Home Nanoporous copper: fabrication techniques and advanced electrochemical applications
Article Publicly Available

Nanoporous copper: fabrication techniques and advanced electrochemical applications

This article has been retracted. Retraction note.
  • Aumber Abbas

    Aumber Abbas received his BSc in metallurgy and materials engineering from Bahauddin Zakariya University, Pakistan, in 2012 and Msc in materials science and engineering from Shandong University, China, in 2016. In September 2016, he became a postgraduate researcher at the School of Chemical Engineering and Advanced Materials, Newcastle University, UK. His current research interests focus on the conversion of biowaste into superior electrode materials for energy storage applications.

    EMAIL logo
    , Saleem Abbas

    Saleem Abbas received his BSc in physics from Bahauddin Zakariya University, Multan, Pakistan, in 2009 and MSc in physics from Quiad-e-Azam University, Islamabad, Pakistan, in 2013. In September 2014, he became a doctoral research student at Hebei Key Laboratory of Boron Nitride Micro and Nano Materials, Hebei University of Technology, Tianjin, China. His current research interests focus on the fabrication of single-walled and multiwalled boron nanotubes and their application in energy devices.

    and Xianli Wang

    Xianli Wang received her Bsc from Dalian Polytechnic University of Materials Chemistry in 2012 and her MSc degree from Shandong University, Jinan, China, in 2015. She is currently a PhD student at the University of Georgia, USA. Her research focuses on functional composite materials and applications in electrical, electromagnetic, and luminescent fields (e.g. semiconductor and wave transparent materials).

Published/Copyright: November 22, 2016

Abstract

Nanoporous copper (NPC), a representative type of nanostructured materials, holds an extensive ability to generate propitious potential for a broad range of highly promising applications. Especially, with the advancement in fabrication techniques, NPC with numerous special and superior properties, such as unique pore structure, large surface-to-volume ratio, enlarged specific surface area, and high electrical and thermal conductivities, has boosted the interest to explore its electrochemical properties and extended its promising applications in energy, sensing, actuation, and catalytic systems. Therefore, timely updates of such a type of material are highly demanding and appealing for a broad audience. This review summarizes the latest advances in the development of NPC with a special focus on synthesis methods and state-of-the-art electrochemical applications such as electrocatalysts, sensors, and energy conversion/storage systems. The important scientific disputes and future research directions are also presented.

1 Introduction

Porous materials consist of various types of organic and inorganic ones, including metals, polymers, metal oxides, and organic-inorganic composite materials (Erlebacher & Seshadri, 2009; Zhao et al., 2006). According to the International Union of Pure and Applied Chemistry (IUPAC), porous materials can be divided into three categories based on their pore size: macroporous (>50 nm), mesoporous (50 nm >pore size >2 nm), and microporous (<2 nm; Zhao et al., 2006). Nanoscience has taught us that any material fabricated by the building units in the size range of 1–100 nm can cause better properties than their bulk counterparts. Therefore, it is necessary to tailor and introduce the pore structures in this size range. Nanoporous copper (NPC), as a subset of nanoporous materials, with substantially ordered porous network, narrow pore size distribution, and high conductivity, is of particular interest for a broad variety of potential applications. Electrochemical catalysts, for example, are of substantial interest because the porous network not only increases the active surface area available to the reactant molecules but also improves the mobility of electrons in the solid ligaments (Bae et al., 2012; Erlebacher & Seshadri, 2009; Qiao & Li, 2011; Yamauchi et al., 2009).

Copper (Cu) is of considerable interest in the electronic industry due to its higher electrical conductivity, lower electrical resistivity, good resistance toward electromigration, and considerable mass transfer ability (Krongelb et al., 1998; Riveros et al., 2005). Furthermore, Cu nanowires and nanorods can be suitable candidates for various promising applications, such as interconnects in nanoelectronic devices (Krongelb et al., 1998; Liu & Bando, 2003; Pena et al., 2000). Other potential applications of NPC have been investigated from electrochemical catalysis and sensors to the energy conversion and storage devices (Lang et al., 2011; Yamauchi et al., 2009; Zheng & Li, 2012). However, the fabrication of well-ordered NPC faces many challenges, as metals at the nanoscale level tend to prefer the low surface areas for minimizing the surface energy. In the last several decades, numerous advanced approaches have been proposed and developed to optimize and rationally design the pore structure of nanoporous metals (NPMs; Braun & Wiltzius, 1999; Chen et al., 2009; Velev et al., 1999). The various successfully synthesized NPCs (Aburada et al., 2011; Chen et al., 2009; Dan et al., 2012, 2013; Li et al., 2015; Liu et al., 2011, 2013, 2015; Luo et al., 2012; Luo et al., 2013; Mao et al., 2012; Qi et al., 2009b; Zhang et al., 2009a,b; Zhang et al., 2011a; Zhao et al., 2009) enable the systematic investigation of their physiochemical and electrochemical properties and the fundamental understanding, leading to the invention of a broad range of innovative functional devices, such as high-performance lithium ion batteries, ultrathin supercapacitors, and highly sensitive biosensors (Chen et al., 2010; Hou et al., 2013; Lang et al., 2011; Liu et al., 2013; Srivastava et al., 2011; Zhang & Li, 2012).

Currently, the field of NPC fabrication and application is on the way of exciting development and increasing achievements in practical life. There is a high need to provide timely updates of this advanced nanomaterial, including its extensive fabrication methods, fundamental properties, and potential applications. Various fundamental and commonly used fabrication techniques as well as complex advanced synthetic strategies and state-of-the-art electrochemical application in sensors, catalysis, and energy systems are discussed here. In addition, crucial scientific challenges and perspective directions of research are also presented.

2 Techniques to fabricate NPC

2.1 Template method

The template method is a commonly used approach for synthesizing NPMs over a wide range of pore size distribution from macroporous to microporous and further to ordered hierarchical nanoporous structures. The peak popular template approach for preparing NPC is the sacrificial porous mold technique as follows: (1) selecting a proper template with the required porous structure, (2) filling the target precursor into the pores of the template and then transforming it into a solid phase via chemical or electrochemical reduction, and (3) fabricating nanoporous structure by removing template (Kulinowski et al., 2000; Stein et al., 2007; Velev et al., 1999; Yamauchi & Kuroda, 2008; Zhao et al., 2006).

A large number of templates with a wide range of pore size distributions and pore densities are available at present. These include alumina templates prepared from the anodic oxidation of aluminum (Li et al., 1999; Masuda & Satoh, 1996; Sulka et al., 2002), polycarbonate membranes (Molares et al., 2001; Schuchert et al., 2003), mica crystals (Possin, 1970) containing etched nuclear particle tracks, and diblock copolymers (Thurn-Albrecht et al., 2000). Compared to other templates, the alumina templates have decent characteristics for synthesizing porous structures because they are mechanically and thermally stable, which can be produced in a range from ~10 nm to several hundred nanometers. Mijangos et al. (2016) recently reported a comprehensive review of the progress of polymer nanostructures with modulated morphologies and properties using nanoporous AAO templates. Another report was also recently presented on the facile fabrication of mesoporous CaO sorbents using simple salt as a pore template in a template-assisted and spray-drying synthesis method (Ebrahimi et al., 2016). Furthermore, anodization conditions can be tuned easily to adjust the alumina pore diameter from ~10 to >100 nm (Li et al., 2000; Metzger et al., 2000). Moreover, electrodeposition is one of the most commonly used techniques for filling semiconducting and conducting materials into the nanoporous template to produce continuous nanowires. This technique is called “bottom-up” fashion and is taken as the most promising and efficient method for the fabrication of nanostructures (Zhang et al., 2004). Nanotubes, nanorods, and nanowires can be fabricated using this technique.

Riveros et al. (2005) described this bottom-up template method for the fabrication of Cu nanowires. He reported that highly crystallographically oriented single-crystalline Cu nanowires can be fabricated by electrochemically depositing Cu from 0.5 m CuSO4·5H2O solution into the nanopores of a commercial alumina template. The kinetics of propagating process was studied via current versus time (it) curves and a pore filling of 80%–90% was obtained. Cu nanowires with a diameter of 150 nm with face-centered cubic structure highly oriented along the [100] direction parallel to the wire axis were obtained after the removal of template. This preferential growth was obtained for the first time.

Figure 1 demonstrates the schematic representation of the different stages of the growth process and current response during the electrodeposition of Cu at constant potential of 50 mV versus Ag/AgCl. Figure 2 shows the scanning electron microscopy (SEM) images of the Cu nanowires obtained after the removal of the template. It can be seen that highly ordered Cu nanowires with high packing density in bulk quantities were fabricated and that Cu nanowires with diameters ranging from 190 to 230 nm, with an average diameter of 210 nm, were fabricated (Riveros et al., 2005).

Figure 1: 
						Current density versus time curve for the potentiostatic electrodeposition of Cu nanowires (η=−50 mV) into pores of anodic alumina membrane (20 nm nominal pore diameter; 0.5 m CuSO4·5H2O solution, pH 1, room temperature).
						The scheme displays five different stages of the growth process: in region (A), double layer charging; in region (B), copper wires grow into the pores; in region (C), the pores are just completely filled; in region (D), hemispherical caps are formed; and in region (E), film formation commences over the whole of the alumina surface. Reproduced from Riveros et al. (2005), with permission from Springer.
Figure 1:

Current density versus time curve for the potentiostatic electrodeposition of Cu nanowires (η=−50 mV) into pores of anodic alumina membrane (20 nm nominal pore diameter; 0.5 m CuSO4·5H2O solution, pH 1, room temperature).

The scheme displays five different stages of the growth process: in region (A), double layer charging; in region (B), copper wires grow into the pores; in region (C), the pores are just completely filled; in region (D), hemispherical caps are formed; and in region (E), film formation commences over the whole of the alumina surface. Reproduced from Riveros et al. (2005), with permission from Springer.

Figure 2: 
						SEM micrographs of the exposed Cu nanowires after the alumina template membrane was dissolved away.
						Electrodeposition has been stopped into region (b). See Figure 2. (A) ×10,000 and (B) ×30,000 (η=−250 mV). Reproduced from Riveros et al. (2005), with permission from Springer.
Figure 2:

SEM micrographs of the exposed Cu nanowires after the alumina template membrane was dissolved away.

Electrodeposition has been stopped into region (b). See Figure 2. (A) ×10,000 and (B) ×30,000 (η=−250 mV). Reproduced from Riveros et al. (2005), with permission from Springer.

2.2 Magnetron sputtering method

Magnetron sputtering has been fundamentally used for the fabrication of NPC with tunable nanopore size for various promising applications in sensors and lithium ion batteries (Hou et al., 2013; Su et al., 2013). NPC is fabricated through the combination of deposition of an alloy and selective chemical dealloying. Magnetron sputtering with specific input power is used to deposit an alloy of Cu and other relatively less noble metal on a substrate of Cu or other metals (Su et al., 2013). Several reports have been recently presented for the fabrication of various nanostructures through magnetron sputtering. For example, Masłyk et al. (2016) reported on the growth of Zn/ZnO nanostructures obtained by DC reactive magnetron sputtering. Hakamada et al. (2016) also reported on the electrical resistivity of nanoporous gold (NPG) modified with thiol self-assembled monolayers.

Figure 3 shows the schematic representation of the fabrication mechanism of NPC in forming thin films on a substrate, with columnar and ligament-channel structures (Su et al., 2013). It can be seen from Figure 3A that a Cu film consisting of pseudo-pyramid-shaped tiny particles was predeposited on a glass substrate. When this Cu/glass substrate was subjected to external negative biased voltage, the electrons generated by the static electricity preferentially gathered on the high curvature surfaces of particles especially on their tips or tops. The new coming Cu atoms were polarized by the electric field of the particles tops and absorbed thereby preferentially on the tops. As a result, columnar growth took place on the predeposited Cu film (Su et al., 2013).

Figure 3: 
						Schematic illustrations show the formation mechanisms of NPC films with (A) a columnar structure and (B) a ligament-channel structure.
						Reprinted from Su et al. (2013), with permission from Elsevier.
Figure 3:

Schematic illustrations show the formation mechanisms of NPC films with (A) a columnar structure and (B) a ligament-channel structure.

Reprinted from Su et al. (2013), with permission from Elsevier.

On the contrary, Figure 3B shows a Cu film consisting of large asymmetrical particles having randomly aligned tips, which deposited directly on the glass slide. The selective accumulation of the static electrons on tips of the particles lead toward a rod-like growth process of Cu along the tip position orientation. Following the deposition process, the growth of these randomly designed rods lead to an adhesion of rods at the active adsorption sites when they combined together, and as a result, a bicontinuous ligament-channel structure formed (Su et al., 2013).

Hou et al. (2013) described the formation of bicontinuous NPC by following the deposition of Cu30Mn70 alloy thin film onto the Cu foil substrate via magnetron sputtering and subsequent chemical dealloying the deposited film in 10 mm HCl solution for 5 h. During this chemical dealloying, less noble metal Mn from Cu30Mn70 alloy is selectively dissolved, whereas the remained Cu forms a bicontinuous nanoporous structure. Figure 4 shows the top-view SEM image of as-dealloyed NPC supported on Cu substrate. It can be seen that a 800-nm-thick layer of bicontinuous NPC consisting of quasi-periodic ligament-channel structure is seamlessly jointed with Cu foil substrate.

Figure 4: 
						(A) Top-view and (B) cross-sectional SEM images of S/NP Cu with a characteristic length of 50 nm.
						Reprinted from Hou et al. (2013), with permission from Nature Publishing Group.
Figure 4:

(A) Top-view and (B) cross-sectional SEM images of S/NP Cu with a characteristic length of 50 nm.

Reprinted from Hou et al. (2013), with permission from Nature Publishing Group.

2.3 Electrodeposition

The fabrication of NPMs, such as Cu, nickel (Ni), and cobalt, by electrodeposition has been a very intriguing topic of research for a number of decades from both technological and scientific points of view (Ben-Jacob & Garik, 1990; Schetty, 2001). Particularly, 3D nanoporous structures produced by electrodeposition are well suited for various applications in batteries, supercapacitors, fuel cells, and sensors. Recently, a promising and green method based on the gas bubble template was proposed for the fabrication of 3D NPMs and alloys with thoroughly porous dendritic wall (Shin & Liu, 2004, 2005; Shin et al., 2004; Tong et al., 2009). The electrodeposition process can efficiently generate a 3D nanoporous structures; however, it is very challenging to control the microstructures with desired characteristics (Shin et al., 2003). The branches in the porous structures are usually not able to bear the weight of a number of subbranches and results in the collapse of itself in the solution. Moreover, dendritic structures usually overgrow very easily, resulting in a thick film with quite a little porosity, lowering the gas/liquid transport and its potential for electrochemical applications.

Recently, several reports have been presented on the fabrication of nanoporous structures through electrodeposition. For example, Ashassi-Sorkhabi et al. (2016) recently reported on the fabrication of porous Ni through electrodeposition using the hydrogen gas bubble template method. They deposited Zn and Ni(OH)2 on the surface of porous Ni followed by dealloying of Zn to prepare a 3D porous Ni/Ni(OH)2 nanocomposite for supercapacitor applications. Bozzini et al. (2016) reported on the electrochemical fabrication of NPG decorated with manganese oxide nanowires from eutectic urea/choline chloride ionic liquid.

Figure 5 schematically shows the procedure of NPC formation by electrochemical deposition using the hydrogen bubble dynamic template (Shin et al., 2003). Hydrogen bubbles emerging from the cathodic reaction at metal substrate create a uniform path from substrate to electrolyte. Where there is a bubble, there will be no deposition of metal due to the unavailability of metal ions. Hence, a large number of hydrogen bubbles generated at the substrate during deposition process will result in the pores in the deposit. As the moving hydrogen bubble creates a pathway, the growth process of metal takes this route among the gas bubbles. As a result, a chain of gas bubbles behaves as a template for the fabrication of NPC.

Figure 5: 
						Simplified description of the formation process of the nanoporous structures.
						Reproduced from Li et al. (2007a,b) with permission from American Chemical Society.
Figure 5:

Simplified description of the formation process of the nanoporous structures.

Reproduced from Li et al. (2007a,b) with permission from American Chemical Society.

Shin et al. (2003, 2004) reported on the formation of dendritic structure of Cu with highly porous walls. Figure 6 shows the SEM images of electrochemically deposited NPC. The highly porous walls were composed of a number of small-ramified Cu and overgrowing was not noticed. The characteristic size of branches in NPC was of the order of several hundreds of nanometers. It was found that porous deposits were cross-linked in the form of a network, resulting in a mechanically sound supported structure.

Figure 6: 
						Typical SEM images of porous Cu deposits created by electrodeposition for different periods of time: (A) 5 s, (B) 10 s, and (C) 20 s.
						The surface pore size of the 3D foam structure increased with the time of deposition (or the distance from the substrate). Reproduced from Shin et al. (2003) with permission from Wiley.
Figure 6:

Typical SEM images of porous Cu deposits created by electrodeposition for different periods of time: (A) 5 s, (B) 10 s, and (C) 20 s.

The surface pore size of the 3D foam structure increased with the time of deposition (or the distance from the substrate). Reproduced from Shin et al. (2003) with permission from Wiley.

Li et al. (2007a,b) also reported on the fabrication of NPC by electrodeposition using the hydrogen bubble template. Figure 7 shows the typical SEM micrographs of NPC obtained by electrochemical deposition in a 0.1 m CuSO4 and 0.5 m H2SO4 solution for 45 s. Cu films consisted of 3D interconnected porous structures. The porous structures consisting of numerous dendrites in all directions can be vividly observed in magnified images.

Figure 7: 
						Top-view (A, B, C) and side-view (D) SEM images with different magnifications of the porous Cu film electrochemically deposited at a 0.8 A cm−2 cathodic current density in a solution of 0.1 m CuSO4 and 0.5 m H2SO4 for 45 s.
						Reproduced from Li et al. (2007a,b) with permission from American Chemical Society.
Figure 7:

Top-view (A, B, C) and side-view (D) SEM images with different magnifications of the porous Cu film electrochemically deposited at a 0.8 A cm−2 cathodic current density in a solution of 0.1 m CuSO4 and 0.5 m H2SO4 for 45 s.

Reproduced from Li et al. (2007a,b) with permission from American Chemical Society.

2.4 Dealloying

Dealloying, also known as selective dissolution, has been acknowledged as the most promising approach for the fabrication of NPMs (Erlebacher, 2004; Erlebacher & Seshadri, 2009; Erlebacher et al., 2001). Raney Ni, for example, a porous metallic powder produced through dealloying of NiAl3, has been used for more than 90 years as a heterogeneous catalyst (Raney, 1927). Raney Cu has also been fabricated a long ago through the dealloying of CuAl2 in NaOH solution (Friedrich et al., 1983; Marsden et al., 1980). With the development of modern microscopic characterization techniques and thorough investigation of porosity evolution process, this promising approach has been extensively implemented to synthesize NPMs with intriguing unique properties.

Dealloying is basically a corrosion process, in which the selective dissolution of an active element out of a single-phase alloy consisting of two or multiple elements having dissimilar chemical activities generates a porous structure (Chang et al., 2008; Chen et al., 2009; Erlebacher et al., 2001; Huang & Sun, 2004; Wittstock et al., 2010a,b). Alloy systems with solid solubility across a wide range of composition give the possibility of tailoring the pore size over a broad range. The fundamental condition for dealloying process to occur is that both elements of a two-phase alloy system should necessarily have sufficiently dissimilar equilibrium potentials, allowing the more active one to dissolve away while the second one remain intact. Furthermore, exposure to corrosion conditions should not result in the formation of stable oxides. The nanoporous structure could be formed by the selective dissolution of the more active component either naturally in base/acid solution (called chemical corrosion) or more speedily upon the application of external potential (called electrochemical corrosion).

2.4.1 Chemical dealloying

Chemical dealloying has been extensively used for the fabrication of a number of NPMs (Jiang et al., 2013; Liu et al., 2011; Luo et al., 2013; Qi et al., 2009a,b; Sun et al., 2004; Wang et al., 2009; Zhang et al., 2009a,b; Zhao et al., 2009; Zhao et al., 2010). NPC has been synthesized by chemical dealloying a long ago. For example, Raney Cu was synthesized by the chemical dealloying of CuAl2 binary alloys in NaOH solution (Smith et al., 1999). Through numerous studies on the formation process of NPC, it has been found that dealloying conditions have a major effect on the final porous structure of NPC (Hayes et al., 2006; Jiang et al., 2013; Liu et al., 2011; Luo et al., 2013; Qi et al., 2009b; Zhao et al., 2009). However, there is still a lack of systematic investigation on the effect of dealloying conditions on the formation and evolution of nanoporosity of NPC. Erlebacher et al. (2001) described in detail the physical mechanism behind dealloying and the evolution of nanoporosity in metals and suggested that an inherent dynamical pattern formation mechanism was involved in it.

To produce a uniform nanoporous structure by dealloying, two basic requirements of Cu-containing precursors are needed to be fulfilled: (1) homogeneous single-phase solid solution and (2) sufficient electrochemical difference between Cu and other alloying elements (Erlebacher, 2004). Chen et al. (2009) presented a systematic study on the formation of NPC with tunable nanoporosity. They carried out the selective dissolution of Mn out of Cu30Mn70 binary alloy precursor in HCl solution. Figure 8 represents the formation of uniform NPC obtained by dealloying Cu30Mn70 alloy in 0.025 m HCl solution. It can be seen from the cross-sectional SEM image that the ligament-channel structure was uniformly arranged throughout the sample. The uniform nanoporosity was further revealed by the transmission electron microscopy (TEM) images in Figure 8B.

Figure 8: 
							(A) Cross-sectional SEM image and (B) TEM micrograph of NPC produced by dealloying Cu30Mn70 ribbons in 0.025 m HCl solution for 12 h.
							Reproduced from Chen et al. (2009), with permission from Wiley.
Figure 8:

(A) Cross-sectional SEM image and (B) TEM micrograph of NPC produced by dealloying Cu30Mn70 ribbons in 0.025 m HCl solution for 12 h.

Reproduced from Chen et al. (2009), with permission from Wiley.

Chen et al. (2009) also described the effect of dealloying time on the formation nanoporosity. The coarsening of nanopores occurred with the increase in dealloying time. Figure 9 shows NPC formed with different pore sizes according to the subjected dealloying time. The average nanopores size could be tailored from ~15 to ~120 nm by monitoring the dealloying time. In particular, the bicontinuous structure of all samples was necessarily the same apart from the variation in the specific length of pores and ligaments, which is same as that of NPG.

Figure 9: 
							SEM images of NPC produced by dealloying Cu30Mn70 in 0.025 m HCl for (A) 0.5 h, (B) 2 h, (C) 6 h, (D) 12 h, (E) 20 h, and (F) 32 h.
							Reproduced from Chen et al. (2009), with permission from Wiley.
Figure 9:

SEM images of NPC produced by dealloying Cu30Mn70 in 0.025 m HCl for (A) 0.5 h, (B) 2 h, (C) 6 h, (D) 12 h, (E) 20 h, and (F) 32 h.

Reproduced from Chen et al. (2009), with permission from Wiley.

Asghari et al. (2016) recently reported on NPC fabrication through dealloying for application in catalysis. Cu sheets were coated with a layer of Cu-Zn alloy. The zinc content of the alloy was decreased during immersion in HCl solution and an NPC electrode with lower zinc content was formed. Tin oxide nanoparticles were synthesized through a galvanostatic pathway on polypyrole (PPy)-coated NPC. In methanol solution, a significant enhancement in the catalytic action of PPy was observed after decoration of tin oxide nanoparticles. This enhancement is attributed to the effect of tin oxide on the adsorption of intermediates of methanol oxidation as well as the oxidation of by-products such as CO.

It has been reported that dealloying of one single-phase solid solutions in different electrolytes generated nanoporous structures with considerably different morphologies and length scales; however, all of the electrolytes produced bicontinuous ligament-channel structures. Hayes et al. (2006) studied the effect of different electrolytes on the morphology of NPC obtained by dealloying Mn0.7Cu0.3 in HCl, citric acid, H2SO4, MnSO4, and (NH4)2SO4+MnSO4 aqueous solutions. Figure 10A–D shows the SEM images of NPC obtained from dealloying Mn0.7Cu0.3 in (A) pH1.3 HCl, (B) 1 m citric acid, (C) 0.01 m H2SO4+0.001 m MnSO4, and (D) 1 m (NH4)2SO4+0.01 m MnSO4. Hayes et al. suggested that the NPC obtained with significant different ligament sizes and structures is attributed to the difference in dealloying rates and surface motilities of Cu in different electrolytes. Thus, the selection of suitable electrolyte is highly important to generate a nanoporous structure with desirable characteristics.

Figure 10: 
							Dealloying electrolytes affect both the ligament size and morphology.
							These SEM micrographs show the structure resulting from dealloying Mn0.7Cu0.3 in (A) pH 1.3 HCl, (B) 1 m citric acid, (C) 0.01 m H2SO4+0.001 m MnSO4, and (D) 1 m (NH4)2SO4+0.01 m MnSO4. Reprinted from Hayes et al. (2006), with permission from Materials Research Society.
Figure 10:

Dealloying electrolytes affect both the ligament size and morphology.

These SEM micrographs show the structure resulting from dealloying Mn0.7Cu0.3 in (A) pH 1.3 HCl, (B) 1 m citric acid, (C) 0.01 m H2SO4+0.001 m MnSO4, and (D) 1 m (NH4)2SO4+0.01 m MnSO4. Reprinted from Hayes et al. (2006), with permission from Materials Research Society.

2.4.2 Electrochemical dealloying

One of the promising strategies to fabricate NPMs is electrochemical dealloying (Luo et al., 2013; Sun et al., 2004; Zhang et al., 2009a,b; Zhao et al., 2010). The morphology of NPMs can be adjusted by fine-tuning the initial alloy composition and typical dealloying conditions. Numerous researchers studied the effect of initial precursor alloy composition on the final microstructure of NPMs (Aburada et al., 2011; Dan et al., 2013; Liu et al., 2011; Mao et al., 2012).

A number of reports have been published on the fabrication of NPC through the electrochemical dealloying of binary alloys (Dan et al., 2012, 2013; Hayes et al., 2006; Li et al., 2015; Luo et al., 2013; Qi et al., 2009b; Zhao et al., 2010). Tang et al. (2016) reported on the fabrication of well-ordered NPC by dealloying Cu-Mn. Wang et al. (2016) reported on the fabrication of nanosheets Co3O4 by oxidation-assisted dealloying method for high-capacity supercapacitors. Luo et al. (2013) described the nucleation and growth of NPC ligaments during the electrochemical dealloying of Mg-based metallic glasses. Zhang et al. (2009a,b) reported on the fabrication of a number of NPMs including Cu through the electrochemical dealloying of Al-based alloys. Zhang et al. found that interactions among the coexistent phases take place during the dealloying process of a two-phase alloy.

Zhao et al. (2010) described the electrochemical dealloying of Mg-Cu binary alloys in 0.2 m NaCl aqueous solutions and suggested that the electrochemical activities of this binary alloy were dependent not only on the initial alloy compositions but also on the phase constitutions. The critical potential and open circuit potential for MgCu2 phase are much higher than those of Mg2Cu phase by about 450 mV. Furthermore, the dissolution of less noble Mg2Cu determines the critical and open circuit potential of biphasic alloy rather than being proportional to the amounts of constitutive elements or phases. The dealloying manners of single-phase and biphasic Mg-Cu alloys depend on the electrochemical properties of constitutive phases.

Figure 11 shows the cyclic voltammetry (CV)-based investigation of the dealloying behavior of Mg50Cu50 biphasic alloy under the applied potential of −1.0 to −0.3 V versus Ag/AgCl (Zhao et al., 2010). Both quenched alloys were composed of Mg2Cu and MgCu2 phases; however, the dealloying behaviors of the two alloys were different. In the first 10 cycles, a type of pitting corrosion was observed for the dealloying of biphasic Mg50Cu50 alloy (Figure 11A). That is, the current density of the forward scan is lower than that of the reverse scan. It was suggested that the dealloying process of Mg2Cu is dominant at this initial stage. A crossover appeared in the 18th cycle but quickly disappeared with increasing cycles (Figure 11B). At lower potential section, the dealloying of more active Mg2Cu phase took place because the current density of the reverse scan is higher than that of the forward scan (pitting corrosion), whereas at higher potential higher current density can be observed for the forward scan, which is linked with the dealloying of the less active MgCu2 phase. Upon the dissolution of Mg2Cu, the crossover disappears. The dealloying of the less active MgCu2 phase is dominant at this stage. In addition, it can be observed that the dissolution rate of Cu is higher during the dealloying of MgCu2 than that during the dealloying of Mg2Cu with in the same potential zone.

Figure 11: 
							Evolution of CV curves of the rapidly solidified Mg50Cu50 alloy subjected to 80 cycles of CV measurements.
							(A) Indicates 1, 3, 10 cycles of CV measurements and (B) indicates 18, 25, 50 and 80 cycles of CV measurements. Reprinted from Zhao et al. (2010), with permission from Elsevier.
Figure 11:

Evolution of CV curves of the rapidly solidified Mg50Cu50 alloy subjected to 80 cycles of CV measurements.

(A) Indicates 1, 3, 10 cycles of CV measurements and (B) indicates 18, 25, 50 and 80 cycles of CV measurements. Reprinted from Zhao et al. (2010), with permission from Elsevier.

Figure 12 represents the CV plots of Mg33Cu67 alloy subjected to the CV measurements between −0.6 and −0.2 V versus Ag/AgCl. A pitting corrosion was observed at higher potential during the initial dealloying stage, which indicates that the dissolution of Cu is unavoidable even at the start. The current density values during the initial 25 cycles (Figure 12A) are at least one order smaller than that of Mg50Cu50 alloy. It was noticed that current density gradually increases with the increase in CV cycles up to 200, even at a lower potential regime. In addition, a crossover was observed on all loops after the 50th cycle and reducing peak becomes indiscernible. Furthermore, with increasing cycles, the crossover transfers to the higher potential, suggesting that the dealloying process is converted from pitting-type corrosion to bulk corrosion (Figure 12B; Zhao et al., 2010).

Figure 12: 
							Evolution of CV curves of the rapidly solidified Mg33Cu67 alloy subjected to 200 cycles of CV measurements.
							Reprinted from Zhao et al. (2010), with permission from Elsevier.
Figure 12:

Evolution of CV curves of the rapidly solidified Mg33Cu67 alloy subjected to 200 cycles of CV measurements.

Reprinted from Zhao et al. (2010), with permission from Elsevier.

Figure 13 shows the section-view SEM images of as-dealloyed Mg50Cu50 and Mg33Cu67 alloys in 0.2 m NaCl solution at an applied potential of −0.3 V versus Ag/AgCl. The microstructure of as-dealloyed Mg50Cu50 consists of a bicontinuous interpenetrating ligament-channel structure with an average size of 188±45 nm (Figure 13A). Although two phases were present in the starting precursor Mg50Cu50, the final microstructure is homogeneous throughout the sample.

Figure 13: 
							Section-view SEM images showing the microstructures of NPC through potentiostatic dealloying of the rapidly solidified Mg-Cu (A) Mg50Cu50, (B) Mg33Cu67 alloys in the 0.2 m NaCl aqueous solution at the potential of −0.3 V versus Ag/AgCl.
							Inset in (B): the entire section-view SEM of the Mg33Cu67 alloy ribbons. Reprinted from Zhao et al. (2010), with permission from Elsevier.
Figure 13:

Section-view SEM images showing the microstructures of NPC through potentiostatic dealloying of the rapidly solidified Mg-Cu (A) Mg50Cu50, (B) Mg33Cu67 alloys in the 0.2 m NaCl aqueous solution at the potential of −0.3 V versus Ag/AgCl.

Inset in (B): the entire section-view SEM of the Mg33Cu67 alloy ribbons. Reprinted from Zhao et al. (2010), with permission from Elsevier.

Figure 13B shows the cross-section-view microstructure of the as-dealloyed Mg33Cu67 alloy. A typical nanostructure is obtained after electrochemical dealloying. It is worth noting that ligaments are flake like, with sizes much smaller than those of the channels (120±30). On the contrary, ligaments observed in Figure 13A are rod-like and similar in size to that of channels. These microstructures are the result of typical electrochemical dealloying mechanism described in detail in the above section (Zhao et al., 2010).

McCue et al. (2016a) presented a comprehensive review of dealloying and dealloyed materials. They discussed the fundamental material principles underlying the formation of dealloyed materials and then looked at two major applications: catalysis and nanomechanics. McCue et al. (2016b) also studied the size effects in the mechanical properties of bulk bicontinuous Ta/Cu nanocomposites made by liquid metal dealloying.

3 Electrochemical applications

3.1 Electrochemical sensors

NPMs have been successfully and widely used in electrochemical sensors (Howorka & Siwy, 2009; Liu & Searson, 2006; Liu et al., 2009; Meng et al., 2011; Wanunu et al., 2010; Wittstock et al., 2010a,b). NPC has attracted huge attention for biosensors, especially in the form of bimetallic core-shell structures, due to its low material cost and high conductivity (Chen et al., 2010; Wang et al., 2013). The enzymatic sensors have low long-term stability and enzyme-induced intermediate H2O2 may also interfere with the sensitivity (Jena & Raj, 2006; Zen et al., 2003). Therefore, it is highly desired to synthesize a stable nonenzymatic electrochemical glucose sensor. Chen et al. (2010) conducted a systematic study on the synthesis of nonenzymatic glucose biosensor based on Au@NPC composite. Figure 14 represents the CV curves of glucose oxidation at NPG and Au@NPC electrodes in a buffer solution containing 0.05 m glucose at a scan rate of 20 mV s−1. Two anodic peaks were observed for the CV curve of Au@NPC during a positive potential scan. The intermediate formed as a result of electrosorption of glucose, which may correspond to the first peak, is adsorbed on the active sites on Au-coated NPC ligaments. Moving toward higher positive potentials, AuOH forms, which catalyzes the intermediate oxidation and forms a new gold active sites for the direct oxidation of glucose that relates to the second peak at 0.27 V (Chen et al., 2010).

Figure 14: 
						CV curves of Au@NPC and NPG electrodes in 0.1 m PBS (pH 7.4) containing 0.05 m glucose and with the presence of 0.1 m KCl (scan rate: 20 mV s−1).
						Reprinted from Chen et al. (2010), with permission from Wiley.
Figure 14:

CV curves of Au@NPC and NPG electrodes in 0.1 m PBS (pH 7.4) containing 0.05 m glucose and with the presence of 0.1 m KCl (scan rate: 20 mV s−1).

Reprinted from Chen et al. (2010), with permission from Wiley.

The oxidation of glucose over NPG was also conducted to compare the electrochemical activity of NPG to that of Au@NPC composite. The same phenomenon as presented in Figure 14 was noticed, which is in agreement with the literature (Li et al., 2007a,b). It was observed that higher potential was required for the oxidation of glucose over NPG compared to Au@NPC, signifying that Au@NPC has superior electrocatalytic activity than NPG.

Figure 15A represents the steady-state amperometric response of glucose on NPG and Au@NPG electrodes. It was observed that current signal lifted up rapidly and with high sensitivity when 0.001 m glucose was successively added in the stirring solution. The current density corresponding to NPG decreases a little upon the addition of glucose, whereas the current density response at Au@NPC rises linearly with the successive addition of glucose. Figure 15B shows this behavior more clearly, which suggests that Au@NPC electrode has far better glucose detection sensitivity than that of NPG electrode with similar porosity (Chen et al., 2010).

Figure 15: 
						(A) Amperometric response of Au@NPC and NPG electrodes to successive addition of 0.001 m glucose into stirred 0.1 m PBS (pH 7.4) with 0.1 m KCl and (B) calibration curves of the sensors.
						Reproduced from Chen et al. (2010), with permission from Wiley.
Figure 15:

(A) Amperometric response of Au@NPC and NPG electrodes to successive addition of 0.001 m glucose into stirred 0.1 m PBS (pH 7.4) with 0.1 m KCl and (B) calibration curves of the sensors.

Reproduced from Chen et al. (2010), with permission from Wiley.

3.2 Electrochemical energy systems

3.2.1 Supercapacitors

Supercapacitors, also called electrochemical capacitors, have been recognized for several decades and are known as one of the promising potential energy storage devices in addition to batteries (Conway, 1999). Based on the electrochemical response, there are two types of electrochemical supercapacitors. One category of capacitor materials is that showing high surface reactivity resulting in the formation of a double layer, also basically known as electrical double-layer (EDL) capacitors (EDLC). The second ones are those exhibiting Faradic electrochemical reactions taking place on the surface (Lokhande et al., 2011). Supercapacitors can store considerably more energy per unit volume compared to conventional capacitors due to two reasons: (1) the charge separation occurs over a very little distance in the EDL, which is the distance between the electrolyte and the adjacent electrode, and (2) the substantially enlarged specific surface area of the electrode generated by the huge number of nanopores offers the ability to store an increased amount of charge. The process of energy storage is inherently rapid due to the simple movement of ions to and from electrode surfaces. The supercapacitors display a quite high degree of reversibility in charging-discharging processes and exhibit a cycle life of more than 500,000 cycles (Lokhande et al., 2011). These excellent properties make the supercapacitors promising for a number of applications in portable electronic market, power quality systems, and particularly low-emission buses, cars, and trucks (Winter & Brodd, 2004).

Supercapacitors consist of two nonreactive porous plates suspended in an electrolyte under an applied voltage across the two porous plates (Divyashree & Hegde, 2015; Khan et al., 2000). When the potential is applied on two plates, the positive plate (anode) attracts positive charges from the electrolyte, whereas the negative plate (cathode) attracts electrons as displayed in Figure 16. This phenomenon efficiently provides sufficient space to the two layers for storing charge (Brenna et al., 2009). Therefore, supercapacitors are well suited for the electrostatic storage of charges effectively.

Figure 16: 
							Schematic diagram of charge distribution in a supercapacitor.
							Reproduced from Divyashree and Hegde (2015), with permission from The Royal Society of Chemistry.
Figure 16:

Schematic diagram of charge distribution in a supercapacitor.

Reproduced from Divyashree and Hegde (2015), with permission from The Royal Society of Chemistry.

The charging and discharging mechanism of EDL supercapacitors is presented in Figure 17 (Divyashree & Hegde, 2015). It represents the accumulation of the charges around an electrode during the charging/discharging cycle of the supercapacitors.

Figure 17: 
							Structure of double-layer capacitor during charging and discharging.
							Reproduced from Divyashree and Hegde (2015), with permission from The Royal Society of Chemistry.
Figure 17:

Structure of double-layer capacitor during charging and discharging.

Reproduced from Divyashree and Hegde (2015), with permission from The Royal Society of Chemistry.

To better understand the specific power density and energy density of supercapacitors in comparison to batteries and electrochemical energy conversion systems, a Ragone plot is presented in Figure 18 (Simon & Gogotsi, 2008). It can be observed clearly that supercapacitors occupy a specific and significant position in the Ragone plot. Today’s need of energy storage is met by supercapacitors because of its extraordinary energy density and huge power capability. Supercapacitors can work in a broad temperature range; they exhibit long cycle life and can supply high-power density.

Figure 18: 
							Ragone plot of energy storage devices.
							Reproduced from Simon and Gogotsi (2008), with permission from Nature Publishing Group.
Figure 18:

Ragone plot of energy storage devices.

Reproduced from Simon and Gogotsi (2008), with permission from Nature Publishing Group.

A large number of articles and reviews have been published on the high capacitance electrode materials for supercapacitors (Burke, 2000; Chmiola et al., 2010; Conway, 1999; Divyashree & Hegde, 2015; Kotz and Carlen, 2000; Lang et al., 2011; Mahon & Drummond, 2001; Miller & Burke, 2008; Nishino, 1996; Shukla & Martha, 2001; Simon & Burke, 2008; Wang et al., 2012; Zhang & Zhao, 2009; Zheng et al., 1997). Numerous reports have been presented on the fabrication of nanoporous materials for capacitors applications. For example, Nguyen et al. (2016) recently reported on hydrogen bubbling-induced microporous/nanoporous MnO2 films prepared by electrodeposition for pseudocapacitor electrodes. Capacitance depends largely on the features of electrode materials, especially surface areas and pore size distribution. Cupric oxide (CuO) has been widely used for high capacitance supercapacitors. NPC in the form of composite materials has been successfully used for high-energy density supercapacitors.

Prasad et al. (2016) recently reported on the fabrication and characterization of nanoporous carbon electrodes embedded with CuO nanoparticles for supercapacitors. Zhang et al. (2011b) demonstrated the porous CuO nanobelts for high-energy density pseudocapacitors. Moosavifard et al. (2015) recently reported on the fabrication of 3D highly ordered nanoporous CuO electrode for high-performance asymmetric supercapacitors. They described that CuO with highly ordered and interconnected nanopores and nanowalls have a specific surface area of about 149 m2 g−1 and exhibited a specific capacitance of 431 F g−1 at 3.5 mA cm−2. They assembled the electrode in an asymmetric cell that demonstrated energy density of about 19.7 W h kg−1 and a very high cycle life.

3.2.2 Lithium ion batteries

Batteries are electrochemical rechargeable energy storage devices that convert chemical energy into electrical energy as a result of a redox reaction taking place at anode and cathode (Becker, 1957; Linden, 1984; Samuel, 1964). This principle was first presented by Becker (1957). When an external load is applied to the battery, a flow of electrons takes place from anode to cathode resulting in the discharging of the battery (Appleby, 1988; Henne, 2007). A schematic of a conventional battery with cathode, anode, and electrolyte is presented in Figure 19. When a battery has to be recharged, the electrons’ flow direction is reversed using an external power source. This involves the restoring of the original states of cathode and anode for further redox reactions (Ogasawara et al., 2006).

Figure 19: 
							Cross-section of a conventional rechargeable battery with anode, electrolyte, and cathode connected using an external electrically powered device.
							Reproduced from Divyashree and Hegde (2015), with permission from The Royal Society of Chemistry.
Figure 19:

Cross-section of a conventional rechargeable battery with anode, electrolyte, and cathode connected using an external electrically powered device.

Reproduced from Divyashree and Hegde (2015), with permission from The Royal Society of Chemistry.

Nanostructured electroactive materials can effectively enhance the rate of lithium ion insertion/extraction to increase the power output of batteries. However, sluggish ion and electron transport kinetics combined with conventional approaches for the fabrication of nanostructured metal oxide results in considerable reductions in the energy capacities. Hou et al. (2013) reported on the 3D bicontinuous NPC/MnO2 hybrid electrode for high-performance lithium ion batteries. The solid/nanoporous hybrid enhanced the electron transport of MnO2, facilitated the fast ion diffusion, and exhibited ultrahigh charge/discharge rates and a stable capacity of 1100 mAh g−1 for 1000 cycles. Figure 20 shows the fabrication process of seamlessly integrated solid/NPC/MnO2 bulk electrode.

Figure 20: 
							Scheme showing the fabrication of seamlessly integrated S/NP Cu/MnO2 bulk electrode: (A) cleaned Cu foil substrate; (B) Cu30Mn70 alloy film deposited on the Cu foil by magnetron sputtering; (C) NPC layer on the Cu foil produced by chemically dealloying Cu30Mn70 in diluted HCl solution; (D) nanocrystalline MnO2 directly grown onto S/NP Cu skeleton using electroless plating; (E) photograph of a flexible S/NP Cu/MnO2 hybrid bulk electrode being bent (2×3 cm2); (F) four batteries assembled with S/NP Cu/MnO2 bulk electrodes power blue, red, and green LEDs; (G) schematic battery constructed with S/NP Cu/MnO2 and lithium foil as electrodes, 1 m LiPF6 in EC/EMC/DMC as electrolyte, and porous polymer as separator.
							Reprinted from Hou et al. (2013), with permission from Nature Publishing Group.
Figure 20:

Scheme showing the fabrication of seamlessly integrated S/NP Cu/MnO2 bulk electrode: (A) cleaned Cu foil substrate; (B) Cu30Mn70 alloy film deposited on the Cu foil by magnetron sputtering; (C) NPC layer on the Cu foil produced by chemically dealloying Cu30Mn70 in diluted HCl solution; (D) nanocrystalline MnO2 directly grown onto S/NP Cu skeleton using electroless plating; (E) photograph of a flexible S/NP Cu/MnO2 hybrid bulk electrode being bent (2×3 cm2); (F) four batteries assembled with S/NP Cu/MnO2 bulk electrodes power blue, red, and green LEDs; (G) schematic battery constructed with S/NP Cu/MnO2 and lithium foil as electrodes, 1 m LiPF6 in EC/EMC/DMC as electrolyte, and porous polymer as separator.

Reprinted from Hou et al. (2013), with permission from Nature Publishing Group.

Liu et al. (2013) reported on the fabrication of NPC-supported cuprous oxide (Cu2O) electrode for high-performance lithium ion batteries. The hybrid electrode was fabricated by the selective etching of Cu50Al50 alloy followed by in situ thermal oxidation. During the first cycle, a high lithium storage capacity of about 2.35 mA h cm−2 and a high reversible capacity of 1.45 mA h cm−2 were attained after 120 cycles.

Figure 21A presents the fabrication mechanism of NPC-supported Cu2O schematically. Ingots of Cu50Al50 alloy were first prepared by electron beam melting and Al was selectively dissolved in NaOH aqueous solution to produce NPC (Figure 21B and C). Finally, to prepare a Cu2O film on the surface of NPC, thermal treatment was carried out for 3 min in 140°C in air (Liu et al., 2013).

Figure 21: 
							(A) Schematic showing the fabrication process of 3D NPC@Cu2O anodes, (B) and (C) SEM images of NPC fabricated by dealloying Cu50Al50, and (D) a digital photograph showing a Cu50Al50 ingot and NPC fabricated by dealloying Cu50Al50.
							Reprinted from Liu et al. (2013), with permission from The Royal Society of Chemistry.
Figure 21:

(A) Schematic showing the fabrication process of 3D NPC@Cu2O anodes, (B) and (C) SEM images of NPC fabricated by dealloying Cu50Al50, and (D) a digital photograph showing a Cu50Al50 ingot and NPC fabricated by dealloying Cu50Al50.

Reprinted from Liu et al. (2013), with permission from The Royal Society of Chemistry.

3.3 Electrochemical catalysis

NPC has been extensively used as electrocatalyst for a wide variety of applications, such as click chemistry, water-gas shift reaction, surface-enhanced Raman spectroscopy, renewable energies, organic synthesis, and reduction of carbon dioxide (CO2) or carbon monoxide (CO) into useful chemicals and liquid fuels (Chen et al., 2009; Feng et al., 2012; Jin et al., 2011, 2012; Li, et al., 2014; Sen et al., 2014). Here, we will discuss two very promising applications of NPC as a catalyst for the conversion of CO/CO2 into useful hydrocarbons and for organic synthesis in detail.

3.3.1 CO2 reduction

The production of hydrocarbons and liquid fuels from the electrochemical reduction of CO2 has been extensively investigated because CO2 is the huge and sustainable carbon feedstock and the conversion of CO2 into hydrocarbons and liquid fuels could provide an incentive for the CO2 capture (Gattrell et al., 2007; Hori, 2008). However, efficient and durable electrocatalysts for the electroreduction of CO2 and its derivatives into desirable fuels are not available at the present (Benson et al., 2009; Costentin et al., 2012; Costentin et al., 2013). Many catalysts can reduce the CO2 to CO (Chen et al. 2012; DiMeglio & Rosenthal, 2013; Ebbesen & Mogensen, 2009; Sen et al., 2014; Tornow et al., 2012), but the synthesis of liquid fuels requires that CO should be further reduced using H2O as H+ source. CO is a derivative of CO2, which acts as an intermediate during the production of hydrocarbons.

Since the discovery of Hori et al. in 1985, Cu has been known as a decent catalyst for the electrochemical reduction of CO2/CO to hydrocarbons (Hori et al., 1985, 1986). However, in bulk form, its efficiency and selectivity for the production of higher hydrocarbons and liquid fuels are far too low for commercial applications. A mixture of compounds is produced on the polycrystalline Cu foil in CO saturated aqueous solutions, which at low overpotentials are dominated by H2 and at high overpotentials are dominated by CO and HCO2 and by multicarbon oxygenates and hydrocarbons at very extreme potentials (Hori et al., 1989; Kuhl et al., 2012). Particularly, H2O reduction to H2 outcompetes CO reduction on the Cu electrode at low overpotentials, and at higher overpotentials, gaseous hydrocarbons are the major products of CO2 reduction (DiMeglio & Rosenthal, 2013; Ebbesen & Mogensen, 2009). Hence, there is a high need of a catalyst that can work efficiently at low overpotentials.

NPC has been reported as an exclusive novel material for the electrochemical reduction of CO2 into hydrocarbons with high Faraday efficiency and selectivity (Jia et al., 2014; Li & Kanan, 2012; Li et al., 2014; Manthiram et al., 2014; Sen et al., 2014; Tang et al., 2012). Ponnurangam et al. (2016) recently reported on a poly(4-vinylpyridine)-Cu electrode as a robust catalyst for the electroreduction of CO2. The cathodic current relating to the production of hydrocarbons is significantly improved on NPC during the electroreduction of CO/CO2 at very low overpotentials (Jia et al., 2014). The onset potential for the reduction of CO2 at porous Cu foam was −1.0 V versus Ag/AgCl with the formation of formic acid (HCOOH) initially (Figure 22). The Faraday efficiency of HCOOH rises from 4% to 26% at −1.1 V on porous Cu foam, which is appreciably higher than that for smooth Cu (i.e. <1% at −1.1 V). This value increases to 37% at −1.5 V, which is the highest value of Faraday efficiency reported to date for HCOOH at a Cu electrode (Sen et al., 2014). Furthermore, the production of propylene from CO2 reduction has been observed for the first time on a high surface area nanostructured Cu foam electrode (Figure 22).

Figure 22: 
							Product distribution as a function of applied potential obtained during the electrochemical reduction of CO2. The working electrode was Cu nanofoam electrodeposited for 15 s. Data for the electrochemical reduction of CO2 to formate at a smooth Cu electrode was also included for comparison.
							Reproduced from Sen et al. (2014), with permission from American Chemical Society.
Figure 22:

Product distribution as a function of applied potential obtained during the electrochemical reduction of CO2. The working electrode was Cu nanofoam electrodeposited for 15 s. Data for the electrochemical reduction of CO2 to formate at a smooth Cu electrode was also included for comparison.

Reproduced from Sen et al. (2014), with permission from American Chemical Society.

The main primary reactions occurring during the electrochemical reduction of CO2 at a Cu electrode are given below. A series of steps (1) to (6) are predicted for the electroreduction of CO2 at the Cu electrode (Sen et al., 2014). The asterisk in any reaction indicates either a surface-bound species or a vacant active site.

(1) CO 2 + H + + e HCOO  (F-intermediate pathway)
(2) CO 2 + H + + e COOH  (C-intermediate pathway)
(3) HCOO /HCOOH + H + + e HCOOH
(4) HCOOH HCOOH +
(5) COOH + H + + e CO + H 2 O
(6) CO + H + + e CHO
(7) H + + e + H
(8) H + H + + e H 2 +
(9) H + H H 2 + 2

From the above reaction steps, we can conclude the complete reactions for the formation of hydrocarbons and liquid fuels.

(10) CO 2 + 2H + + 2e CO  +  H 2 O
(11) CO 2 + 2H + + 2e HCOOH
(12) CO 2 + 6H + + 6e CH 3 OH  +  H 2 O
(13) CO 2 + 8H + + 8e H 4 + 2H 2 O
(14) 2CO 2 + 12H + + 12e C 2 H 5 OH + 3H 2 O
(15) 2CO 2 + 12H + + 12e C 2 H 4 + 4H 2 O
(16) 3CO 2 + 18H + + 12e C 3 H 6 + 6H 2 O
(17) 2H + + 2e H 2

The formation of HCOOH, CO, and alcohols during the electroreduction of CO2 has been reported for various Cu electrodes; however, the porous Cu nanofoam is the only catalyst that generates propylene in detectable quantities. Although propylene has been measured at very low yields on porous Cu nanofoam, it has not been observed on any other Cu electrodes reported to date (Sen et al., 2014). The production of ethylene and propylene suggests that the NPC foam can provide both the nanostructured surfaces and cavities that promote the reaction between CO2 and hydrogen to produce the higher hydrocarbons (ethylene, propylene, etc.) during the electroreduction of CO2.

It has been reported by many researchers that that electroreduction of CO2 on Cu is very sensitive to its surface structure (Goncalves et al., 2013; Hori et al., 2002, 2003; Montoya et al., 2013). Hori et al. (2002, 2003) described that the formation of C2H4 occurs preferentially at Cu(100) facets at potentials lower than that for Cu foil, whereas the formation of CH4 took place selectively at Cu(111) at same potentials. The most negative potentials were required for the reduction of CO2 at Cu(110) facets, and as a result, C2 and C3 products were obtained. These investigations indicate that the surface structure of the electrode plays a crucial role in the electroreduction of CO2.

Sen et al. (2014) suggested that the mechanism of CO2 electroreduction was changed by porous Cu foam based on the observation of following points:

  1. Enlarged Faraday efficiency of formate at all potentials

  2. Reduced Faraday efficiency of CO and CH4

  3. Formation of saturated hydrocarbons (e.g. C2H6)

  4. Generation of novel higher hydrocarbons, namely, C3H6.

Insight into the importance of I to III was provided in detail by Norskov’s group (Durand et al., 2011; Peterson et al., 2010). They reported the detailed description of chemical processes occurring during the CO2 electroreduction at the Cu-water interface. Finally, point IV, the generation of novel C3-hydrocarbons (namely, propylene) might be caused by the hierarchical porous structure of Cu foam.

The production of saturated hydrocarbons, such as ethane, and C3 products, such as propylene, on hierarchical porous Cu foam suggested that the mechanism of CO2 electroreduction on Cu nanofoam might be changed, as these products were not observed on smooth Cu surface by previous studies. The generation of the saturated hydrocarbons and C3 products may be caused by the hierarchical porous structure of the Cu foam because the residence time of the various intermediates within these confined spaces of pores may increase, allowing for the generation of products that were not observed on smooth Cu electrodes. The idea that nanoporous electrodes favor the reaction pathways that are different from those observed at smooth electrodes via a confinement mechanism has been studied in detail in the electroreduction of O2 at nanoporous Pt electrodes (Bae et al., 2012; Boo et al., 2004).

Reaction pathways and intermediates: Formate and CO are formed at first before the production of hydrocarbons at higher cathodic potentials. Cu nanoparticles prepared by traditional vapor condensation produce about 96% CO from the reduction of CO2 (Rosen et al., 2011). This might be due to the reason that CO is first produced before the production of alcohols. Although the conversion mechanism of CO2 to liquid fuels is not very clear at present, but it is widely accepted that CO2 is first reduced to CO and then it is further converted to alcohols and other hydrocarbons through multistep electron transfer pathways. Earlier studies have found that CO reduction on Cu results in similar products as obtained from CO2 reduction, proposing that CO was obtained as an intermediate during the electroreduction of CO2 (DeWulf et al., 1989; Hori et al., 1987; Laitar et al., 2005; Sullivan et al., 2012). Further reactions of formate do not produce any quantifiable products (Cook, et al., 1989).

Two voltage-dependent pathways were predicted for the electrochemical reduction of CO2 at a Cu electrode (Durand et al., 2011; Peterson et al., 2010).

  1. The formate (OCHO) or F-intermediate pathway and

  2. The carboxyl (COOH) or C-intermediate pathway

The F-intermediate pathway leads exclusively to formic acid and the C-intermediate pathway leads to formic acid and high-order hydrocarbons. The calculations of the products obtained from both pathways have been performed on (111), (100), and (211) surfaces of Cu. The F-intermediate pathway dominates on (111) and (100) surfaces, because it has the lowest change in free energy of the two pathways (Durand et al., 2011; Schouten et al., 2012). It was proposed that (100) and (200) surfaces of Cu are assumed to be equivalent in density functional theory calculations, and the production of HCOOH via the F-intermediate pathway is expected to be enhanced at Cu nanofoam as 22% more (200) surface was observed comparatively (Sen et al., 2014).

These earlier studies signify that CO is adsorbed on the electrode surface during CO2 reduction and this CO acts as an intermediate during the hydrocarbons production. The second point was well documented in the report of Smith et al. (1997), where saturated solutions of both CO and CO2 showed similar spectra. Due to high reportage of CO during the reduction of CO2, the adsorbed CO reduction was proposed to be the rate-determining step for the entire reaction to hydrocarbon formation (Kim et al., 1988). Therefore, many researches have been carried out on the CO reduction, as this takes a benefit of only a few reaction steps to be considered (Hori et al., 1997; Li et al., 2014; Schouten et al., 2013; Verdaguer-Casadevall et al., 2015; Zhang et al., 2014).

3.3.2 CO reduction

CO reduction at Cu electrode was studied in detail by Hori’s group (Hori et al., 1997). CO reduction reactions were conducted at different pH values using borate buffer or phosphate buffer. Watanabe et al. (1994) evaluated the state of adsorbed CO anion radical on Cu using ab initio calculations. They predicted a little decrease in Cu-C bonds and an uplift of the C-O bonds to about 1.25 Å; hence, they estimated a mostly double-bond character (Watanabe et al., 1993, 1994). When Cu is supplied with only CO in the absence of CO2, higher hydrocarbons and multicarbon oxygenates are produced; however, to compete with H2 evolution, much negative potentials are still required to stimulate CO reduction (Hori et al., 1987, 1997). Energetic efficient electrolysis is impeded by large overpotentials and favor the production of hydrocarbons over multicarbon oxygenates.

Recently, it was discovered that oxide-derived nanocopper has a very high selectivity for CO reduction over H2 evolution than that of polycrystalline Cu foil (Li & Kanan, 2012). It was reported in a current letter to Nature (Li et al., 2014) that CO can be reduced with a very high Faraday efficiency (57%) over oxide-derived Cu at a very low overpotential of −0.3 V versus reversible hydrogen electrode (RHE), which is >0.6 V lower compared to polycrystalline Cu foil. It was proposed that high CO reduction activity was due to the grain boundaries participating in catalysis. The major products obtained from the reduction of CO were ethanol and acetate, which related to 8e and 4e reductions using H2O as H+ ion source (Li et al., 2014).

(18) 2 CO + 7 H 2 O + 8 e CH 3 CH 2 OH + 8 HO E = 0.18  V vs . RHE
(19) 2 CO + 3 H 2 O + 2 e CH 3 CO 2 + 3 HO E = 0.50  V vs . RHE

The activity of CO reduction was measured using steady-state conditions through constant potential electrolysis in CO saturated (1 atm) 0.1 m KOH solution at ambient temperature. Polycrystalline Cu foil showed lower current density and H2 was the only measureable product under these similar conditions. Another study has shown 22% overall Faraday efficiency for the reduction of CO at very negative potentials (about −0.7 V vs. RHE) and its Faraday efficiency for oxygenates was only 7% (Hori et al., 1987). A maximum of 65% Faraday efficiency from CO reduction have been stated for polycrystalline Cu foil at −0.9 V versus RHE, but Faraday efficiency of oxygenates was only 10% (Hori et al., 1997).

CO reduction over oxide-derived Cu exhibited quite higher geometric current densities than that of over Cu nanoparticles because of their roughness factor (Figure 23). Faraday efficiency was ≥94% for H2 evolution at all potentials for Cu nanoparticle electrode, and small remaining current was related to ethanol, acetate, and ethylene formation. On the contrary, a much higher tendency of CO reduction was shown by oxide-derived Cu electrodes. Oxide-derived Cu-1 electrode showed an overall CO reduction efficiency of ~57% at −0.3 V versus RHE, whereas oxide-derived Cu-2 showed 48% Faraday efficiency at −0.4 V versus RHE. With increase in the negative potentials, these values decreased because of catalyst reaching mass-transport-current density for the CO reduction. Oxide-derived Cu electrodes represented higher intrinsic CO reduction activity due to its higher Faraday efficiency compared to Cu nanoparticles (Li et al., 2014).

Figure 23: 
							Faraday efficiencies for CO reduction products ethanol (EtOH), acetate (AcO-), n-propanol (n-PrOH), ethylene (C2H4), and ethane (C2H6) at specific potentials versus RHE.
							Reprinted from Li et al. (2014), with permission from Nature Publishing Group.
Figure 23:

Faraday efficiencies for CO reduction products ethanol (EtOH), acetate (AcO-), n-propanol (n-PrOH), ethylene (C2H4), and ethane (C2H6) at specific potentials versus RHE.

Reprinted from Li et al. (2014), with permission from Nature Publishing Group.

Electrolysis data (Figure 24) give some insights into the CO reduction mechanism on oxide-derived Cu. The absence of C1 products indicates that C-C coupling was very fast during the start of CO reduction (Montoya et al., 2013) or maybe because initial electron transfer was combined to the C-C bond formation between a CO from solution and surface-bound CO (Calle-Vallejo & Koper, 2013). It is estimated that acetate formation results from the attack of HO on a carbonyl-containing intermediate or a surface-bound ketene after the C-C bond formation. This argument was signified from the observation of enlarged acetate formation when electrolysis was executed in 1 m KOH solution (Figure 24).

Figure 24: 
							Faraday efficiency of various products for CO reduction bulk electrolysis on oxide-derived Cu-1 in 1 m KOH solution saturated with 1 atm CO.
							Reprinted from Li et al. (2014), with permission from Nature Publishing Group.
Figure 24:

Faraday efficiency of various products for CO reduction bulk electrolysis on oxide-derived Cu-1 in 1 m KOH solution saturated with 1 atm CO.

Reprinted from Li et al. (2014), with permission from Nature Publishing Group.

Detailed mechanism can be observed vividly from Figure 25. In our previous report (Abbas et al., 2016), we suggested that CO2 is first converted into CO (ads) active species at Cu or possibly at another electrode and binds to the catalyst. This CO (ads) may dimerize in the presence of an alkaline media and form C2O2 (ads). The hydrogenation process of these species would produce CH2CHO (ads). Two pathways originate from this intermediate toward the formation of ethylene and ethanol. (1) The hydrogenation of the CH2 group leads to the formation of acetaldehyde, which is eventually reduced to ethanol. (2) The hydrogenation of C in the carbonyl group leads to the formation of ethylene and O (ads), which quickly forms water.

Figure 25: 
							Schematic of the mechanism for the electroreduction of CO2 into hydrocarbons at NPC electrode.
							Reprinted from Abbas et al. (2016), with permission from Taylor & Francis Group.
Figure 25:

Schematic of the mechanism for the electroreduction of CO2 into hydrocarbons at NPC electrode.

Reprinted from Abbas et al. (2016), with permission from Taylor & Francis Group.

Influence of electrolyte, temperature, and pressure on the CO/CO2 reduction: The Gouy-Chapman theory describes the relation between the Debye length of EDL and electrolyte concentration (Boo et al., 2004). They proposed that with the increase in electrolyte concentration the EDL thickness decreases and at a critical value the EDL becomes thin enough that the electric fields of adjacent pores do not overlap anymore. In this way, the electric field maps the exact shape of the pore and hence provides the extra surface area to participate in electrochemical reactions.

In Figure 26, chronoamperometric measurements show that at a smooth Cu electrode the current density gradually increases from 7 to 31 mA cm−2 (~4.5×) with the increase of electrolyte concentration from 0.1 to 1 m. On the contrary, at NPC foam (60 s), the current density increases sharply from 10 to 82 mA cm−2 (~8×) above a critical concentration of 0.5 m KHCO3 (Sen et al., 2014). Thus, this data shows that nanoscale pores present inside the Cu foam are only accessible at a concentration above 0.5 m KHCO3 (i.e. at this point, the thickness of the double layer is minimum). These 3D small electroactive areas enable the production of C2 and C3 products (e.g. propylene due to increase of residence time of surface intermediates; Sen et al., 2014).

Figure 26: 
							Chronoamperometric measurements of CO2 electroreduction at smooth Cu and NPC foam plotted versus electrolyte concentration.
							Reproduced from Sen et al. (2014), with permission from American Chemical Society.
Figure 26:

Chronoamperometric measurements of CO2 electroreduction at smooth Cu and NPC foam plotted versus electrolyte concentration.

Reproduced from Sen et al. (2014), with permission from American Chemical Society.

The product distribution from CO2 reduction is also influenced by the reaction temperature. It was reported in galvanostatic measurements being carried out over the range of 0°C–40°C that lower temperature results in a variation in current density for formate, ethylene, and methane. Current efficiencies for ethylene formation and hydrogen evolution were decreased and the efficiency of methane was increased, reaching a value of 65% at 0°C (Hori et al., 1986). Some other workers have reported, in addition to the increase in current efficiency for methane, a shift in Tafel slop from 93 mV decade−1 at 22°C to 520 mV decade−1 at 0°C (Kim et al., 1988).

It was confirmed by Kanan’s group (Li et al., 2014) that geometric current densities of CO reduction increase with the increase of CO pressure. The geometric current density was 1.8–2.4-fold higher at 2.4 atm CO compared to 1 atm CO at potentials between −0.3 and −0.5 V (Figure 27). The Faraday efficiency for CO reduction was also significantly improved at potential E<−0.3V under these conditions. These results point out that the practical current densities may be possibly obtained with the more increase in the CO transport to the catalyst.

Figure 27: 
							Comparison of CO reduction in 0.1 m KOH saturated with 1 atm CO versus 2.4 atm CO.
							Reprinted from Li et al. (2014), with permission from Nature Publishing Group.
Figure 27:

Comparison of CO reduction in 0.1 m KOH saturated with 1 atm CO versus 2.4 atm CO.

Reprinted from Li et al. (2014), with permission from Nature Publishing Group.

3.3.3 Organic synthesis using NPC

NPC-catalyzed click reaction: Click reaction has recently emerged as a promising method for organic synthesis (Hu et al., 2013; Markiewicz et al., 2010; Milles et al., 2012; Yamamoto, 2014). The Huisgen [3+2] cycloaddition of terminal alkynes and organic azides, namely, click reaction, has been widely catalyzed by homogeneous Cu(1) salts for organic synthesis. Heterogeneous Cu(1) has also been used to catalyze the click reaction, such as Cu-in-charcoal nanoparticles, immobilized Cu nanoclusters and CuO nanoparticles, Cu3N nanoparticles, and Cu zeolites. Researchers also reported that click reaction also be catalyzed by Cu metal; however, the reaction was very slow, which required high catalytic loading.

Jin et al. (2011) carried out the click reaction using NPC as a catalyst. Click reaction catalyzed by NPC progressed very smoothly and catalyst efficiency was greatly dependent on its nanoporosity (Figure 28). The highest trizole yield was obtained when cat-3 with a ligament size of 40 nm was used. The specific surface area of NPC-cat-3 measured by BET method was 14 m2 g−1 and the turnover frequency (TOF) extended to 0.26/s.

Figure 28: 
							Click reaction dependence on nanoporosity.
							Reproduced from Jin et al. (2011), with permission from Wiley.
Figure 28:

Click reaction dependence on nanoporosity.

Reproduced from Jin et al. (2011), with permission from Wiley.

Yamamoto (2014) carried out click reaction using cat-3 catalyst for a wide range of organic azides and its scope is presented in Figure 29. Triazoles substituted by various functional groups could be synthesized in high chemical yields. Catalyst loading could be reduced to 0.1 mol% under clean conditions, mean without using toluene solvent, and the reaction efficiency was still similar to that of 2 mol% NPC in toluene solvent and TOF was lifted up to 5.2 s−1. Moreover, cat-3 NPC can be reused more than 10 times under conditions presented in Figure 28.

Figure 29: 
							Scope of click reaction using cat-3 NPC.
							Reproduced from Yamamoto (2014), with permission from Elsevier.
Figure 29:

Scope of click reaction using cat-3 NPC.

Reproduced from Yamamoto (2014), with permission from Elsevier.

4 Summary and perspectives

With the extensive advancement in nanotechnology and material sciences, numerous strategies toward the preparation of NPC have been developed, starting from fundamental fabrication techniques such as dealloying. With these robust and reliable strategies, NPC with a wide range of pore size distribution from macroporous to microporous can be efficiently and selectively fabricated and designed for various promising potential electrochemical applications. NPC has been extensively investigated for sensing and catalysis due to its inherent catalytic activities for oxygen reduction reaction, methanol oxidation, CO/CO2 reduction, and so on. In particular, tunable porous structures with enlarged specific surface areas show higher catalytic activity, sensitivity, and selectivity toward electrochemical applications. The large-scale fabrication of multifunctional transducers for high-performance biosensors has become possible due to the good compatibility of NPC with enzymes. The binary NPM systems present high synergetic effect of each constituent to further enhance the catalytic activity. The extensive improvements in the reversible energy storage capability and cycling ability have been achieved due to the large surface area provided by the unique porous structure of NPC and its ability to accommodate the volume changes.

In spite of the abundant success in the synthesis of NPC through a number of strategies and sufficiently accumulative knowledge on its physiochemical properties, various challenges may need to be addressed to examine their full potential. Presently available techniques are preferable to synthesize noble NPMs for various high-performance electrochemical applications. The future inclination is to expand the applications of NPC and as a replacement for noble metals, obviously for cost effectiveness. There is a great need for the development of general methodologies for the preparation of high-performance and highly efficient electrochemical devices using nonnoble NPMs, such as Cu, Fe, Co, and Ni. Alternatively, the fabrication of noble/nonnoble metal composite systems could be very promising for the systematic evaluation of catalytic activity and efficiency of nonnoble metals for noble tasks. Furthermore, an in-depth understanding of the scientific fundamental conditions, such as structural effects in nanorange and the dependence of higher catalytic activities on different reaction kinetics, are highly challenging. Much deeper experimental and theoretical studies are highly needed to effectively design NPC with required pore size and intriguing properties for particular applications. One of the key barriers toward the practical application of NPMs in their limited stability because coarsening usually takes place due to their affinity toward lowering the surface free energy during the reaction processes. The stability of NPMs is very likely to be improved by coating another metal or self-assembly of surfactant monolayer on the surface of NPMs. For electrochemical application, however, the complete removal of surfactant is of high significance to generate enlarged intrinsic catalytic activities.

Due to the highly increasing demand from numerous industry sectors, the development of various fabrication strategies of NPMs and exploration of their-high performance applications will remain a topic of high research interests. Due to the limited resources of fossil fuels and environmental issues, substantial efforts have been devoted to the development of clean energy technologies, such as solar cells, fuel cells, and CO/CO2 conversion into liquid fuels.

To make the CO2 reduction commercially applicable, it is required that highest Faraday efficiency, highest energetic efficiency, and highest current density should be achieved for a single product. Currently, many reports have been presented for each of the CO2 reduction products, exhibiting a high Faraday efficiency, high energetic efficiency, or high current density, but combining all these figures of merit has been a challenge (Hori et al., 1987, 1989, 1997; Kuhl et al., 2012).

The catalyst for the selective production of various products from CO2 reduction has been established, but the catalysts that exhibit low overpotentials (i.e. <0.2 V) combined with high current densities (i.e. >100 mA cm−2) required for practical applications are still missing. These catalysts can be developed by the deep and fundamental studies focusing on reaction mechanism occurring on catalyst during CO2 reduction, an area in which only a few reports are presented (Chandrasekaran & Bockris, 1987; Schouten et al., 2012; Schouten et al., 2013).

The low catalyst activity, selectivity, and stability are few major technological challenges behind the commercialization of CO2 electroreduction into liquid fuels. To deal with these challenges, we propose some research directions for the future:

  1. Improvement in the catalytic activity and stability of electrocatalysts by innovation of new materials: Almost all of the pure metals and their associated compounds have been tried as a catalyst material for CO2 electroreduction, and even some improvements have been achieved in catalytic activity, selectivity, and stability, but still these advances are not enough for commercial applications. Developing new materials, therefore, for required performance should be emphasized. We propose two important types of materials here: (i) composite materials and (ii) nanostructured materials. Composite materials, which are usually synthesized by the combination of different materials, should have different properties and catalytic performance than that of their individual parts because each of the individual parts in composite experiences a synergistic effect on its properties. This effect comes from optimizing particle size, active sites, porosity, specific surface area, electron conductivity, and protection from mechanical and chemical degradation. In this way, the obtained composites may have optimum catalytic activity, selectivity, and stability in CO2 electroreduction. Furthermore, nanostructured materials such as nanoporous materials, nanoaerogels, nanotubes/rods, and nanoplates/sheets, due to their exceptional properties, such as higher specific surface area and 2D and 3D structures, can offer easy pathways for electron/proton, leading to faster reaction kinetics, more efficient electrolyte ions contact, and increased active sites for the catalytic process.

  2. Fundamental understanding of the mechanism and theoretical and experimental modeling: For designing, synthesizing, and optimizing new catalysts to improve catalytic activity, selectivity, and stability, a significant fundamental understanding of the mechanism by theoretical and experimental modeling is necessary. For example, the fundamental understanding of the CO2 electroreduction mechanism and its relationship to the catalyst structure, composition, and active sites, using both theoretical calculations and experimental techniques, are highly needed to develop new catalysts.

  3. Optimizing the electrode/reactor design for commercial applications: For the industrial-scale CO2 electroreduction, the big limitation is the catalyst degradation and slow transfer of CO2 to the electrode surface, both of which might be related to the electrode/reactor design. It is widely accepted that catalyst degradation is related to the material itself but also related to the environment. Optimizing the electrode/reactor design, besides optimizing catalyst stability, to optimize the operating conditions is also highly important for better performance.

In summary, a significant key opportunity exists in the optimization of electrode structure and composition. Based on the study of different researchers, CO2 electroreduction is very sensitive to the structure and composition of microporous/nanoporous layer structure materials. We believe that, with further efforts focused on developing innovative composite materials and nanostructured materials to overcome the challenge of inadequate catalyst activity, selectivity, and stability, the CO2 electroreduction will become commercially applicable in the near future.

The improvements on the sensitivity, selectivity, and stability for electrochemical sensors may considerably benefit from the shape- and size-controlled synthesis of NPMs in the coming days. NPMs with the combination of biocompatibility, intrinsic high electrical conductivity, and flexibility of surface functionalization may offer a good platform for building smart and high-performance nanodevices.

About the authors

Aumber Abbas

Aumber Abbas received his BSc in metallurgy and materials engineering from Bahauddin Zakariya University, Pakistan, in 2012 and Msc in materials science and engineering from Shandong University, China, in 2016. In September 2016, he became a postgraduate researcher at the School of Chemical Engineering and Advanced Materials, Newcastle University, UK. His current research interests focus on the conversion of biowaste into superior electrode materials for energy storage applications.

Saleem Abbas

Saleem Abbas received his BSc in physics from Bahauddin Zakariya University, Multan, Pakistan, in 2009 and MSc in physics from Quiad-e-Azam University, Islamabad, Pakistan, in 2013. In September 2014, he became a doctoral research student at Hebei Key Laboratory of Boron Nitride Micro and Nano Materials, Hebei University of Technology, Tianjin, China. His current research interests focus on the fabrication of single-walled and multiwalled boron nanotubes and their application in energy devices.

Xianli Wang

Xianli Wang received her Bsc from Dalian Polytechnic University of Materials Chemistry in 2012 and her MSc degree from Shandong University, Jinan, China, in 2015. She is currently a PhD student at the University of Georgia, USA. Her research focuses on functional composite materials and applications in electrical, electromagnetic, and luminescent fields (e.g. semiconductor and wave transparent materials).

Acknowledgments

This work was supported by the National Natural Science Foundation of China (NSFC no. 51301096) and the Shandong Provincial Natural Science Foundation (ZR2013EMQ012). The work was also financially supported by the Fundamental Research Funds of Shandong University (no. 2014JC017).

References

Abbas A, Ullah M, Ali Q, Zahid I, Abbas S, Wang XL. An overview of CO2 electroreduction into hydrocarbons and liquid fuels on nanostructured copper catalysts. Green Chem Lett Rev 2016; 9: 166–178.10.1080/17518253.2016.1188162Search in Google Scholar

Aburada T, Fitz-Gerald JM, Scully JR. Synthesis of nanoporous copper by dealloying of Al-Cu-Mg amorphous alloys in acidic solution: the effect of nickel. Corros Sci 2011; 53: 1627–1632.10.1016/j.corsci.2011.01.033Search in Google Scholar

Appleby AJ. Fuel cell handbook, New York, NY; Van Nostrand Reinhold Co. Inc., 1988.Search in Google Scholar

Asghari E, Ashassi-Sorkhabi H, Vahed A, Rezaei-Moghadam B, Charmi GR. The use of a hierarchically platinum-free electrode composed of tin oxide decorated polypyrrole on nanoporous copper in catalysis of methanol electrooxidation. Thin Solid Films 2016; 598: 6–15.10.1016/j.tsf.2015.12.005Search in Google Scholar

Ashassi-Sorkhabi H, Badakhshan PL, Asghari E. Electrodeposition of three dimensional-porous Ni/Ni(OH)2 hierarchical nano composite via etching the Ni/Zn/Ni(OH)2 precursor as a high performance pseudocapacitor. Chem Eng J 2016; 299: 282–291.10.1016/j.cej.2016.04.069Search in Google Scholar

Bae JH, Han J-H, Chung TD. Electrochemistry at nanoporous interfaces: new opportunity for electrocatalysis. Phys Chem Chem Phys 2012; 14: 448–463.10.1039/C1CP22927CSearch in Google Scholar PubMed

Becker HI. Low voltage electrolytic capacitor. Google Patents, 1957.Search in Google Scholar

Ben-Jacob E, Garik P. The formation of patterns in non-equilibrium growth. Nature 1990; 343: 523–530.10.1038/343523a0Search in Google Scholar

Benson EE, Kubiak CP, Sathrum AJ, Smieja JM. Electrocatalytic and homogeneous approaches to conversion of CO2 to liquid fuels. Chem Soc Rev 2009; 38: 89–99.10.1039/B804323JSearch in Google Scholar

Boo H, Park S, Ku B, Kim Y, Park JH, Kim HC, Chung TD. Ionic strength-controlled virtual area of mesoporous platinum electrode. J Am Chem Soc 2004; 126: 4524–4525.10.1021/ja0398316Search in Google Scholar PubMed

Bozzini B, Busson B, Humbert C, Mele C, Tadjeddine A. Electrochemical fabrication of nanoporous gold decorated with manganese oxide nanowires from eutectic urea/choline chloride ionic liquid. Part III-Electrodeposition of Au-Mn: a study based on in situ sum-frequency generation and Raman spectroscopies. Electrochim Acta 2016; 218: 208–215.10.1016/j.electacta.2016.09.077Search in Google Scholar

Braun PV, Wiltzius P. Microporous materials: electrochemically grown photonic crystals. Nature 1999; 402: 603–604.10.1038/45137Search in Google Scholar

Brenna M, Lazaroiu G, Rotaru R, Tironi E. Interconnection of electrical energy storage systems for power quality improvement. PowerTech, 2009 IEEE Bucharest, IEEE, 2009.10.1109/PTC.2009.5281993Search in Google Scholar

Burke A. Ultracapacitors: why, how, and where is the technology. J Power Sources 2000; 91: 37–50.10.1016/S0378-7753(00)00485-7Search in Google Scholar

Calle-Vallejo F, Koper M. Theoretical considerations on the electroreduction of CO to C2 species on Cu(100) electrodes. Angew Chem 2013; 125: 7423–7426.10.1002/ange.201301470Search in Google Scholar

Chandrasekaran K, Bockris LM. In-situ spectroscopic investigation of adsorbed intermediate radicals in electrochemical reactions: CO2 on platinum. Surf Sci 1987; 185: 495–514.10.1016/S0039-6028(87)80173-5Search in Google Scholar

Chang J-K, Hsu S-H, Sun I-W, Tsai W-T. Formation of nanoporous nickel by selective anodic etching of the nobler copper component from electrodeposited nickel-copper alloys. J Phys Chem C 2008; 112: 1371–1376.10.1021/jp0772474Search in Google Scholar

Chen LY, Yu JS, Fujita T, Chen MW. Nanoporous copper with tunable nanoporosity for SERS applications. Adv Funct Mater 2009; 19: 1221–1226.10.1002/adfm.200801239Search in Google Scholar

Chen LY, Fujita T, Ding Y, Chen MW. A three-dimensional gold-decorated nanoporous copper core-shell composite for electrocatalysis and nonenzymatic biosensing. Adv Funct Mater 2010; 20: 2279–2285.10.1002/adfm.201000326Search in Google Scholar

Chen Y, Li CW, Kanan MW. Aqueous CO2 reduction at very low overpotential on oxide-derived Au nanoparticles. J Am Chem Soc 2012; 134: 19969–19972.10.1021/ja309317uSearch in Google Scholar PubMed

Chmiola J, Largeot C, Taberna P-L, Simon P, Gogotsi Y. Monolithic carbide-derived carbon films for micro-supercapacitors. Science 2010; 328: 480–483.10.1126/science.1184126Search in Google Scholar PubMed

Conway B. Electrochemical supercapacitors – scientific fundamentals and technological applications. New York: Plenum Press, 1999.Search in Google Scholar

Cook RL, MacDuff RC, Sammells AF. Evidence for formaldehyde, formic acid, and acetaldehyde as possible intermediates during electrochemical carbon dioxide reduction at copper. J Electrochem Soc 1989; 136: 1982–1984.10.1149/1.2097110Search in Google Scholar

Costentin C, Drouet S, Robert M, Savéant J-M. A local proton source enhances CO2 electroreduction to CO by a molecular Fe catalyst. Science 2012; 338: 90–94.10.1126/science.1224581Search in Google Scholar PubMed

Costentin C, Robert M, Savéant J-M. Catalysis of the electrochemical reduction of carbon dioxide. Chem Soc Rev 2013; 42: 2423–2436.10.1039/C2CS35360ASearch in Google Scholar

Dan Z, Qin F, Sugawara Y, Muto I, Hara N. Fabrication of nanoporous copper by dealloying amorphous binary Ti-Cu alloys in hydrofluoric acid solutions. Intermetallics 2012; 29: 14–20.10.1016/j.intermet.2012.04.016Search in Google Scholar

Dan Z, Qin F, Sugawara Y, Muto I, Hara N. Effects of the initial microstructure of Ti-Cu alloys on final nanoporous copper via dealloying. J Alloys Compd 2013; 557: 166–171.10.1016/j.jallcom.2013.01.010Search in Google Scholar

DeWulf DW, Jin T, Bard AJ. Electrochemical and surface studies of carbon dioxide reduction to methane and ethylene at copper electrodes in aqueous solutions. J Electrochem Soc 1989; 136: 1686–1691.10.1149/1.2096993Search in Google Scholar

DiMeglio JL, Rosenthal J. Selective conversion of CO2 to CO with high efficiency using an inexpensive bismuth-based electrocatalyst. J Am Chem Soc 2013; 135: 8798–8801.10.1021/ja4033549Search in Google Scholar PubMed PubMed Central

Divyashree A, Hegde G. Activated carbon nanospheres derived from bio-waste materials for supercapacitor applications – a review. RSC Adv 2015; 5: 88339–88352.10.1039/C5RA19392CSearch in Google Scholar

Durand WJ, Peterson AA, Studt F, Abild-Pedersen F, Nørskov JK. Structure effects on the energetics of the electrochemical reduction of CO2 by copper surfaces. Surf Sci 2011; 605: 1354–1359.10.1016/j.susc.2011.04.028Search in Google Scholar

Ebbesen SD, Mogensen M. Electrolysis of carbon dioxide in solid oxide electrolysis cells. J Power Sources 2009; 193: 349–358.10.1016/j.jpowsour.2009.02.093Search in Google Scholar

Ebrahimi A, Arab M, Saffari M, Minett AI, Langrish T. Facile fabrication of mesoporous CaO sorbents using simple salt as a pore template in a template-assisted and spray-drying synthesis method. Chem Eng J 2016; 291: 1–11.10.1016/j.cej.2016.01.094Search in Google Scholar

Erlebacher J. An atomistic description of dealloying porosity evolution, the critical potential, and rate-limiting behavior. J Electrochem Soc 2004; 151: C614–C626.10.1149/1.1784820Search in Google Scholar

Erlebacher J, Seshadri R. Hard materials with tunable porosity. MRS Bull 2009; 34: 561–568.10.1557/mrs2009.155Search in Google Scholar

Erlebacher J, Aziz MJ, Karma A, Dimitrov N, Sieradzki K. Evolution of nanoporosity in dealloying. Nature 2001; 410: 450–453.10.1038/35068529Search in Google Scholar PubMed

Feng Q, Liu S, Wang X, Jin G. Nanoporous copper incorporated platinum composites for electrocatalytic reduction of CO2 in ionic liquid BMIMBF 4. Appl Surf Sci 2012; 258: 5005–5009.10.1016/j.apsusc.2012.01.030Search in Google Scholar

Friedrich J, Wainwright M, Young D. Methanol synthesis over Raney copper-zinc catalysts: I. Activities and surface properties of fully extracted catalysts. J Catal 1983; 80: 1–13.10.1016/0021-9517(83)90223-3Search in Google Scholar

Gattrell M, Gupta N, Co A. Electrochemical reduction of CO2 to hydrocarbons to store renewable electrical energy and upgrade biogas. Energy Convers Manage 2007; 48: 1255–1265.10.1016/j.enconman.2006.09.019Search in Google Scholar

Goncalves M, Gomes A, Condeco J, Fernandes T, Pardal T, Sequeira C, Branco J. Electrochemical conversion of CO2 to C2 hydrocarbons using different ex situ copper electrodeposits. Electrochim Acta 2013; 102: 388–392.10.1016/j.electacta.2013.04.015Search in Google Scholar

Hakamada M, Kato N, Mabuchi M. Electrical resistivity of nanoporous gold modified with thiol self-assembled monolayers. 2016; 387:1088–1092.10.1016/j.apsusc.2016.07.059Search in Google Scholar

Hayes J, Hodge A, Biener J, Hamza A, Sieradzki K. Monolithic nanoporous copper by dealloying Mn-Cu. J Mater Res 2006; 21: 2611–2616.10.1557/jmr.2006.0322Search in Google Scholar

Henne R. Solid oxide fuel cells: a challenge for plasma deposition processes. J Therm Spray Technol 2007; 16: 381–403.10.1007/s11666-007-9053-4Search in Google Scholar

Hori Y. Electrochemical CO2 reduction on metal electrodes. Mod Aspects Electrochem 2008; 42: 89–189.10.1007/978-0-387-49489-0_3Search in Google Scholar

Hori Y, Kikuchi K, Suzuki S. Production of CO and CH4 in electrochemical reduction of CO2 at metal electrodes in aqueous hydrogen carbonate solution. Chem Lett 1985; 11: 1695–1698.10.1246/cl.1985.1695Search in Google Scholar

Hori Y, Kikuchi K, Murata A, Suzuki S. Production of methane and ethylene in electrochemical reduction of carbon dioxide at copper electrode in aqueous hydrogencarbonate solution. Chem Lett 1986; 15: 897–898.10.1246/cl.1986.897Search in Google Scholar

Hori Y, Murata A, Takahashi R, Suzuki S. Electroreduction of carbon monoxide to methane and ethylene at a copper electrode in aqueous solutions at ambient temperature and pressure. J Am Chem Soc 1987; 109: 5022–5023.10.1021/ja00250a044Search in Google Scholar

Hori Y, Murata A, Takahashi R. Formation of hydrocarbons in the electrochemical reduction of carbon dioxide at a copper electrode in aqueous solution. J Chem Soc Faraday Trans 1 1989; 85: 2309–2326.10.1039/f19898502309Search in Google Scholar

Hori Y, Takahashi R, Yoshinami Y, Murata A. Electrochemical reduction of CO at a copper electrode. J Phys Chem B 1997; 101: 7075–7081.10.1021/jp970284iSearch in Google Scholar

Hori Y, Takahashi I, Koga O, Hoshi N. Selective formation of C2 compounds from electrochemical reduction of CO2 at a series of copper single crystal electrodes. J Phys Chem B 2002; 106: 15–17.10.1021/jp013478dSearch in Google Scholar

Hori Y, Takahashi I, Koga O, Hoshi N. Electrochemical reduction of carbon dioxide at various series of copper single crystal electrodes. J Mol Catal A Chem 2003; 199: 39–47.10.1016/S1381-1169(03)00016-5Search in Google Scholar

Hou C, Lang X-Y, Han G-F, Li Y-Q, Zhao L, Wen Z, Zhu Y-F, Zhao M, Li J-C, Lian J-S. Integrated solid/nanoporous copper/oxide hybrid bulk electrodes for high-performance lithium-ion batteries. Sci Rep 2013; 3: 2878.10.1038/srep02878Search in Google Scholar PubMed PubMed Central

Howorka S, Siwy Z. Nanopore analytics: sensing of single molecules. Chem Soc Rev 2009; 38: 2360–2384.10.1039/b813796jSearch in Google Scholar PubMed

Hu X, Yan L, Xiao H, Li X, Jing X. Application of microwave-assisted click chemistry in the preparation of functionalized copolymers for drug conjugation. J Appl Polym Sci 2013; 127: 3365–3373.10.1002/app.37662Search in Google Scholar

Huang J-F, Sun I-W. Formation of nanoporous platinum by selective anodic dissolution of PtZn surface alloy in a Lewis acidic zinc chloride-1-ethyl-3-methylimidazolium chloride ionic liquid. Chem Mater 2004; 16: 1829–1831.10.1021/cm030462mSearch in Google Scholar

Jena BK, Raj CR. Enzyme-free amperometric sensing of glucose by using gold nanoparticles. Chem Eur J 2006; 12: 2702–2708.10.1002/chem.200501051Search in Google Scholar PubMed

Jia F, Yu X, Zhang L. Enhanced selectivity for the electrochemical reduction of CO2 to alcohols in aqueous solution with nanostructured Cu-Au alloy as catalyst. J Power Sources 2014; 252: 85–89.10.1016/j.jpowsour.2013.12.002Search in Google Scholar

Jiang H, Jie L, Haoran G, Qinglei W. Influence of cooling rate and addition of lanthanum and cerium on formation of nanoporous copper by chemical dealloying of Cu15Al85 alloy. J Rare Earths 2013; 31: 1119–1124.10.1016/S1002-0721(12)60414-1Search in Google Scholar

Jin T, Yan M, Minato T, Bao M, Yamamoto Y. Nanoporous copper metal catalyst in click chemistry: nanoporosity-dependent activity without supports and bases. Adv Synth Catal 2011; 353: 3095–3100.10.1002/adsc.201100760Search in Google Scholar

Jin T, Yan M, Yamamoto Y. Click chemistry of alkyne-azide cycloaddition using nanostructured copper catalysts. ChemCatChem 2012; 4: 1217–1229.10.1002/cctc.201200193Search in Google Scholar

Khan N, Mariun N, Zaki M, Dinesh L. Transient analysis of pulsed charging in supercapacitors. TENCON 2000. Proc IEEE 2000.Search in Google Scholar

Kim J, Summers D, Frese K. Reduction of CO2 and CO to methane on Cu foil electrodes. J Electroanal Chem Interfacial Electrochem 1988; 245: 223–244.10.1016/0022-0728(88)80071-8Search in Google Scholar

Kotz R, Carlen M. Principles and applications of electrochemical capacitors. Electrochim Acta 2000; 45: 2483.10.1016/S0013-4686(00)00354-6Search in Google Scholar

Krongelb S, Romankiw LT, Tornello JA. Electrochemical process for advanced package fabrication. IBM J Res Dev 1998; 42: 575–586.10.1147/rd.425.0575Search in Google Scholar

Kuhl KP, Cave ER, Abram DN, Jaramillo TF. New insights into the electrochemical reduction of carbon dioxide on metallic copper surfaces. Energy Environ Sci 2012; 5: 7050–7059.10.1039/c2ee21234jSearch in Google Scholar

Kulinowski KM, Jiang P, Vaswani H, Colvin VL. Porous metals from colloidal templates. Adv Mater 2000; 12: 833–838.10.1002/(SICI)1521-4095(200006)12:11<833::AID-ADMA833>3.0.CO;2-XSearch in Google Scholar

Laitar DS, Müller P, Sadighi JP. Efficient homogeneous catalysis in the reduction of CO2 to CO. J Am Chem Soc 2005; 127: 17196–17197.10.1021/ja0566679Search in Google Scholar

Lang X, Hirata A, Fujita T, Chen M. Nanoporous metal/oxide hybrid electrodes for electrochemical supercapacitors. Nat Nanotechnol 2011; 6: 232–236.10.1038/nnano.2011.13Search in Google Scholar

Li CW, Kanan MW. CO2 reduction at low overpotential on Cu electrodes resulting from the reduction of thick Cu2O films. J Am Chem Soc 2012; 134: 7231–7234.10.1021/ja3010978Search in Google Scholar

Li AP, Müller F, Birner A, Nielsch K, Gösele U. Fabrication and microstructuring of hexagonally ordered two-dimensional nanopore arrays in anodic alumina. Adv Mater 1999; 11: 483–487.10.1002/(SICI)1521-4095(199904)11:6<483::AID-ADMA483>3.3.CO;2-9Search in Google Scholar

Li Y, Meng G, Zhang L, Phillipp F. Ordered semiconductor ZnO nanowire arrays and their photoluminescence properties. Appl Phys Lett 2000; 76: 2011–2013.10.1063/1.126238Search in Google Scholar

Li Y, Jia W-Z, Song Y-Y, Xia X-H. Superhydrophobicity of 3D porous copper films prepared using the hydrogen bubble dynamic template. Chem Mater 2007a; 19: 5758–5764.10.1021/cm071738jSearch in Google Scholar

Li Y, Song Y-Y, Yang C, Xia X-H. Hydrogen bubble dynamic template synthesis of porous gold for nonenzymatic electrochemical detection of glucose. Electrochem Commun 2007b; 9: 981–988.10.1016/j.elecom.2006.11.035Search in Google Scholar

Li CW, Ciston J, Kanan MW. Electroreduction of carbon monoxide to liquid fuel on oxide-derived nanocrystalline copper. Nature 2014; 508: 504–507.10.1038/nature13249Search in Google Scholar

Li J, Jiang H, Yu N, Xu C, Geng H. Fabrication and characterization of bulk nanoporous copper by dealloying Al-Cu alloy slices. Corros Sci 2015; 90: 216–222.10.1016/j.corsci.2014.10.014Search in Google Scholar

Linden D. Handbook of batteries and fuel cells. New York: McGraw-Hill Book Co., 1984: 1.Search in Google Scholar

Liu Z, Bando Y. A novel method for preparing copper nanorods and nanowires. Adv Mater 2003; 15: 303–305.10.1002/adma.200390073Search in Google Scholar

Liu Z, Searson PC. Single nanoporous gold nanowire sensors. J Phys Chem B 2006; 110: 4318–4322.10.1021/jp056940tSearch in Google Scholar PubMed

Liu Z, Du J, Qiu C, Huang L, Ma H, Shen D, Ding Y. Electrochemical sensor for detection of p-nitrophenol based on nanoporous gold. Electrochem Commun 2009; 11: 1365–1368.10.1016/j.elecom.2009.05.004Search in Google Scholar

Liu W, Zhang S, Li N, Zheng J, Xing Y. Influence of phase constituent and proportion in initial Al-Cu alloys on formation of monolithic nanoporous copper through chemical dealloying in an alkaline solution. Corros Sci 2011; 53: 809–814.10.1016/j.corsci.2010.11.017Search in Google Scholar

Liu D, Yang Z, Wang P, Li F, Wang D, He D. Preparation of 3D nanoporous copper-supported cuprous oxide for high-performance lithium ion battery anodes. Nanoscale 2013; 5: 1917–1921.10.1039/c2nr33383jSearch in Google Scholar PubMed

Liu W, Chen L, Yan J, Li N, Shi S, Zhang S. Dealloying solution dependence of fabrication, microstructure and porosity of hierarchical structured nanoporous copper ribbons. Corros Sci 2015; 94: 114–121.10.1016/j.corsci.2015.01.043Search in Google Scholar

Lokhande C, Dubal D, Joo O-S. Metal oxide thin film based supercapacitors. Curr Appl Phys 2011; 11: 255–270.10.1016/j.cap.2010.12.001Search in Google Scholar

Luo X, Li R, Liu Z, Huang L, Shi M, Xu T, Zhang T. Three-dimensional nanoporous copper with high surface area by dealloying Mg-Cu-Y metallic glasses. Mater Lett 2012; 76: 96–99.10.1016/j.matlet.2012.02.028Search in Google Scholar

Luo X, Li R, Huang L, Zhang T. Nucleation and growth of nanoporous copper ligaments during electrochemical dealloying of Mg-based metallic glasses. Corros Sci 2013; 67: 100–108.10.1016/j.corsci.2012.10.010Search in Google Scholar

Mahon PJ, Drummond CJ. Supercapacitors-nanostructured materials and nanoscale processes contributing to the next mobile generation. Aust J Chem 2001; 54: 473–476.10.1071/CH01142Search in Google Scholar

Manthiram K, Beberwyck BJ, Alivisatos AP. Enhanced electrochemical methanation of carbon dioxide with a dispersible nanoscale copper catalyst. J Am Chem Soc 2014; 136: 13319–13325.10.1021/ja5065284Search in Google Scholar PubMed

Mao R, Liang S, Wang X, Yang Q, Han B. Effect of preparation conditions on morphology and thermal stability of nanoporous copper. Corros Sci 2012; 60: 231–237.10.1016/j.corsci.2012.03.032Search in Google Scholar

Markiewicz JT, Wiest O, Helquist P. Synthesis of primary aryl amines through a copper-assisted aromatic substitution reaction with sodium azide. J Org Chem 2010; 75: 4887–4890.10.1021/jo101002pSearch in Google Scholar PubMed

Marsden WL, Wainwright MS, Friedrich JB. Zinc-promoted Raney copper catalysts for methanol synthesis. Ind Eng Chem Prod Res Dev 1980; 19: 551–556.10.1021/i360076a014Search in Google Scholar

Masłyk M, Borysiewicz MA, Wzorek M, Wojciechowski T, Kwoka M, Kamińska E. Influence of absolute argon and oxygen flow values at a constant ratio on the growth of Zn/ZnO nanostructures obtained by DC reactive magnetron sputtering. Appl Surf Sci 2016; 389: 287–293.10.1016/j.apsusc.2016.07.098Search in Google Scholar

Masuda H, Satoh M. Fabrication of gold nanodot array using anodic porous alumina as an evaporation mask. Jpn J Appl Phys 1996; 35: L126.10.1143/JJAP.35.L126Search in Google Scholar

McCue I, Benn E, Gaskey B, Erlebacher J. Dealloying and dealloyed materials. Annu Rev Mater Res 2016a; 46: 263–286.10.1146/annurev-matsci-070115-031739Search in Google Scholar

McCue I, Ryan S, Hemker K, Xu XD, Li N, Chen MW, Erlebacher J. Size effects in the mechanical properties of bulk bicontinuous Ta/Cu nanocomposites made by liquid metal dealloying. Adv Eng Mater 2016b; 18: 46–50.10.1002/adem.201500219Search in Google Scholar

Meng F, Yan X, Liu J, Gu J, Zou Z. Nanoporous gold as non-enzymatic sensor for hydrogen peroxide. Electrochim Acta 2011; 56: 4657–4662.10.1016/j.electacta.2011.02.105Search in Google Scholar

Metzger RM, Konovalov VV, Sun M, Xu T, Zangari G, Xu B, Benakli M, Doyle W. Magnetic nanowires in hexagonally ordered pores of alumina. Magnet IEEE Trans 2000; 36: 30–35.10.1109/20.824421Search in Google Scholar

Mijangos C, Hernandez R, Martin J. A review on the progress of polymer nanostructures with modulated morphologies and properties, using nanoporous AAO templates. Prog Polym Sci 2016; 54: 148–182.10.1016/j.progpolymsci.2015.10.003Search in Google Scholar

Miller JR, Burke AF. Electrochemical capacitors: challenges and opportunities for real-world applications. Electrochem Soc Interface 2008; 17: 53.10.1149/2.F08081IFSearch in Google Scholar

Milles S, Tyagi S, Banterle N, Koehler C, VanDelinder V, Plass T, Neal AP, Lemke EA. Click strategies for single-molecule protein fluorescence. J Am Chem Soc 2012; 134: 5187–5195.10.1021/ja210587qSearch in Google Scholar PubMed

Molares MT, Brötz J, Buschmann V, Dobrev D, Neumann R, Scholz R, Schuchert I, Trautmann C, Vetter J. Etched heavy ion tracks in polycarbonate as template for copper nanowires. Nucl Instrum Methods Phys Res Sect B 2001; 185: 192–197.10.1016/S0168-583X(01)00755-8Search in Google Scholar

Montoya JH, Peterson AA, Nørskov JK. Insights into C-C Coupling in CO2 electroreduction on copper electrodes. ChemCatChem 2013; 5: 737–742.10.1002/cctc.201200564Search in Google Scholar

Moosavifard SE, El-Kady MF, Rahmanifar MS, Kaner RB, Mousavi MF. Designing 3D highly ordered nanoporous CuO electrodes for high-performance asymmetric supercapacitors. ACS Appl Mater Interf 2015; 7: 4851–4860.10.1021/am508816tSearch in Google Scholar PubMed

Nguyen T, Boudard M, Carmezim MJ, Montemor MF. Hydrogen bubbling-induced micro/nano porous MnO2 films prepared by electrodeposition for pseudocapacitor electrodes. Electrochim Acta 2016; 202: 166–174.10.1016/j.electacta.2016.03.204Search in Google Scholar

Nishino A. Capacitors: operating principles, current market and technical trends. J Power Sources 1996; 60: 137–147.10.1016/S0378-7753(96)80003-6Search in Google Scholar

Ogasawara T, Débart A, Holzapfel M, Novák P, Bruce PG. Rechargeable Li2O2 electrode for lithium batteries. J Am Chem Soc 2006; 128: 1390–1393.10.1021/ja056811qSearch in Google Scholar PubMed

Pena DJ, Razavi B, Smith PA, Mbindyo JK, Natan MJ, Mayer TS, Mallouk TE, Keating CD. Electrochemical synthesis of multi-material nanowires as building blocks for functional nanostructures. MRS Proc, Cambridge, England: Cambridge Univ Press, 2000.10.1557/PROC-636-D4.6.1Search in Google Scholar

Peterson AA, Abild-Pedersen F, Studt F, Rossmeisl J, Nørskov JK. How copper catalyzes the electroreduction of carbon dioxide into hydrocarbon fuels. Energy Environ Sci 2010; 3: 1311–1315.10.1039/c0ee00071jSearch in Google Scholar

Ponnurangam S, Yun CM, Chernyshova IV. Robust electroreduction of CO2 at a poly(4-vinylpyridine)-copper electrode. ChemElectroChem 2016; 3: 74–82.10.1002/celc.201500421Search in Google Scholar

Possin GE. A method for forming very small diameter wires. Rev Sci Instrum 1970; 41: 772–774.10.1063/1.1684640Search in Google Scholar

Prasad KPS, Dhawale DS, Sivakumar T, Aldeyab SS, Zaidi JSM, Ariga K, Vinu A. Fabrication and textural characterization of nanoporous carbon electrodes embedded with CuO nanoparticles for supercapacitors. Sci Technol Adv Mater 2016; 12: 044602.10.1088/1468-6996/12/4/044602Search in Google Scholar PubMed PubMed Central

Qi Z, Zhang Z, Jia H, Qu Y, Liu G, Bian X. Alloy composition dependence of formation of porous Ni prepared by rapid solidification and chemical dealloying. J Alloys Compd 2009a; 472: 71–78.10.1016/j.jallcom.2008.04.017Search in Google Scholar

Qi Z, Zhao C, Wang X, Lin J, Shao W, Zhang Z, Bian X. Formation and characterization of monolithic nanoporous copper by chemical dealloying of Al-Cu alloys. J Phys Chem C 2009b; 113: 6694–6698.10.1021/jp810742zSearch in Google Scholar

Qiao Y, Li CM. Nanostructured catalysts in fuel cells. J Mater Chem 2011; 21: 4027–4036.10.1039/C0JM02871ASearch in Google Scholar

Raney M. Method of producing finely-divided nickel, Google Patents, 1927.Search in Google Scholar

Riveros G, Gomez H, Cortes A, Marotti R, Dalchiele E. Crystallographically-oriented single-crystalline copper nanowire arrays electrochemically grown into nanoporous anodic alumina templates. Appl Phys A 2005; 81: 17–24.10.1007/s00339-004-3112-1Search in Google Scholar

Rosen BA, Salehi-Khojin A, Thorson MR, Zhu W, Whipple DT, Kenis PJ, Masel RI. Ionic liquid-mediated selective conversion of CO2 to CO at low overpotentials. Science 2011; 334: 643–644.10.1126/science.1209786Search in Google Scholar PubMed

Samuel E. Fuel cell system. Google Patents, 1964.Search in Google Scholar

Schetty R. Minimization of tin whisker formation for lead-free electronics finishing. Circuit World 2001; 27: 17–20.10.1108/03056120110367131Search in Google Scholar

Schouten KJP, Qin Z, Gallent EPr, Koper MT. Two pathways for the formation of ethylene in CO reduction on single-crystal copper electrodes. J Am Chem Soc 2012; 134: 9864–9867.10.1021/ja302668nSearch in Google Scholar PubMed

Schouten KJP, Pérez Gallent E, Koper MT. Structure sensitivity of the electrochemical reduction of carbon monoxide on copper single crystals. ACS Catal 2013; 3: 1292–1295.10.1021/cs4002404Search in Google Scholar

Schuchert I, Molares MT, Dobrev D, Vetter J, Neumann R, Martin M. Electrochemical copper deposition in etched ion track membranes. Experimental results and a qualitative kinetic model. J Electrochem Soc 2003; 150: C189–C194.10.1149/1.1554722Search in Google Scholar

Sen S, Liu D, Palmore GTR. Electrochemical reduction of CO2 at copper nanofoams. ACS Catal 2014; 4: 3091–3095.10.1021/cs500522gSearch in Google Scholar

Shin H-C, Liu M. Copper foam structures with highly porous nanostructured walls. Chem Mater 2004; 16: 5460–5464.10.1021/cm048887bSearch in Google Scholar

Shin HC, Liu M. Three-dimensional porous copper-tin alloy electrodes for rechargeable lithium batteries. Adv Funct Mater 2005; 15: 582–586.10.1002/adfm.200305165Search in Google Scholar

Shin HC, Dong J, Liu M. Nanoporous structures prepared by an electrochemical deposition process. Adv Mater 2003; 15: 1610–1614.10.1002/adma.200305160Search in Google Scholar

Shin HC, Dong J, Liu M. Porous tin oxides prepared using an anodic oxidation process. Adv Mater 2004; 16: 237–240.10.1002/adma.200305660Search in Google Scholar

Shukla A, Martha S. Electrochemical power sources. Resonance 2001; 6: 52–63.10.1007/BF02835270Search in Google Scholar

Simon P, Burke A. Nanostructured carbons: double-layer capacitance and more. Electrochem Soc Interface 2008; 17: 38.10.1149/2.F05081IFSearch in Google Scholar

Simon P, Gogotsi Y. Materials for electrochemical capacitors. Nat Mater 2008; 7: 845–854.10.1038/nmat2297Search in Google Scholar PubMed

Smith B, Irish D, Kedzierzawski P, Augustynski J. A Surface enhanced Raman scattering study of the intermediate and poisoning species formed during the electrochemical reduction of CO2 on copper. J Electrochem Soc 1997; 144: 4288–4296.10.1149/1.1838180Search in Google Scholar

Smith A, Tran T, Wainwright M. Kinetics and mechanism of the preparation of Raney® copper. J Appl Electrochem 1999; 29: 1085–1094.10.1023/A:1003637410133Search in Google Scholar

Srivastava R, Prathap MA, Kore R. Morphologically controlled synthesis of copper oxides and their catalytic applications in the synthesis of propargylamine and oxidative degradation of methylene blue. Colloids Surf A 2011; 392: 271–282.10.1016/j.colsurfa.2011.10.004Search in Google Scholar

Stein A, Li F, Denny NR. Morphological control in colloidal crystal templating of inverse opals, hierarchical structures, and shaped particles. Chem Mater 2007; 20: 649–666.10.1021/cm702107nSearch in Google Scholar

Su J, Wang H, Jiang M, Li L. Bias deposition of nanoporous Cu thin films. Mater Lett 2013; 102: 72–75.10.1016/j.matlet.2013.03.114Search in Google Scholar

Sulka G, Stroobants S, Moshchalkov V, Borghs G, Celis J-P. Nanostructuring of aluminium and synthesis of porous aluminium membranes by anodising. Bull Cercle Etud Métaux 2002; 17: P1.Search in Google Scholar

Sullivan BP, Krist K, Guard H. Electrochemical and electrocatalytic reactions of carbon dioxide. New York: Elsevier, 2012.Search in Google Scholar

Sun L, Chien C-L, Searson PC. Fabrication of nanoporous nickel by electrochemical dealloying. Chem Mater 2004; 16: 3125–3129.10.1021/cm0497881Search in Google Scholar

Tang W, Peterson AA, Varela AS, Jovanov ZP, Bech L, Durand WJ, Dahl S, Nørskov JK, Chorkendorff I. The importance of surface morphology in controlling the selectivity of polycrystalline copper for CO2 electroreduction. Phys Chem Chem Phys 2012; 14: 76–81.10.1039/C1CP22700ASearch in Google Scholar

Tang B, Zhou M, Zhou R, Bai PF, Zhou GF. Well-ordered nanoporous copper fabricated by dealloying Cu-Mn and its characterizations. Key Eng Mater 2016; 693: 662–668.10.4028/www.scientific.net/KEM.693.662Search in Google Scholar

Thurn-Albrecht T, Schotter J, Kästle G, Emley N, Shibauchi T, Krusin-Elbaum L, Guarini K, Black C, Tuominen M, Russell T. Ultrahigh-density nanowire arrays grown in self-assembled diblock copolymer templates. Science 2000; 290: 2126–2129.10.1126/science.290.5499.2126Search in Google Scholar PubMed

Tong X, Zhao G, Liu M, Cao T, Liu L, Li P. Fabrication and high electrocatalytic activity of three-dimensional porous nanosheet Pt/boron-doped diamond hybrid film. J Phys Chem C 2009; 113: 13787–13792.10.1021/jp9029503Search in Google Scholar

Tornow CE, Thorson MR, Ma S, Gewirth AA, Kenis PJ. Nitrogen-based catalysts for the electrochemical reduction of CO2 to CO. J Am Chem Soc 2012; 134: 19520–19523.10.1021/ja308217wSearch in Google Scholar PubMed

Velev OD, Tessier P, Lenhoff A, Kaler E. Materials: a class of porous metallic nanostructures. Nature 1999; 401: 548–548.10.1038/44065Search in Google Scholar

Verdaguer-Casadevall A, Li CW, Johansson TP, Scott SB, McKeown JT, Kumar M, Stephens IE, Kanan MW, Chorkendorff I. Probing the active surface sites for CO reduction on oxide-derived copper electrocatalysts. J Am Chem Soc 2015; 137: 9808–9811.10.1021/jacs.5b06227Search in Google Scholar PubMed

Wang X, Qi Z, Zhao C, Wang W, Zhang Z. Influence of alloy composition and dealloying solution on the formation and microstructure of monolithic nanoporous silver through chemical dealloying of Al-Ag alloys. J Phys Chem C 2009; 113: 13139–13150.10.1021/jp902490uSearch in Google Scholar

Wang G, Zhang L, Zhang J. A review of electrode materials for electrochemical supercapacitors. Chem Soc Rev 2012; 41: 797–828.10.1039/C1CS15060JSearch in Google Scholar

Wang G, He X, Wang L, Gu A, Huang Y, Fang B, Geng B, Zhang X. Non-enzymatic electrochemical sensing of glucose. Microchim Acta 2013; 180: 161–186.10.1007/s00604-012-0923-1Search in Google Scholar

Wang R, Qi JQ, Sui YW, Chang Y, He YZ, Wei FX, Meng QK, Sun Z, Zhao YL. Fabrication of nanosheets Co3O4 by oxidation-assisted dealloying method for high capacity supercapacitors. Mater Lett 2016; 184: 181–184.10.1016/j.matlet.2016.08.049Search in Google Scholar

Wanunu M, Dadosh T, Ray V, Jin J, McReynolds L, Drndić M. Rapid electronic detection of probe-specific microRNAs using thin nanopore sensors. Nat Nanotechnol 2010; 5: 807–814.10.1038/nnano.2010.202Search in Google Scholar PubMed

Watanabe K, Nagashima U, Hosoya H. An ab initio study of interactions of carbon monoxide and metal electrodes. Chem Phys Lett 1993; 209: 109–110.10.1016/0009-2614(93)87210-TSearch in Google Scholar

Watanabe K, Nagashima U, Hosoya H. An ab initio study of adsorbed carbon monoxide on a metal electrode by cluster model. Appl Surf Sci 1994; 75: 121–124.10.1016/0169-4332(94)90147-3Search in Google Scholar

Winter M, Brodd RJ. What are batteries, fuel cells, and supercapacitors? Chem Rev 2004; 104: 4245–4270.10.1021/cr020730kSearch in Google Scholar PubMed

Wittstock A, Biener J, Bäumer M. Nanoporous gold: a new material for catalytic and sensor applications. Phys Chem Chem Phys 2010a; 12: 12919–12930.10.1039/c0cp00757aSearch in Google Scholar PubMed

Wittstock A, Zielasek V, Biener J, Friend C, Bäumer M. Nanoporous gold catalysts for selective gas-phase oxidative coupling of methanol at low temperature. Science 2010b; 327: 319–322.10.1126/science.1183591Search in Google Scholar PubMed

Yamamoto Y. Perspectives on organic synthesis using nanoporous metal skeleton catalysts. Tetrahedron 2014; 70: 2305–2317.10.1016/j.tet.2013.09.065Search in Google Scholar

Yamauchi Y, Kuroda K. Rational design of mesoporous metals and related nanomaterials by a soft-template approach. Chem Asian J 2008; 3: 664–676.10.1002/asia.200700350Search in Google Scholar PubMed

Yamauchi Y, Suzuki N, Radhakrishnan L, Wang L. Breakthrough and future: nanoscale controls of compositions, morphologies, and mesochannel orientations toward advanced mesoporous materials. Chem Rec 2009; 9: 321–339.10.1002/tcr.200900022Search in Google Scholar PubMed

Zen J-M, Kumar AS, Chung C-R. A glucose biosensor employing a stable artificial peroxidase based on ruthenium purple anchored cinder. Anal Chem 2003; 75: 2703–2709.10.1021/ac020542uSearch in Google Scholar PubMed

Zhang J, Li CM. Nanoporous metals: fabrication strategies and advanced electrochemical applications in catalysis, sensing and energy systems. Chem Soc Rev 2012; 41: 7016–7031.10.1039/c2cs35210aSearch in Google Scholar PubMed

Zhang LL, Zhao X. Carbon-based materials as supercapacitor electrodes. Chem Soc Rev 2009; 38: 2520–2531.10.1039/b813846jSearch in Google Scholar PubMed

Zhang M, Lenhert S, Wang M, Chi L, Lu N, Fuchs H, Ming N-B. Regular arrays of copper wires formed by template-assisted electrodeposition. Adv Mater 2004; 16: 409–413.10.1002/adma.200305577Search in Google Scholar

Zhang Q, Wang X, Qi Z, Wang Y, Zhang Z. A benign route to fabricate nanoporous gold through electrochemical dealloying of Al-Au alloys in a neutral solution. Electrochim Acta 2009a; 54: 6190–6198.10.1016/j.electacta.2009.05.089Search in Google Scholar

Zhang Z, Wang Y, Qi Z, Zhang W, Qin J, Frenzel J. Generalized fabrication of nanoporous metals (Au, Pd, Pt, Ag, and Cu) through chemical dealloying. J Phys Chem C 2009b; 113: 12629–12636.10.1021/jp811445aSearch in Google Scholar

Zhang S, Xing Y, Jiang T, Du Z, Li F, He L, Liu W. A three-dimensional tin-coated nanoporous copper for lithium-ion battery anodes. J Power Sources 2011a; 196: 6915–6919.10.1016/j.jpowsour.2010.12.021Search in Google Scholar

Zhang X, Shi W, Zhu J, Kharistal DJ, Zhao W, Lalia BS, Hng HH, Yan Q. High-power and high-energy-density flexible pseudocapacitor electrodes made from porous CuO nanobelts and single-walled carbon nanotubes. ACS Nano 2011b; 5: 2013–2019.10.1021/nn1030719Search in Google Scholar PubMed

Zhang Y-J, Sethuraman V, Michalsky R, Peterson AA. Competition between CO2 reduction and H2 evolution on transition-metal electrocatalysts. ACS Catal 2014; 4: 3742–3748.10.1021/cs5012298Search in Google Scholar

Zhao X, Su F, Yan Q, Guo W, Bao XY, Lv L, Zhou Z. Templating methods for preparation of porous structures. J Mater Chem 2006; 16: 637–648.10.1039/B513060CSearch in Google Scholar

Zhao C, Qi Z, Wang X, Zhang Z. Fabrication and characterization of monolithic nanoporous copper through chemical dealloying of Mg-Cu alloys. Corros Sci 2009; 51: 2120–2125.10.1016/j.corsci.2009.05.043Search in Google Scholar

Zhao C, Wang X, Qi Z, Ji H, Zhang Z. On the electrochemical dealloying of Mg-Cu alloys in a NaCl aqueous solution. Corros Sci 2010; 52: 3962–3972.10.1016/j.corsci.2010.08.005Search in Google Scholar

Zheng XT, Li CM. Single cell analysis at the nanoscale. Chem Soc Rev 2012; 41: 2061–2071.10.1039/C1CS15265CSearch in Google Scholar

Zheng J, Huang J, Jow T. The limitations of energy density for electrochemical capacitors. J Electrochem Soc 1997; 144: 2026–2031.10.1149/1.1837738Search in Google Scholar

Received: 2016-04-23
Accepted: 2016-10-11
Published Online: 2016-11-22
Published in Print: 2016-12-01

©2016 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 18.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/corrrev-2016-0023/html
Scroll to top button