin-Forbidden Two-Electron Reduction of O2. J. Am. Chem. Soc. 2001, 123, 49384950,  DOI: 10.1021/ja004166b

    4
    梅茨,M.;所罗门,EI 双氧结合脱氧血蓝蛋白:O2 的自旋禁止双电子还原的电子结构和机制J. Am. Chem. Soc. 200112349384950 DOI: 10.1021/ja004166b
  5. 5
    Gligorovski, S.; Strekowski, R.; Barbati, S.; Vione, D. Environmental Implications of Hydroxyl Radicals (OH). Chem. Rev. 2015, 115, 1305113092,  DOI: 10.1021/cr500310b

    5
    格里戈罗夫斯基,S.;斯特雷科夫斯基,R.;巴巴蒂,S.;维奥内 D. 羟基自由基的环境影响 (OH)。化学修订版 20151151305113092 DOI: 10.1021/cr500310b
    内容
  6. 6
    Hou, J.-T.; Zhang, M.; Liu, Y.; Ma, X.; Duan, R.; Cao, X.; Yuan, F.; Liao, Y.-X.; Wang, S.; Ren, W. X. Fluorescent Detectors for Hydroxyl Radical and their Applications in Bioimaging: A Review. Coord. Chem. Rev. 2020, 421, 213457  DOI: 10.1016/j.ccr.2020.213457

    6
    侯 J.-T.;张 M.;刘 Y.;马,X.;段 R.;曹 X.;袁 F.;廖 Y.-X.;王 S.;任,W. X. 羟基自由基荧光检测器及其在生物成像中的应用:综述Coord. Chem. Rev. 2020421213457  DOI: 10.1016/j.ccr.2020.213457
    内容
  7. 7
    Wang, B.; Su, H.; Zhang, B. Hydrodynamic Cavitation as a Promising Route for Wastewater Treatment–A Review. Chem. Eng. J. 2021, 412, 128685  DOI: 10.1016/j.cej.2021.128685

    7
    王 B.;苏,H.;张斌 流体动力空化作为废水处理的有前途的路线——综述化学工程学杂志 2021412128685  DOI: 10.1016/j.cej.2021.128685
    内容
  8. 8
    Cheng, C.; Ren, W.; Miao, F.; Chen, X.; Chen, X.; Zhang, H. Generation of FeIV═O and its Contribution to Fenton-Like Reactions on a Single-Atom Iron-N-C Catalyst. Angew. Chem., Int. Ed. 2023, 62, e202218510  DOI: 10.1002/ange.202218510

    8
    程 C.;任, W.;苗, F.;陈 X.;陈 X.;张 H. FeIV═O 的生成及其对单原子铁-N-C 催化剂上 Fenton 类反应的贡献Angew. Chem., 国际教育 202362e202218510  DOI: 10.1002/ange.202218510
    内容
  9. 9
    Chen, Z.; Zhou, H.; Kong, F.; Wang, M. Piezocatalytic Oxidation of 5-Hydroxymethylfurfural to 5-Formyl-2-Furancarboxylic Acid over Pt Decorated Hydroxyapatite. Appl. Catal., B 2022, 309, 121281  DOI: 10.1016/j.apcatb.2022.121281

    9
    陈 Z.;周, H.;孔,F.;王 M. 5-羟甲基糠醛在 Pt 修饰的羟基磷灰石上压电催化氧化为 5-甲酰基-2-呋喃羧酸应用加泰罗尼亚, B 2022309121281  DOI: 10.1016/j.apcatb.2022.121281
    内容
  10. 10
    Liu, Y.; Wang, M.; Zhang, B.; Yan, D.; Xiang, X. Mediating the Oxidizing Capability of Surface-Bound Hydroxyl Radicals Produced by Photoelectrochemical Water Oxidation to Convert Glycerol into Dihydroxyacetone. ACS Catal. 2022, 12, 69466957,  DOI: 10.1021/acscatal.2c01319
  11. 11
    Zope, B. N.; Hibbitts, D. D.; Neurock, M.; Davis, R. J. Reactivity of the Gold/Water Interface During Selective Oxidation Catalysis. Science 2010, 330, 7478,  DOI: 10.1126/science.1195055
  12. 12
    Kung, M. C.; Davis, R. J.; Kung, H. H. Understanding Au-Catalyzed Low-Temperature CO Oxidation. J. Phys. Chem. C 2007, 111, 1176711775,  DOI: 10.1021/jp072102i
  13. 13
    Takanabe, K.; Iglesia, E. Rate and Selectivity Enhancements Mediated by OH Radicals in the Oxidative Coupling of Methane Catalyzed by Mn/Na2WO4/SiO2. Angew. Chem., Int. Ed. 2008, 47, 76897693,  DOI: 10.1002/anie.200802608
  14. 14
    Jiang, D.; Ni, D.; Rosenkrans, Z. T.; Huang, P.; Yan, X.; Cai, W. Nanozyme: New Horizons for Responsive Biomedical Applications. Chem. Soc. Rev. 2019, 48, 36833704,  DOI: 10.1039/C8CS00718G
  15. 15
    Wu, J.; Wang, X.; Wang, Q.; Lou, Z.; Li, S.; Zhu, Y.; Qin, L.; Wei, H. Nanomaterials with Enzyme-Like Characteristics (Nanozymes): Next-Generation Artificial Enzymes (II). Chem. Soc. Rev. 2019, 48, 10041076,  DOI: 10.1039/C8CS00457A
  16. 16
    Xu, B.; Wang, H.; Wang, W.; Gao, L.; Li, S.; Pan, X.; Wang, H.; Yang, H.; Meng, X.; Wu, Q.; Zheng, L.; Chen, S.; Shi, X.; Fan, K.; Yan, X.; Liu, H. A Single-Atom Nanozyme for Wound Disinfection Applications. Angew. Chem., Int. Ed. 2019, 58, 49114916,  DOI: 10.1002/anie.201813994
  17. 17
    Zhu, S.; Li, Z.; Zhang, F.; Liu, F.; Ni, P.; Chen, C.; Jiang, Y.; Lu, Y. Single-Atom Cobalt Catalysts as Highly Efficient Oxidase Mimics for Time-Based Visualization Monitoring the TAC of Skin Care Products. Chem. Eng. J. 2023, 456, 141053  DOI: 10.1016/j.cej.2022.141053
  18. 18
    Ye, Y.; Xiao, L.; Bin, H.; Zhang, Q.; Nie, T.; Yang, X.; Wu, D.; Cheng, H.; Li, P.; Wang, Q. Oxygen-Tuned Nanozyme Polymerization for the Preparation of Hydrogels with Printable and Antibacterial Properties. J. Mater. Chem. B 2017, 5, 15181524,  DOI: 10.1039/C6TB03317B
  19. 19
    Mondelli, C.; Gözaydın, G.; Yan, N.; Pérez-Ramírez, J. Biomass Valorisation over Metal-Based Solid Catalysts from Nanoparticles to Single Atoms. Chem. Soc. Rev. 2020, 49, 37643782,  DOI: 10.1039/D0CS00130A
  20. 20
    Liu, S.-S.; Sun, K.-Q.; Xu, B.-Q. Specific Selectivity of Au-Catalyzed Oxidation of Glycerol and Other C3-Polyols in Water without the Presence of a Base. ACS Catal. 2014, 4, 22262230,  DOI: 10.1021/cs5005568
  21. 21
    Zhang, J.; Nagamatsu, S.; Du, J.; Tong, C.; Fang, H.; Deng, D.; Liu, X.; Asakura, K.; Yuan, Y. A Study of FeNx/C Catalysts for the Selective Oxidation of Unsaturated Alcohols by Molecular Oxygen. J. Catal. 2018, 367, 1626,  DOI: 10.1016/j.jcat.2018.08.004
  22. 22
    Tu, C.; Nie, X.; Chen, J. G. Insight into Acetic Acid Synthesis from the Reaction of CH4 and CO2. ACS Catal. 2021, 11, 33843401,  DOI: 10.1021/acscatal.0c05492
  23. 23
    Thomas, J. M.; Raja, R.; Sankar, G.; Bell, R. G. Molecular Sieve Catalysts for the Regioselective and Shape-Selective Oxyfunctionalization of Alkanes in Air. Acc. Chem. Res. 2001, 34, 191200,  DOI: 10.1021/ar970020e
  24. 24
    Wang, Q.; Hou, W.; Li, S.; Xie, J.; Li, J.; Zhou, Y.; Wang, J. Hydrophilic Mesoporous Poly(ionic liquid)-Supported Au–Pd Alloy Nanoparticles towards Aerobic Oxidation of 5-Hydroxymethylfurfural to 2,5-Furandicarboxylic Acid under Mild Conditions. Green Chem. 2017, 19, 38203830,  DOI: 10.1039/C7GC01116D
  25. 25
    Liu, X.; Chen, L.; Xu, H.; Jiang, S.; Zhou, Y.; Wang, J. Straightforward Synthesis of Beta Zeolite Encapsulated Pt Nanoparticles for the Transformation of 5-Hydroxymethyl Furfural into 2,5-Furandicarboxylic Acid. Chin. J. Catal. 2021, 42, 9941003,  DOI: 10.1016/S1872-2067(20)63720-2
  26. 26
    Zhang, M.; Ma, H.; Liu, X.; Zhang, S.; Luo, Y.; Gao, J.; Xu, J. Control in Local Coordination Environment Boosting Activating Molecular Oxygen with an Atomically Dispersed Binary Mn–Co Catalyst. ACS Appl. Mater. Interfaces 2022, 14, 1853918549,  DOI: 10.1021/acsami.2c01858
  27. 27
    Liu, X.; Luo, Y.; Ma, H.; Zhang, S.; Che, P.; Zhang, M.; Gao, J.; Xu, J. Hydrogen-Binding-Initiated Activation of O–H Bonds on a Nitrogen-Doped Surface for the Catalytic Oxidation of Biomass Hydroxyl Compounds. Angew. Chem., Int. Ed. 2021, 133, 1825118258,  DOI: 10.1002/ange.202103604
  28. 28
    Sun, Y.; Ma, H.; Jia, X.; Ma, J.; Luo, Y.; Gao, J.; Xu, J. A High-Performance Base-Metal Approach for the Oxidative Esterification of 5-Hydroxymethylfurfural. ChemCatChem 2016, 8, 29072911,  DOI: 10.1002/cctc.201600484
  29. 29
    Sun, Y.; Ma, H.; Luo, Y.; Zhang, S.; Gao, J.; Xu, J. Activation of Molecular Oxygen Using Durable Cobalt Encapsulated with Nitrogen-Doped Graphitic Carbon Shells for Aerobic Oxidation of Lignin-Derived Alcohols. Chem. - Eur. J. 2018, 24, 46534661,  DOI: 10.1002/chem.201705824
  30. 30
    Wang, T.; Ma, H.; Liu, X.; Luo, Y.; Zhang, S.; Sun, Y.; Wang, X.; Gao, J.; Xu, J. Ultrahigh-Content Nitrogen-Doped Carbon Encapsulating Cobalt NPs as Catalyst for Oxidative Esterification of Furfural. Chem. - Asian J. 2019, 14, 15151522,  DOI: 10.1002/asia.201900099
  31. 31
    Chen, C.; Wang, L.; Zhu, B.; Zhou, Z.; El-Hout, S. I.; Yang, J.; Zhang, J. 2, 5-Furandicarboxylic Acid Production via Catalytic Oxidation of 5-Hydroxymethylfurfural: Catalysts, Processes and Reaction Mechanism. J. Energy Chem. 2021, 54, 528554,  DOI: 10.1016/j.jechem.2020.05.068
  32. 32
    Yang, G.; Ma, Y.; Xu, J. Biomimetic Catalytic System Driven by Electron Transfer for Selective Oxygenation of Hydrocarbon. J. Am. Chem. Soc. 2004, 126, 1054210543,  DOI: 10.1021/ja047297b
  33. 33
    Dijkman, W. P.; Groothuis, D. E.; Fraaije, M. W. Enzyme-Catalyzed Oxidation of 5-Hydroxymethylfurfural to Furan-2, 5-Dicarboxylic Acid. Angew. Chem., Int. Ed. 2014, 53, 65156518,  DOI: 10.1002/anie.201402904
  34. 34
    Jia, H. Y.; Zong, M. H.; Zheng, G. W.; Li, N. One-Pot Enzyme Cascade for Controlled Synthesis of Furancarboxylic Acids from 5-Hydroxymethylfurfural by H2O2 Internal Recycling. ChemSusChem 2019, 12, 47644768,  DOI: 10.1002/cssc.201902199
  35. 35
    Liu, W.; Zhang, L.; Yan, W.; Liu, X.; Yang, X.; Miao, S.; Wang, W.; Wang, A.; Zhang, T. Single-Atom Dispersed Co–N–C Catalyst: Structure Identification and Performance for Hydrogenative Coupling of Nitroarenes. Chem. Sci. 2016, 7, 57585764,  DOI: 10.1039/C6SC02105K
  36. 36
    Jagadeesh, R. V.; Junge, H.; Pohl, M.-M.; Radnik, Jr.; Brückner, A.; Beller, M. Selective Oxidation of Alcohols to Esters Using Heterogeneous Co3O4–N@C Catalysts under Mild Conditions. J. Am. Chem. Soc. 2013, 135, 1077610782,  DOI: 10.1021/ja403615c
  37. 37
    Madden, K. P.; Taniguchi, H. The Role of the DMPO-Hydrated Electron Spin Adduct in DMPO-OH Spin Trapping. Free Radical Biol. Med. 2001, 30, 13741380,  DOI: 10.1016/S0891-5849(01)00540-8
  38. 38
    Yamazaki, I.; Piette, L. H. EPR Spin-Trapping Study on the Oxidizing Species Formed in the Reaction of the Ferrous Ion with Hydrogen Peroxide. J. Am. Chem. Soc. 1991, 113, 75887593,  DOI: 10.1021/ja00020a021
  39. 39
    Qi, Y.; Li, J.; Zhang, Y.; Cao, Q.; Si, Y.; Wu, Z.; Akram, M.; Xu, X. Novel Lignin-Based Single Atom Catalysts as Peroxymonosulfate Activator for Pollutants Degradation: Role of Single Cobalt and Electron Transfer Pathway. Appl. Catal., B 2021, 286, 119910  DOI: 10.1016/j.apcatb.2021.119910
  40. 40
    Tran, P.; Kopel, J.; Risitic, B.; Marsh, H.; Fralick, J.; Reid, T. Antimicrobial Seleno-Organic Coatings and Compounds Acting Primarily on the Plasma Membrane: A Review. Adv. Redox Res. 2022, 4, 100031  DOI: 10.1016/j.arres.2022.100031
  41. 41
    Shiibashi, T.; Iida, T. J. D. NADPH and NADH Serve as Electron Donor for the Superoxide-Generating Enzyme in Tilapia (Oreochromis Niloticus) Neutrophils. Dev. Comp. Immunol. 2001, 25, 461465,  DOI: 10.1016/S0145-305X(01)00005-2
  42. 42
    Gleason, J. E.; Galaleldeen, A.; Peterson, R. L.; Taylor, A. B.; Holloway, S. P.; Waninger-Saroni, J.; Cormack, B. P.; Cabelli, D. E.; Hart, P. J.; Culotta, V. C. Candida Albicans SOD5 Represents the Prototype of an Unprecedented Class of Cu-only Superoxide Dismutases Required for Pathogen Defense. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, 58665871,  DOI: 10.1073/pnas.1400137111
  43. 43
    Jiao, L.; Wu, J.; Zhong, H.; Zhang, Y.; Xu, W.; Wu, Y.; Chen, Y.; Yan, H.; Zhang, Q.; Gu, W. Densely Isolated FeN4 Sites for Peroxidase Mimicking. ACS Catal. 2020, 10, 64226429,  DOI: 10.1021/acscatal.0c01647
  44. 44
    Urbański, N. K.; Beręsewicz, A. Generation of OH Initiated by Interaction of Fe2+ and Cu+ with Dioxygen; Comparison with the Fenton Chemistry. Acta Biochim. Pol. 2000, 47, 951962,  DOI: 10.18388/abp.2000_3950
  45. 45
    Chen, C.; Li, X.; Wang, L.; Liang, T.; Wang, L.; Zhang, Y.; Zhang, J. Highly Porous Nitrogen- and Phosphorus-Codoped Graphene: An Outstanding Support for Pd Catalysts to Oxidize 5-Hydroxymethylfurfural into 2, 5-Furandicarboxylic Acid. ACS Sustainable Chem. Eng. 2017, 5, 1130011306,  DOI: 10.1021/acssuschemeng.7b02049
  46. 46
    Senthamarai, T.; Chandrashekhar, V. G.; Rockstroh, N.; Rabeah, J.; Bartling, S.; Jagadeesh, R. V.; Beller, M. A “Universal” Catalyst for Aerobic Oxidations to Synthesize (Hetero)aromatic Aldehydes, Ketones, Esters, Acids, Nitriles, and Amides. Chem 2022, 8, 508531,  DOI: 10.1016/j.chempr.2021.12.001
  47. 47
    Xu, D.; Li, J.; Li, B.; Zhao, H.; Zhu, H.; Kou, J.; Zhang, F.; Dong, Z.; Ma, J. Selective Oxidation of Alcohols to High Value-Added Carbonyl Compounds using Air over Co-Co3O4@NC Catalysts. Chem. Eng. J. 2022, 434, 134545  DOI: 10.1016/j.cej.2022.134545
  48. 48
    Mate, V.; Jha, A.; Joshi, U.; Patil, K.; Shirai, M.; Rode, C. Effect of Preparation Parameters on Characterization and Activity of Co3O4 Catalyst in Liquid Phase Oxidation of Lignin Model Substrates. Appl. Catal., A 2014, 487, 130138,  DOI: 10.1016/j.apcata.2014.08.023
  49. 49
    Zhou, H.; Xu, H.; Wang, X.; Liu, Y. Convergent Production of 2,5-Furandicarboxylic Acid from Biomass and CO2. Green Chem. 2019, 21, 29232927,  DOI: 10.1039/C9GC00869A

Cited By

Click to copy section linkSection link copied!
Citation Statements
  • Supporting
    Supporting0
  • Mentioning
    Mentioning1
  • Contrasting
    Contrasting0
Explore this article's citation statements on scite.ai

This article is cited by 1 publications.

  1. Xinying Tan, Shenghe Si, Difei Xiao, Xiaolei Bao, Kepeng Song, Zeyan Wang, Yuanyuan Liu, Zhaoke Zheng, Peng Wang, Ying Dai, Baibiao Huang, Hefeng Cheng. Single Cobalt Atoms Induced Molecular O2 Activation for Enhanced Photocatalytic Biomass Upgrading on ZnIn2S4 Nanosheets. ACS Catalysis 2023, 13 (21) , 14395-14403. https://doi.org/10.1021/acscatal.3c03332
Open PDF 打开 PDF

ACS Catalysis ACS 催化

Cite this: ACS Catal. 2023, 13, 16, 11104–11116
引用此文档:ACS Catal.2023, 13, 16, 11104–11116
Click to copy citation 点击复制引文Citation copied!
https://doi.org/10.1021/acscatal.3c02736
Published August 7, 2023
已发布 2023 年 8 月 7 日
Copyright © 2023 American Chemical Society
版权所有 © 2023 美国化学会

Article Views 文章浏览量

2304

Altmetric

-

Citations 引文

Learn about these metrics
了解这些指标

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract 抽象

    Scheme 1 方案 1

    Scheme 1. Comparison of OH Generation in Different Systems (a); Design of the Biomimetic System for In Situ and Continuous Generation of OH (b); Conversion of RCH2OH in Different OH Generation Systems: Oxidative Degradation of RCH2OH with Uncontrollable Production of OH (c, Left), Selective Oxidation of RCH2OH with Controllable Production of OH (c, Right)
    方案 1.不同系统中 OH 生成的比较 (a);用于原位和连续生成 OH 的仿生系统设计 (b);RCH2OH 在不同 OH 生成系统中的转化:RCH2OH 的氧化降解与 OH 的不可控产生(c,左),RCH2OH 的选择性氧化与 OH 的可控产生(c,右)

    Figure 1 图 1

    Figure 1. TEM image (a), HAADF STEM image (b), and corresponding EDX maps for the overlapping Co, N, C, and O (c, d) of Co–N–C; normalized XANES spectra (e), EXAFS spectra (f), and wavelet transforms for the EXAFS spectra (g) of Co–N–C and Co foil; N 1s XPS spectrum (h), AC HAADF-STEM image (i) of Co–N–C; TEM images of Co/N–C-1 (j), Co/N–C-2 (k), and Co/AC (l); XRD patterns (m), ratio of the intensity of DMPO-OH adducts in the electron spin resonance (ESR) spectra at the reaction times of 20 min (n), and the catalytic performance of HMF oxidation (o) of different Co-based catalysts. Reaction conditions: 0.5 mmol HMF, 1.25 mmol NaOH, 0.032 mmol Co in the catalyst, 5 mL of H2O, 40 mL/min O2, 60 °C, and 5 h.
    图 1.TEM 图像 (a)、HAADF STEM 图像 (b) 以及 Co-N-C 重叠的 Co、N、C 和 O (c, d) 的相应 EDX 图;Co-N-C 和 Co 箔的 EXAFS 光谱 (g) 的归一化 XANES 光谱 (e)、EXAFS 光谱 (f) 和小波变换;N 1s XPS 谱图 (h),Co–N–C 的 AC HAADF-STEM 图像 (i);Co/N-C-1 (j)、Co/N-C-2 (k) 和 Co/AC (l) 的 TEM 图像;XRD 图谱 (m),反应时间为 20 min 时电子自旋共振 (ESR) 光谱中 DMPO-OH 加合物强度的比值 (n),以及不同钴基催化剂的 HMF 氧化催化性能 (o)。反应条件:催化剂中加入 0.5 mmol HMF、1.25 mmol NaOH、0.032 mmol Co、5 mL H2O、40 mL/min O2、60 °C 和 5 h。

    Figure 2 图 2

    Figure 2. Cascade process of OH generation by the activation of O2 and H2O over Co–N–C (a); ESR spectrum of the DMPO-O2•– in the toluene phase of step 1 (b); MS spectrum of triphenylphosphine oxide formed in the water phase of step 2 (c); ESR spectrum of the DMPO-OH in the water phase of step 3 (d); and effect of solvents and inhibitors on the generation of ROS (e, f). Reaction conditions: for the detection of H2O2 and OH in water, 0.032 mmol Co in Co–N–C, 1.25 mmol NaOH, 1 mmol scavenger, 2.5 mL of H2O, 40 mL/min O2, and 60 °C and for the detection of O2•– in toluene, 0.032 mmol Co in Co–N–C, 1.25 mmol NaOH, 1 mmol scavenger, 2.5 mL of toluene, 0.6 MPa O2, and 100 °C.
    图 2.通过 Co-N-C 上 O2 和 H2O 的活化产生 OH 的级联过程 (a);DMPO-O 的 ESR 谱2•– 在步骤 1 (b) 的甲苯相中;在步骤 2 (c) 的水相中形成的三苯基氧化膦的 MS 谱图;步骤 3 (d) 水中 DMPO-OH 的 ESR 谱图;以及溶剂和抑制剂对 ROS 生成的影响 (e, f)。反应条件:用于检测水中 H2O2 和 OH、0.032 mmol Co 的 Co-N-C 溶液、1.25 mmol NaOH、1 mmol 清除剂、2.5 mL 的 H2O、40 mL/min O2 和 60 °C,以及用于检测甲苯中的 O2•--0,0.032 mmol Co 的 Co-N-C 溶液, 1.25 mmol NaOH、1 mmol 清除剂、2.5 mL 甲苯、0.6 MPa O2 和 100 °C。

    Figure 3

    Figure 3. Mass spectra of the OH trapping experiments by DMPO in isotope-labeled H2O and O2. Reaction conditions: 1.25 mmol NaOH, 0.032 mmol Co in Co–N–C, 5 mL of H2O, 60 °C, 40 mL/min O2, and 30 min.

    Figure 4

    Figure 4. Time curve of the intensity of DMPO-OH adducts in the ESR spectra of the Co–N–C system with HMF addition (a, above), the concentrations of HMF and FDCA during the HMF oxidation process in the Co–N–C system (a, below); the high-performance liquid chromatography (HPLC) traces for HMF oxidation in the Co–N–C system (b); the time curve of the intensity of DMPO-OH adducts in the ESR spectra of the Co–N–C system without HMF addition (c); the time curve of the intensity of DMPO-OH adducts in the ESR spectra of the H2O2 + OH system with and without HMF addition (d). Reaction conditions: in the Co–N–C system, 1.25 mmol NaOH, 0.032 mmol Co in Co–N–C, 5 mL of H2O, 40 mL/min O2, and 60 °C and in the H2O2 + OH system, 1.25 mmol NaOH, 1 mL of 30 wt % H2O2, 4 mL of H2O, and 60 °C. In these two systems, 0.05 mmol HMF was added in the 11th min, 25th min, 39th min, and 53rd min.

    Figure 5

    Figure 5. Reaction time profile of HMF oxidation over Co–N–C (a) and inhibited by IPA (b) and the reaction time profile of HMFCA oxidation over Co–N–C (c) and inhibited by IPA (d). Reaction conditions: 0.5 mmol HMF or HMFCA, 0.032 mmol Co in Co–N–C, 1.25 mmol NaOH, 5 mL of H2O, 1.0 mmol IPA, 40 mL/min O2, and 60 °C.

    Figure 6

    Figure 6. UV–vis spectra of the HMF oxidation process in the Co–N–C system (a) and H2O2 + OH system (b); evolution of FDCA concentration in HMF oxidation in the Co–N–C system with O2 bubbling (A), with H2O2 addition 10 times (B), with H2O2 addition 1 time (C), and in the H2O2 + OH system (D) (c); ESR spectra of DMPO-OH adducts at the reaction time of 1 min in the Co–N–C system with O2 bubbling, with H2O2 addition 10 times and 1 time and in the H2O2 + OH system (d). Reaction conditions: 0.5 mmol HMF, 0.032 mmol Co in Co–N–C, 1.25 mmol NaOH, 5 mL of H2O, 60 °C, and 40 mL/min O2 (A); 0.1 mL of 30 wt % H2O2 was added 10 times at equal intervals (B); 1 mL of 30 wt % H2O2 was added 1 time (C); 1 mL of 30 wt % H2O2 was added 1 time without Co–N–C addition (D). Reaction conditions of UV–vis monitoring in (a, b) were consistent with those of (A) and (D) in (c), respectively.

    Figure 7

    Figure 7. ESR spectra of DMPO-OH adducts at the reaction time of 30 min in the H2O2 + OH system at a pH value of 12.79 and Co–N–C system with different pH values (a); the intensity of the DMPO-OH adducts in the ESR spectra (b); Arrhenius plots of HMF oxidation with and without OH in the Co–N–C system (c); the product yield in HMF oxidation in the Co–N–C system with different pH values and in the H2O2 + OH system at a pH value of 12.79 (d); and the schematic illustration of the HMF oxidation with OH formation in the Co–N–C system (e). Reaction conditions: 0.5 mmol HMF, 0.032 mmol Co in Co–N–C, 5 mL of H2O, 60 °C, 40 mL/min O2, 5 h, at different pH values; a0.5 mmol HMF, 1 mL of 30 wt % H2O2, 4 mL of H2O, 60 °C, 5 h, and pH = 12.79.

  • References


    This article references 49 other publications.

    1. 1
      Pereira, M. M.; Dias, L. D.; Calvete, M. J. F. Metalloporphyrins: Bioinspired Oxidation Catalysts. ACS Catal. 2018, 8, 1078410808,  DOI: 10.1021/acscatal.8b01871
    2. 2
      Piera, J.; Bäckvall, J.-E. Catalytic Oxidation of Organic Substrates by Molecular Oxygen and Hydrogen Peroxide by Multistep Electron Transfer─a Biomimetic Approach. Angew. Chem., Int. Ed. 2008, 47, 35063523,  DOI: 10.1002/anie.200700604
    3. 3
      Li, Q.; Li, F.-t. Recent Advances in Molecular Oxygen Activation via Photocatalysis and its Application in Oxidation Reactions. Chem. Eng. J. 2021, 421, 129915  DOI: 10.1016/j.cej.2021.129915
    4. 4
      Metz, M.; Solomon, E. I. Dioxygen Binding to Deoxyhemocyanin: Electronic Structure and Mechanism of the Spin-Forbidden Two-Electron Reduction of O2. J. Am. Chem. Soc. 2001, 123, 49384950,  DOI: 10.1021/ja004166b
    5. 5
      Gligorovski, S.; Strekowski, R.; Barbati, S.; Vione, D. Environmental Implications of Hydroxyl Radicals (OH). Chem. Rev. 2015, 115, 1305113092,  DOI: 10.1021/cr500310b
    6. 6
      Hou, J.-T.; Zhang, M.; Liu, Y.; Ma, X.; Duan, R.; Cao, X.; Yuan, F.; Liao, Y.-X.; Wang, S.; Ren, W. X. Fluorescent Detectors for Hydroxyl Radical and their Applications in Bioimaging: A Review. Coord. Chem. Rev. 2020, 421, 213457  DOI: 10.1016/j.ccr.2020.213457
    7. 7
      Wang, B.; Su, H.; Zhang, B. Hydrodynamic Cavitation as a Promising Route for Wastewater Treatment–A Review. Chem. Eng. J. 2021, 412, 128685  DOI: 10.1016/j.cej.2021.128685
    8. 8
      Cheng, C.; Ren, W.; Miao, F.; Chen, X.; Chen, X.; Zhang, H. Generation of FeIV═O and its Contribution to Fenton-Like Reactions on a Single-Atom Iron-N-C Catalyst. Angew. Chem., Int. Ed. 2023, 62, e202218510  DOI: 10.1002/ange.202218510
    9. 9
      Chen, Z.; Zhou, H.; Kong, F.; Wang, M. Piezocatalytic Oxidation of 5-Hydroxymethylfurfural to 5-Formyl-2-Furancarboxylic Acid over Pt Decorated Hydroxyapatite. Appl. Catal., B 2022, 309, 121281  DOI: 10.1016/j.apcatb.2022.121281
    10. 10
      Liu, Y.; Wang, M.; Zhang, B.; Yan, D.; Xiang, X. Mediating the Oxidizing Capability of Surface-Bound Hydroxyl Radicals Produced by Photoelectrochemical Water Oxidation to Convert Glycerol into Dihydroxyacetone. ACS Catal. 2022, 12, 69466957,  DOI: 10.1021/acscatal.2c01319
    11. 11
      Zope, B. N.; Hibbitts, D. D.; Neurock, M.; Davis, R. J. Reactivity of the Gold/Water Interface During Selective Oxidation Catalysis. Science 2010, 330, 7478,  DOI: 10.1126/science.1195055
    12. 12
      Kung, M. C.; Davis, R. J.; Kung, H. H. Understanding Au-Catalyzed Low-Temperature CO Oxidation. J. Phys. Chem. C 2007, 111, 1176711775,  DOI: 10.1021/jp072102i
    13. 13
      Takanabe, K.; Iglesia, E. Rate and Selectivity Enhancements Mediated by OH Radicals in the Oxidative Coupling of Methane Catalyzed by Mn/Na2WO4/SiO2. Angew. Chem., Int. Ed. 2008, 47, 76897693,  DOI: 10.1002/anie.200802608
    14. 14
      Jiang, D.; Ni, D.; Rosenkrans, Z. T.; Huang, P.; Yan, X.; Cai, W. Nanozyme: New Horizons for Responsive Biomedical Applications. Chem. Soc. Rev. 2019, 48, 36833704,  DOI: 10.1039/C8CS00718G
    15. 15
      Wu, J.; Wang, X.; Wang, Q.; Lou, Z.; Li, S.; Zhu, Y.; Qin, L.; Wei, H. Nanomaterials with Enzyme-Like Characteristics (Nanozymes): Next-Generation Artificial Enzymes (II). Chem. Soc. Rev. 2019, 48, 10041076,  DOI: 10.1039/C8CS00457A
    16. 16
      Xu, B.; Wang, H.; Wang, W.; Gao, L.; Li, S.; Pan, X.; Wang, H.; Yang, H.; Meng, X.; Wu, Q.; Zheng, L.; Chen, S.; Shi, X.; Fan, K.; Yan, X.; Liu, H. A Single-Atom Nanozyme for Wound Disinfection Applications. Angew. Chem., Int. Ed. 2019, 58, 49114916,  DOI: 10.1002/anie.201813994
    17. 17
      Zhu, S.; Li, Z.; Zhang, F.; Liu, F.; Ni, P.; Chen, C.; Jiang, Y.; Lu, Y. Single-Atom Cobalt Catalysts as Highly Efficient Oxidase Mimics for Time-Based Visualization Monitoring the TAC of Skin Care Products. Chem. Eng. J. 2023, 456, 141053  DOI: 10.1016/j.cej.2022.141053
    18. 18
      Ye, Y.; Xiao, L.; Bin, H.; Zhang, Q.; Nie, T.; Yang, X.; Wu, D.; Cheng, H.; Li, P.; Wang, Q. Oxygen-Tuned Nanozyme Polymerization for the Preparation of Hydrogels with Printable and Antibacterial Properties. J. Mater. Chem. B 2017, 5, 15181524,  DOI: 10.1039/C6TB03317B
    19. 19
      Mondelli, C.; Gözaydın, G.; Yan, N.; Pérez-Ramírez, J. Biomass Valorisation over Metal-Based Solid Catalysts from Nanoparticles to Single Atoms. Chem. Soc. Rev. 2020, 49, 37643782,  DOI: 10.1039/D0CS00130A
    20. 20
      Liu, S.-S.; Sun, K.-Q.; Xu, B.-Q. Specific Selectivity of Au-Catalyzed Oxidation of Glycerol and Other C3-Polyols in Water without the Presence of a Base. ACS Catal. 2014, 4, 22262230,  DOI: 10.1021/cs5005568
    21. 21
      Zhang, J.; Nagamatsu, S.; Du, J.; Tong, C.; Fang, H.; Deng, D.; Liu, X.; Asakura, K.; Yuan, Y. A Study of FeNx/C Catalysts for the Selective Oxidation of Unsaturated Alcohols by Molecular Oxygen. J. Catal. 2018, 367, 1626,  DOI: 10.1016/j.jcat.2018.08.004
    22. 22
      Tu, C.; Nie, X.; Chen, J. G. Insight into Acetic Acid Synthesis from the Reaction of CH4 and CO2. ACS Catal. 2021, 11, 33843401,  DOI: 10.1021/acscatal.0c05492
    23. 23
      Thomas, J. M.; Raja, R.; Sankar, G.; Bell, R. G. Molecular Sieve Catalysts for the Regioselective and Shape-Selective Oxyfunctionalization of Alkanes in Air. Acc. Chem. Res. 2001, 34, 191200,  DOI: 10.1021/ar970020e
    24. 24
      Wang, Q.; Hou, W.; Li, S.; Xie, J.; Li, J.; Zhou, Y.; Wang, J. Hydrophilic Mesoporous Poly(ionic liquid)-Supported Au–Pd Alloy Nanoparticles towards Aerobic Oxidation of 5-Hydroxymethylfurfural to 2,5-Furandicarboxylic Acid under Mild Conditions. Green Chem. 2017, 19, 38203830,  DOI: 10.1039/C7GC01116D
    25. 25
      Liu, X.; Chen, L.; Xu, H.; Jiang, S.; Zhou, Y.; Wang, J. Straightforward Synthesis of Beta Zeolite Encapsulated Pt Nanoparticles for the Transformation of 5-Hydroxymethyl Furfural into 2,5-Furandicarboxylic Acid. Chin. J. Catal. 2021, 42, 9941003,  DOI: 10.1016/S1872-2067(20)63720-2
    26. 26
      Zhang, M.; Ma, H.; Liu, X.; Zhang, S.; Luo, Y.; Gao, J.; Xu, J. Control in Local Coordination Environment Boosting Activating Molecular Oxygen with an Atomically Dispersed Binary Mn–Co Catalyst. ACS Appl. Mater. Interfaces 2022, 14, 1853918549,  DOI: 10.1021/acsami.2c01858
    27. 27
      Liu, X.; Luo, Y.; Ma, H.; Zhang, S.; Che, P.; Zhang, M.; Gao, J.; Xu, J. Hydrogen-Binding-Initiated Activation of O–H Bonds on a Nitrogen-Doped Surface for the Catalytic Oxidation of Biomass Hydroxyl Compounds. Angew. Chem., Int. Ed. 2021, 133, 1825118258,  DOI: 10.1002/ange.202103604
    28. 28
      Sun, Y.; Ma, H.; Jia, X.; Ma, J.; Luo, Y.; Gao, J.; Xu, J. A High-Performance Base-Metal Approach for the Oxidative Esterification of 5-Hydroxymethylfurfural. ChemCatChem 2016, 8, 29072911,  DOI: 10.1002/cctc.201600484
    29. 29
      Sun, Y.; Ma, H.; Luo, Y.; Zhang, S.; Gao, J.; Xu, J. Activation of Molecular Oxygen Using Durable Cobalt Encapsulated with Nitrogen-Doped Graphitic Carbon Shells for Aerobic Oxidation of Lignin-Derived Alcohols. Chem. - Eur. J. 2018, 24, 46534661,  DOI: 10.1002/chem.201705824
    30. 30
      Wang, T.; Ma, H.; Liu, X.; Luo, Y.; Zhang, S.; Sun, Y.; Wang, X.; Gao, J.; Xu, J. Ultrahigh-Content Nitrogen-Doped Carbon Encapsulating Cobalt NPs as Catalyst for Oxidative Esterification of Furfural. Chem. - Asian J. 2019, 14, 15151522,  DOI: 10.1002/asia.201900099
    31. 31
      Chen, C.; Wang, L.; Zhu, B.; Zhou, Z.; El-Hout, S. I.; Yang, J.; Zhang, J. 2, 5-Furandicarboxylic Acid Production via Catalytic Oxidation of 5-Hydroxymethylfurfural: Catalysts, Processes and Reaction Mechanism. J. Energy Chem. 2021, 54, 528554,  DOI: 10.1016/j.jechem.2020.05.068
    32. 32
      Yang, G.; Ma, Y.; Xu, J. Biomimetic Catalytic System Driven by Electron Transfer for Selective Oxygenation of Hydrocarbon. J. Am. Chem. Soc. 2004, 126, 1054210543,  DOI: 10.1021/ja047297b
    33. 33
      Dijkman, W. P.; Groothuis, D. E.; Fraaije, M. W. Enzyme-Catalyzed Oxidation of 5-Hydroxymethylfurfural to Furan-2, 5-Dicarboxylic Acid. Angew. Chem., Int. Ed. 2014, 53, 65156518,  DOI: 10.1002/anie.201402904
    34. 34
      Jia, H. Y.; Zong, M. H.; Zheng, G. W.; Li, N. One-Pot Enzyme Cascade for Controlled Synthesis of Furancarboxylic Acids from 5-Hydroxymethylfurfural by H2O2 Internal Recycling. ChemSusChem 2019, 12, 47644768,  DOI: 10.1002/cssc.201902199
    35. 35
      Liu, W.; Zhang, L.; Yan, W.; Liu, X.; Yang, X.; Miao, S.; Wang, W.; Wang, A.; Zhang, T. Single-Atom Dispersed Co–N–C Catalyst: Structure Identification and Performance for Hydrogenative Coupling of Nitroarenes. Chem. Sci. 2016, 7, 57585764,  DOI: 10.1039/C6SC02105K
    36. 36
      Jagadeesh, R. V.; Junge, H.; Pohl, M.-M.; Radnik, Jr.; Brückner, A.; Beller, M. Selective Oxidation of Alcohols to Esters Using Heterogeneous Co3O4–N@C Catalysts under Mild Conditions. J. Am. Chem. Soc. 2013, 135, 1077610782,  DOI: 10.1021/ja403615c
    37. 37
      Madden, K. P.; Taniguchi, H. The Role of the DMPO-Hydrated Electron Spin Adduct in DMPO-OH Spin Trapping. Free Radical Biol. Med. 2001, 30, 13741380,  DOI: 10.1016/S0891-5849(01)00540-8
    38. 38
      Yamazaki, I.; Piette, L. H. EPR Spin-Trapping Study on the Oxidizing Species Formed in the Reaction of the Ferrous Ion with Hydrogen Peroxide. J. Am. Chem. Soc. 1991, 113, 75887593,  DOI: 10.1021/ja00020a021
    39. 39
      Qi, Y.; Li, J.; Zhang, Y.; Cao, Q.; Si, Y.; Wu, Z.; Akram, M.; Xu, X. Novel Lignin-Based Single Atom Catalysts as Peroxymonosulfate Activator for Pollutants Degradation: Role of Single Cobalt and Electron Transfer Pathway. Appl. Catal., B 2021, 286, 119910  DOI: 10.1016/j.apcatb.2021.119910
    40. 40
      Tran, P.; Kopel, J.; Risitic, B.; Marsh, H.; Fralick, J.; Reid, T. Antimicrobial Seleno-Organic Coatings and Compounds Acting Primarily on the Plasma Membrane: A Review. Adv. Redox Res. 2022, 4, 100031  DOI: 10.1016/j.arres.2022.100031
    41. 41
      Shiibashi, T.; Iida, T. J. D. NADPH and NADH Serve as Electron Donor for the Superoxide-Generating Enzyme in Tilapia (Oreochromis Niloticus) Neutrophils. Dev. Comp. Immunol. 2001, 25, 461465,  DOI: 10.1016/S0145-305X(01)00005-2
    42. 42
      Gleason, J. E.; Galaleldeen, A.; Peterson, R. L.; Taylor, A. B.; Holloway, S. P.; Waninger-Saroni, J.; Cormack, B. P.; Cabelli, D. E.; Hart, P. J.; Culotta, V. C. Candida Albicans SOD5 Represents the Prototype of an Unprecedented Class of Cu-only Superoxide Dismutases Required for Pathogen Defense. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, 58665871,  DOI: 10.1073/pnas.1400137111
    43. 43
      Jiao, L.; Wu, J.; Zhong, H.; Zhang, Y.; Xu, W.; Wu, Y.; Chen, Y.; Yan, H.; Zhang, Q.; Gu, W. Densely Isolated FeN4 Sites for Peroxidase Mimicking. ACS Catal. 2020, 10, 64226429,  DOI: 10.1021/acscatal.0c01647
    44. 44
      Urbański, N. K.; Beręsewicz, A. Generation of OH Initiated by Interaction of Fe2+ and Cu+ with Dioxygen; Comparison with the Fenton Chemistry. Acta Biochim. Pol. 2000, 47, 951962,  DOI: 10.18388/abp.2000_3950
    45. 45
      Chen, C.; Li, X.; Wang, L.; Liang, T.; Wang, L.; Zhang, Y.; Zhang, J. Highly Porous Nitrogen- and Phosphorus-Codoped Graphene: An Outstanding Support for Pd Catalysts to Oxidize 5-Hydroxymethylfurfural into 2, 5-Furandicarboxylic Acid. ACS Sustainable Chem. Eng. 2017, 5, 1130011306,  DOI: 10.1021/acssuschemeng.7b02049
    46. 46
      Senthamarai, T.; Chandrashekhar, V. G.; Rockstroh, N.; Rabeah, J.; Bartling, S.; Jagadeesh, R. V.; Beller, M. A “Universal” Catalyst for Aerobic Oxidations to Synthesize (Hetero)aromatic Aldehydes, Ketones, Esters, Acids, Nitriles, and Amides. Chem 2022, 8, 508531,  DOI: 10.1016/j.chempr.2021.12.001
    47. 47
      Xu, D.; Li, J.; Li, B.; Zhao, H.; Zhu, H.; Kou, J.; Zhang, F.; Dong, Z.; Ma, J. Selective Oxidation of Alcohols to High Value-Added Carbonyl Compounds using Air over Co-Co3O4@NC Catalysts. Chem. Eng. J. 2022, 434, 134545  DOI: 10.1016/j.cej.2022.134545
    48. 48
      Mate, V.; Jha, A.; Joshi, U.; Patil, K.; Shirai, M.; Rode, C. Effect of Preparation Parameters on Characterization and Activity of Co3O4 Catalyst in Liquid Phase Oxidation of Lignin Model Substrates. Appl. Catal., A 2014, 487, 130138,  DOI: 10.1016/j.apcata.2014.08.023
    49. 49
      Zhou, H.; Xu, H.; Wang, X.; Liu, Y. Convergent Production of 2,5-Furandicarboxylic Acid from Biomass and CO2. Green Chem. 2019, 21, 29232927,  DOI: 10.1039/C9GC00869A
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acscatal.3c02736.

    • Details of experimental procedures and data processing, comparison of Co–N–C and reported non-noble catalysts for HMF oxidation, additional characterization data of Co-based catalysts, ESR spectra of DMPO-radical adducts, UV–vis spectra of KI oxidation, GC spectra of Ph3P oxidation, mass spectrum of CO2, reaction time profiles for HMF and HMFCA oxidations, recycling experiments of Co–N–C, kinetics analysis for benzyl alcohol oxidation, and supplementary catalytic evaluation (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

I found these useful today (select all that apply):