Applied Catalysis B: Environment and Energy
《应用催化 B 辑:环境与能源》
第 366 卷,2025 年 6 月 5 日,125075 页
Electrically driven gaseous ammonia decomposition on Co-based SiC composite catalysts for low-temperature H2 production
钴基碳化硅复合催化剂电驱动气态氨分解用于低温制氢
Highlights 研究亮点
- •An electrically driven gaseous ammonia decomposition strategy for low-temperature hydrogen production was proposed.
提出了一种电驱动气态氨分解策略用于低温制氢。 - •The ammonia decomposition reaction was carried out in the absence of external heating using Co-based SiC composite catalysts.
在无外部加热条件下,采用钴基碳化硅复合催化剂实现了氨分解反应。 - •An NH3 conversion of 73 % was achieved at ca. 200℃ over SiC-supported catalyst (Co/SiC-Al2O3).
在约 200℃条件下,碳化硅负载催化剂(Co/SiC-Al 2 O 3 )实现了 73%的氨转化率。 - •Hydrogen species migration is important for NHx dehydrogenation and N-N recombination.
氢物种迁移对于 NH x 的脱氢和 N-N 重组至关重要。
Abstract 摘要
氨(NH₃)因其易于储存的特性被视为一种极具前景的无碳氢(H₂)载体。然而,由于反应动力学缓慢,在 300℃以下实现 NH₃高效分解制氢难以实现,这阻碍了其多场景实际应用。本文提出了一种采用钴基碳化硅复合催化剂的电驱动 NH₃分解新策略。该电驱动策略无需外部加热即可在气态 NH₃中实施,焦耳热效应使系统维持在低温状态。研究对比了两种主要碳化硅复合催化剂(包括碳化硅负载型催化剂 Co/SiC-Al₂O₃和碳化硅介导型催化剂 Co/Al₂O₃-SiC)的催化活性与反应机制差异。Co/SiC-Al₂O₃与电场具有更显著的协同效应,在约 200℃条件下可实现 73%的 NH₃转化率,性能优于传统热催化 NH₃分解的尖端催化剂。系统表征表明,电场激发的电荷迁移对反应中间体的演变具有重要作用。 本研究表明,利用地球储量丰富的过渡金属在低温条件下从氨气中高效制氢具有巨大潜力。
Graphical Abstract 图文摘要
Keywords 关键词
氢气 | 氨分解 | 电场 | 钴基催化剂 | 碳化硅
1. Introduction 1. 引言
在过去两百年间,碳基化石燃料的使用向大气排放了大量二氧化碳,其温室效应严重影响了人类社会发展。为此,研究者们正寻求用零碳燃料替代化石燃料[1]。通过可再生能源制备的绿氢(H 2 )被视为零碳排放的终极能源[2][3]。然而传统高压储氢存在质量储氢密度低、能耗高及泄漏风险等技术瓶颈[4]。作为替代方案,H 2 可通过材料储存在高温条件下按需释放。其中极具前景的储氢材料氨(NH 3 )具有 17.6 wt%的高重量储氢密度,且能在温和条件(20℃、0.8 MPa)下液化储存[5]。此外,利用风能/太阳能驱动 NH 3 分解制取的 H 2 全程零碳排,未反应的 NH 3 可通过吸附剂从 H 2 流中高效脱除[6],完全满足燃料电池汽车(FCVs)的用氢需求[7]。这些特性使 NH 3 成为理想的储氢介质。
氨分解反应需要高温(>400℃),因为该反应为吸热过程(NH₃ → 1/2N₂ + 3/2H₂,ΔH = +46 kJ mol⁻¹)。高温环境导致现场制氢能效降低、应用场景受限,同时会缩短催化剂使用寿命。学界普遍认为钌(Ru)是氨分解活性最高的金属催化剂[8]。研究表明,钾(K)促进的 Ru/CNTs 催化剂在低至 450℃时即可接近热力学平衡转化率[9]。然而钌的稀缺性和高昂成本促使研究者致力于开发更廉价的地球丰产元素催化剂。近期 Hassina 等人[10]报道的 CoNi 双金属催化剂在 450℃下可实现氨的完全转化,展现了钴基催化剂的潜力。但温度进一步降至 400℃时,其氨转化率仅约 40%。若通过复杂催化剂设计进一步提升低温性能,则需面临成本不可控的难题。 为此,研究人员尝试引入外场以增强催化活性[11][12],例如等离子体催化[13][14]和电催化[15]。其中,在适当直流电场辅助下的 NH3 分解在低温条件下展现出显著提升催化活性的潜力[16][17]。Ofuchi Y 等学者通过在 Ru/CeO2 催化剂上施加电场,实现了显著低温条件下的高 NH3 转化率,并发现表面质子传导改变了速率决定步骤[18]。这种电场辅助的 NH3 分解是一种非法拉第过程,其电能输入需求低于 NH3 电解[15]和等离子体催化[13]。
常见的氧化物载体如 Al₂O₃、SiO₂、ZrO₂和 MgO 因具有高电阻会严重阻碍电流传输(表 S1),而其他氧化物半导体(如先前研究中常用于传输电场的 CeO₂[12,16])则存在导电不稳定性和无外部加热时严重放电的问题(图 S1)。额外的外部加热限制了反应温度的进一步降低,使得电场与催化剂的高效催化协同效应无法充分发挥。SiC 作为电子工业中广泛应用的半导体材料,凭借其宽带隙和高导热性[19]展现出优异的电场传输能力,因此可用于增强催化剂在低温下的电场传输性能。此外,由于 SiC 具有出色的耐热性[20,21],也常被用于传统热催化领域。本研究提出采用钴基 SiC 复合催化剂实现无需外部加热的电驱动 NH₃分解策略。 由于稳定的电激活和高效的热激活催化过程,在约 200℃下实现了高效的 NH3 分解。比较了两种主要碳化硅复合催化剂(包括碳化硅负载型催化剂 Co/SiC-Al2O3 和碳化硅介导型催化剂 Co/Al2O3-SiC)在催化活性与反应机理上的差异。
2. Experimental 2. 实验部分
2.1. Catalyst preparation
2.1. 催化剂制备
采用沉积-沉淀法(DP)[22]与高压研磨两步简易工艺制备钴基碳化硅复合电场催化剂。具体步骤为:将 2.0 克γ-Al₂O₃(分析纯级,阿拉丁试剂)悬浮于 100 毫升去离子水中,同时将 1.975 克 Co(NO₃)₂·6H₂O(金属基纯度 99.99%,麦克林试剂)溶于 50 毫升去离子水。随后将钴前驱体溶液以 15.7 wt%的标称载钴量逐滴加入悬浮液,剧烈搅拌 3 小时后通过氨水逐滴调节 pH 至 9.5。所得悬浮液室温下持续搅拌过夜,经抽滤分离沉淀物,去离子水洗涤后于 70℃空气烘箱中干燥 8 小时。研磨后的样品以 5℃/分钟升温速率在 600℃空气氛围中煅烧 4 小时,最终产物标记为 Co/Al₂O₃。 采用与 Co/Al₂O₃相同的制备方法,以 SiC(40 纳米,麦克林)和 Co(NO₃)₂·6H₂O 为原料合成 Co/SiC 催化剂。不同之处在于,由于 SiC 在高温下易被氧化导致导电性下降,煅烧过程需在氩气保护下以 5℃/分钟的升温速率升至 600℃并保持 4 小时。为进行对比,按照 Co/SiC 相同方法,将 Co 负载于 Al₂O₃与 SiC(质量比 1:1)的混合物上,制得 Co/(Al₂O₃-SiC)复合催化剂。
高电阻的 Co/Al₂O₃严重阻碍电流传输,而导电性过高的 Co/SiC 则限制了电场强度的提升(表 S1),并表现出较低的活性(图 S2)。为获得具有适中导电性的电场催化剂,通过高压研磨引入催化剂第二组分来控制电阻。两组分比例的确定方法详见表 S1。对于 SiC 介导的催化剂,简要步骤如下:将充分研磨的 Co/Al₂O₃与 SiC(质量比=1:1)通过涡旋振荡器混合 30 分钟,随后在不锈钢研钵中强力研磨 30 分钟。所得样品在 15 MPa 压力下压制 20 分钟,经破碎筛分获得 178-425μm 颗粒。制得的 SiC 介导催化剂标记为 Co/Al₂O₃-SiC。同理,采用相同方法使用充分研磨的 Co/SiC 与 Al₂O₃(质量比=2:1)制备了 SiC 负载型催化剂(Co/SiC-Al₂O₃)。由于单独的 Al₂O₃和 SiC 对 NH₃分解的催化活性可忽略不计(图 S3),本研究中进行的活性测试与表征均基于相同质量的 Co/Al 2 O 3 、Co/SiC 及 Co/(Al 2 O 3 -SiC)催化剂。
2.2. Catalyst characterization
2.2. 催化剂表征
采用 Avio 500 型电感耦合等离子体发射光谱仪(ICP-OES)测定各催化剂钴含量。催化剂 X 射线衍射(XRD)测试在配备 Cu Kα辐射源(λ=1.5406 Å)的 D8 ADVANCE DaVinci 型 X 射线衍射仪上进行。拉曼光谱通过配备 532 nm 激发激光源的 Renishaw inVia Qontor 共聚焦拉曼光谱仪采集。采用 Quantachrome NOVA 2000e 全自动气体吸附仪在液氮温度下进行 N₂吸附-脱附测试,测定平均孔径、孔容及 BET 比表面积。X 射线光电子能谱(XPS)测试使用配备 Al Kα辐射源的 Thermo Fisher ESCALAB Xi+多功能光电子能谱仪完成,并以 284.8 eV 的 C1s 峰位进行荷电校正。透射电子显微镜(TEM)及高分辨透射电子显微镜(HRTEM)分析在加速电压 200 kV 的 JEOL JEM-2100 型分析型透射电镜上完成。
采用配备 TCD 检测器的 BUILDER PCA-1200 仪器进行氢气程序升温还原(H2-TPR)测试。首先将 100 mg 样品在氩气氛围下 200℃脱水 30 分钟,冷却至 30℃后切换为 10 vol% H2/Ar 混合气(流速 30 mL/min),并以 10℃/min 的升温速率加热至 800℃。
采用相同仪器进行一氧化碳程序升温脱附(CO-TPD)测试。将 100 mg 样品置于反应器中,在 10 vol% H₂/Ar(30 mL/min)气氛下 600℃还原 1 小时,随后用 He(30 mL/min)吹扫去除残留氢气。接着在 30℃下通入纯 CO(30 mL/min)进行 1 小时预吸附,再用 He(30 mL/min)吹扫 30 分钟去除物理吸附的 CO。最后以 10℃/min 升温速率在 He 气流(30 mL/min)中进行 TPD 测试。电场处理样品制备方法为:实验前对催化剂施加 8 W 功率通电处理 3 小时。
氨气程序升温脱附(NH₃-TPD)测试流程与 CO-TPD 相同,但预吸附气体改为 10 vol% NH₃/Ar 混合气。电场处理样品制备方法为:实验前对催化剂施加 8 W 功率通电处理 3 小时。
氢气程序升温脱附(H₂-TPD)测试流程同样参照 CO-TPD 方法,但预吸附气体采用 10 vol% H₂/Ar 混合气,且程序升温脱附阶段载气更换为 Ar 气。
氢电驱动脱附(H-EDD)实验采用自主设计的配备质谱仪(Hiden Analytical,HPR-20 EGA)的电场反应器进行。预处理步骤与 H-TPR 相同。待系统冷却至 30℃后,向反应器通入 1 vol% H2/Ar 混合气。系统稳定 20 分钟后,对催化剂施加电场,并通过质谱监测脱附气体。数分钟后关闭电场。分别在 3、5、7 和 9 W 四种不同电输入功率下进行了四次 H-EDD 测试。
氨表面反应实验采用自主设计的程序升温反应器配合质谱仪完成。预活化流程与 CO-TPD 相同。对于热表面反应,催化剂温度维持在 300℃;而电表面反应则对催化剂施加 8 W 电场并持续至测试结束。随后向反应器通入 2 vol% NH3/Ar 混合气(流速 100 mL/min),利用质谱仪监测 NH3 表面反应过程。
对使用过的催化剂进行产物脱附时,采用了与 NH 3 表面反应相同的仪器。首先,在 400℃纯 NH 3 气氛中对催化剂进行 2 小时测试。冷却至 30℃后,切换为氩气(30 mL/min)以去除残留 NH 3 。随后进行程序升温脱附(TPD)时,以 10℃/min 速率升温至 600℃;而电驱动脱附实验则是在短时间内对催化剂施加 8W 电场。所有脱附产物均通过质谱仪进行实时监测。
采用原位漫反射红外傅里叶变换光谱(DRIFT)在傅里叶变换红外光谱仪(Nicolet 6700)上进行实验,以研究反应过程中催化剂表面物种的演变。光谱采集参数设置为扫描次数 32 次,分辨率 4 cm −1 。实验前,每个样品均在 20% H 2 /Ar(50 mL/min)气氛中于 400℃下还原 1 小时。测量温度下的背景谱图在纯 Ar 气流中采集,随后将气氛切换为 5% NH 3 /Ar(50 mL/min)。
2.3. Catalytic activity tests
2.3. 催化活性测试
如图 S4 所示,电场反应器由柱状石英管反应器(外径 12 毫米,内径 7 毫米)改造而成。将 100 毫克催化剂(Co/载体)填充于反应池内两块泡沫铜(作为集流体)制备的滤网之间。两根独立铜管作为电极贯穿反应器两端,连接至电源(Teslaman TD2200)。铜(Cu)通常被认为是氨分解催化活性较差的金属,因此泡沫铜和电极对催化反应的影响可忽略不计。除非另有说明,反应器的热扩散条件为自然对流且无外部加热。焦耳热效应导致一定温升,通过红外热像仪(HIKMICRO HM-TPK20–3AQF/W)进行原位监测。水冷测试采用双层石英反应器,5℃循环冷却水流经反应器外层。作为对比实验,传统热催化氨分解采用程序升温管式炉加热。 产物通过配备热导检测器(TCD)的气相色谱仪(GC-9860)进行监测。
活性测试通过向反应器中通入纯气态 NH₃(99.999%)并在常压下进行评价。对于常规热 NH₃分解,催化剂需先在 600℃、70 vol% H₂/Ar 混合气中预还原 1 小时。随后测试温度从 400℃逐步升至 650℃,升温间隔为 50℃。对于电驱动 NH₃分解,催化剂需先在 8W 电功率的 NH₃气氛中预处理 15 分钟,随后将功率调节至 5-11W 范围内进行活性测试。NH₃转化率根据公式(1)计算得出,H₂产率(mmol/g/s)则通过公式(2)求得。(1)(2)Where and are the NH3 flow rate for inlet and outlet gas, respectively. mcat. is the mass of Co/support.
其中 和 分别代表进出口气体中 NH 3 的流量,m cat. 为钴基载体催化剂的质量。
2.4. Kinetic studies 2.4 动力学研究
通过拟合阿伦尼乌斯曲线(如公式(3)所示)获得表观活化能。在通电条件下,维持电输入功率为 4.6W 的同时施加外部加热以构建温度梯度。通过调节流量和反应温度,将 NH 3 转化率控制在 25%以下以排除传质和传热限制。(3)Where k and k0 are reaction constant (mol/gcat./s) and frequency factor (mol/gcat./s), respectively; R is universal gas constant (kJ mol−1 K−1); T is Kelvin temperature (K).
式中 k 和 k 0 分别为反应常数(mol/g cat. /s)和频率因子(mol/g cat. /s);R 为通用气体常数(kJ mol −1 K −1 );T 为开尔文温度(K)。
通过改变混合气体(NH₃/H₂/N₂/Ar)浓度测定反应级数。热条件下反应级数在 500℃下测定,电条件下反应级数在 4.6 W 电功率下测定,并通过外部热源将温度控制在约 500℃。结果通过式(4)和式(5)获得。(4)(5)Where is NH3 reaction rate (mol/gcat./s); α, β, γ are the reaction orders of NH3, H2, and N2, respectively; [NH3], [H2] and [N2] are partial pressure. It is widely acknowledged that the reaction order of N2 (γ) is approximately zero and the presence of N2 does not influence the activity [23], [24], which was also found in the electric field (Fig. S5), so the N2 dependence was not involved in the reaction order analysis.
式中 表示 NH₃反应速率(mol/g·cat/s);α、β、γ分别为 NH₃、H₂和 N₂的反应级数;[NH₃]、[H₂]和[N₂]为分压。学界普遍认为 N₂的反应级数(γ)近似为零,且 N₂的存在不影响反应活性[23,24],电场条件下也发现此现象(图 S5),因此反应级数分析未考虑 N₂依赖性。
3. Results and discussion
3. 结果与讨论
3.1. Comparison of NH3 decomposition activity
3.1. NH₃分解活性对比
所得 SiC 载体催化剂(Co/SiC-Al 2 O 3 )与 SiC 介导催化剂(Co/Al 2 O 3 -SiC)的电阻率相当(表 S1),因此可在相同电场强度下研究电场与催化剂的协同效应。电驱动 NH 3 分解反应在无需外部加热与绝热的条件下进行(图 S4 所示),这使得装置具有便携性。然而由于电场中的焦耳热效应,催化剂存在温升现象。如图 1a-h 所示,反应器的红外热成像图表明电场中的温度始终低于约 200℃,远低于传统热催化 NH 3 分解温度(通常>400℃)。NH 3 分解的能耗包含化学能转化(即 NH 3 →1/2N 2 +3/2H 2 ,ΔH=+46 kJ mol −1 )与热耗散两部分。当 NH 3 转化率相同时,化学能转化值为定值,而电场中较低的反应温度意味着更少的热耗散,从而降低总能耗。如图示传统热催化 NH 3 分解过程中... 2a 图中显示,钴基催化剂在 400℃下活化,需升温至约 600℃才能实现 70%的氨气转化率(文献 13),而完全转化氨气(文献 14)则需要超过 650℃。催化活性顺序为:Co/Al₂O₃-SiC > Co/(Al₂O₃-SiC) > Co/SiC-Al₂O₃。其中碳化硅介导的催化剂(Co/Al₂O₃-SiC)表现出最佳活性,这表明直接将钴负载在 Al₂O₃上可获得更优的热活化性能。

Fig. 1. (a) Visible infrared thermogram of the reactor at an electrical input power of 5 W. (b)-(h) Infrared rainbow thermograms of the reactor at an electrical input power of (b) 5 W, (c) 6 W, (d) 7 W, (e) 8 W, (f) 9 W, (g) 10 W, (h) 11 W.
图 1. (a) 反应器在 5W 电输入功率下的可见光-红外热成像图。(b)-(h) 反应器在不同电输入功率下的红外彩虹热成像图:(b) 5W,(c) 6W,(d) 7W,(e) 8W,(f) 9W,(g) 10W,(h) 11W。

Fig. 2. (a) Catalytic activity of conventional thermal NH3 decomposition, WHSV: 12,000 mL gcat.−1 h−1; (b) Catalytic activity of electrically driven NH3 decomposition, WHSV: 12,000 mL gcat.−1 h−1. (c) Comparison of T70 with other reported catalysts: 1: 1.22 %Na-Co3O4 [25], 2: Ru/La0.03Ce0.67 [22], 3: 10Ni/Al1Ce0.05Ox [26], 4: Ru/MgO(111) [23], 5: HEA-Co25Mo45 [27], 6: K-CoNialloy-MgO-CeO2-SrO [10], 7: Cs0.018Co3Mo3N [28], 8: 1.0Ru/CeO2 [29], 9: Ru/Sm2O3 [30]. E represents the electrical condition.
图 2. (a)传统热催化 NH₃分解活性,重时空速:12,000 mL·g⁻¹·h⁻¹;(b)电驱动 NH₃分解催化活性,重时空速:12,000 mL·g⁻¹·h⁻¹。(c)与其他报道催化剂的 T₅₀对比:1: 1.22%Na-Co₃O₄[25],2: Ru/La₂Ce₂[22],3: 10Ni/Al₂Ce₃O₇[26],4: Ru/MgO(111)[23],5: HEA-Co₄Mo₃[27],6: K-CoNi-MgO-CeO₂-SrO[10],7: Cs₂Co₃Mo₂N[28],8: 1.0Ru/CeO₂[29],9: Ru/Sm₂O₃[30]。E 代表电驱动条件。
相比之下,如图 2b 所示,电驱动氨分解在低温条件下仍表现出优异性能。但与常规热催化氨分解不同,Co/SiC-Al₂O₃在电场(7-11 W)中展现出最佳催化活性,而 Co/Al₂O₃-SiC 性能最差,这表明直接负载于 SiC 上的钴能与电场产生更好的协同效应。当电输入功率增至 11 W 时,这些催化剂的活性逐渐趋同,因为反应接近相应温度下的热力学平衡状态,此时 Co/SiC-Al₂O₃实现了 73%的氨转化率。为便于直观对比,图 2c 进一步将其活性与精选的常规热催化氨分解前沿催化剂进行对照。由于低温热力学平衡的限制,氨无法完全转化,因此采用 T₇₀(实现 70%氨转化所需的温度)作为活性比较指标。 本研究基于非贵金属的电驱动策略在性能上超越了包括钌基催化剂在内的已报道最先进催化剂,在常规热 NH3 分解中将反应温度降低了至少 150℃。
3.2. Possible mechanism 3.2. 可能的反应机理
催化剂内部建立的电场将导致载流子迁移和焦耳热效应。有理由认为 Al₂O₃和 SiC 在热活化与载流子增强活化过程中发挥不同作用,因为 SiC 介导与 SiC 负载催化剂在传统热 NH₃分解和电驱动 NH₃分解中的活性存在差异。为探究载流子增强活化的本征活性,本研究构建了恒温(5℃)水循环包裹的自制反应器(图 S6a),以最大限度消除焦耳热效应,使催化剂在电场中保持低温状态(图 S6b)。如图 3a 所示,与传统热 NH₃分解相比,催化活性虽有所下降但仍保持较高水平。该结果表明电场中的催化活性包含载流子增强活性和部分焦耳热活性,前者是中间体活化的主要来源,后者则进一步对催化剂进行热活化。 与热 NH3 分解(>400℃)相比,电场中(<200℃)存在的部分焦耳热活性表明,焦耳加热产生的热活化比传统外部加热对活性位点激活更为高效。这一现象类似于热载流子驱动光催化中的光热效应——由局域表面等离子体非辐射衰变激发的热载流子,其能量可高于直接光激发产生的载流子[31]。另一方面,在水冷条件下 Co/SiC-Al2O3 的活性仍优于 Co/Al2O3-SiC,证明碳化硅载体催化剂能实现更优异的载流子增强活化。值得注意的是,Co/(Al2O3-SiC)在 5-6W 功率下表现出最佳活性(图 2b),说明低电输入功率时 Al2O3 部分取代可改善碳化硅载体催化剂的热活化性能,但会限制高电输入功率下的载流子增强活化效应。

Fig. 3. Comparison of Co/Al2O3-SiC and Co/SiC-Al2O3 in (a) Electrically driven NH3 decomposition under water cooling, the temperature of cooling water was 5℃; (b) Arrhenius plots. T: thermal NH3 decomposition; E: electrically driven NH3 decomposition, the electrical input power was 4.6 W.
3.3. Mechanism characterization
Table 1. Textural properties of Co-based catalysts.
| Catalysts | Co content -(wt%)a | SBET (m2 / gcat.)b | Vpore (cm3 / gcat.)b |
|---|---|---|---|
| SiC | ̸ | 23.34 | 0.083 |
| Al2O3 | ̸ | 123.65 | 0.823 |
| Co/SiC | 16.02 | 44.89 | 0.411 |
| Co/Al2O3 | 14.15 | 108.64 | 0.619 |
| Co/(Al2O3-SiC) | 15.13 | 79.71 | 0.327 |
- a
- Determined by ICP-OES analysis.
- b
- The specific area (SBET), and the total pore volume (Vpore) were measured by N2 physisorption analysis.

Fig. 4. (a) XRD patterns, (b) Raman results of the Co-based catalysts.

Fig. 5. (a) HRTEM images, (b) TEM images and (c-f) Energy-dispersive X-ray spectroscopy elemental mapping of Co/Al2O3-SiC.

Fig. 6. (a) HRTEM images, (b) TEM images and (c-f) Energy-dispersive X-ray spectroscopy elemental mapping of Co/SiC-Al2O3.

Fig. 7. XPS spectra of Co/Al2O3-SiC and Co/SiC-Al2O3 for fresh and the electric field-treated samples (termed as, E) in the electric field at 8 W in Ar: (a) Co 2p; (b) Si 2p.
Si=O and Si-C, respectively [44]. The proportion of O
Si=O in pristine Co/SiC-Al2O3 is higher than that in pristine Co/Al2O3-SiC, possibly because the interaction between SiC and oxygenated species (CoxOy or Al2O3) in Co/SiC-Al2O3 is stronger. Interestingly, after being treated by the electric field in Ar, the O
Si=O ratio of both catalysts increased, especially for Co/SiC-Al2O3 (from 49.3 % to 65.7 %). This result indicates that electric field strengthened the interaction of different catalyst components, which is conducive to current carriers’ transportation. This result is also consistent with the increase in surface oxygen (OA) as illustrated by the XPS of O 1 s (Fig. S9). In conventional thermal NH3 decomposition, it is reported that the enhanced metal-support interaction is beneficial for ammonia decomposition by improving H2/N2 desorption [22], [24]. Since Co/SiC-Al2O3 shows better activity in the electric field, it is believed that the enhanced interaction is one of the mechanism by which the electric field promotes NH3 decomposition [45].
Fig. 8. Comparison for SiC-mediated and SiC-supported catalysts (a) NH3-TPD; (b) H2-TPR of fresh and the electric field-treated samples (identified as, E).

Fig. 9. NH3 surface reaction for Co/SiC-Al2O3 and Co/Al2O3-SiC (a) at 300℃ without the electric field; (b)in the electric field (8 W).

Fig. 10. Product desorption of the used catalysts (a) TPD profiles; (b) Electrical desorption profiles. The fresh catalysts were tested at 400℃ for 2 hours in pure NH3 to obtain the used catalysts.

Fig. 11. Comparison for Co/SiC-Al2O3 and Co/Al2O3-SiC catalysts (a) H2-TPD profiles; (b) CO2-TPD of pristine and the electric field treated samples (termed as, E).

Fig. 12. Time-resolved in situ DRIFTS over Co/Al2O3-SiC at (a) 200℃ and (c) 400℃; Co/SiC-Al2O3 at (b) 200℃ and (d) 400℃.

Fig. 13. H2-EDD at different electrical input power. Top: H2 and H2O signal on Co/Al2O3-SiC; Middle: H2 and H2O signal on Co/SiC-Al2O3; Bottom: Temperature evolution. Abscissa is the relative time.
3.4. Long-term stability and WHSV effect

Fig. 14. (a) Stability test for electrically driven NH3 decomposition on Co/SiC-Al2O3 at 12,000 mL gcat.−1 h−1 in the electric field. Top: activity performance; bottom: corresponding electrical input power; (b) The conversion rate of electrically driven NH3 decomposition at different WHSV on Co/SiC-Al2O3.
4. Conclusion
Declaration of Competing Interest
Acknowledgements
Appendix A. Supplementary material
Supplementary material
References
- [1]A review on clean ammonia as a potential fuel for power generatorsRenew. Sustain. Energy Rev., 103 (2019), pp. 96-108
- [2]Ru-based catalysts for ammonia decomposition: a mini-reviewEnergy Fuels, 35 (2021), pp. 11693-11706
- [3]A systemic review of hydrogen supply chain in energy transitionFront. Energy, 17 (2023), pp. 102-122
- [4]Techno-economic and environmental assessment of hydrogen production through ammonia decompositionAppl. Energy, 358 (2024), Article 122605
- [5]Ammonia decomposition by ruthenium nanoparticles loaded on inorganic electride C12A7:e−Chem. Sci., 4 (2013), pp. 3124-3130
- [6]High performance hydrogen fuel cells with ultralow pt loading carbon nanotube thin film catalystsJ. Phys. Chem. C., 111 (2007), pp. 17901-17904
- [7]Carbon-free energy: a review of ammonia- and hydrazine-based electrochemical fuel cellsEnergy Environ. Sci., 4 (2011), pp. 1255-1260
- [8]Low-temperature ammonia decomposition catalysts for hydrogen generationAppl. Catal. B: Environ., 226 (2018), pp. 162-181
- [9]A mini-review on ammonia decomposition catalysts for on-site generation of hydrogen for fuel cell applicationsAppl. Catal. A: Gen., 277 (2004), pp. 1-9
- [10]Hydrogen generation via ammonia decomposition on highly efficient and stable Ru-free catalysts: approaching complete conversion at 450 °C., Energy Environ. Sci., 15 (2022), pp. 4190-4200
- [11]Catalysis under electric-/magnetic-/electromagnetic-field couplingChem. Soc. Rev. (2025)
- [12]Enhanced activity of catalysts on substrates with surface protonic current in an electrical field – a reviewChem. Commun., 57 (2021), pp. 5737-5749
- [13]NH3 decomposition for H2 generation: effects of cheap metals and supports on plasma–catalyst synergyACS Catal., 5 (2015), pp. 4167-4174
- [14]Nonthermal-PLasma-catalytic Ammonia Synthesis Using Fe2O3/CeO2 mechanically mixed with Al2O3: insights into the promoting effect of plasma discharge enhancement on the role of catalystsACS Sustain. Chem. Eng., 12 (2024), pp. 14349-14362
- [15]A rigorous electrochemical ammonia electrolysis protocol with in operando quantitative analysisJ. Mater. Chem. A, 9 (2021), pp. 11571-11579
- [16]Achieving equilibrium conversions for low-temperature NH3 decomposition: synergy between catalyst and electric field effectsChem. Eng. J., 476 (2023), Article 146715
- [17]Synergistic effect between electric field and Ce-doped catalysts to promote hydrogen production from ammonia decompositionFuel, 351 (2023), Article 128796
- [18]Hydrogen production by NH3 decomposition at low temperatures assisted by surface protonicsChem. Sci., 15 (2024), pp. 15125-15133
- [19]Status of silicon carbide (SiC) as a wide-bandgap semiconductor for high-temperature applications: a reviewSolid-State Electron., 39 (1996), pp. 1409-1422
- [20]Mesoporous SiC with potential catalytic application by electrochemical dissolution of polycrystalline 3C-SiCACS Appl. Nano Mater., 1 (2018), pp. 2609-2620
- [21]Ni nanoparticles supported on high surface area carborundum for enhanced hydrogen production by ammonia decompositionACS Appl. Nano Mater., 6 (2023), pp. 9892-9900
- [22]Ru-supported lanthania-ceria composite as an efficient catalyst for COx-free H2 production from ammonia decompositionAppl. Catal. B: Environ., 285 (2021), Article 119831
- [23]Dispersed surface Ru ensembles on MgO(111) for catalytic ammonia decompositionNat. Commun., 14 (2023), p. 647
- [24]Interfacial regulation of Ru-based catalysts for the enhanced activity of ammonia decompositionACS Sustain. Chem. Eng., 12 (2024), pp. 15024-15032
- [25]Enhanced ammonia decomposition activity over unsupported Co3O4: unravelling the promotion effect of alkali metalAppl. Catal. B: Environ., 330 (2023), Article 122644
- [26]Facile one-pot synthesis of Ni-based catalysts by cation-anion double hydrolysis method as highly active Ru-free catalysts for green H2 production via NH3 decompositionAppl. Catal. B: Environ., 307 (2022), Article 121167
- [27]Highly efficient decomposition of ammonia using high-entropy alloy catalystsNat. Commun., 10 (2019), p. 4011
- [28]Hydrogen production by ammonia decomposition over Cs-modified Co3Mo3N catalystsAppl. Catal. B: Environ., 218 (2017), pp. 1-8
- [29]Ceria-supported ruthenium clusters transforming from isolated single atoms for hydrogen production via decomposition of ammoniaAppl. Catal. B: Environ., 268 (2020), Article 118424
- [30]Catalytic properties of trivalent rare-earth oxides with intrinsic surface oxygen vacancyNat. Commun., 15 (2024), p. 5751
- [31]Quantifying hot carrier and thermal contributions in plasmonic photocatalysisScience, 362 (2018), pp. 69-72
- [32]Earth-abundant photocatalyst for H2 generation from NH3 with light-emitting diode illuminationScience, 378 (2022), pp. 889-893
- [33]Key factor for the anti-Arrhenius low-temperature heterogeneous catalysis induced by H+ migration: H+ coverage over supportChem. Commun., 56 (2020), pp. 3365-3368
- [34]Geometrical-site-dependent catalytic activity of ordered mesoporous Co-based spinel for benzene oxidation: in situ DRIFTS study coupled with Raman and XAFS spectroscopyACS Catal., 7 (2017), pp. 1626-1636
- [35]MCM-41 supported Co Mo bimetallic catalysts for enhanced hydrogen production by ammonia decompositionChem. Eng. J., 207-208 (2012), pp. 103-108
- [36]CoaSmbOx Catalyst with Excellent Catalytic Performance for NH3 DecompositionChin. J. Chem., 39 (2021), pp. 2359-2366
- [37]Efficient propane low-temperature destruction by Co3O4 crystal facets engineering: Unveiling the decisive role of lattice and oxygen defects and surface acid-base pairsAppl. Catal. B: Environ., 283 (2021), Article 119657
- [38]Raman in situ characterization of the species present in Co/CeO2 and Co/ZrO2 catalysts during the COPrOx reactionInt. J. Hydrog. Energy, 41 (2016), pp. 4993-5002
- [39]Oxidized carbon/nano-SiC supported platinum nanoparticles as highly stable electrocatalyst for oxygen reduction reactionInt. J. Hydrog. Energy, 39 (2014), pp. 16310-16317
- [40]Transition of surface phase of cobalt oxide during CO oxidationPhys. Chem. Chem. Phys., 20 (2018), pp. 6440-6449
- [41]Lean methane oxidation over Co3O4/Ce0.75Zr0.25 catalysts at low-temperature: Synergetic effect of catalysis and electric fieldChem. Eng. J., 369 (2019), pp. 660-671
- [42]Hydrogen generation by ammonia decomposition over Co/CeO2 catalyst: influence of support morphologiesAppl. Surf. Sci., 532 (2020), Article 147335
- [43]Catalytic oxidation of n-octane over cobalt substituted ceria (Ce0.90Co0.10O2−δ) catalystsAppl. Catal. A: Gen., 447-448 (2012), pp. 135-143
- [44]A Simple Highly Efficient Catalyst with NiOX-loaded Reed-based SiC/CNOs for Hydrogen Production by Photocatalytic Water-splittingChemistrySelect, 9 (2024), Article e202303612
- [45]Decreasing the catalytic ignition temperature of diesel soot using electrified conductive oxide catalystsNat. Catal., 4 (2021), pp. 1002-1011
- [46]The enhancement of NH3-SCR performance for CeO2 catalyst by CO pretreatmentEnviron. Sci. Pollut. Res., 27 (2020), pp. 13617-13636
- [47]Influence of CeO2 loading on structure and catalytic activity for NH3-SCR over TiO2-supported CeO2J. Rare Earths, 38 (2020), pp. 883-890
- [48]Preparation of high surface area carborundum-supported cobalt catalysts for hydrogen production by ammonia decompositionCarbon Lett., 33 (2023), pp. 899-908
- [49]Layered double hydroxides derived Nix(MgyAlzOn) catalysts: Enhanced ammonia decomposition by hydrogen spillover effectAppl. Catal. B: Environ., 201 (2017), pp. 451-460
- [50]Trace ammonia removal from air by selective adsorbents reusable with waterACS Appl. Mater. Interfaces, 12 (2020), pp. 15115-15119
- [51]Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis setPhys. Rev. B - Condens. Matter Mater. Phys., 54 (1996), pp. 11169-11186
- [52]Plasma-catalytic ammonia synthesis on Ni catalysts supported on Al2O3, Si-MCM-41 and SiO2Int. J. Hydrog. Energy, 60 (2024), pp. 802-813
- [53]Ni-CeO2 nanocomposite with enhanced metal-support interaction for effective ammonia decomposition to hydrogenChem. Eng. J., 473 (2023), Article 145371
Cited by (2)
Recent Advances in Ammonia Decomposition Technologies for Hydrogen Production
2025, Energy and FuelsSiC-Supported Polymetallic Oxide Catalysts for Enhanced Low-Temperature SCR Denitration
2025, Energy and Fuels

