+ All Categories
Home > Documents > Real-Space Identification of peared as individual molecules or ran- Intermolecular...

Real-Space Identification of peared as individual molecules or ran- Intermolecular...

Date post: 20-Jun-2020
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
7
/ http://www.sciencemag.org/content/early/recent / 26 September 2013 / Page 1 / 10.1126/science.1242603 Intermolecular bonding has been characterized experimentally mainly through crystallography via x-ray and electron diffractions, as well as through infrared, Raman, nuclear magnetic resonance, and near-edge extended absorption fine structure spectroscopy (1, 2). At the single- molecule level, state-of-the-art scanning tunneling microscopy (STM) is a widely used technique to elucidate the molecular structure and chemi- cal specificity of surface-immobilized species (35). The bonding inter- actions between molecules in self-assemblies were also evidenced in scanning tunneling hydrogen microscopy (6). Nevertheless, most of the characterization techniques are so far more sensitive to the covalent structures of the molecules, and in many cases, theoretical calculations of intermolecular interaction are also not as precise as those for covalent- ly bound species. Recently, non-contact atomic force microscopy (NC-AFM) has achieved superior resolution in real-space that has enabled the identifica- tion of the chemical structure, adsorption configurations, and chemical transformation of individual molecules (710). For example, the differ- ence in bond order in aromatic molecules was distinguished via electron- density–dependent Pauli repulsion with CO-functionalized NC-AFM tips (11), and AFM tomography revealed the angular symmetry of a chemical bond on surface (12). We used NC-AFM to investigate the intermolecular interactions in 8-hydroxyquiline (8-hq, Fig. 1A) molecu- lar assemblies formed on Cu(111) at liquid-helium (LHe) and room tem- peratures. The hydrogen bonds (H bonds) formed between 8-hq molecules were characterized via high resolution AFM images, and the local bonding configuration was determined with the atomic precision. We also observe the coordination complex composed of dehydrogenated 8-hq and Cu adatom. The observations were validated with ab initio density functional theory (DFT) calculations. The 8-hq molecules deposited on Cu(111) at LHe temperature ap- peared as individual molecules or ran- domly assembled aggregates (fig. S1) (13). For the single 8-hq molecules, DFT calculations suggest that the mo- lecular plane is slightly tilted with re- spect to the substrate because of the weak interactions between the OH group and N atom of 8-hq and Cu(111) surface. Compared with the calculated total electron density of the molecule shown in Fig. 1B, the STM image (Fig. 1C) exhibits no internal features of the heterocycle because the tunneling cur- rent is primarily sensitive to the local density of states near the Fermi level. In contrast, the AFM images with a CO- functionalized tip revealed the sub- molecular structure of 8-hq through the short-range Pauli repulsive force (Fig. 1, D to F). The calculated electron den- sity map (Fig. 1B) qualitatively repro- duces the observed contrast in frequency shifts in the AFM image. Here, the AFM sensor measured the total force of three components (7, 14): (i) the long-range attractive electrostatic forces, responsible for the overall nega- tive Δf background in the images; (ii) the attractive van der Waals (vdW) force, which contributed to the dark halo surrounding the molecule without atomic corrugation; and (iii) the short- range Pauli repulsion, which contribut- ed to the atomic contrast of molecular structure with respect to the metal substrate. When the tip height was decreased (Fig. 1, D to F), the in- creasing proportion of Pauli repulsion in the total force enhanced con- trast in the AFM images. Although a quantitative understanding of the AFM imaging mechanism is nontrivial, a direct correlation between the AFM images and the chemical structure of a molecule can still be ra- tionalized. In our case, the heterocyclic skeleton and the hydroxyl group of 8-hq were readily distinguished. The pyridine ring in the heterocycle is slightly pronounced, which may be caused by tilting of the molecular plane on the substrate. A further interpretation of the topography need also take into account the difference in electron density of the phenol ring and the pyridine ring. The AFM images of the 8-hq molecular aggregates (Fig. 2, A and B) reveal bonding-like features between adjacent molecules in the assem- blies that were reproduced in all of the observations, whereas these fea- tures were not observed in the corresponding STM images at the same regions [see fig. S3 and (13)]. A close examination of the position and orientation of the bonding-like structures indicated that they coincide very well with the expected locations of H bonds formed between 8-hq molecules (Fig. 2, C and D). The results from recent theoretical and experimental investigations suggest that the H bond has both an electro- static origin and a partly covalent character (15, 16). Despite extensive studies of H bonding in supramolecular and biological systems using various techniques, direct identification of the bonding configuration of H bond in real-space is elusive (17). The formation of covalent bonds in unimolecular reactions has been recently reported (10). The bond contrast in the AFM images has been qualitatively compared with the bond order, where the higher local elec- tron density leads to stronger Pauli repulsion exerted on the tip (10, 11). In our observations, the Δf contrast of these intermolecular bonds is Real-Space Identification of Intermolecular Bonding with Atomic Force Microscopy Jun Zhang, 1 * Pengcheng Chen, 1 * Bingkai Yuan, 1 Wei Ji, 2 † Zhihai Cheng, 1 Xiaohui Qiu 1 1 Key Laboratory of Standardization and Measurement for Nanotechnology, Chinese Academy of Sciences, National Center for Nanoscience and Technology, Beijing 100190, China. 2 Department of Physics, Renmin University of China, Beijing 100872, China. *These authors contributed equally to this work. †Corresponding author. E-mail: [email protected] (X.Q.); [email protected] (Z.C.); [email protected] (W.J.) We report a real-space visualization of the formation of hydrogen bonding in 8- hydroxyquiline (8-hq) molecular assemblies on a Cu(111) substrate using non- contact atomic force microscopy (NC-AFM). The atomically resolved molecular structures enable a precise determination of the characteristics of hydrogen bonding networks, including the bonding sites, orientations, and lengths. The observation of bond contrast was interpreted by ab initio density functional calculations, which indicated the electron density contribution from the hybridized electronic state of the hydrogen bond. Intermolecular coordination between the dehydrogenated 8-hq and Cu adatoms was also revealed by the sub-molecular resolution AFM characterization. The direct identification of local bonding configurations by NC-AFM would facilitate detailed investigations of intermolecular interactions in complex molecules with multiple active sites. on October 9, 2013 www.sciencemag.org Downloaded from on October 9, 2013 www.sciencemag.org Downloaded from on October 9, 2013 www.sciencemag.org Downloaded from on October 9, 2013 www.sciencemag.org Downloaded from on October 9, 2013 www.sciencemag.org Downloaded from on October 9, 2013 www.sciencemag.org Downloaded from on October 9, 2013 www.sciencemag.org Downloaded from
Transcript
Page 1: Real-Space Identification of peared as individual molecules or ran- Intermolecular …szolcsanyi/education/files/Chemia... · 2014-08-21 · forces, responsible for the overall nega-

/ http://www.sciencemag.org/content/early/recent / 26 September 2013 / Page 1 / 10.1126/science.1242603

Intermolecular bonding has been characterized experimentally mainly through crystallography via x-ray and electron diffractions, as well as through infrared, Raman, nuclear magnetic resonance, and near-edge extended absorption fine structure spectroscopy (1, 2). At the single-molecule level, state-of-the-art scanning tunneling microscopy (STM) is a widely used technique to elucidate the molecular structure and chemi-cal specificity of surface-immobilized species (3–5). The bonding inter-actions between molecules in self-assemblies were also evidenced in scanning tunneling hydrogen microscopy (6). Nevertheless, most of the characterization techniques are so far more sensitive to the covalent structures of the molecules, and in many cases, theoretical calculations of intermolecular interaction are also not as precise as those for covalent-ly bound species.

Recently, non-contact atomic force microscopy (NC-AFM) has achieved superior resolution in real-space that has enabled the identifica-tion of the chemical structure, adsorption configurations, and chemical transformation of individual molecules (7–10). For example, the differ-ence in bond order in aromatic molecules was distinguished via electron-density–dependent Pauli repulsion with CO-functionalized NC-AFM tips (11), and AFM tomography revealed the angular symmetry of a chemical bond on surface (12). We used NC-AFM to investigate the intermolecular interactions in 8-hydroxyquiline (8-hq, Fig. 1A) molecu-lar assemblies formed on Cu(111) at liquid-helium (LHe) and room tem-peratures. The hydrogen bonds (H bonds) formed between 8-hq molecules were characterized via high resolution AFM images, and the local bonding configuration was determined with the atomic precision. We also observe the coordination complex composed of dehydrogenated 8-hq and Cu adatom. The observations were validated with ab initio density functional theory (DFT) calculations.

The 8-hq molecules deposited on Cu(111) at LHe temperature ap-

peared as individual molecules or ran-domly assembled aggregates (fig. S1) (13). For the single 8-hq molecules, DFT calculations suggest that the mo-lecular plane is slightly tilted with re-spect to the substrate because of the weak interactions between the OH group and N atom of 8-hq and Cu(111) surface. Compared with the calculated total electron density of the molecule shown in Fig. 1B, the STM image (Fig. 1C) exhibits no internal features of the heterocycle because the tunneling cur-rent is primarily sensitive to the local density of states near the Fermi level. In contrast, the AFM images with a CO-functionalized tip revealed the sub-molecular structure of 8-hq through the short-range Pauli repulsive force (Fig. 1, D to F). The calculated electron den-sity map (Fig. 1B) qualitatively repro-duces the observed contrast in frequency shifts in the AFM image. Here, the AFM sensor measured the total force of three components (7, 14): (i) the long-range attractive electrostatic forces, responsible for the overall nega-tive Δf background in the images; (ii) the attractive van der Waals (vdW) force, which contributed to the dark halo surrounding the molecule without atomic corrugation; and (iii) the short-range Pauli repulsion, which contribut-

ed to the atomic contrast of molecular structure with respect to the metal substrate. When the tip height was decreased (Fig. 1, D to F), the in-creasing proportion of Pauli repulsion in the total force enhanced con-trast in the AFM images. Although a quantitative understanding of the AFM imaging mechanism is nontrivial, a direct correlation between the AFM images and the chemical structure of a molecule can still be ra-tionalized. In our case, the heterocyclic skeleton and the hydroxyl group of 8-hq were readily distinguished. The pyridine ring in the heterocycle is slightly pronounced, which may be caused by tilting of the molecular plane on the substrate. A further interpretation of the topography need also take into account the difference in electron density of the phenol ring and the pyridine ring.

The AFM images of the 8-hq molecular aggregates (Fig. 2, A and B) reveal bonding-like features between adjacent molecules in the assem-blies that were reproduced in all of the observations, whereas these fea-tures were not observed in the corresponding STM images at the same regions [see fig. S3 and (13)]. A close examination of the position and orientation of the bonding-like structures indicated that they coincide very well with the expected locations of H bonds formed between 8-hq molecules (Fig. 2, C and D). The results from recent theoretical and experimental investigations suggest that the H bond has both an electro-static origin and a partly covalent character (15, 16). Despite extensive studies of H bonding in supramolecular and biological systems using various techniques, direct identification of the bonding configuration of H bond in real-space is elusive (17).

The formation of covalent bonds in unimolecular reactions has been recently reported (10). The bond contrast in the AFM images has been qualitatively compared with the bond order, where the higher local elec-tron density leads to stronger Pauli repulsion exerted on the tip (10, 11). In our observations, the Δf contrast of these intermolecular bonds is

Real-Space Identification of Intermolecular Bonding with Atomic Force Microscopy Jun Zhang,1* Pengcheng Chen,1* Bingkai Yuan,1 Wei Ji,2† Zhihai Cheng,1† Xiaohui Qiu1† 1Key Laboratory of Standardization and Measurement for Nanotechnology, Chinese Academy of Sciences, National Center for Nanoscience and Technology, Beijing 100190, China. 2Department of Physics, Renmin University of China, Beijing 100872, China.

*These authors contributed equally to this work.

†Corresponding author. E-mail: [email protected] (X.Q.); [email protected] (Z.C.); [email protected] (W.J.)

We report a real-space visualization of the formation of hydrogen bonding in 8-hydroxyquiline (8-hq) molecular assemblies on a Cu(111) substrate using non-contact atomic force microscopy (NC-AFM). The atomically resolved molecular structures enable a precise determination of the characteristics of hydrogen bonding networks, including the bonding sites, orientations, and lengths. The observation of bond contrast was interpreted by ab initio density functional calculations, which indicated the electron density contribution from the hybridized electronic state of the hydrogen bond. Intermolecular coordination between the dehydrogenated 8-hq and Cu adatoms was also revealed by the sub-molecular resolution AFM characterization. The direct identification of local bonding configurations by NC-AFM would facilitate detailed investigations of intermolecular interactions in complex molecules with multiple active sites.

on

Oct

ober

9, 2

013

ww

w.s

cien

cem

ag.o

rgD

ownl

oade

d fr

om

on

Oct

ober

9, 2

013

ww

w.s

cien

cem

ag.o

rgD

ownl

oade

d fr

om

on

Oct

ober

9, 2

013

ww

w.s

cien

cem

ag.o

rgD

ownl

oade

d fr

om

on

Oct

ober

9, 2

013

ww

w.s

cien

cem

ag.o

rgD

ownl

oade

d fr

om

on

Oct

ober

9, 2

013

ww

w.s

cien

cem

ag.o

rgD

ownl

oade

d fr

om

on

Oct

ober

9, 2

013

ww

w.s

cien

cem

ag.o

rgD

ownl

oade

d fr

om

on

Oct

ober

9, 2

013

ww

w.s

cien

cem

ag.o

rgD

ownl

oade

d fr

om

Page 2: Real-Space Identification of peared as individual molecules or ran- Intermolecular …szolcsanyi/education/files/Chemia... · 2014-08-21 · forces, responsible for the overall nega-

/ http://www.sciencemag.org/content/early/recent / 26 September 2013 / Page 2 / 10.1126/science.1242603

comparable to that of intramolecular covalent bonds in the constant height images, as also evident in the force spectroscopy measurements (fig. S4) (13). Given the clearly identified bonding sites, we could per-form a detailed analysis of the H bond configurations (18). The apparent bond lengths of the intermolecular H bond were measured as summa-rized in fig. S5 (13), and can be used as a reference for data acquired through other characterization methods to understand the effect of the substrate on H bonding (19).

Our study also revealed differences in the H bond configurations in the 8-hq self-assemblies on surface and those in the bulk crystal. In addi-tion to the conventional H bond, which only involves the OH group and N atom of 8-hq, we also observed H bond formation between the aro-matic rings and the OH group or N atom. These results provide direct evidence for the influence of substrate on the intermolecular bonding characteristics (1, 20).

We performed DFT calculations on two types of molecular clusters to further understand the origin of the contrast of intermolecular H bonds in our observation. In Fig. 3A, the paired 8-hq molecules were bonded by two H bonds of O-H…N, as illustrated by the black dotted lines in Fig. 3B (21). The calculated total electron density plotted in Fig. 3C shows a bright spot at the position of the O atom and a protrusion toward the adjacent molecule around the position of the N atom. The in-plane plot of the differential charge density (Fig. 3D), which is defined as

MoleculeDCD Total Substrateρ = ρ − ρ − ρ∑ , reflects the charge redistribution after the bond forms between two 8-hq molecules and indicates that covalent-like characteristics developed from charge reductions near the H and N atoms that led to charge accumulation between them. The enhanced charge density at N and O, as expected, is consistent with the charge transfer from H to N and O atoms, whereas the charge accumulation offers an additional repulsive force to the tip at the region along the H bonding direction. Thus, the observed line feature between the two 8-hq molecules in the AFM image is attributed to a joint effect resulted from both the covalent charge in between H…N and the charge transferred from H to N and O. The above interpretation is primarily based on the simplified mechanism of tip-to-sample Pauli repulsion (22). A quantita-tive understanding of the image contrast of H bond may request further consideration of the relaxation of the CO molecule attached on the tip apex (11). In another configuration (Fig. 3, E to H), DFT calculations also concluded electron density redistribution in the proximity of the newly formed bond. In this unconventional C-H…N hydrogen bond, the effect of charge accumulation between N and H atoms is much weaker compared to that in the O-H…N hydrogen bond (1, 15).

When 8-hq was deposited onto Cu(111) at room temperature (RT), the molecules formed highly ordered dimers and trimers, which are dis-tinct from the H-bonded aggregates formed at LHe temperature. No internal structures of these dimers and trimers could be resolved in the high-resolution STM images (fig. S8). The size of these molecular ag-gregates did not correspond to those of H bonded clusters from DFT calculations. The AFM images of the dimers and trimers (Fig. 4, A and B) allowed the identification of the outer edge of the molecules corre-sponding to the positions of H or C atoms, but the inner edge was blurred.

The dehydrogenation of hydroxyl group on Cu(111) at RT has been widely reported (23). Our experimental and theoretical calculation re-sults also suggested that the individual 8-hq molecules exist as radical species that are weakly bound to the substrate by O and N atoms in a tilted orientation (fig. S7). Alternatively, the highly mobile 8-hq radicals coordinate with Cu adatoms on surface to form organometallic complex (24). The proposed chemical structures of dimer and trimer (Fig. 4, C and D), respectively, in which the dehydrogenated 8-hq is assembled via an O(N)-Cu bond. The calculated geometric sizes of these 8-hq com-plexes agree well with the AFM observations.

When the AFM is operated in the Pauli repulsion regime, the repul-

sive force from the outermost shell valence electrons is also relevant to 2 ( )n r∇ ϕ which is in the form of kinetic energy. The ( )n rϕ is the elec-

tron wavefunction of the nth eigenstate of the sample. The kinetic energy reflects the localization property of electrons which can be estimated by the electron localization function (ELF) (25). The ELF of the dimer and trimer exhibited strongly localized electron density donation from N to Cu, whereas the electrons between the Cu-Cu bonding are rather delocal-ized (Fig. 4, E and F). We attributed the bonding features in the central regions of the dimer and trimer to N-Cu coordination bonding, similar to the observed formation of the metal-organic coordinate bond excited by inelastic electrons (9, 26). According to the ELF, the O-Cu bond is strongly polarized, and the most shared electrons of the bond are local-ized around O, thus the AFM signal is negligible. The C-H…O H bond formed in dimer and trimer. were detectable in AFM because of the lo-calized electrons around O and H.

Structural details, including the molecular conformation, the bond configuration, and the interacting sites on the functional groups, acquired from high-resolution non-contact AFM images provided useful insights into the mechanisms of molecular assembly and recognition. Because the H bond is ubiquitous in nature and central to biological functions, the present technique may provide an important and complementary charac-terization method for unraveling the fundamental aspects of molecular interactions at the single molecule level. The observation of H bonding in real-space may also stimulate theoretical discussion about the nature of this intermolecular interaction.

References and Notes 1. K. Müller-Dethlefs, P. Hobza, Noncovalent interactions: A challenge for

experiment and theory. Chem. Rev. 100, 143–168 (2000). Medline doi:10.1021/cr9900331

2. C. A. Hunter, Quantifying intermolecular interactions: Guidelines for the molecular recognition toolbox. Angew. Chem. Int. Ed. 43, 5310–5324 (2004). Medline doi:10.1002/anie.200301739

3. H. J. Lee, W. Ho, Single-bond formation and characterization with a scanning tunneling microscope. Science 286, 1719–1722 (1999). doi:10.1126/science.286.5445.1719

4. R. Temirov, S. Soubatch, A. Luican, F. S. Tautz, Free-electron-like dispersion in an organic monolayer film on a metal substrate. Nature 444, 350–353 (2006). Medline doi:10.1038/nature05270

5. P. Liljeroth, J. Repp, G. Meyer, Current-induced hydrogen tautomerization and conductance switching of naphthalocyanine molecules. Science 317, 1203–1206 (2007). doi:10.1126/science.1144366

6. C. Weiss, C. Wagner, R. Temirov, F. S. Tautz, Direct imaging of intermolecular bonds in scanning tunneling microscopy. J. Am. Chem. Soc. 132, 11864–11865 (2010). Medline doi:10.1021/ja104332t

7. L. Gross, F. Mohn, N. Moll, P. Liljeroth, G. Meyer, The chemical structure of a molecule resolved by atomic force microscopy. Science 325, 1110–1114 (2009). doi:10.1126/science.1176210

8. L. Gross, F. Mohn, N. Moll, G. Meyer, R. Ebel, W. M. Abdel-Mageed, M. Jaspars, Organic structure determination using atomic-resolution scanning probe microscopy. Nat. Chem. 2, 821–825 (2010). Medline doi:10.1038/nchem.765

9. F. Albrecht, M. Neu, C. Quest, I. Swart, J. Repp, Formation and characterization of a molecule-metal-molecule bridge in real space. J. Am. Chem. Soc. 135, 9200–9203 (2013). Medline doi:10.1021/ja404084p

10. D. G. de Oteyza, P. Gorman, Y. C. Chen, S. Wickenburg, A. Riss, D. J. Mowbray, G. Etkin, Z. Pedramrazi, H. Z. Tsai, A. Rubio, M. F. Crommie, F. R. Fischer, Direct imaging of covalent bond structure in single-molecule chemical reactions. Science 340, 1434–1437 (2013). doi:10.1126/science.1238187

11. L. Gross, F. Mohn, N. Moll, B. Schuler, A. Criado, E. Guitián, D. Peña, A. Gourdon, G. Meyer, Bond-order discrimination by atomic force microscopy. Science 337, 1326–1329 (2012). doi:10.1126/science.1225621

12. J. Welker, F. J. Giessibl, Revealing the angular symmetry of chemical bonds by atomic force microscopy. Science 336, 444–449 (2012). doi:10.1126/science.1219850

13. Supplementary materials are available on Science Online.

Page 3: Real-Space Identification of peared as individual molecules or ran- Intermolecular …szolcsanyi/education/files/Chemia... · 2014-08-21 · forces, responsible for the overall nega-

/ http://www.sciencemag.org/content/early/recent / 26 September 2013 / Page 3 / 10.1126/science.1242603

14. L. Gross, Recent advances in submolecular resolution with scanning probe microscopy. Nat. Chem. 3, 273–278 (2011). Medline doi:10.1038/nchem.1008

15. T. Steiner, The hydrogen bond in the solid state. Angew. Chem. Int. Ed. 41, 48–76 (2002). Medline doi:10.1002/1521-3773(20020104)41:1<48::AID-ANIE48>3.0.CO;2-U

16. G. R. Desiraju, Die Wasserstoffbrücke - eine aktualisierte definition. Angew. Chem. 123, 52–60 (2011). doi:10.1002/ange.201002960

17. The observation of an intramolecular hydrogen bond was tentatively proposed by Gross et al. in the study of an organic compound, cephalandole A (8).

18. The measured bond length may get involved the amplification effect of CO on tip apex, as discussed in (11). It should also be noted that the expected bending of X−H…Y was not resolved in all of the hydrogen bonds shown in Fig. 2, A and B. The exact mechanism was not understood.

19. G. Jones, S. J. Jenkins, D. A. King, Hydrogen bonds at metal surfaces: Universal scaling and quantification of substrate effects. Surf. Sci. 600, 224–228 (2006). doi:10.1016/j.susc.2006.05.053

20. J. Carrasco, A. Hodgson, A. Michaelides, A molecular perspective of water at metal interfaces. Nat. Mater. 11, 667–674 (2012). Medline doi:10.1038/nmat3354

21. We found that it is difficult to obtain a high resolution AFM image on this specific type of 8-hq dimer comparing to single molecules or other molecular aggregates. These dimers often accidentally dislocate during AFM imaging, indicating a weaker interactions between the molecules and the substrate. We suggested that the formation of two hydrogen bonds in the dimer weaken the binding interaction of -OH and N of 8-hq to the metal substrate. It is noted that the optimal imaging parameters for this dimer (see the figure caption of Fig. 3A) are different from those used for other molecular clusters.

22. We conducted a further calculation (13) following the method proposed in (27).

23. X. C. Guo, R. J. Madix, Monolayer structure of phenoxy species on Cu(110): An STM study. Surf. Sci. 341, L1065–L1071 (1995). doi:10.1016/0039-6028(95)00822-5

24. J. V. Barth, G. Costantini, K. Kern, Engineering atomic and molecular nanostructures at surfaces. Nature 437, 671–679 (2005). Medline doi:10.1038/nature04166

25. B. Silvi, A. Savin, Classification of chemical bonds based on topological analysis of electron localization functions. Nature 371, 683–686 (1994). doi:10.1038/371683a0

26. F. Mohn, J. Repp, L. Gross, G. Meyer, M. S. Dyer, M. Persson, Reversible bond formation in a gold-atom-organic-molecule complex as a molecular switch. Phys. Rev. Lett. 105, 266102 (2010). Medline doi:10.1103/PhysRevLett.105.266102

27. N. Moll, L. Gross, F. Mohn, A. Curioni, G. Meyer, A simple model of molecular imaging with noncontact atomic force microscopy. New J. Phys. 14, 083023 (2012). doi:10.1088/1367-2630/14/8/083023

28. B. Hammer, L. B. Hansen, J. K. Nørskov, Improved adsorption energetics within density-functional theory using revised Perdew-Burke-Ernzerhof functionals. Phys. Rev. B 59, 7413–7421 (1999). doi:10.1103/PhysRevB.59.7413

29. P. E. Blöchl, Projector augmented-wave method. Phys. Rev. B 50, 17953–17979 (1994). Medline doi:10.1103/PhysRevB.50.17953

30. G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 59, 1758–1775 (1999). doi:10.1103/PhysRevB.59.1758

31. G. Kresse, J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169–11186 (1996). Medline doi:10.1103/PhysRevB.54.11169

32. S. Grimme, Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 27, 1787–1799 (2006). Medline doi:10.1002/jcc.20495

Acknowledgments: This project is partially supported by the Ministry of Science

and Technology (MOST) of China (grant nos. 2012CB933001, 2012CB932704), the Natural Science Foundation of China (NSFC) (grant nos. 21173058, 21203038, 11274308, 11004244), the Beijing Natural Science Foundation (BNSF) (grant no. 2112019), and the Basic Research Funds in Renmin University of China from the Central Government (grant no. 12XNLJ03). W.J. was supported by the Program for New Century Excellent Talents in Universities. Calculations were performed at the PLHPC of RUC

and Shanghai Super-computing Center. We are indebted to the valuable discussion and technical advices from W. Ho, P. Grutter, and L. Gross.

Supplementary Materials www.sciencemag.org/cgi/content/full/science.1242603/DC1 Materials and Methods Figs. S1 to S9 References (28–32)

1 July 2013; accepted 9 September 2013 Published online 26 September 2013 10.1126/science.1242603

Page 4: Real-Space Identification of peared as individual molecules or ran- Intermolecular …szolcsanyi/education/files/Chemia... · 2014-08-21 · forces, responsible for the overall nega-

/ http://www.sciencemag.org/content/early/recent / 26 September 2013 / Page 4 / 10.1126/science.1242603

Fig. 1. SPM measurements and DFT calculations of single 8-hydroxyquinoline (8-hq) on Cu(111). (A) Chemical structure of 8-hq. (B) DFT-calculated molecular electron density maps at a distance of 150 pm above the molecule. (C) Constant-current STM topography image (V = –100 mV, I = 100 pA) with a CO-functionalized tip. (D to F) Constant-height AFM frequency shift images (V = 0 V, A = 100 p.m.) at different tip heights. The tip height Δz was set with respect to a reference height given by the STM set point above (–100 mV, 100 pA) the bare Cu(111) substrate in the vicinity of the molecule. The plus (minus) sign means the increase (decrease) of tip height. (D): Δz = +30 p.m.; (E): Δz = +10 p.m.; (F): Δz = 0 pm. The size of all images is 1.3 nm × 1.0 nm.

Page 5: Real-Space Identification of peared as individual molecules or ran- Intermolecular …szolcsanyi/education/files/Chemia... · 2014-08-21 · forces, responsible for the overall nega-

/ http://www.sciencemag.org/content/early/recent / 26 September 2013 / Page 5 / 10.1126/science.1242603

Fig. 2. AFM measurements of 8-hq assembled clusters on Cu(111). (A and B) Constant-height frequency shift images of typical molecule assembled clusters, and their corresponding structure models (C and D). Imaging parameters: V = 0 V, A = 100 p.m., Δz = +10 pm. Image size: (A) 2.3 nm × 2.0 nm, (B) 2.5 nm × 1.8 nm. The dashed lines in (C) and (D) indicate likely H bonds between 8-hq molecules.

Page 6: Real-Space Identification of peared as individual molecules or ran- Intermolecular …szolcsanyi/education/files/Chemia... · 2014-08-21 · forces, responsible for the overall nega-

/ http://www.sciencemag.org/content/early/recent / 26 September 2013 / Page 6 / 10.1126/science.1242603

Fig. 3. AFM measurements and DFT calculations of 8-hq dimers on Cu(111). Constant-height frequency shift

image of O-H…N dimer (A) and N…H-Ph dimer (E), and their corresponding DFT-calculated structure model (B and F), electron density map (C and G), charge difference map (D and H). Imaging parameters: V = 0 V, A = 100 p.m., Δz = +50 p.m. (A), Δz = +10 p.m. (E). Image size: (A) 1.6 nm × 1.6 nm, (E) 1.5 nm × 2.0 nm. The dashed frames in (B) and (F) indicate the calculation regions in (D) and (H).

Page 7: Real-Space Identification of peared as individual molecules or ran- Intermolecular …szolcsanyi/education/files/Chemia... · 2014-08-21 · forces, responsible for the overall nega-

/ http://www.sciencemag.org/content/early/recent / 26 September 2013 / Page 7 / 10.1126/science.1242603

Fig. 4. AFM measurements and DFT calculations of coordination complexes (dehydrogenated 8-hq and copper adatom) on Cu(111). Constant-height AFM frequency shift image (A and B), the corresponding DFT-calculated structure model (C and D) and electron localization function map (E and F) of dimer (Cuq2) and timer (Cu3q3) complexes. Imaging parameters: (A) V = 0 V, A = 100 p.m., Δz = +100 p.m., 2.0 nm × 2.0 nm; (B) V = 0 V, A = 100 p.m., Δz = +80 p.m., 2.4 nm × 2.4 nm. The tip height Δz was set with respect to a reference height given by the STM set point (–30 mV, 100 pA) above the bare Cu(111) substrate. The dashed lines in (C) and (D) refer to the H bonds formed in the complexes. The complexes were formed by depositing 8-hq on Cu(111) at RT.


Recommended