view article online chemcomm...halogen bonding8 and metal-ligand coordination interactions,9 offers...

6
rsc.li/chemcomm ChemComm Chemical Communications rsc.li/chemcomm ISSN 1359-7345 COMMUNICATION S. J. Connon, M. O. Senge et al. Conformational control of nonplanar free base porphyrins: towards bifunctional catalysts of tunable basicity Volume 54 Number 1 4 January 2018 Pages 1-112 ChemComm Chemical Communications This is an Accepted Manuscript, which has been through the Royal Society of Chemistry peer review process and has been accepted for publication. Accepted Manuscripts are published online shortly after acceptance, before technical editing, formatting and proof reading. Using this free service, authors can make their results available to the community, in citable form, before we publish the edited article. We will replace this Accepted Manuscript with the edited and formatted Advance Article as soon as it is available. You can find more information about Accepted Manuscripts in the Information for Authors. Please note that technical editing may introduce minor changes to the text and/or graphics, which may alter content. The journal’s standard Terms & Conditions and the Ethical guidelines still apply. In no event shall the Royal Society of Chemistry be held responsible for any errors or omissions in this Accepted Manuscript or any consequences arising from the use of any information it contains. Accepted Manuscript View Article Online View Journal This article can be cited before page numbers have been issued, to do this please use: Q. Liang, Y. Yu, G. Feng, Y. Shen, L. Yang and S. Lei, Chem. Commun., 2020, DOI: 10.1039/D0CC05216G.

Upload: others

Post on 09-Oct-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: View Article Online ChemComm...halogen bonding8 and metal-ligand coordination interactions,9 offers opportunity to achieve highly ordered nanoarchitectures. While a great variety of

rsc.li/chemcomm

ChemCommChemical Communications

rsc.li/chemcomm

ISSN 1359-7345

COMMUNICATIONS. J. Connon, M. O. Senge et al. Conformational control of nonplanar free base porphyrins: towards bifunctional catalysts of tunable basicity

Volume 54Number 14 January 2018Pages 1-112

ChemCommChemical Communications

This is an Accepted Manuscript, which has been through the Royal Society of Chemistry peer review process and has been accepted for publication.

Accepted Manuscripts are published online shortly after acceptance, before technical editing, formatting and proof reading. Using this free service, authors can make their results available to the community, in citable form, before we publish the edited article. We will replace this Accepted Manuscript with the edited and formatted Advance Article as soon as it is available.

You can find more information about Accepted Manuscripts in the Information for Authors.

Please note that technical editing may introduce minor changes to the text and/or graphics, which may alter content. The journal’s standard Terms & Conditions and the Ethical guidelines still apply. In no event shall the Royal Society of Chemistry be held responsible for any errors or omissions in this Accepted Manuscript or any consequences arising from the use of any information it contains.

Accepted Manuscript

View Article OnlineView Journal

This article can be cited before page numbers have been issued, to do this please use: Q. Liang, Y. Yu, G.

Feng, Y. Shen, L. Yang and S. Lei, Chem. Commun., 2020, DOI: 10.1039/D0CC05216G.

Page 2: View Article Online ChemComm...halogen bonding8 and metal-ligand coordination interactions,9 offers opportunity to achieve highly ordered nanoarchitectures. While a great variety of

Journal Name

COMMUNICATION

This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 1

Please do not adjust margins

Please do not adjust margins

a.Tianjin Key Laboratory of Molecular Optoelectronic Science, Department of Chemistry, School of Science & Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin University, Tianjin 300072, P. R. China

b.Jiangxi University of Science and Technology, Ganzhou 341000, P. R. Chinac. MOE of the Key Laboratory of Bioinorganic and Synthetic Chemistry, The Key Lab

of Low-carbon Chemistry and Energy Conservation of Guangdong Province, School of Chemistry, Sun Yat-Sen University, Guangzhou 510275, P. R. China

d.School of Chemistry and Chemical Engineering, Harbin Institute of Technology, Harbin 150080, People's Republic of China

*Email: [email protected]† Electronic Supplementary Information (ESI) available: Detailed experimental procedures and additional experimental data. See DOI: 10.1039/x0xx00000x

Received 00th January 20xx,Accepted 00th January 20xx

DOI: 10.1039/x0xx00000x

www.rsc.org/

Two-dimensional co-crystallization of two carboxylic acid derivatives having dissimilar symmetries at the liquid/solid interfaceQiu Liang,a Yanxia Yu,bc Guangyuan Feng,a Yongtao Shen,a Ling Yangd and Shengbin Lei*a

By co-assembly of two carboxylic acids with distinct symmetry and different number of carboxyl groups, we obtained two novel cocrystal structures at the n-octanoic acid/HOPG interface, one of which sustained by unoptimized R22(8) hydrogen bonding. Benefited from the bias-sensitivity of BTB (1,3,5-tris(4-carboxyphenyl)benzene) molecule, a structure transition between the cocrystal network and a denser BTB lamellae is achieved.

Co-crystallization of multiple components into well-ordered supramolecular structures has gained considerable attention as a promising research interest in the area of nanopatterning, heterojunction materials fabrication, molecular optoelectronics etc.1-3 Assembling of molecular building blocks via noncovalent intermolecular interactions, such as van der Waals interactions,4 hydrogen bonding,5-7 halogen bonding8 and metal-ligand coordination interactions,9 offers opportunity to achieve highly ordered nanoarchitectures. While a great variety of nanostructures have been obtained upon rational molecular design and controlled component assembly, it’s not easy to achieve effective co-crystallization, which strongly depends on the structural match in distinct constituent compounds, the applied methods, and the assembly driving force of the building blocks.10,11

Due to the high directionality and selectivity of hydrogen bonds,12-14 hydrogen bonding is most widely explored by far for the construction of cocrystal nanoarchitectures. Carboxyl

group (COOH) is the most important functional groups employed because it can form two H-bonds (O-H···O). The location and number of carboxyl group in the molecule skeleton will have a huge impact on molecular assembly because they will guide the binding direction between neighbouring molecules.15-19 Lackinger’s group have studied the coadsorption of BTB and TMA (1,3,5-benzenetricarboxylic acid) at the liquid-solid interface in two different solvents. Depending on the stoichiometries of the two solutes, three mixed networks comprising both BTB and TMA molecules were observed.20 Given that BTB is a bias-sensitive molecule, De Feyter et al has further demonstrated that by changing the orientation of an externally applied electric field, one can locally control the mixing behaviour of BTB and TMA physisorbed on a solid surface.21 It is noteworthy that both BTB and TMA exhibit 3-fold symmetry (D3h), while co-crystallization from building blocks with distinct symmetries will enable construction of more complex assembling structures.12,22

Stimuli-induced switchable molecular assemblies on surfaces have attracted great interest in recent years because of their potential in developing artificial smart surfaces. These externally applied stimuli include heat, light, pH, external electric field, etc,22-28 among which an electric field applied via an STM tip is considered a good choice, as it makes the control of the electric field and the observation of the structures synchronous. As a bias-sensitive molecule, the structure transition of BTB between open honeycomb and close-packed 2D supramolecular networks has been achieved by changing the polarity of the bias voltage, and the switching is independent of the solvent used and fully reversible.

In this communication, we report on the co-crystallization and bias induced manipulation of 4,4',4'',4'''-(1,4-phenylenebis(pyridine-6,2,4-triyl))tetrabenzoic acid(PBPTTBA) and BTB. We found that when a solution containing a mixture of the two molecules in an appropriate stoichiometry was deposited onto the HOPG surface, two co-crystallized, supramolecular porous networks were obtained. One is a wide and narrow alternate network structure with two-fold

Page 1 of 5 ChemComm

Che

mC

omm

Acc

epte

dM

anus

crip

t

Publ

ishe

d on

04

Sept

embe

r 20

20. D

ownl

oade

d by

Tia

njin

Uni

vers

ity o

n 9/

4/20

20 8

:05:

27 A

M.

View Article OnlineDOI: 10.1039/D0CC05216G

Page 3: View Article Online ChemComm...halogen bonding8 and metal-ligand coordination interactions,9 offers opportunity to achieve highly ordered nanoarchitectures. While a great variety of

COMMUNICATION Journal Name

2 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Please do not adjust margins

symmetry, the other is a flower-like architecture with three-fold symmetry. The formation of three-fold symmetric cocrystal is interesting since normally the symmetry of cocrystal is controlled by the building block with lower symmetry. Careful inspection on the high resolution STM image reveals that the three-fold cocrystal is sustained by unoptimized R22(8) hydrogen bond, which highlight the flexibility of the hydrogen bonds. Besides, thanks to the bias sensitive characteristic of BTB, we also revealed an electric field-induced switching behaviour between the bicomponent structure and the preferential adsorption of BTB.

Fig. 1 shows the molecular structures of the building blocks used in this study. The reason for choosing PBPTTBA and BTB is twofold: one is that they have matchable size of molecular backbone; the other is because of their distinct symmetries and different number of carboxylic groups, which may facilitate the formation of versatile co-assembly structures. PBPTTBA bears four carboxyl groups with a 2-fold symmetry (D2h), which has been widely used as an organic ligand for MOFs (metal-organic frameworks). 29-32 The solubility of PBPTTBA in n-octanoic acid (OA) is quite low, and the self-assembled structures of PBPTTBA from a saturated solution at the OA/HOPG interface is presented in Fig. 2 and Fig. S1. Two distinctly different assembling structures can be observed in the large-scale image. From high-resolution STM images, we can see that one structure is quadrilateral tiling (Fig. 2a), and the unit cell parameters are a=b=2.3±0.2 nm, a=80±2°. It is discernible that the tilt direction of PBPTTBA is uniform and adjacent molecules are linked together by hydrogen bonds. The other structure is a Kagomé motif (Fig. 2b), and the unit cell parameters are a=b=4.7±0.2 nm, a=55±2°. A Kagomé star (purple lines) is superimposed onto the porous organic network as a visual guide in Fig. 2b. The large hexagon in the middle of the Kagomé structure is enclosed by six PBPTTBA molecules with no 1,4-bis(4-pyridyl)benzene entity acting as its edge, whereas the small surrounding quasi-triangle is formed by three PBPTTBA molecules with alternate edges of 1,4-bis(4-pyridyl)benzene backbone and carboxylic acid–carboxylic acid junction. The white rhombus in the images is the corresponding unit cell, and the molecular model is inserted in the lower right corner of each panel. Both assembling patterns are stable under either positive or negative STM bias, indicating that PBPTTBA isn’t a bias-sensitive molecule (Fig. S2).

Fig. 1 Chemical structures and representative cartoons of (a) 4,4',4'',4'''-(1,4-phenylenebis(pyridine-6,2,4-triyl))tetrabenzoic acid (PBPTTBA) and (b) 1,3,5-tris(4-carboxyphenyl)benzene (BTB).(c) Two possible polygonal networks obtained by combining the PBPTTBA (blue) and BTB (red).

Self-assembly of BTB at n-octanoic/HOPG interface is concentration dependent, similar to that proved by Commeto et al. 33 At a high concentration (4.16×10-4 M), we obtained two different lamellar phases, both are composed by side-by-side arranged BTB rows (Fig. S3). Within the close packed 1D ribbon, two BTB rows are interdigitated to form close packed lamella of 1.70 nm in width. In the other more open lamellar network, the BTBs in the two antiparallel rows are arranged head-to-head to form double H-bonds, where each four molecules border a rectangular cavity. The size of the unit cell amounts to (1.8×3.6) nm2 with an angle of 63°between the lattice vectors and contains two BTB molecules. While at a low concentration (4.16×10-5 M), we obtained a honeycomb porous network, as exhibited in Fig. 2c. Every carboxyl group forms a 2-fold hydrogen bond with a carboxyl group from the adjacent BTB molecule, resulting in six hydrogen bonds per BTB molecule. The rhombus in the image is the corresponding unit cell: in Fig. 2c, a=b=3.2±0.2 nm, a=62±2°, and the molecular model is inserted into each panel.

Fig. 2 High-resolution STM images of self-assembled structures of PBPTTBA and BTB. (a) Quadrilateral tiling of PBPTTBA. (b) Kagomé structure of PBPTTBA. (c) Honeycomb structure of BTB at CBTB= 4.16×10-5M. The insets display the tentative 2D-packing models of the arrangement. Imaging conditions: Vbias = - 600 mV, Iset = 330 pA.

Next, we investigated the cocrystalization behaviour of the binary mixture of PBPTTBA and BTB. Saturated solution of PBPTTBA and BTB (4.16 × 10-5 M) was mixed together with a mole ratio of 1:1 and deposited onto the surface. At this concentration, PBPTTBA forms both oblique and Kagomé structure, while BTB forms honeycomb network exclusively. The assembling of binary mixture of PBPTTBA and BTB results in two bicomponent structures (Fig. 3a). Fig. 3b shows a high-resolution STM image of one of the cocrystal structure at the interface, hereafter called 2D cocrystal A, and a molecular model is displayed in Fig. 3c. This cocrystal is composed by alternatively arranged normal and compressed hexagons, in which each normal hexagon is composed by two PBPTTBA and four BTB molecules, while the compressed hexagon is composed by two PBPTTBA and two BTB (Fig 3c). In the network the building blocks are connected by the energetically

Page 2 of 5ChemComm

Che

mC

omm

Acc

epte

dM

anus

crip

t

Publ

ishe

d on

04

Sept

embe

r 20

20. D

ownl

oade

d by

Tia

njin

Uni

vers

ity o

n 9/

4/20

20 8

:05:

27 A

M.

View Article OnlineDOI: 10.1039/D0CC05216G

Page 4: View Article Online ChemComm...halogen bonding8 and metal-ligand coordination interactions,9 offers opportunity to achieve highly ordered nanoarchitectures. While a great variety of

Journal Name COMMUNICATION

This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 3

Please do not adjust margins

Please do not adjust margins

most favourable homomeric or heteromeric R22(8) hydrogen bonding, where the carboxyl groups are aligned 180° with respect to each other. This cocrystal can be looked at as alternate arrangement of columns of PBPTTBA and BTB, and different combinations can coexist at the interface (Fig. S4 and Fig. S5).

The high-resolution STM image in Fig. 3d reveals the bicomponent supramolecular network, termed as cocrystal B in the following text for simplicity. A large-scale STM image with an area of 100×100 nm2 is shown in Fig. S6a. The molecular arrangement in this flower like structure is of 3-fold symmetry, in which a small quasi-triangle in the centre is surrounded by six quasi-hexagons. The small quasi-triangle in the centre is formed by three PBPTTBA molecules with alternate edges of 1,4-bis(4-pyridyl)benzene backbone and carboxylic acid-carboxylic acid junction, while the surrounding quasi-hexagon consists of three PBPTTBA molecules and two BTB molecules with one edge formed by the 1,4-bis(4-pyridyl)benzene backbone of PBPTTBA. The unit cell is relatively large and contains three molecules of PBPTTBA and two molecules of BTB. It is revealed that, similar to cocrystal A, this network is also sustained by homomeric and heteromeric R22(8) hydrogen bonding between PBPTTBA and PBPTTBA-BTB molecules, respectively. However, their relative orientation is somewhat tentative, and it appears that there is a deviation from the optimal 180°in the R22(8) hydrogen bonding based on the molecular arrangement. However, our DFT simulation indicates that the deviation only brings neglectable effect on the strength of hydrogen bond (1.465 eV vs 1.447 eV) (Fig. S7). The rhombus in Fig. 3d is the corresponding unit cell, and the parameters are a=b=5.4±0.2 nm, a=67±2°. Both cocrystal A and cocrystal B are thermodynamically stable up to 60C. It is worth noting that although the hydrogen bond in cocrystal A is conformationally more favourable, cocrystal B is the dominant structure at the interface according to our STM observation. The ratio of surface coverage for cocrystal A and cocrystal B is 1:5 from our statistical result.

Table 1. Surface coverage () of different networks at different molar ratio.

Molar ratio (PBPTTBA:BTB)

PBPTTBA Cocrystal BTB

2:1 0.85 0.15 -

3:2 0.14 0.86 -

1:1 0.05 0.90 0.05

2:3 - 0.70 0.30

1:2 - 0.19 0.81

In addition, we performed a series of experiments to explore the effect of molar ratio (r, PBPTTBA to BTB) on the co-assembly structure, as demonstrated in Fig. S8. It is found that the quadrilateral tiling of PBPTTBA is the majority on the surface when r = 2:1. When r decreases to 3:2, the cocrystal B appears on the surface, and pure PBPTTBA domains coexist at the same time. Further increasing BTB in the mixing solution (r=2:3) leads to the disappearance of the PBPTTBA quadrilateral tiling in the adlayer, and the honeycomb network

of BTB emerges on the surface and co-exists with the cocrystals. When r is decreased to 1:2, the honeycomb network of BTB is the exclusive structure in the adlayer. A statistical analysis on the ratio of different networks at different molar ratio is summarized in table 1 and Fig. S9.

Fig. 3 (a) Large-scale image of the co-assembly structures of PBPTTBA and BTB. The white dotted line marks out the boundary of the two co-assembly structures. (b) High-resolution STM image of cocrystal A. (d) High-resolution STM image of cocrystal B. (c) and (e) Corresponding tentative 2D-packing models of the cocrystals. Imaging conditions: Vbias = - 600 mV, Iset = 330 pA.

For all the STM characterizations of the coassembly described above, STM imaging was performed with negative sample bias at room temperature. Interestingly, reversing the substrate bias to positive leads to the formation of another phase, which is entirely composed of BTB molecules. The STM images (Fig. 4a–f) show selected snapshots during a continuous scan for 96 mins. Fig. 4a–f shows the dynamic process of the gradually collapse of cocrystal B structure and appearance of lamellae network of BTB, and reappearence of cocrystal B upon reversing the polarity. The surface coverage of the close packed lamellae network of BTB keeps growing with time (Fig. 4c–d). The STM image in Fig. 4c show a typical snapshot during the transition with coexistence of undisturbed co-assembly (domain A), lamellae network of BTB (domain C), and the quadrilateral tiling structure of PBPTTBA molecules (domain B). The appearance of the quadrilateral tiling of PBPTTBA gives a clear proof that the bias induced phase transition is dominated by the bias dependent transition of BTB. STM observations at liquid-solid interface and DFT calculations both support the explanation that the alignment of intrinsic dipole and partial deprotonation under the electric field is responsible for the bias polarity induced phase transition of BTB.33,34 In the present system both these factors also can explain the observed structural transitions. The emergence of quadrilateral tiling of PBPTTBA during the transition further supports the hypothesis of partial deprotonation of BTB, which leads to the rupture of the cocrystal. A zoom-out STM image with a larger scanning area (172 × 172 nm2, Fig. S10) after a sequence of STM imaging (72 × 72 nm2) under positive bias, revealing that the transformed area was localized to the scanned area, further confirmed that

Page 3 of 5 ChemComm

Che

mC

omm

Acc

epte

dM

anus

crip

t

Publ

ishe

d on

04

Sept

embe

r 20

20. D

ownl

oade

d by

Tia

njin

Uni

vers

ity o

n 9/

4/20

20 8

:05:

27 A

M.

View Article OnlineDOI: 10.1039/D0CC05216G

Page 5: View Article Online ChemComm...halogen bonding8 and metal-ligand coordination interactions,9 offers opportunity to achieve highly ordered nanoarchitectures. While a great variety of

COMMUNICATION Journal Name

4 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Please do not adjust margins

the transition is induced by the electric field. Cocrystal A network can also switch to BTB lamellae when the bias translate to positive (Fig. S11).

Fig. 4 Sequential STM images showing the reversible transition between cocrystal B and close packed BTB lamellae achieved by reversing the polarity of the substrate bias. Imaging conditions: Vbias = - 600 mV, Iset = 330 pA. Scale bars: 5 nm. The purple dotted line in (b) and (e) indicates where the polarity of the substrate bias is changed, and the purple arrow depicts the scan direction.

In conclusion, we have studied the cocrystallization of PBPTTBA and BTB at the liquid/solid interface. By virtue of distinct symmetries and different number of carboxyl groups, we successfully obtained two novel co-crystallization structures. High-resolution STM images and DFT simulations suggest that R22(8) intermolecular hydrogen bond is the main force for the stabilization of all the above structures. The PBPTTBA-BTB cocrystal network can be reversibly transformed into a BTB lamellae by switching the polarity of substrate bias, which can be explained by the alignment of intrinsic dipole and partial deprotonation under the electric field. The manipulation of supramolecular networks at the liquid/solid interface by an external electric field could be useful for the controlled drug release, smart surfaces, self-healing materials, etc.

AcknowledgementsThis work was financially supported by the National

Natural Science Foundation of China (21872103, 21901182 and 51633006) and the Ministry of Science and Technology of the People’s Republic of China (Grant 2016YFB0401100)Notes and references1 H. Y. Shi, X. C. Lu, Y. H. Liu, J. Song, K. Deng, Q. D. Zeng and C.

Wang, ACS Nano., 2018, 12, 8781-8790.2 H. Y. Gao, Q. J. Shen, X. R. Zhao, X. Q. Yan, X. Pang and W. J.

Jin, J. Mater. Chem., 2012, 22, 5336-5343.3 Y. Wang, W. G. Zhu, W. N. Du, X. F. Liu, X. T. Zhang, H. L.

Dong and W. P. Hu, Angew. Chem., Int. Ed., 2018, 57, 3963-3967.

4 S. Y. Li, X. Q. Yang, T. Chen, D. Wang, G. S. Zhu and L. J. Wan, Phys. Chem. Chem. Phys. 21, 17846-17851.

5 C. A. Palma, P. Samorì and M. Cecchini, J. Am. Chem. Soc., 2010, 132, 17880-17885.

6 L. X. Cheng, B. Tu, X. W. Xiao, B. Feringán, R. Giménez, X. K. Li, Q. J. Fang, T. Sierra, Y. B. Li, Q. D. Zeng and C. Wang, Langmuir, 2019, 35, 8935-8942.

7 J. Q. Li, B. Tu, X. K. Li, C. Y. Ma, C. Chen, W. B. Duan, X. W. Xiao, Q. D. Zeng, Chem. Commun., 2019, 55, 11599-11602.

8 Q. N. Zheng, X. H. Liu, T. Chen, H. J.Yan, T. Cook, D. Wang, P. J. Stang and L. J. Wan, J. Am. Chem. Soc. 2015, 137, 6128-6131.

9 D. Nieckarz and P. Szabelski, J. Phys. Chem. C., 2013, 117, 11229-11241.

10 J. D. Wuest, Nat. Chem., 2012, 4, 74-75.11 L. Sun, J. Y. Wang, F. X. Yang, X. T. Zhang and W. P. Hu ,

Adv. Mater. 2019, 31, 1902328. 12 F. Cometto, K. Frank, B. Stel, N. Arisnabarreta, K. Kern and

M. Lingenfelder, Chem. Commun., 2017,53, 11430-11432.13 W. Li, S. Y. Qiu, C. Y. Xu, J. P. Hu and X. L. Chen, Phys. Chem.

Chem. Phys., 2019, 21, 8940-8944.14 L. Kampschulte, S. Griessl, W. M. Heckl and M. Lackinger, J.

Phys. Chem. B., 2005, 109, 14074-14078.15 H. Zhou, H. Dang, J. H. Yi, A. Nanci, A. Rochefort and J. D.

Wuest, J. Am. Chem. Soc., 2007, 129, 13774-13775.16 T. Yokoyama, T. Kamikado, S. Yokoyama and S. Mashiko, J.

Chem. Phys., 2004, 121, 11993-11997.17 M. Lackinger and W. M. Heckl, Langmuir., 2009, 25, 11307-

11321.18 J. M. MacLeod,O. Ivasenko, C. Y. Fu, T. Taerum,F. Rosei and

D. F. Perepichka, J. Am. Chem. Soc., 2009, 131, 16844-16850.

19 Y. C. Ye, W. Sun, Y. F. Wang, X. Shao, X. G. Xu, F. Cheng, J. L. Li and K. Wu, J. Phys. Chem. C., 2007, 111, 10138-10141.

20 L. Kampschulte, T. L. Werblowsky, R. S. K. Kishore, M. Schmittel, W. M. Heckl and M. Lackinger, J. Am. Chem. Soc., 2008, 130, 26, 8502–8507.

21 G. Velpula, J. Teyssandier, S. D. Feyter and K. S. Mali, ACS Nano., 2017, 11, 10903-10913.

22 W. Li, S. L. Xu, X. L. Chen, C. Y. Xu, Chinese Chemical. Letter., 2020, DOI: 10.1016/j.cclet.2020.06.012.

23 A. Mahmood, M. Saeed, Y. Chan, A. S. Saleemi, J. T. Guo and S. L. Lee, Langmuir., 2019, 35, 8031-8037.

24 C. H. Liu, W. Zhang, Q. D. Zeng and S. B. Lei, Chem. Eur. J. 2016, 22, 6768-6773.

25 L. L. Cao, Y. X. Yu and S. B. Lei, Chem. Commun., 2018, 54, 12210-12213.

26 S. L. Lee, Y. Fang, G. Velpula, F. P. Cometto, M. Lingenfelder,K. Müllen, K. S. Mali and S. D. Feyter, ACS Nano.,2015, 12, 11608-11617.

27 S. Y. Li, X. Q. Yang, T. Chen, D. Wang, S. F. Wang and L. J. Wan, ACS Nano., 2019, 13, 6751-6759.

28 S. L. Lee, C. H. Lin, K. Y. Cheng, Y. C. Chen, C. H. Chen, J. Phys. Chem. C., 2016, 120, 25505-25510.

29 J. H. Wang, Y. Zhang, M. Li, S. Yan, D. Liand X. M. Zhang, Angew. Chem. Int. Ed., 2017, 56, 6478-6482.

30 J. H. Carter, X. Han, F. Y. Moreau, I. Silva, A. Nevin, H. G.W. Godfrey, C. C. Tang, S. H. Yang and M. Schröder, J. Am. Chem. Soc., 2018, 140, 15564-15567.

31 S. L. Hou, J. Dong, M. H. Tang, X. L. Jiang, Z. H. Jiao and B. Zhao, Anal. Chem., 2019, 91, 5455-5460.

32 X. L. Jiang, S. L. Hou, Z. H. Jiao and B. Zhao, Anal. Chem., 2019, 91, 9754-9759.

33 F. P. Cometto, K. Kern and M. Lingenfelder, ACS Nano., 2015, 5, 5544-5550.

34 M. Ruben, D. Payer, A. Landa, A. Comisso, C. Gattinoni, N. Lin, J. P. Collin, J. P. Sauvage, A. D. Vita and K. Kern, J. Am. Chem. Soc., 2006, 128, 15644-15651.

Page 4 of 5ChemComm

Che

mC

omm

Acc

epte

dM

anus

crip

t

Publ

ishe

d on

04

Sept

embe

r 20

20. D

ownl

oade

d by

Tia

njin

Uni

vers

ity o

n 9/

4/20

20 8

:05:

27 A

M.

View Article OnlineDOI: 10.1039/D0CC05216G

Page 6: View Article Online ChemComm...halogen bonding8 and metal-ligand coordination interactions,9 offers opportunity to achieve highly ordered nanoarchitectures. While a great variety of

Table of Content for

Two-dimensional co-crystallization of two carboxylic acid derivatives having dissimilar symmetries at the liquid/solid interface

Qiu Liang,a Yanxia Yu, bc Guangyuan Feng,a Yongtao Shen,a Ling Yangd and Shengbin Lei*a

Two co-crystallized, supramolecular porous networks were constructed using two carboxylic acid with distinct symmetries and different number of carboxyl groups, and reversible switch is achieved between cocrystal network and BTB lamellae by changing the polarity of the bias voltage.

Page 5 of 5 ChemComm

Che

mC

omm

Acc

epte

dM

anus

crip

t

Publ

ishe

d on

04

Sept

embe

r 20

20. D

ownl

oade

d by

Tia

njin

Uni

vers

ity o

n 9/

4/20

20 8

:05:

27 A

M.

View Article OnlineDOI: 10.1039/D0CC05216G