Chirality-Discriminated Conductivity of Metal–Amino Acid

Aug 24, 2016 - ... Normal University, Zhangzhou 363000, People's Republic of China ... Yuhuan Sun , Chuanqi Zhao , Nan Gao , Jinsong Ren , Xiaogang Qu...
3 downloads 0 Views 4MB Size
Chirality-Discriminated Conductivity of Metal− Amino Acid Biocoordination Polymer Nanowires Jianzhong Zheng,†,§ Yijin Wu,†,§ Ke Deng,† Meng He,† Liangcan He,† Jing Cao,† Xugang Zhang,† Yaling Liu,*,† Shunxing Li,*,§ and Zhiyong Tang*,† †

CAS Key Laboratory of Nanosystem and Hierarchical Fabrication, CAS Center for Excellence in Nanoscience, National Center for Nanoscience and Technology, Beijing 100190, People’s Republic of China § Fujian Province Key Laboratory of Modern Analytical Science and Separation Technology, Minnan Normal University, Zhangzhou 363000, People’s Republic of China S Supporting Information *

ABSTRACT: Biocoordination polymer (BCP) nanowires are successfully constructed through self-assembly of chiral cysteine amino acids and Cd cations in solution. The varied chirality of cysteine is explored to demonstrate the difference of BCP nanowires in both morphology and structure. More interestingly and surprisingly, the electrical property measurement reveals that, although all Cd(II)/cysteine BCP nanowires behave as semiconductors, the conductivity of the Cd(II)/DL-cysteine nanowires is 4 times higher than that of the Cd(II)/L-cysteine or Cd(II)/Dcysteine ones. The origin of such chirality-discriminated characteristics registered in BCP nanowires is further elucidated by theoretical calculation. These findings demonstrate that the morphology, structure, and property of BCP nanostructures could be tuned by the chirality of the bridging ligands, which will shed light on the comprehension of chirality transcription as well as construction of chirality-regulated functional materials. KEYWORDS: biocoordination polymer, self-assembly, nanowire, chirality, conductivity

O

CPs formed with biomolecules as the bridging ligands, namely, biocoordination polymers (BCPs). Notably, in comparison with conventional CPs, BCPs are of more scientific and technical interest because they not only inherit unique biological characteristics including chirality, biocompatibility, and biodegradability but also represent an excellent platform for understanding many biorelated processes such as chirality transcription from molecule to macroworld.21−28 Unfortunately, until now, several BCP nanostructures have been fabricated but seldom has a property investigation been introduced.26−28 The absence of systematic study likely lies in the fact that most biomolecules have multiple liable metal binding nodes,21,27−31 which increases the difficulty and complexity in construction of nanoscale BCPs with welldefined structures. In this article, cysteine (H2SR, R = −CH2CH(NH2)COO¯), one of the typical amino acids with three functional groups

ver the past decades, molecular materials and devices have gained much attention and rapid development due to their potential applications in next-generation electronic or photonic devices.1−9 Particularly, increasing demand on device miniaturization and efficiency requires us to freely and accurately manipulate the electronic structure of these molecular materials at the nanoscale, which presents significant challenges because most used molecules need πconjugated structural features and their selection is limited.10,11 Alternatively, thanks to their infinite structure and function tailorability that originates from the abundant building blocks, diversiform coordination polymers (CPs) fabricated by metal ion connectors and organic bridging ligands could become the emerging candidate materials.12−15 Indeed, CPs have been found to possess considerable electrical conductivity if the ligand π* or σ* orbitals and the metal dπ orbitals are designed to be overlapped.16 Nevertheless, design and property study of CPs at the nanoscale are still in its infancy; for instance, to our knowledge, only a few works have reported the carrier transport property in nanoshaped CPs,17−20 and none is aimed at the © 2016 American Chemical Society

Received: June 9, 2016 Accepted: August 24, 2016 Published: August 24, 2016 8564

DOI: 10.1021/acsnano.6b03833 ACS Nano 2016, 10, 8564−8570

Article

www.acsnano.org

Article

ACS Nano

Figure 1. Scanning electron microscopy images of (a) Cd(II)/L-cysteine, (b) Cd(II)/D-cysteine, and (c) Cd(II)/DL-cysteine BCPs produced after reaction for 180 h, 180 h, and 60 days, respectively. (d) Powder X-ray diffraction patterns of all three Cd(II)/cysteine BCPs.

Table 1. Summary of Key Crystallographic Parameters for All Three Cd(II)/Cysteine BCPs Synthesized empirical formula molecular weight crystal structure lattice parameters calculated from XRD data

lattice parameters optimized by DFT calculation

Cd(II)/L-cysteine

Cd(II)/D-cysteine

Cd(II)/DL-cysteine

C3H5CdNO2S 231.54 orthorhombic 5.6509 9.5187 9.5731 509.477 5.4715 9.3255 9.4395 481.646

C3H5CdNO2S 231.54 orthorhombic 5.6488 9.5211 9.5777 509.477 5.4715 9.3255 9.4395 481.646

C3H5CdNO2S 231.54 orthorhombic 5.6673 9.5404 9.5990 516.809 5.5050 9.4885 9.5715 499.960

a/Å b/Å c/Å V/Å3 a/Å b/Å c/Å V/Å3

(−SH, −COOH, and −NH2), is selected as the bridging ligands for preparation of chiral BCP nanowires with Cd(II) ions. Impressively, both experiment and simulation results reveal that the chirality of the cysteine building blocks can be conferred to the Cd(II)/cysteine BCP nanowires, resulting in a large influence on their morphology, structure, and electrical property.

bundle-like structure (called nanopetals), in which each petal is a nanowire with a length of ∼4 μm and a width of ∼40 nm (Figure 1c). The structures of all three products are thoroughly investigated by a series of characterization techniques, and several important conclusions are drawn as follows. First, both energy-dispersive X-ray (EDX) elemental analysis and inductively coupled plasma mass spectrometry (ICP-MS) indicate that the Cd/S atomic ratio in all products is around 1:1 (Figure S1 and Table S1), so the molecular formula of asprepared BCPs should be assigned to {CdSR}n. Also, a similar 1:1 ratio of Cd and cysteine in the BCPs is achieved based on the thermogravimetric analysis (TGA) (Figure S2; the rapid weight loss of ∼36% is assigned to loss of R in the Cd(II)/ cysteine BCPs). Second, disappearance of the S−H stretching peak at ∼2526 cm−1 in the Fourier transform infrared (FTIR) spectra reveals formation of Cd−S coordination bonds between Cd(II) ions and cysteine molecules in BCPs. Meanwhile, the obvious red shift of both the NH2 asymmetric bending at 1624

RESULTS AND DISCUSSION In brief, synthesis of Cd(II)/cysteine BCP nanowires follows the below recipe: 0.025 M cysteine and 0.01 M Cd(ClO4)2· 6H2O were mixed in deionized water under vigorous stirring with a final pH value of around 8.0 and then heated to 37 °C (human body temperature). After different reaction times, the Cd(II)/cysteine BCP assemblies were formed (Figure 1a−c). Interestingly, for both the Cd(II)/L-cysteine and Cd(II)/Dcysteine systems, rather uniform nanowires more than 10 μm in length and about 50 nm in width are produced (Figure 1a,b), whereas the obtained Cd(II)/DL-cysteine BCPs show a petal 8565

DOI: 10.1021/acsnano.6b03833 ACS Nano 2016, 10, 8564−8570

Article

ACS Nano cm−1 and the COO asymmetric stretching at 1744 cm−1 of cysteine denotes associated interactions between {CdSR}n in three-dimensional (3D) lattices for all of the BCPs (Figure S3). Third, according to the powder X-ray diffraction (XRD) patterns given in Figure 1d, one can see that all three BCP nanowires belong to the orthorhombic crystal system; however, the Cd(II)/L-cysteine and Cd(II)/D-cysteine BCPs share the same lattice parameters, slightly different from those of the Cd(II)/DL-cysteine BCPs (Table 1). Note that the analogous orthorhombic structure has been reported in the bulk compound of Cd(L-cysteinate).32 The selected area electron diffraction patterns (Figure S4) further demonstrate that all BCP nanostructures, regardless of nanowires or nanopetals, grow along the [100] direction corresponding to the a direction within the Cd(II)/cysteine BCP lattice deduced from the XRD data. It is worth pointing out that the chirality of cysteine molecules can be well transferred to the BCP products. As presented in Figure 2, the circular dichroism (CD) responses of

the nanowires more than 10 μm in length and about 50 nm in width are obtained (Figure S6f). For comparison, the growth of the Cd(II)/cysteine BCP nanostructures using racemic cysteine as building blocks is much slower (Figure S7). Under the same reaction conditions, the olivary-shaped microrods are formed after 18 days (Figure S7a). Then, nanofibrils are unselectively generated on the entire surface of the microrods and further slowly grown longer (Figure S7b,c). After 50 days of reaction, the petal bundle-like nanostructures are obtained (Figure S7d). Evidently, the chirality of the cysteine amino acid is the key to influence the growth kinetics of the Cd(II)/cysteine BCPs. The transport properties of all three Cd(II)/cysteine BCPs are investigated via deposition onto the Cr/Au gap electrode arrays, which are fabricated on the Si/SiO2 substrates by photolithography with an electrode gap width of 2 μm. It should be noticed that the Cd(II)/cysteine BCPs are dispersed onto the gap electrodes with the controlled density, so that the devices based on single nanowires can be obtained in order to avoid the influence of disorder and grain boundary on the electrical properties. Furthermore, for the Cd(II)/DL-cysteine BCPs, the petal bundle-like structures are broken into individual nanowires by ultrasonication (as shown in Figure S8, ultrasonication does not change the structure and electrical property of the Cd(II)/DL-cysteine BCPs) before dispersion onto the gap electrodes in order to facilitate comparing the difference from the structure rather than the morphology. Figure 3a,c,e illustrates the typical current−voltage (I−V) characteristics and the corresponding SEM images of the devices fabricated based on single nanowires of all three Cd(II)/cysteine BCPs. Unexpectedly, the nanowires fabricated with chiral or racemic cysteine display distinct electrical transport properties, though all of these devices behave as semiconductors. Quantitative comparison of nanowire conductivity is performed according to the equation Rtotal = Rcontact + RBCP = Rcontact + ρL/A, where Rtotal is the total resistance, Rcontact is the combined contact resistance associated with the Cr/Au electrodes in two contact terminals, RBCP is the channel resistance of BCP single nanowires, and ρ, L, and A are the resistivity, the length across the Cr/Au gap electrodes, and the cross-sectional area of BCP single nanowires, respectively (part S3 in the Supporting Information).33 Figure 3b,d,f presents the corresponding statistical results by plotting Rtotal against L/A under the voltage of 10 V for more than 10 devices based on all three Cd(II)/cysteine BCP single nanowires, in which Rcontact and ρ can be extracted from the intercept and slope, respectively, of the linear regression line. The results show that Rcontact is (7.10 ± 6.35) × 1010 Ω, (5.66 ± 3.56) × 1010 Ω, and (2.07 ± 1.86) × 1010 Ω for the devices based on single Cd/ L-cysteine, Cd/D-cysteine, and Cd/DL-cysteine BCP nanowires, respectively. One may notice that all the Rcontact values are 1−2 orders of magnitude lower than the corresponding Rtotal. This result illustrates that the device resistance is dominated by the Cd(II)/cysteine BCP single nanowires; namely, the effect of Rcontact on overall device performance is little and can be ignored. Thereafter, the resistivity (ρ) is calculated to be 2107.9 ± 121.2 Ω·m, 2110.6 ± 80.4 Ω·m, and 528.9 ± 63.4 Ω·m for single Cd(II)/L-cysteine, Cd(II)/D-cysteine, and Cd(II)/DLcysteine BCP nanowires, respectively. Significantly, the conductivity of the Cd(II)/DL-cysteine BCPs is 4 times higher than that of both the Cd(II)/L-cysteine or the Cd(II)/Dcysteine BCPs. Such differences in the electrical transport property should originate from the varied band gap between three BCPs. As discerned from the UV−vis absorption spectra,

Figure 2. CD spectra of Cd(II)/L-cysteine (blue curve), Cd(II)/Dcysteine (black curve), and Cd(II)/DL-cysteine (red curve) BCP products.

opposite symmetry centered at 236 nm are distinguished for Cd(II)/L-cysteine and Cd(II)/D-cysteine BCP nanowires, whereas the Cd(II)/DL-cysteine BCP nanopetals exhibit no chiral signal in CD spectroscopy. How can the BCP nanostructures of different morphologies be generated when chiral or racemic cysteine molecules are mixed with Cd(II) ions? To answer this question, scanning electron microscopy (SEM) imaging is used to investigate the intermediate products during self-assembly process of these three BCPs. It is found that the morphological evolution process is analogous for both Cd(II)/L-cysteine and Cd(II)/Dcysteine BCPs (Figures S5 and S6; the formation rate of the Cd(II)/L-cysteine BCP nanowires is slightly faster than that of the Cd(II)/D-cysteine BCP ones). Hence, we take the Cd(II)/ D-cysteine BCP nanowires as an example to describe their formation processes. As shown in Figure S6a, the olivaryshaped microrods with a width of ∼1 μm and a length of ∼4 μm are formed after 42 h reaction at 37 °C. Soon afterward, many nanofibrils are epitaxially grown at both ends of the microrods due to the high surface energy (Figure S6b). Subsequently, the nanofibrils are further evolved into longer nanowires accompanied by the gradual reduction of the width of the middle microrods (Figure S6c). As the reaction time is prolonged to 72 h, the nanowires start to be split off from the “mother” microrods (Figure S6d) and continue being grown longer with the reaction time extension (Figure S6e). Finally, 8566

DOI: 10.1021/acsnano.6b03833 ACS Nano 2016, 10, 8564−8570

Article

ACS Nano

Figure 3. (a,c,e) I−V curves and corresponding SEM images (insets) of devices based on (a) Cd(II)/L-cysteine, (c) Cd(II)/D-cysteine, and (e) Cd(II)/DL-cysteine BCP single nanowires. Notably, in each curve, error bars are presented only at points with applied voltages of ±5, ±10, ±15, and ±20 V. (b,d,f) Device resistance (Rtotal) as a function of L/A under the voltage of 10 V for more than 10 devices based on (b) Cd(II)/L-cysteine, (d) Cd(II)/D-cysteine, and (f) Cd(II)/DL-cysteine BCP single nanowires, in which the red lines represent a linear fit.

the band gap is estimated to be 4.27 eV for both the Cd(II)/Lcysteine and Cd(II)/D-cysteine BCPs, while the value is 2.89 eV for the Cd(II)/DL-cysteine BCP (Figure S9). These results clearly demonstrate that introduction of racemic cysteine into the BCP nanostructures lowers the band gap and enhances the conductivity. The last but not the least question is why the chirality of the cysteine building blocks determines the growth and the electrical transport property of the Cd(II)/cysteine BCPs? Density functional theory (DFT) calculation is then performed to understand the interaction between Cd(II) ions and cysteine molecules of different chirality.34,35 On the basis of the lattice parameters of all three Cd(II)/cysteine BCPs calculated from the XRD data (Table 1), the corresponding lattices as well as the geometry of the BCPs in the lattices could be optimized. Obviously, the refined lattice parameters are in good agreement with the experimental ones (Table 1). Figure 4 summarizes the DFT-optimized cells and structures of all three Cd(II)/cysteine BCPs. It is clear that there are four Cd-cysteine units in each BCP cell, and every two Cd-cysteine units form one chain along the a direction. In each Cd-cysteine chain, Cd and S atoms present a one-dimensional (1D) ladder arrangement via bridging of Cd centers by the thiol moiety of cysteine. These 1D units are further held together in a regular fashion due to coordination of the carboxylate group of cysteine to Cd centers in two neighboring ladders via two oxygens. The coordination of Cd atoms with the amine, carboxylate, and thiolate groups of

cysteine constitutes the 3D lattices of all three Cd(II)/cysteine BCPs. The key difference among three Cd(II)/cysteine BCP cells is that in the BCP cells with the enantiomeric form (L- or D-) of cysteine, both of the chains constitute the same Cdcysteine (L- or D-) units; meanwhile, in the BCP cells with racemic cysteine as building blocks, the two chains are composed of one Cd(II)/L-cysteine chain and one Cd(II)/Dcysteine chain (Figure 4c). According to the DFT-optimized structures, the interaction energy closely associated with formation of the 3D lattices of BCP products may be calculated. As demonstrated in Table 2, both the Cd(II)/L-cysteine and Cd(II)/D-cysteine BCP lattices possess the same interaction energy, including the interaction between the units within one chain, the interchain interaction, and even the total interaction between all units within one cell. This result discloses that spontaneous formation of these two BCPs needs to overcome the same energy barrier. As a comparison, the interaction energy between the units within one chain (either the Cd(II)/L-cysteine chain or the Cd(II)/Dcysteine chain) in the Cd(II)/DL-cysteine BCP lattice is slightly higher than that of Cd(II)/L-cysteine and Cd(II)/D-cysteine BCP lattices, demonstrating that the assembly units of homochirality tend to form a chain within the Cd(II)/DLcysteine BCP lattice. However, both the interchain interaction energy and the total interaction energy between all units within one cell in the Cd(II)/DL-cysteine BCP lattice are considerably smaller than those of the Cd(II)/L-cysteine and Cd(II)/D8567

DOI: 10.1021/acsnano.6b03833 ACS Nano 2016, 10, 8564−8570

Article

ACS Nano

Figure 4. DFT-optimized cell and structure of (a) Cd(II)/L-cysteine, (b) Cd(II)/D-cysteine, and (c) Cd(II)/DL-cysteine BCPs (Cd, pink; S, yellow; O, red; N, blue; C, cyan; H, gray). The symbols, Cd/L- and Cd/D-, in (c) are denoted as Cd(II)/L-cysteine chain and Cd(II)/Dcysteine chain, respectively.

Table 2. Comparison of Interaction Energy within All Three Cd(II)/Cysteine BCP Lattices interaction energy (eV)

Cd(II)/L-cysteine

Cd(II)/D-cysteine

Cd(II)/DL-cysteine

interaction between the units within one chain interchain interaction total interaction between all units within one cell

−10.552 −10.550 −7.192 −28.294

−10.552 −10.550 −7.192 −28.294

−10.708 (L-cysteine chain) −10.874 (D-cysteine chain) −5.012 −26.594

the Cd−S−Cd backbone of the chain, we thus deduce that stronger interaction between the Cd−S−Cd in the Cd(II)/DLcysteine BCP is responsible for its better conductivity. To verify this conjecture, band structures and partial density of states (PDOS) of all three Cd(II)/cysteine BCPs are further computed by DFT calculation (Figure S10). The band gap is estimated to be 4.027 eV for both the Cd(II)/L-cysteine and Cd(II)/D-cysteine BCPs and 2.993 eV for the Cd(II)/DLcysteine BCP, which are in good agreement with the experimental data (Figure S9). We also notice that in PDOS, both the valence band and conductance band (near the Fermi energy) mainly arise from the p electrons, so the electrons mainly contributing to the conductivity should be the p electrons.

cysteine BCP lattices. The decreased interaction energy originates from the incomplete coordination interaction between the Cd-cysteine units within the neighboring two chains in the Cd(II)/DL-cysteine BCP cell because there is chiral mismatch of the neighboring cysteine building blocks. The above calculation result well explains why the growth rate of the Cd(II)/cysteine BCPs with racemic cysteine as building blocks is evidently slower than that of the Cd(II)/cysteine BCPs with the enantiomeric form (L- or D-) of cysteine as the building blocks (Figure S7 vs Figures S5 and S6). The origin of chirality-discriminated conductivity of the Cd(II)/cysteine BCPs can be also elucidated by DFT calculation. Seen from the TEM and XRD data (Figure 1d and Figure S4), it is deduced that all three Cd(II)/cysteine BCP nanostructures grow along the [100] direction, just corresponding to the direction of the 1D Cd−S ladder arrangement, as shown in Figure 4. Therefore, in order to distinguish the conductivity difference in all three Cd(II)/ cysteine BCPs, the bond length of Cd−S is calculated. It is found that in the Cd(II)/DL-cysteine BCP lattice, the average bond length of Cd−S is 2.607 Å, which is shorter than that of the Cd(II)/L-cysteine and Cd(II)/D-cysteine BCPs (2.653 Å). Considering that the electrons are mainly transformed through

CONCLUSIONS In conclusion, we constructed a series of BCP nanostructures composed of cysteine with different chiralities and Cd(II) cations by a self-assembly technique. Coupling the chirality of amino acid molecules into coordination frameworks allows us to tune the morphology, structure, and property of the Cd(II)/ cysteine BCP assemblies. Furthermore, both the chirality transcription from cysteine to BCP lattices and the chirality8568

DOI: 10.1021/acsnano.6b03833 ACS Nano 2016, 10, 8564−8570

Article

ACS Nano

ASSOCIATED CONTENT

discriminated conductivity of the Cd(II)/cysteine BCPs are elucidated by experimental observation and theoretical analysis. This work is significant for understanding many biorelated selfassembly processes as well as opens the avenue for creation of the chirality-regulated functional materials with applications in molecular electronics.

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsnano.6b03833. Materials and characterization, additional experimental data, and the detailed calculation process of the contact resistance and the resistivity (PDF)

METHOD AUTHOR INFORMATION

Synthesis of Cd(II)/Cysteine BCPs. The synthesis procedure of all three Cd(II)/cysteine BCPs was same. Taking the Cd(II)/Lcysteine BCPs as an example, the reaction conditions were as follows: the reaction solution was prepared in a vial by dissolving 0.025 mol/L L-cysteine and 0.010 mol/L Cd(ClO4)2·6H2O in deionized water (15 mL) under vigorous stirring with a final pH value of around 8.0 (adjusted by 2 mol/L NaOH carefully). Then, the vial was transferred immediately into an oven with a fixed temperature at 37 °C (human body temperature) until white precipitates were produced. Finally, the products were centrifuged and washed with additional deionized water several times. Device Fabrication. Devices based on all three Cd(II)/cysteine BCP single nanowires were fabricated using the following two steps. The first step involved preparation of the Cr/Au gap electrode arrays with the electrode gap width of 2 μm on Si/SiO2 substrates using the classical photolithography method. The specific steps were as follows: (1) cleaning Si/SiO2 substrates with a hot solution of concentrated sulfuric acid and hydrogen peroxide (sulfuric acid/hydrogen peroxide = 2:1), pure acetone, pure ethanol, pure isopropyl alcohol, and pure deionized water step by step; (2) drying Si/SiO2 substrates with a N2 gun and then baking them at 150 °C for 10 min on a hot plate (dehydration); (3) spin-coating hexamethyldisilazane (HMDS) primer at 7500 rpm for 35 s on the surface of Si/SiO2 substrates and then baking Si/SiO2/HMDS at 120 °C for 60 s on a hot plate; (4) spincoating the S1813 photoresist at 5000 rpm for 35 s on the surface of Si/SiO2/HMDS and then baking the Si/SiO2/HMDS/S1813 at 100 °C for 60 s on a hot plate; (5) being exposed under mask for 4 s at 19 mW/cm2; (6) being developed in undiluted MIF 300 developer for 30 s and then baking at 110 °C for 3 min on a hot plate; (7) depositing Cr (10 nm) and Au (20 nm) in turn by vacuum deposition; (8) liftingoff the photoresist with acetone. The second step was deposition of well-dispersed Cd(II)/cysteine BCP single nanowires onto the Cr/Au gap electrode arrays preprepared by photolithography. It should be noted that all three Cd(II)/cysteine BCP nanowires were dispersed via ultrasonic treatment, and their deposition density onto the Cr/Au gap electrode arrays could be precisely controlled by adjusting the ultrasonic power and time. Computational Details. On the basis of the lattice parameters calculated from the XRD data, the corresponding lattices as well as the geometry of all the BCPs in the unit cell were first optimized. Then, using the refined lattice parameters, we performed the DFT calculations by employing the DMol3 code.34 The periodic boundary conditions were used to describe the 3D periodic structure in this work. To describe exchange and correlation, we applied Perdew and Wang parametrization35 of the local exchange correlation energy in local spin density approximation. The all-electron spin-unrestricted Kohn−Sham wave functions were expanded in a local atomic orbital basis. For the large system, the numerical basis set was applied. All calculations were all-electron ones and were performed with the medium mesh. In the self-consistent field procedure, the convergence criterion was 10−5 au on the energy and electron density. Combined with the experimental data, we optimized the unit lattice parameters and the geometry of all three BCPs in the unit cell. When the energy and density convergence criterion were reached, the optimized parameters and the interaction energy were obtained. Furthermore, the band structures and partial density of states were analyzed for these BCPs, and the Brillouin zone of the unit cells was sampled by a 5 × 3 × 3 k-point mesh.

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. *E-mail: [email protected]. Notes

The authors declare no competing financial interest.

ACKNOWLEDGMENTS The authors acknowledge financial support from National Key Basic Research Program of China (2014CB931801 to Z.T.; 2012CB933001 to K.D.), National Natural Science Foundation of China (21475029 and 91427302 to Z.T.; 21371038 to Y.L.; 21475055, 21175115, and 21507054 to S.L.), Instrument Developing Project of the Chinese Academy of Sciences (YZ201311 to Z.T.; YZ201318 to K.D.), CAS-CSIRO Cooperative Research Program (GJHZ1503 to Z.T.), “Strategic Priority Research Program” of Chinese Academy of Sciences (XDA09040100 to Z.T.; XDA09030308 to Y.L.), and Youth Innocation Promotion Association of the Chinese Academy of Sciences (Y.L.). REFERENCES (1) Hipps, K. W. It’s All about Contacts. Science 2001, 294, 536−537. (2) Whitesides, G. M.; Mathias, J. P.; Seto, C. T. Molecular SelfAssembly and Nanochemistry: A Chemical Strategy for the Synthesis of Nanostructures. Science 1991, 254, 1312−1319. (3) Stupp, S. I.; LeBonheur, V.; Walker, K.; Li, L. S.; Huggins, K. E.; Keser, M.; Amstutz, A. Supramolecular Materials: Self-Organized Nanostructures. Science 1997, 276, 384−389. (4) Jia, C. C.; Guo, X. F. Molecule-Electrode Interfaces in Molecular Electronic Devices. Chem. Soc. Rev. 2013, 42, 5642−5660. (5) Wang, C. Y.; Zhang, J.; Long, G. K.; Aratani, N.; Yamada, H.; Zhao, Y.; Zhang, Q. C. Synthesis, Structure, and Air-stable N-type Field-Effect Transistor Behaviors of Functionalized Octaazanonacene8,19-dione. Angew. Chem., Int. Ed. 2015, 54, 6292−6296. (6) Yao, H. B.; Fang, H. Y.; Wang, X. H.; Yu, S. H. Hierarchical Assembly of Micro-/Nano-Building Blocks: Bio-inspired Rigid Structural Functional Materials. Chem. Soc. Rev. 2011, 40, 3764−3785. (7) Shieh, F. K.; Wang, S. C.; Yen, C. I.; Wu, C. C.; Dutta, S.; Chou, L. Y.; Morabito, J. V.; Hu, P.; Hsu, M. H.; Wu, K. C. W.; et al. Imparting Functionality to Biocatalysts via Embedding Enzymes into Nanoporous Materials by a de Novo Approach: Size-Selective Sheltering of Catalase in Metal-Organic Framework Microcrystals. J. Am. Chem. Soc. 2015, 137, 4276−4279. (8) Huang, M. H.; Schilde, U.; Kumke, M.; Antonietti, M.; Cölfen, H. Polymer-Induced Self-Assembly of Small Organic Molecules into Ultralong Microbelts with Electronic Conductivity. J. Am. Chem. Soc. 2010, 132, 3700−3707. (9) Hu, Y. X.; Sun, Y. G. A Generic Approach for the Synthesis of Dimer Nanoclusters and Asymmetric Nanoassemblies. J. Am. Chem. Soc. 2013, 135, 2213−2221. (10) Moriuchi, T.; Hirao, T. Design and Redox Function of Conjugated Complexes with Polyanilines or Quinonediimines. Acc. Chem. Res. 2012, 45, 347−360. (11) Hasegawa, M.; Iyoda, M. Conducting Supramolecular Nanofibers and Nanorods. Chem. Soc. Rev. 2010, 39, 2420−2427. 8569

DOI: 10.1021/acsnano.6b03833 ACS Nano 2016, 10, 8564−8570

Article

ACS Nano

lattices Based on an Amino Acid Ligand. Inorg. Chem. 2008, 47, 9390− 9399. (33) Cao, Q.; Han, S.-J.; Tulevski, G. S.; Franklin, A. D.; Haensch, W. Evaluation of Field-Effect Mobility and Contact Resistance of Transistors that Use Solution-Processed Single-Walled Carbon Nanotubes. ACS Nano 2012, 6, 6471−6477. (34) Delley, B. From Molecules to Solids with the DMol3 Approach. J. Chem. Phys. 2000, 113, 7756−7764. (35) Perdew, J. P.; Wang, Y. Accurate and Simple Analytic Representation of the Electron-Gas Correlation Energy. Phys. Rev. B: Condens. Matter Mater. Phys. 1992, 45, 13244−13249.

(12) Falcaro, P.; Ricco, R.; Doherty, C. M.; Liang, K.; Hill, A. J.; Styles, M. J. MOF Positioning Technology and Device Fabrication. Chem. Soc. Rev. 2014, 43, 5513−5560. (13) Givaja, G.; Amo-Ochoa, P.; Gómez-García, C. J.; Zamora, F. Electrical Conductive Coordination Polymers. Chem. Soc. Rev. 2012, 41, 115−147. (14) Campbell, M. G.; Sheberla, D.; Liu, S. F.; Swager, T. M.; Dincã, M. Cu3(hexaiminotriphenylene)2: An Electrically Conductive 2D Metal-Organic Framework for Chemiresistive Sensing. Angew. Chem., Int. Ed. 2015, 54, 4349−4352. (15) Talin, A. A.; Centrone, A.; Ford, A. C.; Foster, M. E.; Stavila, V.; Haney, P.; Kinney, R. A.; Szalai, V.; El Gabaly, F.; Yoon, H. P.; et al. Tunable Electrical Conductivity in Metal-Organic Framework ThinFilm Devices. Science 2014, 343, 66−69. (16) Hendon, C. H.; Tiana, D.; Walsh, A. Conductive Metal-Organic Frameworks and Networks: Fact or Fantasy? Phys. Chem. Chem. Phys. 2012, 14, 13120−13132. (17) Low, K.; Roy, V. A. L.; Chui, S. S.; Chan, S. L.; Che, C. Highly Conducting Two-Dimensional Copper(I) 4-Hydroxythiophenolate Network. Chem. Commun. 2010, 46, 7328−7330. (18) Welte, L.; Calzolari, A.; Di Felice, R.; Zamora, F.; GómezHerrero, J. Highly Conductive Self-Assembled Nanoribbons of Coordination Polymers. Nat. Nanotechnol. 2010, 5, 110−115. (19) Liu, Y. L.; Tang, Z. Y. Nanoscale Biocoordination Polymers: Novel Materials from an Old Topic. Chem. - Eur. J. 2012, 18, 1030− 1037. (20) Okubo, T.; Tanaka, N.; Kim, K. H.; Yone, H.; Maekawa, M.; Kuroda-Sowa, T. Magnetic and Conducting Properties of New HalideBridged Mixed-Valence CuI−CuII 1D Coordination Polymers Including a Hexamethylene Dithiocarbamate Ligand. Inorg. Chem. 2010, 49, 3700−3702. (21) An, J.; Geib, S. J.; Rosi, N. L. Cation-Triggered Drug Release from a Porous Zinc-Adeninate Metal-Organic Framework. J. Am. Chem. Soc. 2009, 131, 8376−8377. (22) Rabone, J.; Yue, Y. F.; Chong, S. Y.; Stylianou, K. C.; Bacsa, J.; Bradshaw, D.; Darling, G. R.; Berry, N. G.; Khimyak, Y. Z.; Ganin, A. Y.; et al. An Adaptable Peptide-Based Porous Material. Science 2010, 329, 1053−1057. (23) Liu, Y.; Ma, W.; Liu, W.; Li, C.; Liu, Y. L.; Jiang, X. Y.; Tang, Z. Y. Silver(I)-Glutathione Biocoordination Polymer Hydrogel: Effectibe Antibacterial Activity and Improved Cytocompatibility. J. Mater. Chem. 2011, 21, 19214−19218. (24) Crassous, J. Transfer of Chirality from Ligands to Metal Centers: Recent Examples. Chem. Commun. 2012, 48, 9687−9695. (25) Marino, N.; Armentano, D.; Pardo, E.; Vallejo, J.; Neve, F.; Di Donna, L.; De Munno, G. Homochiral Self-Assembly of Biocoordination Polymers: Anion-Triggered Helicity and Absolute Configuration Inversion. Chem. Sci. 2015, 6, 4300−4305. (26) Imaz, I.; Rubio-Martínez, M.; Saletra, W. J.; Amabilino, D. B.; Maspoch, D. Amino Acid Based Metal-Organic Nanofibers. J. Am. Chem. Soc. 2009, 131, 18222−18223. (27) Li, C.; Deng, K.; Tang, Z. Y.; Jiang, L. Twisted Metal-Amino Acid Nanobelts: Chirality Transcription from Molecules to Frameworks. J. Am. Chem. Soc. 2010, 132, 8202−8209. (28) Imaz, I.; Rubio-Martínez, M.; An, J.; Solé-Font, I.; Rosi, N. L.; Maspoch, D. Metal-Biomolecule Frameworks (MBioFs). Chem. Commun. 2011, 47, 7287−7302. (29) Vaidhyanathan, R.; Bradshaw, D.; Rebilly, J.; Barrio, J.; Gould, J.; Berry, N.; Rosseinsky, M. A Family of Nanoporous Materials Based on an Amino Acid Backbone. Angew. Chem., Int. Ed. 2006, 45, 6495− 6499. (30) Lee, H. Y.; Kampf, J. W.; Park, K. S.; Marsh, E. N. G. Covalent Metal-Peptide Framework Compounds That Extend in One and Two Dimensions. Cryst. Growth Des. 2008, 8, 296−303. (31) Mantion, A.; Massüger, L.; Rabu, P.; Palivan, C.; McCusker, L. B.; Taubert, A. Metal-Peptide Frameworks (MPFs): “Bioinspired” Metal Organic Frameworks. J. Am. Chem. Soc. 2008, 130, 2517−2526. (32) Rebilly, J. N.; Gardner, P. W.; Darling, G. R.; Bacsa, J.; Rosseinsky, M. J. Chiral II-VI Semiconductor Nanostructure Super8570

DOI: 10.1021/acsnano.6b03833 ACS Nano 2016, 10, 8564−8570