24.01.2013 Views

Noncontact Atomic Force Microscopy - Yale School of Engineering ...

Noncontact Atomic Force Microscopy - Yale School of Engineering ...

Noncontact Atomic Force Microscopy - Yale School of Engineering ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

12 th International Conference on<br />

<strong>Noncontact</strong> <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong><br />

and<br />

Casimir 2009 Workshop<br />

CONFERENCE PROGRAM<br />

August 10-14, 2009 � <strong>Yale</strong> University, New Haven, CT, USA


Institutional Sponsors<br />

<strong>Yale</strong> Center for Research<br />

on Interface Structures<br />

and Phenomena<br />

Past Conferences<br />

Osaka, Japan (1998)<br />

Pontresina, Switzerland (1999)<br />

Hamburg, Germany (2000)<br />

Kyoto, Japan (2001)<br />

Montreal, Canada (2002)<br />

Dingle, Ireland (2003)<br />

Seattle, USA (2004)<br />

Osnabrück, Germany (2005)<br />

Kobe, Japan (2006)<br />

Antalya, Turkey (2007)<br />

Madrid, Spain (2008)<br />

Forthcoming Conference<br />

Kanazawa, Japan (2010)<br />

Udo D. Schwarz<br />

Dept. <strong>of</strong> Mechanical <strong>Engineering</strong><br />

<strong>Yale</strong> University<br />

New Haven, CT 06515, USA<br />

udo.schwarz@yale.edu<br />

Cover Page: Banner <strong>of</strong> <strong>Yale</strong><br />

University (1.8 µm � 1.8 µm),<br />

written into a PMMA film by<br />

dynamic force microscopy.<br />

Image by M. Heyde &<br />

U. D. Schwarz<br />

Left: Harkness Tower on<br />

<strong>Yale</strong> Campus.<br />

• <strong>Yale</strong> <strong>School</strong> <strong>of</strong> <strong>Engineering</strong><br />

• Departments <strong>of</strong> Mechanical, Electrical,<br />

and Chemical <strong>Engineering</strong><br />

• Department <strong>of</strong> Physics<br />

Endorsed<br />

Topical<br />

Conference<br />

www.avs.org


Welcome<br />

Dear conference participant:<br />

Welcome to the 12th International Conference on Non-Contact <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong><br />

(NC-AFM 2009), which is held on historic <strong>Yale</strong> University campus in New Haven, CT, USA.<br />

It continues a series <strong>of</strong> international conferences constituted 1998 in Osaka, Japan. Since<br />

then, the annual NC-AFM conferences have been established as the leading events for<br />

NC-AFM related topics. This year’s meeting has again attracted a large number <strong>of</strong><br />

participants (over 130 attendees from 17 countries), who are contributing 43 oral and 71<br />

poster presentations. 15 additional talks are provided by the Satellite Workshop on Casimir<br />

<strong>Force</strong>s and Their Measurement, and seven companies feature their newest products in an<br />

exhibition.<br />

Like in other years, the scientific program showcases the rapid development that NC-AFM<br />

enjoys. Various experimental improvements such as high-stability measurements<br />

performed at low temperatures, novel stiff self-sensing oscillators with atomically controlled<br />

tips that allow chemical identification and tunneling current collection, drift compensation by<br />

forward-feedbacking, atom tracking, and post-acquisition drift correction, and all-digital<br />

high-speed, low-noise electronics enable a new level <strong>of</strong> sophistication in imaging, analysis,<br />

and atom manipulation. Progress is particularly remarkable for high-resolution data<br />

acquisition in liquids, demonstrating that NC-AFM is leaving its ultrahigh vacuum niche.<br />

The main conference is complemented by a Satellite Workshop on Casimir <strong>Force</strong>s and<br />

Their Measurement (Casimir 2009). Designed to stimulate discussion between the Casimir<br />

and the NC-AFM communities, the Casimir 2009 workshop has been well received (87<br />

participants). Credit for initiating this event goes to Woo-Joong Kim, who was aided in its<br />

organization by Alex Sushkov and Steven K. Lamoreaux.<br />

Contributions by many key players were indispensable in realizing this conference, among<br />

them the ones by the members <strong>of</strong> the Local Organizing Committee (Eric I. Altman, Hong X.<br />

Tang, and Woo-Joong Kim) and by the team <strong>of</strong> the <strong>Yale</strong> Conference Service under the<br />

supervision <strong>of</strong> Susan Adler. Special thanks go to Mehmet Baykara for his help with various<br />

practical aspects such as the design <strong>of</strong> this abstract booklet. We were also fortunate to<br />

receive significant support from numerous <strong>Yale</strong> entities, whose generous financial<br />

assistance made this conference possible, from the European Science Foundation through<br />

its “New Trends and Applications <strong>of</strong> the Casimir Effect” program, and from the American<br />

Vacuum Society, which arranged for the publication <strong>of</strong> the conference proceedings. The<br />

complete list <strong>of</strong> institutional sponsors can be found opposite to this page; corporative<br />

sponsors and exhibitors are listed on page 9.<br />

We hope that you will enjoy the conference and have a wonderful time in New Haven.<br />

Udo D. Schwarz, Conference Chair<br />

1


2<br />

Contents<br />

Welcome 1<br />

Topics 3<br />

Committees 4<br />

General Information 6<br />

Local Maps 8<br />

Exhibitors 9<br />

Proceedings 10<br />

Conference Program 13<br />

Oral Presentations 28<br />

Poster Presentations 91<br />

Author Index 164<br />

List <strong>of</strong> Participants 169<br />

Advertisement 174


Topics<br />

• <strong>Atomic</strong> resolution imaging <strong>of</strong> surfaces, thin films, and molecular<br />

systems<br />

• Measuring tip-sample interaction potentials and mapping force fields<br />

• High-resolution imaging <strong>of</strong> clusters, biomolecules, and biological<br />

systems<br />

• High-resolution imaging and spectroscopy in liquid environments<br />

• Novel instrumentation and measurement techniques in dynamic AFM<br />

• Small amplitude measurements<br />

• Lateral force measurements using dynamic methods<br />

• <strong>Atomic</strong>- and molecular-scale manipulation<br />

• Theoretical analysis <strong>of</strong> contrast mechanisms; forces & tunnelling<br />

phenomena<br />

• Simulation <strong>of</strong> images and virtual AFM systems<br />

• Mechanisms for damping and energy dissipation<br />

• Amplitude modulation versus frequency modulation imaging<br />

• Measuring nanoscale charges, work functions, and magnetic<br />

properties<br />

• Characterization and modification <strong>of</strong> force microscopy tips at the<br />

atomic scale<br />

3


Committees<br />

Conference Chairman<br />

Udo Schwarz <strong>Yale</strong> University (USA)<br />

Local Organizing Committee<br />

Eric I. Altman <strong>Yale</strong> University (USA)<br />

Hong X. Tang <strong>Yale</strong> University (USA)<br />

Woo-Joong Kim <strong>Yale</strong> University (USA)<br />

Satellite Workshop Organization<br />

Woo-Joong Kim <strong>Yale</strong> University (USA)<br />

Udo Schwarz <strong>Yale</strong> University (USA)<br />

Alex Sushkov <strong>Yale</strong> University (USA)<br />

Steven K. Lamoreaux <strong>Yale</strong> University (USA)<br />

International Steering Committee<br />

Franz Giessibl University <strong>of</strong> Regensburg (Germany)<br />

Peter Grütter McGill University (Canada)<br />

Ernst Meyer University <strong>of</strong> Basel (Switzerland)<br />

Seizo Morita Osaka University (Japan)<br />

Hiroshi Onishi Kobe University (Japan)<br />

Ahmet Oral Sabanci University (Turkey)<br />

Rubén Peréz Universidad Autonoma de Madrid (Spain)<br />

Michael Reichling University <strong>of</strong> Osnabrück (Germany)<br />

Alexander Schwarz University <strong>of</strong> Hamburg (Germany)<br />

Udo Schwarz <strong>Yale</strong> University (USA)<br />

Alexander Shluger University College London (UK)<br />

4


Committees<br />

Program Committee<br />

Jaime Colchero Universidad de Murcia (Spain)<br />

Oscar Custance National Institute for Material Science (Japan)<br />

Takeshi Fukuma Kanazawa University (Japan)<br />

Hendrik Hölscher Karlsruhe Research Laboratory (Germany)<br />

Sascha Sadewasser Helmholtz Institute Berlin (Germany)<br />

André Schirmeisen University <strong>of</strong> Münster (Germany)<br />

Santiago Solares University <strong>of</strong> Maryland (USA)<br />

Yasuhiro Sugawara Osaka University (Japan)<br />

Masaru Tsukada Waseda University (Japan)<br />

Hir<strong>of</strong>umi Yamada Kyoto University (Japan)<br />

Conference Proceedings Editors<br />

Hendrik Hölscher Karlsruhe Research Laboratory (Germany)<br />

Udo D. Schwarz <strong>Yale</strong> University (USA)<br />

Conference Organization<br />

Tara Schule, Joanne Dupee, Roberta Hudson, and Susan Adler, <strong>Yale</strong> Conference<br />

Services<br />

5


General Information<br />

Conference Venue:<br />

Talks will be held in Davies Auditorium in the basement <strong>of</strong> <strong>Yale</strong> University’s Becton Center,<br />

15 Prospect Street, New Haven, CT 06511 (see the maps page 8 for the exact location).<br />

For easiest access, follow the signs around the building to the lower courtyard located<br />

between Becton Center and the rear building (Dunham Laboratory, 10 Hillhouse Avenue).<br />

There, you will find the main entrance to the conference area, which is indicated by the red<br />

arrow in the satellite picture. For the auditorium, please turn left after entering the doors. In<br />

contrast, turning right twice will lead you to the J. Robert Mann Jr. <strong>Engineering</strong> Student<br />

Center, where c<strong>of</strong>fee breaks and parts <strong>of</strong> the exhibition are taking place. Poster sessions<br />

and luncheons will be held in Commons Hall at the corner <strong>of</strong> Grove and College Streets<br />

(the building with the dome opposite Becton Center). Please keep in mind that smoking is<br />

not allowed in any <strong>Yale</strong> University building, including Davies Auditorium, the Mann Student<br />

Center, and Commons Hall.<br />

Registration:<br />

The registration desk in the foyer <strong>of</strong> Davies Auditorium (see above) will be staffed 07:45-<br />

18:00 (Monday), 08:00-12:00 (Tuesday), and 08:30-09:00 (Wednesday through Friday).<br />

Welcome Party:<br />

We invite all conference participants to a casual welcome party with reception and dinner<br />

(barbeque) on Monday, August 10, 2009, between 18:30 and 20:30. The exact location will<br />

be announced ahead <strong>of</strong> time via email and on the web.<br />

Conference Banquet:<br />

Conference banquet will be held on Thursday, Aug. 13 th , 2009, 18:30-22:00 at Sterling<br />

Memorial Library (SML) on <strong>Yale</strong> campus. Please see the maps on page 8 for the location <strong>of</strong><br />

SML.<br />

6


General Information (cont.)<br />

Oral Presentations:<br />

Oral presentations will be scheduled for 20 minutes each including a 5-minute discussion<br />

for sessions during the NC-AFM conference (August 11-14). For the specialized Casimir<br />

workshop sessions on August 10, talks will be 25 minutes each. An LCD projector is<br />

available in the conference room, but all speakers are required to bring their own laptops.<br />

Poster Presentations:<br />

Two Poster sessions will be held during the conference on Tuesday, August 11, and<br />

Wednesday, August 12, 2009. The available size for posters will be 1.2 m x 1.2 m. Pins for<br />

mounting the poster on the board will be provided.<br />

Internet Access:<br />

Wireless internet access will be available in Davies Auditorium and the exhibition areas.<br />

Detailed instructions for connecting to the internet will be distributed with your conference<br />

material. Also, a number <strong>of</strong> desktop computers will be available in the exhibition area and a<br />

nearby computer cluster room. A guest login ID and a personal password, which allows you<br />

to connect to the internet from these computers, will be handed to each participant at the<br />

registration desk.<br />

7


Local Maps<br />

8


Exhibitors<br />

9


Proceedings<br />

Both NC-AFM Conference and Casimir Workshop participants are invited to contribute an<br />

article to the conference proceedings, which will be published in the Journal <strong>of</strong> Vacuum<br />

Science and Technology B (JVST B). Publishing your work in the conference proceedings<br />

represents an excellent opportunity to promote your work, as NC-AFM proceedings have in<br />

the past always been highly visible and widely cited. This is in particular the case as every<br />

conference participant (main conference AND Casimir Workshop) will receive a bound<br />

complementary copy <strong>of</strong> the proceedings via mail.<br />

The proceedings will be edited by Hendrik Hölscher, Karlsruhe Research Laboratory,<br />

Germany, and Udo D. Schwarz, <strong>Yale</strong> University. Submission deadline is October 16,<br />

2009, but please contact us at your earliest convenience to let us know about your plans to<br />

submit a contribution so that we can plan ahead. Publication <strong>of</strong> the proceedings is<br />

scheduled for the May/June issue <strong>of</strong> JVST B in 2010.<br />

Instructions for preparing manuscripts for the proceedings <strong>of</strong> the NC-AFM 2009<br />

Conference:<br />

1. Submission: Papers will be reviewed according to criteria set by the JVST and must<br />

meet JVST standards for both technical content and written English. All manuscripts<br />

for the proceedings must be submitted as a PDF file along with a cover letter<br />

containing your complete address to the guest editor via email<br />

(Hendrik.Hoelscher@imt.fzk.de).<br />

2. Program number: It is important that the submitted manuscript has the program<br />

number <strong>of</strong> your contribution to the NC-AFM conference or the satellite conference<br />

entered in the top-right corner <strong>of</strong> the title page.<br />

10


Proceedings (cont.)<br />

3. Length: Papers can contain up to a nominal total <strong>of</strong> 20-25 pages <strong>of</strong> double-spaced<br />

text, tables, figures, references, abstract, etc. if the submitted paper is longer than<br />

25 pages, the final decision on acceptance for publication will be made by the JVST<br />

editors based on the recommendations <strong>of</strong> the reviewers.<br />

4. Format: The text should be double-spaced; the font recommended by JVST is<br />

Times New Roman 12pt. Pages should have one-inch margins on all sides.<br />

5. Figures and Tables: Each figure and table should be displayed on a separate<br />

page. Table and figure captions should be listed on a separate page and NOT below<br />

the table or figure; only the figure/table number must appear with the figure/table.<br />

6. Footnotes: Authors may include their e-mail addresses along with all other footnotes<br />

in the following format: Electronic mail: smith@yale.edu<br />

7. Costs: JVST publishes articles authored by members <strong>of</strong> the American Vacuum<br />

Society (AVS) free <strong>of</strong> charge. Therefore, let us know when submitting whether any <strong>of</strong><br />

the authors is presently AVS member. For submissions where no author is AVS<br />

member, we will provide a complementary 1-year membership for the paper's<br />

corresponding author paid for form conference funds. As a result, there will be no<br />

page charges or other costs for your submission.<br />

8. Color: Your figures can appear online in color for free. To take advantage <strong>of</strong> this<br />

<strong>of</strong>fer, you will need to send the electronic file(s) <strong>of</strong> your figure(s) within 24 hours <strong>of</strong><br />

receiving your acceptance email from JVST directly to the publisher (the American<br />

Institute <strong>of</strong> Physics, AIP). Please do not send anything until you receive notice the<br />

paper is accepted and has a 9 digit AIP number. Sending color figures in the pdf file<br />

11


Proceedings (cont.)<br />

used for the review process does not allow for color reproduction (either online or<br />

print) at the publishing stage. Only sending AIP the files directly will enable this. A<br />

“Publisher’s Announcements” email with more details will be sent to you once your<br />

paper has been accepted.<br />

9. Materials Index Entries: If applicable, please provide us with at least three material<br />

terms, which should indicate what materials (i.e., chemical species) are discussed in<br />

the article. These terms are needed as entries for the JVST Materials Index.<br />

For more details, please refer to:<br />

JVST author instructions: www.avs.org/literature.information.aspx<br />

JVST Word template: www.avs.org/pdf/JVST/eC.doc<br />

Instructions for this template: www.avs.org/pdf/JVST/eCEinstructions.pdf<br />

Manuscript deadline: October 16, 2009<br />

Please submit your manuscript as PDF together with a cover letter containing your<br />

complete address and information whether or not one <strong>of</strong> the authors is presently AVS<br />

member via email to guest editor Dr. Hendrik Hölscher at Hendrik.Hoelscher@imt.fzk.de<br />

12


Conference<br />

Program<br />

13


NC-AFM 2009<br />

Sessions<br />

Monday<br />

August 10<br />

• Liquids I<br />

Tuesday 9:20-10:40<br />

• <strong>Force</strong> Spectroscopy I<br />

Tuesday 11:20-12:40<br />

• Electronic, photonic, & Casimir forces<br />

Tuesday 14:30-15:50<br />

• Oxides<br />

Wednesday 9:00-10:40<br />

• Carbon-based materials<br />

Wednesday 11:20-12:40<br />

Tuesday<br />

August 11<br />

14<br />

Wednesday<br />

August 12<br />

• Molecules on insulators<br />

Wednesday 14:30-15:50<br />

• <strong>Force</strong> spectroscopy II<br />

Thursday 9:00-10:40<br />

• <strong>Force</strong>s & charges<br />

Thursday 11:20-13:00<br />

• Method development<br />

Friday 9:00-10:40<br />

• Liquids II<br />

Friday 11:20-12:40<br />

Thursday<br />

August 13<br />

Friday<br />

August 14<br />

8:30 Registration Registration Registration Registration<br />

9:00 Opening Remarks M. Heyde A. Campbellova Y. Naitoh<br />

9:20 T. Fukuma A. Yurtsever F. J. Giessibl S. Torbrügge<br />

9:40 H. Asakawa R. Pérez G. Langewisch Y. Hosokawa<br />

10:00 N. Oyabu R. Bechstein A. Barat<strong>of</strong>f E. Tsunemi<br />

10:20 M. Tsukada S. Sadewasser S. Kawai U. Zerweck<br />

10:40 C<strong>of</strong>fee Break C<strong>of</strong>fee Break C<strong>of</strong>fee Break C<strong>of</strong>fee Break<br />

11:20 D. Ebeling U. D. Schwarz A. Sweetman S. Nishida<br />

11:40 A. Schirmeisen P. Jelínek L. Gross S. Ido<br />

12:00 Y. Sugimoto M. Ashino Y. Miyahara E. T. Herruzo<br />

12:20 Y. Sugawara P. Pou T. Trevethan K. Fukui<br />

12:40 A. Bettac Closing Remarks<br />

13:00<br />

Lunch<br />

110 min.<br />

(12:40-14:30)<br />

Lunch<br />

110 min.<br />

(12:40-14:30)<br />

14:30<br />

A. Parsegian<br />

A. Kühnle<br />

14:50<br />

(invited)<br />

A. Schwarz<br />

15:10 Th. Glatzel Ch. Loppacher<br />

15:30 H. Hölscher C. J. Gómez<br />

15:50<br />

|<br />

18:00<br />

C<br />

A<br />

S<br />

I<br />

M<br />

I<br />

R<br />

W<br />

O<br />

R<br />

K<br />

S<br />

H<br />

O<br />

P<br />

18:30-20:30<br />

Welcome<br />

Reception/Dinner<br />

Poster Session I Poster Session II<br />

Lunch<br />

90 min.<br />

(13:00-14:30)<br />

18:30-22:00<br />

Conference<br />

Banquet


Program Summary <strong>of</strong> the NC-AFM 2009 Satellite Workshop on<br />

Casimir <strong>Force</strong>s and Their Measurement<br />

Monday sessions only; Tuesday sessions are joint with the NC-AFM Conference (see the main<br />

conference program for listing). Note in particular that the Casimir workshop’s second invited talk,<br />

given by Adrian Parsegian (NIH, USA), will take place Tuesday 14:30-15:10.<br />

Monday, August 10<br />

7:45 Registration<br />

8:50 Opening Remarks<br />

9:00 Rudi Porgornik (University <strong>of</strong> Ljubljana, Slovenia), invited<br />

9:30 Noah Graham (Middlebury College, USA)<br />

9:55 Kimball Milton (University Oklahoma, USA)<br />

10:20<br />

C<strong>of</strong>fee Break<br />

10:50 Raul Esquivel-Sirvent (Universidad Nacional Autónoma, Mexico)<br />

11:15 Diego Dalvit (Los Alamos National Laboratory, USA)<br />

11:40 Rick Rajter (MIT, USA)<br />

12:05 Maarten de Boer (Sandia National Laboratory, USA)<br />

12:30<br />

Lunch (12:30-14:30)<br />

14:30 Hong Tang (<strong>Yale</strong> University, USA)<br />

14:55 Alessandro Siria (CNRS, France)<br />

15:20 Arvind Narayanaswamy (Columbia University, USA)<br />

15:45<br />

C<strong>of</strong>fee Break<br />

16:15 Sven de Man (VU University Amsterdam, Netherlands)<br />

16:40 Gauthier Torricelli (University <strong>of</strong> Leicester, UK)<br />

17:05 Ho Bun Chan (University <strong>of</strong> Florida, USA)<br />

17:30 Peter van Zwol (University <strong>of</strong> Groningen, Netherlands)<br />

17:55 Woo-Joong Kim (<strong>Yale</strong> University, USA)<br />

18:30-20:30<br />

Welcome Reception with Dinner<br />

15


Monday, August 10<br />

Satellite Workshop on Casimir <strong>Force</strong>s and<br />

Their Measurement<br />

8:50 Opening Remarks<br />

Session I Chair: W.-J. Kim<br />

9:00 Non-retarded and retarded interactions between dielectric cylinders<br />

Rudi Podgornik<br />

9:30 Casimir <strong>Force</strong>s from Scattering Theory<br />

Noah Graham<br />

9:55 Multiple Scattering Casimir <strong>Force</strong> Calculations between Layered and Corrugated Materials<br />

Kimball Milton<br />

10:20 C<strong>of</strong>fee Break<br />

Session II Chair: R. Podgornik<br />

10:50 Drude corrections to Casimir force calculations in liquids<br />

Raul Esquivel-Sirvent<br />

11:15 Dispersive Casimir interactions between atoms and surfaces<br />

Diego Dalvit<br />

11:40 Van der Waals with a twist: how nanotube chirality impacts interaction strength<br />

Rick Rajter<br />

12:05 Using Casimir and capillary forces to model adhesion <strong>of</strong> MEMS cantilevers<br />

Maarten de Boer<br />

12:30 Lunch Break<br />

Session III Chair: A. Parsegian<br />

14:30 Measuring “Virtual Photon” <strong>Force</strong>s with “Real Photon” <strong>Force</strong>s<br />

Hong Tang<br />

14:55 Near-field radiative heat transfer<br />

Alessandro Siria<br />

15:20 Near-field radiative transfer measurements and implications for Casimir force<br />

measurements<br />

Arvind Narayanaswamy<br />

15:45 C<strong>of</strong>fee Break<br />

16


Monday, August 10<br />

Satellite Workshop on Casimir <strong>Force</strong>s and Their Measurement (cont.)<br />

Session IV<br />

17<br />

Chair: D. Dalvit<br />

16:15 Tricks and facts in a high precision measurement <strong>of</strong> the Casimir force with transparent<br />

conductors<br />

Sven de Man<br />

16:40 Measurements <strong>of</strong> the Casimir force gradient by AFM for Different Materials<br />

Gauthier Torricelli<br />

17:05 Measuring the topological dependence <strong>of</strong> the Casimir force on nanostructured silicon<br />

surfaces<br />

Ho Bun Chan<br />

17:30 Short range Casimir force measurements under ambient conditions and liquid<br />

environments<br />

Peter van Zwol<br />

17:55 Contact potential difference in a Casimir force measurement: How do we deal with it?<br />

Woo-Joong Kim<br />

Welcome Reception with Dinner (18:30-20:30)


Tuesday, August 11<br />

9:00 Opening Remarks<br />

Liquids I<br />

9:20 3D Scanning <strong>Force</strong> <strong>Microscopy</strong> at Solid/Liquid Interface<br />

T. Fukuma, Y. Ueda<br />

18<br />

Chair: H. Hölscher<br />

9:40 Anisotropic Hydration <strong>of</strong> Biological Molecules Visualized by Three-Dimensional Scanning<br />

<strong>Force</strong> <strong>Microscopy</strong><br />

H. Asakawa,Y. Ueda, T. Fukuma<br />

10:00 Experimental and Theoretical Studies on 3D Hydration Structures on Muscovite Mica<br />

Surfaces in Aqueos Solution<br />

N. Oyabu, K. Kimura, S. Ido, K. Suzuki, K. Kobayashi, T. Imai, H. Yamada<br />

10:20 Theory <strong>of</strong> Tip-Sample Interaction <strong>Force</strong> Mediated by Water<br />

M. Tsukada, M.Harada, K.Tagami<br />

10:40 C<strong>of</strong>fee Break<br />

<strong>Force</strong> Spectroscopy I<br />

11:20 Dynamic <strong>Force</strong> Spectroscopy <strong>of</strong> Single Chain-like Molecules<br />

D. Ebeling, H. Fuchs, F. Oesterhelt, H. Hölscher<br />

11:40 Spatial force fields above a single atom defect<br />

A. Schirmeisen, D. Weiner<br />

12:00 Simultaneous measurement <strong>of</strong> force and tunneling current<br />

D. Sawada, Y. Sugimoto, K.-I. Morita, M. Abe, S. Morita<br />

12:20 Atom Manipulation on Cu(110)-O Surface with LT-AFM<br />

Y. Sugawara, Y. Kinoshita, Y. J. Li, Y. Naitoh, M. Kageshima<br />

12:40 Lunch Break<br />

Chair: F. J. Giessibl<br />

Electronic, Photonic, & Casimir <strong>Force</strong>s Chair: S. Morita<br />

14:30 Water, Ions, Membranes, Real Metals, Finite Temperature: Is there ever a pure Casimir<br />

force? (Invited)<br />

Adrian Parsegian<br />

15:10 Short and medium range electrostatic forces analyzed by Kelvin probe force microscopy<br />

Th. Glatzel, S. Kawai, S. Koch, A. Barat<strong>of</strong>f, E. Meyer<br />

15:30 The Effective Quality Factor in Dynamic <strong>Force</strong> Microscopes with Fabry-Perot<br />

Interferometer Detection<br />

H. Hölscher, P. Milde, U. Zerweck, L. M. Eng, R. H<strong>of</strong>fmann<br />

Poster Session I (15:50 – 18:00)<br />

Chair: A. Oral


Wednesday, August 12<br />

Oxides Chair: A. Schwarz<br />

9:00 Imaging Single Atoms on Oxide Surfaces – Gold on Alumina/NiAl(110)<br />

M. Heyde, G. H. Simon, Th. König, H.-J. Freund<br />

9:20 Site-specific force spectroscopy on TiO2 (110) surface at low-temperature<br />

A. Yurtsever, A. Pratama, Y. Sugimoto, M. Abe, S. Morita<br />

9:40 Character <strong>of</strong> the short-range interaction between a silicon based tip and the TiO2(110)<br />

surface: a DFT study<br />

C. Gozalez, P. Jelínek, R. Pérez<br />

10:00 True atomic resolution imaging on an application-oriented system: Understanding<br />

photocatalytic reactivity <strong>of</strong> transition-metal doped TiO2<br />

R. Bechstein, M. Kitta, J. Schütte, H. Onishi, A. Kühnle<br />

10:20 Local surface photovoltage spectroscopy <strong>of</strong> molecular clusters using Kelvin probe force<br />

microscopy<br />

S. Sadewasser, M. Ch. Lux-Steiner<br />

10:40 C<strong>of</strong>fee Break<br />

Carbon-based Materials Chair: M. Reichling<br />

11:20 Why is Graphite so Slippery? Gathering Clues from Three-Dimensional Lateral <strong>Force</strong>s<br />

Measurements<br />

U. D. Schwarz, M. Z. Baykara, T. C. Schwendemann, B. J. Albers, N. Pilet, E. I. Altman<br />

11:40 Theoretical DFT simulations <strong>of</strong> the STM/AFM atomic scale imaging on graphite<br />

V. Rozsíval, M. Ondráček, P. Jelínek<br />

12:00 <strong>Atomic</strong>-Resolution Damping <strong>Force</strong> Spectroscopy on Nanotube Peapods with Different<br />

Tube Diameters<br />

M. Ashino, R. Wiesendanger, A. N. Khlobystov, S. Berber, D. Tománek<br />

12:20 Theoretical study <strong>of</strong> the forces and atomic configurations <strong>of</strong> NC-AFM experiments on lowdimension<br />

carbon materials<br />

P. Pou, R. Perez<br />

12:40 Lunch Break<br />

Molecules on Insulators Chair: R. Perez<br />

14:30 Anchoring highly polar molecules onto an ionic crystal<br />

J. Schütte, R. Bechstein, M. Rohlfing, M. Reichling, A. Kühnle<br />

14:50 Steering the formation <strong>of</strong> molecular nanowires and compact nanocrystallites on NaCl(001)<br />

S. Fremy, A. Schwarz, K. Laemmle, R. Wiesendanger, M. Prosenc<br />

15:10 2-Dimensional growth <strong>of</strong> phenylenediboronic acid assisted by H-bonding<br />

R. Pawlak, L. Nony, F. Bocquet, M. Sassi, V. Oison, J.-M. Debierre, Ch. Loppacher, L.<br />

Porte<br />

15:30 Molecular scale dissipation in oligothiophene monolayers measured by dynamic force<br />

microscopy<br />

C. J. Gómez, N. F. Martínez, W. Kamiński, C. Albonetti, F. Biscarini, R. Perez, R. Garcia<br />

Poster Session II (15:50 – 18:00) Chair: Y. Sugawara<br />

19


Thursday, August 13<br />

<strong>Force</strong> Spectroscopy II<br />

20<br />

Chair: P. Grütter<br />

9:00 Dependence <strong>of</strong> the atomic scale image <strong>of</strong> a Si adatom on the tip apex termination: a DFT<br />

study<br />

A. Campbellova, P. Pou , R. Pérez, P. Klapetek, P. Jelínek<br />

9:20 AFM probe tips with a small front atom<br />

T. H<strong>of</strong>mann, J. Welker, M. Ternes, C. P. Lutz, A. J. Heinrich, F. J. Giessibl<br />

9:40 Intramolecular features <strong>of</strong> organic molecules characterized by force field spectroscopy:<br />

The case <strong>of</strong> PTCDA on Cu and Ag<br />

G. Langewisch, D.-A. Braun, D. Weiner, B. Such, H. Fuchs, A. Schirmeisen<br />

10:00 Analysis <strong>of</strong> bimodal and higher mode small-amplitude near-contact AFM and energy<br />

dissipation<br />

S. Kawai, A. Barat<strong>of</strong>f, Th. Glatzel, S. Koch, B. Such, E. Meyer<br />

10:20 Adhesion-induced energy dissipation and atom-tracked tip changes<br />

S. Kawai, Th. Glatzel, S. Koch, B. Such, A. Barat<strong>of</strong>f, E. Meyer<br />

10:40 C<strong>of</strong>fee Break<br />

<strong>Force</strong>s & Charges<br />

Chair: H. Onishi<br />

11:20 Combined Qplus-AFM and STM imaging <strong>of</strong> the Si(100) surface: Activating the c(4x2) to<br />

p(2x1) transition with subnanometre oscillation amplitudes<br />

A. Sweetman, S. Gangopadhyay, R. Danza, P. Moriarty<br />

11:40 Measuring <strong>Atomic</strong> Charge States by nc-AFM<br />

L. Gross, F. Mohn, P. Liljeroth, J. Repp, F. J. Giessibl, G. Meyer<br />

12:00 Mechanism <strong>of</strong> Dissipative Interaction by Tunneling Single-Electrons<br />

Y. Miyahara, L. Cockins, S. D. Bennett, A. A. Clerk, S. A. Studenikin, P. Poole, A.<br />

Sachrajda, P. Grütter<br />

12:20 Controlling electron transfer processes on insulating surfaces with the NC-AFM<br />

Th. Trevethan, A. Shluger<br />

12:40 NC-AFM imaging with atomic resolution in a temperature range between 5 K and 1083 K<br />

A. Bettac, A. Feltz<br />

13:00 Lunch Break<br />

Guided <strong>Yale</strong> Campus Walking Tour (16:00-17:30)<br />

Conference Banquet (18:30 – 22:00)


Friday, August 14<br />

Method Development Chair: A. Schirmeisen<br />

9:00 <strong>Atomic</strong> scale elasticity mapping <strong>of</strong> Ge(001) surface by multifrequency FM-AFM<br />

Y. Naitoh, Z. Ma, Y. Li, M. Kageshima, Y. Sugawara<br />

9:20 Small amplitude atomic resolution NC-AFM imaging and force spectroscopy experiments<br />

using a stiff piezoelectric force sensor<br />

S. Torbrügge, J. Rychen, O. Schaff<br />

9:40 Determination <strong>of</strong> the Optimum Spring Constant and Oscillation Amplitude for<br />

<strong>Atomic</strong>/Molecular-Resolution FM-AFM<br />

Y. Hosokawa, K. Kobayashi, H. Yamada, K. Matsushige<br />

10:00 Visualization <strong>of</strong> Anisotropic Conductance in Polydiacetylene Crystal by Two-probe FM-<br />

AFM/KFM<br />

E. Tsunemi, K. Kobayashi, K. Matsushige, H. Yamada<br />

10:20 Scattering Scanning Near-Field Optical <strong>Microscopy</strong> performed by NC-AFM<br />

U. Zerweck, S. C. Schneider, M. T. Wenzel, H.-G. von Ribbeck, S. Grafström, R. Jacob, S.<br />

Winnerl, M. Helm, L. M. Eng<br />

10:40 C<strong>of</strong>fee Break<br />

Liquids II Chair: T. Fukuma<br />

11:20 <strong>Atomic</strong> resolution dynamic lateral force microscopy in liquid<br />

S. Nishida, D. Kobayashi, N. Okabe, H. Kawakatsu<br />

11:40 Molecular-scale Investigations <strong>of</strong> Biomolecules in Liquids by FM-AFM<br />

S. Ido, N. Oyabu, K. Kobayashi, Y. Hirata, M. Tsukada, K. Matsushige, H. Yamada<br />

12:00 Bimodal AFM imaging <strong>of</strong> antibodies and chaperonins in liquids<br />

E. T. Herruzo, C. Dietz, J. R. Lozano, R.Garcia<br />

12:20 Redox-state Dependent Reversible Change <strong>of</strong> Molecular Ensembles in Water Solution by<br />

Electrochemical FM-AFM<br />

K.-I. Umeda, Y. Yokota, K.-I. Fukui<br />

12:40 Closing Remarks<br />

21


POSTER Session I (Tuesday)<br />

<strong>Force</strong> Spectroscopy<br />

P.I-01 <strong>Force</strong> and Tunneling Current Measurements on the Semiconductor Surface<br />

D. Sawada, Y. Sugimoto, K.-I. Morita, M. Abe, S. Morita<br />

P.I-02 <strong>Force</strong> Map <strong>of</strong> <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong> on Si(111)-(5x5)-DAS Surface<br />

A. Masago, M. Tsukada<br />

22<br />

Chair: A. Oral<br />

P.I-03 From non-contact to atomic scale contact between a Si tip and a Si surface analyzed using<br />

an nc-AFM and nc-AFS based instrument<br />

T. Arai, K. Kiyohara, T. Sato, S. Kushida, M. Tomitori<br />

P.I-04 Improved atomic-scale contrast via bimodal dynamic force microscopy<br />

S. Kawai, Th. Glatzel, S. Koch, B. Such, A. Barat<strong>of</strong>f, E. Meyer<br />

P.I-05 Static cantilever deflection in dynamic force microscopy<br />

S. Kawai, Th. Glatzel, S. Koch, B. Such, A. Barat<strong>of</strong>f, E. Meyer<br />

P.I-06 Resonance frequency shift due to tip-sample interaction in the thermal oscillations regime<br />

G. Malegori, G. Ferrini<br />

P.I-07 Influence <strong>of</strong> thermal noise on measurements <strong>of</strong> chemical bonds in UHV-AFM<br />

P. M. H<strong>of</strong>fmann<br />

P.I-08 <strong>Atomic</strong> force microscope cantilever resonance frequency shift based thermal metrology<br />

A. Narayanaswamy, C. Canetta, N. Gu<br />

Kelvin Probe <strong>Microscopy</strong><br />

P.I-09 Contact potential difference on the atomic-scale probed by Kelvin Probe <strong>Force</strong> <strong>Microscopy</strong>:<br />

an imaging scenario<br />

L. Nony, A. Foster, F. Bocquet, Ch. Loppacher<br />

P.I-10 Self-assembled Boronitride Nanomesh on Rh(111) Investigated by Means <strong>of</strong> Kelvin Probe<br />

<strong>Force</strong> <strong>Microscopy</strong><br />

S. Koch, M. Langer, J. Lobo-Checa, Th. Brugger, S. Kawai, B. Such, E. Meyer, Th. Glatzel<br />

P.I-11 Kelvin probe force microscopy in application to organic thin films: frequency modulation,<br />

amplitude modulation, and hover mode KPFM<br />

B. Moores, F. Hane, L. M. Eng, Z. Leonenko<br />

P.I-12 Resolution enhanced multifrequency electrostatic force microscopy under ambient<br />

conditions<br />

X. D. Ding, J. B. Xu, J. X. Zhang<br />

P.I-13 Deconvolution and Tip Geometry Effects in <strong>Atomic</strong>- and Nanoscale Kelvin probe <strong>Force</strong><br />

<strong>Microscopy</strong><br />

G. Elias, Y. Rosenwaks, A. Boag, E. Meyer, Th. Glatzel<br />

P.I-14 Kelvin <strong>Force</strong> <strong>Microscopy</strong> Dynamic Behavior and Noise Propagation<br />

H. Diesinger, D. Deresmes, J.-P. Nys, Th. Mélin<br />

P.I-15 Charge transfer from doped silicon nanocrystals<br />

Ł. Borowik, Th. Nguyen-Tran, P. Roca i Cabarrocas, K. Kusiaku, D. Theron, H. Diesinger,<br />

D. Deremes, Th. Mélin


Poster Session I cont. (Tuesday)<br />

Instrumentation<br />

P.I-16 Open source scanning probe microscopy control and data analysis s<strong>of</strong>tware package<br />

Gxsm<br />

P. Zahl<br />

P.I-17 2 nd generation Dynamic Nanostencil AFM/DFM/STM for in-situ (UHV) resistless patterning<br />

<strong>of</strong> nanostructures<br />

P. Zahl, P. Sutter<br />

P.I-18 Design <strong>of</strong> a Variable Temperature Variable Magnetic Field <strong>Noncontact</strong> Scanning <strong>Force</strong><br />

Microscope for the Characterization <strong>of</strong> Nanoscale Electronic and Magnetic Phenomena<br />

P. Staffier, M. Liebmann, J. Falter, N. Pilet, Ch. Ahn, U. D. Schwarz<br />

P.I-19 An Active Q Control System in Scanning <strong>Force</strong> <strong>Microscopy</strong><br />

J. Kim, M. Zech, J. E. H<strong>of</strong>fman<br />

P.I-20 Besocke style quartz tuning fork FM-AFM/STM for use in UHV and low temperatures<br />

S. M. Huston, R. T. Port, K. M. Andrews, Th. P. Pearl<br />

P.I-21 Design <strong>of</strong> a Low Temperature <strong>Noncontact</strong> <strong>Atomic</strong> <strong>Force</strong> Microscope Combined with a Field<br />

Ion Microscope<br />

J. Falter, D.-A. Braun, H. Hölscher, U. D. Schwarz, A. Schirmeisen, H. Fuchs<br />

P.I-22 A homebuilt low-temperature STM / tuning fork AFM combination<br />

M. Lange, J. Schaffert, N. Wintjes, R. Möller<br />

P.I-23 Development <strong>of</strong> quartz force sensors for noncontact atomic force microscopy/spectroscopy<br />

K. Hori, T. Arai, M. Tomitori<br />

P.I-24 High-Speed Frequency Modulation <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong> using Wideband Digital<br />

Phase-Locked Loop Detector<br />

T. Fukuma, Y. Mitan<br />

Method Development<br />

P.I-25 Recent Advances in Multi-Spectral <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong><br />

S. Jesse, N. Balke, P. Maksymovych, O. Ovchinnikov, A.P. Baddorf, S.V. Kalinin<br />

P.I-26 Deciphering Nanoscale Interactions: Artificial Neural Networks and Scanning Probe<br />

<strong>Microscopy</strong><br />

M. Nikiforov, S. Jesse, O. Ovchinnikov, S. V. Kalinin<br />

P.I-27 NC-AFM study <strong>of</strong> a cleaved InAs (110) surface using modified Si probes under ambient<br />

atmospheric pressure<br />

Y. Jeong, M. Hirade, R. Kokawa, H. Yamada, K. Kobayashi, N. Oyabu, H. Yamatani, T.<br />

Arai, A. Sasahara, M. Tomitori<br />

P.I-28 Dual-Frequency-Modulation AFM Spectroscopy<br />

G. Chawla, C. A. Wright, S. D. Solares<br />

P.I-29 Theory <strong>of</strong> Multifrequency Method in FM-AFM<br />

Z. Ma, Y. Naitoh, Y. Li, M. Kageshima, Y. Sugawara<br />

P.I-30 Internal Resonances and Spatio-Temporal Instabilities in Nonlinear Multi-mode NC-AFM<br />

Dynamics<br />

O. Gottlieb, S. Hornstein, W. Wu, A. Shavit<br />

23


Poster Session I cont. (Tuesday)<br />

P.I-31 Frequency Noise in Frequency Modulation <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong><br />

K. Kobayashi, H. Yamada, K. Matsushige<br />

P.I-32 Relation between lateral forces and dissipation in FM-AFM<br />

M. Klocke, D. E. Wolf<br />

P.I-33 Experimental Study <strong>of</strong> Dissipation Mechanisms in AFM Cantilevers<br />

F. Zypman<br />

Nanolithography<br />

P.I-34 Ultrasonic Nanolithography on Hard Substrates<br />

M. T. Cuberes<br />

P.I-35 Silicon nanowire transistors with a channel width <strong>of</strong> 4 nm fabricated by atomic force<br />

microscope nanolithography<br />

J. Martinez, R. V. Martinez, R. Garcia<br />

P.I-36 Contacting self-ordered molecular wires by nanostencil lithography<br />

L. Gross, R. R. Schlittler, G. Meyer, Th. Glatzel, S. Kawai, S. Koch, E. Meyer<br />

24


POSTER Session II (Wednesday)<br />

Molecules<br />

25<br />

Chair: Y. Sugawara<br />

P.II-01 Molecular Structures <strong>of</strong> Organic Single Crystals Investigated by Frequency Modulation<br />

<strong>Atomic</strong> <strong>Force</strong> Microscopes<br />

T. Minato, H. Aoki, T. Wagner, K. Itaya<br />

P.II-02 Imaging <strong>of</strong> aromatic molecules by tuning-fork based LT-NC-AFM<br />

B. Such, Th. Glatzel, S. Kawai, S. Koch, A. Barat<strong>of</strong>f, E. Meyer, C. H. M. Amijs, P. de<br />

Mendoza, A. M. Echavarren<br />

P.II-03 One and Two Dimensional Structure <strong>of</strong> Water on Cu(110) and O/Cu(110)-(2x1) Surface<br />

B. Y. Choi, Y. Shi, T. Duden, M. Salmeron<br />

P.II-04 Dynamical simulations <strong>of</strong> truxene molecules adsorbed on the KBr (001) surface<br />

T. Trevethan, A. Shluger<br />

P.II-05 Temperature-dependent growth <strong>of</strong> C60 on CaF2(111)<br />

F. Loske, P. Maaß, J. Schütte, A. Kühnle<br />

P.II-06 Creating 1D nanostructures: Heptahelicene-carboxylic acid on Calcite<br />

P. Rahe, M. Nimmrich, J. Schütte, I. G. Stara, A. Kühnle<br />

P.II-07 <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong> Study <strong>of</strong> Cross-Linked C32H66 Monolayer by Low-Energy (10eV)<br />

Hyperthermal Bombardment<br />

Y. Liu, H.Y. Nie, D.Q. Yang, M.W. Lau, J. Yang<br />

P.II-08 Towards a molecule-based Ferroelectric-OFET: surface modification <strong>of</strong> PZT mediated<br />

through functionalized thiophene derivates<br />

P. Milde, K. Haubner, E. Jaehne, D. Köhler, U. Zerweck, L. M. Eng<br />

P.II-09 Imaging and Detection <strong>of</strong> Single Molecule Recognition Events on Organic Semiconductor<br />

Surfaces<br />

N. S. Losilla, J. Preiner, A. Ebner, P. Annibale, F. Biscarini, R. Garcia, P. Hinterdorfer<br />

P.II-10 Transverse conductance image <strong>of</strong> DNA probed by current-feedback noncontact AFM<br />

T. Matsumoto, Y. Maeda, T. Kawai<br />

P.II-11 Imaging Schwann Cell NGF Receptors using <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong><br />

R. Williamson, Cheryl Miller<br />

Liquids<br />

P.II-12 Comparative Studies on Water Structures on Hydrophilic and Hydrophobic Surfaces by<br />

FM-AFM<br />

K. Suzuki, N. Oyabu, K. Kobayashi, K. Matsushige, H. Yamada<br />

P.II-13 Molecular Resolution Investigation <strong>of</strong> Lysozyme Crystal in Liquid by Frequency-Modulation<br />

AFM<br />

K. Nagashima, M. Abe, S. Morita, N. Oyabu, K. Kobayashi, H. Yamada, R. Murai, H.<br />

Adachi, K. Takano, H. Matsumura, S. Murakami, T. Inoue, Y. Mori, M. Ohta, R. Kokawa<br />

P.II-14 <strong>Noncontact</strong> <strong>Atomic</strong> <strong>Force</strong> Microscope Observation <strong>of</strong> TiO2(110) Surface in Pure Water<br />

A. Sasahara, Y. Jeong, M. Tomitori<br />

P.II-15 Development <strong>of</strong> Multifrequency High-speed NC-AFM in Liquid<br />

Y. J. Li, K. Takahashi, N. Kobayashi, Y. Naitoh, M. Kageshima, Y. Sugawara


Poster Session II cont. (Wednesday)<br />

P.II-16 Frequency-Domain and Time-Domain Analyses <strong>of</strong> S<strong>of</strong>t-Matter Dynamics Using Wide-Band<br />

Magnetic Excitation AFM<br />

M. Kageshima, T. Ogawa, S. Kurachi, Y. Naitoh, Y. J. Li, Y. Sugawara<br />

P.II-17 <strong>Noncontact</strong> observation in liquid with van der Pol-type FM-AFM<br />

M. Kuroda, H. Yabuno, T. Someya, R. Kokawa, M. Ohta<br />

P.II-18 Development <strong>of</strong> a NC – AFM for Ambient and Liquid Environments<br />

H. I. Rasool, S. Sharma, J. K. Gimzewski<br />

P.II-19 Cantilever Holder Design for Spurious-Free Cantilever Excitation in Liquid by Piezoactuator<br />

H. Asakawa, T. Fukuma<br />

Oxides and Insulators<br />

P.II-20 The different faces <strong>of</strong> the calcite (10-14) surface<br />

J. Schütte, L. Tröger, P. Rahe, R. Bechstein, A. Kühnle<br />

P.II-21 Manipulation Mechanism <strong>of</strong> Single Cu Atoms on Cu(110)-O Surface with Low Temperature<br />

Non-Contact AFM<br />

Y. Kinoshita, T. Satoh, Y. J. Li, Y. Naitoh, M. Kageshima, Y. Sugawara<br />

P.II-22 Simultaneous NC-AFM/STM Imaging <strong>of</strong> the Surface Oxide Layer on Cu(100) and<br />

Identification <strong>of</strong> Lattice Sites<br />

M. Z. Baykara, T. C. Schwendemann, E. I. Altman, U. D. Schwarz<br />

P.II-23 nc-AFM Investigations <strong>of</strong> Metal Nanoclusters on α-alumina<br />

K. Venkataramani, M. C. R. Jensen, M. Reichling, F. Besenbacher, J. V. Lauritsen<br />

P.II-24 Atom-resolved AFM studies <strong>of</strong> the polar MgAl2O4 (001) surface<br />

M. K. Rasmussen, J. V. Lauritsen, F. Besenbacher<br />

P.II-25 Contrast formation on cross-linked (1x2) reconstructed titania (110)<br />

H. H. Pieper, S. Torbrügge, S. Bahr, K. Venkataramani, A. Kühnle, M. Reichling<br />

P.II-26 Nano volcanoes – the surface structure <strong>of</strong> antimony-doped TiO2(110)<br />

R. Bechstein, M. Kitta, J. Schütte, H. Onishi, A. Kühnle<br />

Electronic and Magnetic Properties<br />

P.II-27 Non-contact scanning nonlinear dielectric microscopy imaging <strong>of</strong> TiO2 (110) surfaces<br />

N. Kin, Y. Cho<br />

P.II-28 Characterizations <strong>of</strong> Carbon Material by Non-contact Scanning Non-linear Dielectric<br />

<strong>Microscopy</strong><br />

S. Kobayashi, Y. Cho<br />

P.II-29 Local Dielectric Spectroscopy <strong>of</strong> Nanocomposite Materials Interfaces<br />

M. Labardi, D. Prevosto, S. Capaccioli, M. Lucchesi, P.A. Rolla<br />

P.II-30 Probing Local Bias-Induced Phase Transitions on the Single Defect Level: from Imaging to<br />

Deterministic Mechanisms<br />

N. Balke, S. Jesse, P. Maksymovych, Y.H. Chu, R. Ramesh, S. Choudhury, L.Q. Chen,<br />

S.V. Kalinin<br />

P.II-31 Polarization-dependent electron tunneling into ferroelectric surfaces<br />

P. Maksymovych, S. Jesse, P. Yu, R. Ramesh, A. P. Baddorf, S. V. Kalinin<br />

P.II-32 Local ferroelectric and magnetic investigations on multiferroic thin films<br />

U. Zerweck, D. Köhler, P. Milde, Ch. Loppacher, S. Geprägs, S.T.B. Goennenwein, R.<br />

Gross, L.M. Eng<br />

26


Poster Session II cont. (Wednesday)<br />

P.II-33 Magnetic Resonance <strong>Force</strong> <strong>Microscopy</strong> in Anisotropic Systems<br />

T. Fan, V. I. Tsifrinovich<br />

P.II-34 Sub-10 nm resolution in Magnetic <strong>Force</strong> <strong>Microscopy</strong> (MFM) at ambient conditions<br />

Ö. Karcı, H. Atalan, M. Dede, Ü. Çelik, A.Oral<br />

P.II-35 Enhancement <strong>of</strong> the Exchange-bias Effect based on Quantitative Magnetic <strong>Force</strong><br />

<strong>Microscopy</strong> Results<br />

N. Pilet, M.A. Marioni, S. Romer, N. Joshi, S. Özer, H.J. Hug<br />

27


Oral<br />

Presentations<br />

Monday, 10 August<br />

28


Non-retarded and retarded interactions between dielectric cylinders<br />

Mo-0900<br />

R. Podgornik 1,2 , Antonio Šiber 3 , Rick Rajter 4 , Roger H. French 5 , W.Y. Ching 6 , and V.<br />

Adrian Parsegian 2<br />

1 Faculty <strong>of</strong> Mathematics and Physics, University <strong>of</strong> Ljubljana, Ljubljana, Slovenia and Department<br />

<strong>of</strong> Theoretical Physics, J. Stefan Institute, SI-1000 Ljubljana, Slovenia<br />

2 Laboratory <strong>of</strong> Physical and Structural Biology, NICHD, National Institutes <strong>of</strong> Health, Bldg. 9,<br />

Room 1E116, Bethesda, Maryland 20892-0924, USA.<br />

3 Institute <strong>of</strong> Physics, P.O. Box 304, 10001 Zagreb, Croatia<br />

4 Department <strong>of</strong> Materials Science and <strong>Engineering</strong>, Massachusetts Institute <strong>of</strong> Technology, Room 13-<br />

5046 Cambridge, Massachusetts 02139, USA<br />

5 DuPont Co. Central Research, Experimental Station, E400-5207 Wilmington, Delaware 19880, USA<br />

6 Department <strong>of</strong> Physics, University <strong>of</strong> Missouri-Kansas City, Kansas City, Missouri, 64110, USA<br />

I will present a complete theory <strong>of</strong> non-retarded and retarded van der Waals<br />

interactions between dielectric cylinders. It is based on the Lifshitz formulation <strong>of</strong> the<br />

interactions between two anisotropic semiinfinite media as was worked out by Yu.<br />

Barash [1] in the complete retarded case. One then recasts the two semiinfinite media<br />

problem into two anisotropic cylinders problem by expanding the dielectric response<br />

function as a function <strong>of</strong> the density <strong>of</strong> the dielectric cylinders that constitute the two<br />

media. The first order in the density expansion yields the semiinfinite plane-cylinder<br />

interaction and the second order term the cylinder-cylinder interaction. Explicit formulae<br />

are obtained for this interaction and the interaction is evaluated for different examples <strong>of</strong><br />

ab initio carbon nanotube dielectric response functions. The non-retarded case has<br />

already been discussed in our previous publications [2,3]. I will thus concentrate on the<br />

features <strong>of</strong> the van der Waals interaction that are introduced by the retardation and<br />

orientation effects.<br />

[1] Yu. S. Barash, Izv. Vyssh. Uchebn. Zaved. Radi<strong>of</strong>iz. 21 163 (1978). J. N. Munday, D. Iannuzzi, Y.<br />

Barash and F. Capasso, Phys. Rev. A 71, 042102 (2005). Erratum <strong>of</strong> the paper J. N. Munday, D.<br />

Iannuzzi, Y. Barash and F. Capasso, Phys. Rev. A 71, 042102 (2005).<br />

[2] Rick F. Rajter, Rudi Podgornik, V. Adrian Parsegian, Roger H. French, and W. Y. Ching: van der<br />

Waals-London dispersion interactions for optically anisotropic cylinders: Metallic and<br />

semiconducting single-wall carbon nanotubes, Phys. Rev B 76, 045417 (2007).<br />

[3] Rick Rajter, Roger H. French, Rudi Podgornik, W. Y. Ching, and V. Adrian Parsegian, Spectral mixing<br />

formulations for van der Waals–London dispersion interactions between multicomponent carbon<br />

nanotubes, Journal <strong>of</strong> Applied Physucs 104, 053513 (2008).<br />

29


Noah Graham 1<br />

Casimir <strong>Force</strong>s from Scattering Theory<br />

1 Department <strong>of</strong> Physics, Middlebury College, Middlebury, VT USA<br />

Mo-0930<br />

Because the parallel-plate geometry <strong>of</strong> Casimir's original calculation is one <strong>of</strong> the most<br />

difficult configurations in which to study the Casimir force experimentally, it is important<br />

to be able to extend this calculation to other situations. I will describe a general set <strong>of</strong><br />

techniques that allow one to obtain precise theoretical predictions <strong>of</strong> the Casimir force for<br />

a wide range <strong>of</strong> geometries and materials. It applies to any situation where the scattering<br />

matrix <strong>of</strong> each individual object can be calculated (or measured), in any basis for which<br />

the decomposition <strong>of</strong> a plane wave is known. Although this approach is simplest at large<br />

distances, it also yields precise results at any separation, as long as the objects do not<br />

overlap in the radial coordinates <strong>of</strong> the bases used for each object. In addition to the<br />

usual situation where the objects are outside <strong>of</strong> each other, these results extend to the case<br />

<strong>of</strong> one object inside another.<br />

The work I will report on has been done in collaboration with Thorsten Emig<br />

(Paris/Cologne) Robert L. Jaffe (MIT), Mehran Kardar (MIT), S. Jamal Rahi (MIT), and<br />

Saad Zaheer (MIT).<br />

30


Mo-0955<br />

Multiple Scattering Casimir <strong>Force</strong> Calculations between Layered and<br />

Corrugated Materials<br />

Kimball A. Milton 1 , Inés Cavero-Peláez 2 , Prachi Parashar 1 , K.V. Shajesh 3 ,<br />

and Jef Wagner 1<br />

1 H.L. Dodge Department <strong>of</strong> Physics and Astronomy, University <strong>of</strong> Oklahoma, Norman, OK 73019, USA<br />

2 Laboratoire Kastler Brossel, Université Pierre et Marie Curie, F-75252 Paris, France<br />

3 St Edward's <strong>School</strong>, Vero Beach, FL 32963, USA<br />

Multiple scattering methods have recently proven useful in calculating quantum vacuum<br />

forces between distinct bodies. In fact, 40 years ago such an approach was used to derive<br />

the Lifshitz formula for the force between parallel dielectric slabs. More generally,<br />

numerical results can be readily obtained in many cases, but more remarkably, for weak<br />

coupling (e.g., for dilute dielectrics), closed-form exact expressions can be derived,<br />

reflecting the summation <strong>of</strong> Casimir-Polder forces or their analogues. We have recently<br />

used such methods to derive forces between corrugated planar and cylindrical surfaces.<br />

(See Fig. 1.) For scalar fields with Dirichlet boundary conditions, the forces can be<br />

computed perturbatively in the corrugation amplitudes; 4 th order perturbative results<br />

agree closely with the exact results for weak coupling. We are now extending such<br />

calculations to electromagnetism and to general multilayered surfaces (see Fig. 2), where<br />

again exact results can be obtained in many cases. These results will have important<br />

applications to experimentally accessible situations, and to nanomachinery. The<br />

extension to finite temperature will be explored in the near future.<br />

Fig. 1 Fig. 2<br />

Figure 1: Concentric corrugated gears. The corrugations have equal spatial frequency. The stable<br />

equilibrium configuration occurs when the outer troughs align with the inner peaks. If the gears<br />

are misaligned, a torque is exerted on one cylinder due to the other.<br />

Figure 2: Multilayered surface. The Casimir energy between two such layered potentials can be<br />

calculated in closed form.<br />

31


Drude corrections to Casimir force calculations in liquids<br />

Raul Esquivel-Sirvent 1<br />

1 Instituto de Física, Universidad Nacioanl Autónoma de México, México D.F. 01000.<br />

Mo-1050<br />

The Casimir force for metals immersed in different fluids has been measured recently<br />

[1,2,3]. The comparison with theory is based on the Lifshitz formula and the knowledge<br />

<strong>of</strong> the dielectric function <strong>of</strong> the involved materials. In the case <strong>of</strong> the reported<br />

experiments, Au is a common metal used in the force measurements. The data for the<br />

dielectric function <strong>of</strong> Au can be obtained from tabulated data with a low frequency<br />

interpolation using the Drude model. The data for the optical properties <strong>of</strong> Au are usually<br />

measured in air or partial vacuum. However, changes in the Drude parameters <strong>of</strong> metals<br />

when immersed in fluids have been reported in the literature. Gugger [4] using the<br />

method <strong>of</strong> total attenuated internal reflection measured the dielectric function <strong>of</strong> Ag films<br />

in contact with different liquids. The measurements showed a change in the dielectric<br />

function <strong>of</strong> the Ag film depending on the index <strong>of</strong> refraction <strong>of</strong> the fluids. The same<br />

conclusion was reached in the work <strong>of</strong> Chen [5] that by means <strong>of</strong> spectroscopic<br />

ellipsometry measured the Drude parameters for Au and Ag films immersed in liquids,<br />

confirming the change in the Drude parameters <strong>of</strong> the metals.<br />

In this paper we examine the change in the Casimir force between two Au surfaces<br />

immersed in different fluids and how the change in the Drude parameters affects the<br />

Casimir force calculations. In particular, we show that variations <strong>of</strong> the order <strong>of</strong> 8% are<br />

possible and observed discrepancies between theory and experiment can be explained<br />

by this effect. The change <strong>of</strong> the Drude parameters <strong>of</strong> a metal immersed in a fluid are<br />

described by a Bruggeman effective medium theory. Finally, we discuss the possible<br />

applications <strong>of</strong> effective medium theories within the Lifshitz formalism.<br />

[1] J. N. Munday and F. Capasso, Phys. Rev. A 75, 060102(R) (2007).<br />

[2] J. N. Munday, F. Capasso, V. A. Parseguian and S. M. Bezrukov, Phys. Rev. A 78,<br />

029906 (2008).<br />

[3] P. J. van Zwol, G. Palasantzas and J. Th. M. De Hosson, Phys. Rev. B 79, 195428<br />

(2009).<br />

[4] H. Guger, M. Jurich and J. D. Swalen, Phys. Rev. B 30, 4189 (1984).<br />

[5] L. Y. Chen and D. W. Lynch, Phys. Rev. B 36, 1425 (1987).<br />

32


Diego Dalvit 1<br />

Dispersive Casimir interactions between atoms and surfaces<br />

1 Theoretical Devision, MS B213 , Los Alamos National Laboratory, Los Alamos, NM 87545, USA.<br />

Mo-1115<br />

Casimir atom-surface interactions may be important in experiments involving atoms in<br />

proximity to surfaces, such as atom chips for quantum information processing. In this talk<br />

I will review a general scattering approach to Casimir atom-surface forces, and give some<br />

examples <strong>of</strong> its utility in the calculation <strong>of</strong> non-trivial geometrical effects <strong>of</strong> the quantum<br />

vacuum, such as lateral Casimir-Polder forces. I will then briefly describe two proposals<br />

to use Bose-Einstein condensates (BEC) to probe lateral Casimir forces for atoms above<br />

corrugated surfaces: a "BEC cantilever" sensitive to gradients <strong>of</strong> the atom-surface force,<br />

and BEC Bragg spectroscopy directly sensitive to the atom-surface potential.<br />

33


Mo-1140<br />

Van der Waals with a twist: how nanotube chirality impacts interaction<br />

Rick F. Rajter 1<br />

strength.<br />

1Department <strong>of</strong> Materials Science and <strong>Engineering</strong>, Massachusetts Institute <strong>of</strong> Technology, Cambridge,<br />

MA, USA<br />

The van der Waals - London dispersion interaction depends on the geometries and optical<br />

properties <strong>of</strong> all materials or components present within a given system. The ongoing<br />

investigation <strong>of</strong> the material property contribution to this interaction continues to yield<br />

novel and sometimes surprising phenomenon. A chronological review <strong>of</strong> these<br />

developments helps underline the importance <strong>of</strong> material property thinking and why these<br />

effects must be considered. When first discovered, the differences in interaction strength<br />

were largely attributed to variations in atomic composition, such as that between various<br />

noble gases. Once Lifshitz quantified the connection <strong>of</strong> interaction strength to the<br />

system’s optical properties, differences in said interactions were also predicted and<br />

confirmed for allotropes. Perhaps the most obvious example is that <strong>of</strong> carbon, which can<br />

take the form <strong>of</strong> diamond, graphite, graphene, nanotubes, buckyballs, ribbons, etc. Here,<br />

it was the bond type that was the differentiating factor. The most recent discovery was<br />

that bond twisting or bending within a given allotrope (i.e. carbon nanotubes) could lead<br />

to equally dramatic changes. One <strong>of</strong> the unique features <strong>of</strong> carbon nanotubes is that a<br />

very small change in the chirality or twist can greatly change the electronic structure<br />

properties, which are the key determining factor a material's optical properties. This<br />

work will show many examples <strong>of</strong> chirality-dependent optical properties and how this<br />

information can be used to create experiments to exploit these differences for a variety <strong>of</strong><br />

applications.<br />

34


Using Casimir and capillary forces to model adhesion <strong>of</strong> MEMS<br />

cantilevers<br />

Maarten P. de Boer 1 and Frank W. DelRio 2<br />

1 MEMS Technology Dept. Sandia National Laboratories, Albuquerque, NM, USA<br />

2 National Institute <strong>of</strong> Standards and Technology, Gaithersburg, MD, USA<br />

Mo-1205<br />

Long-range Casimir forces set the lower limit <strong>of</strong> unwanted adhesion <strong>of</strong> microscale<br />

cantilevers to a substrate [1]. To quantify this adhesion, we measure deflections <strong>of</strong><br />

actuated MEMS cantilevers (Fig. 1) by interferometry (Fig. 2), and compare to an<br />

energy-release rate model. To understand the values, we develop a detailed model <strong>of</strong> the<br />

interface including surface roughness, contact mechanics and dispersion forces. We find<br />

that when surface roughness is less than 3 nm root mean square, the dispersion forces<br />

dominate adhesion. As surface roughness increases, the real area <strong>of</strong> contact dominates.<br />

Agreement between theory and model is within ± 20% when correlations between the<br />

upper and lower surfaces are taken into account [2]. Adhesion due to capillary forces is<br />

much greater than that from van der Waals forces. In that case, constitutive laws that<br />

incorporate disjoining pressure are needed to explain these very high values [3]. Linking<br />

single asperity models to rough surface models will further develop our understanding in<br />

these areas.<br />

Fig. 1 Cantilever adhesion geometry Fig. 2 Interferograms <strong>of</strong> microcantilevers<br />

References:<br />

[1] F. W. DelRio, M. P. de Boer, J. A. Knapp, E. D. Reedy, P. J. Clews and M. L. Dunn, “The role <strong>of</strong> van<br />

der Waals forces in adhesion <strong>of</strong> micromachined surfaces”, Nature Materials, 4 629 (2005).<br />

[2] F. W. DelRio, M. L. Dunn, L. M. Phinney, C. J. Bourdon and M. P. de Boer, “Rough surface adhesion<br />

in the presence <strong>of</strong> capillary adhesion”, Applied Physics Letters, 90 163104 (2007).<br />

[3] F. W. DelRio, M. L. Dunn and M. P. de Boer, “Capillary adhesion model for contacting<br />

micromachined surfaces. Scripta Materialia, (2008).<br />

Sandia is a multiprogram laboratory operated by Sandia Corporation, a Lockheed Martin<br />

Company, for the United States Department <strong>of</strong> Energy’s National Nuclear Security Administration<br />

under contract DE-AC04-94AL85000.<br />

35


Measuring “Virtual Photon” <strong>Force</strong>s with “Real Photon” <strong>Force</strong>s<br />

Hong X. Tang<br />

Department <strong>of</strong> Electrical <strong>Engineering</strong>, <strong>Yale</strong> University, New Haven, USA<br />

Mo-1430<br />

We present an integrated, all-optical nanomechanical device platform to study Casimir<br />

effect. A versatile optical force 1,2 arising from guided lightwaves (or real photons)<br />

structure is harnessed to provide accurate counter measure <strong>of</strong> the Casimir force (the force<br />

arising from virtual photons). This optical NEMS platform eliminates the electrical<br />

connections to the devices. The interacting surface can be insulator, semiconducting or<br />

metallic. The residual potential problem, which has been recently identified to be a major<br />

source <strong>of</strong> errors in Casimir measurement, is circumvented by fabricating an electrical link<br />

between two interacting surface to null out contact potentials and trapped charges.<br />

c)<br />

Si Si<br />

a)<br />

b)<br />

c)<br />

si<br />

metal metal si<br />

actuator<br />

Metal<br />

Metal<br />

sensor<br />

Figures: All-optical schemes for on-chip measurement <strong>of</strong> Casimir forces. Figure 1: Repulsive optical force<br />

is used to counter the attractive Casimir force between two silicon beams. Figure 2: Optical force is used to<br />

balance the Casimir force between two metal beams.<br />

[1] Mo Li , W. Pernice, C. Xiong, T. Baehr-Jones, M. Hochberg, H. Tang , “Harnessing optical forces in<br />

integrated photonic circuits.”, Nature , 456, 480(2008)<br />

[2] M. Li. W. Pernice, H. Tang, “Tunable bipolar interactions between guided lightwaves” Nature<br />

Photonics (under review, 2009), see also: arxiv:0903.5117<br />

36


Near-field radiative heat transfer<br />

Mo-1455<br />

Alessandro Siria 1, 2 , Emmanuel Rousseau 3 , Jean-Jacques Greffet 3 and Joel Chevrier 1<br />

1 Institut Néel, CNRS and Université Joseph Fourier, 38042 Grenoble France<br />

2 CEA-LETI/MINATEC, 17 avenue des Martyrs 38042 Grenoble France<br />

3 Laboratoire Charles Fabry de L’Institut d’Optique, CNRS UMR, 91127 Palaiseau France<br />

Near-field force and energy exchange between two objects due to quantum and thermal<br />

induced electrodynamic fluctuations give rise to interesting phenomena, such as Casimir<br />

force and thermal radiative transfer exceeding Plank’s theory <strong>of</strong> blackbody radiation. A<br />

theoretical explanation, in the framework <strong>of</strong> stochastic electrodynamics introduced by<br />

Rytov [1] in the late sixties, accounts for quantum and thermodynamic fluctuations. This<br />

theory has been successfully applied to model Casimir forces [2] and radiative heat<br />

transfer [3]. While Casimir force has its origin in quantum fluctuations, related to zero<br />

point energy, near-field radiative heat transfer is only due to classical thermodynamics<br />

fluctuations.<br />

Although significant progresses have been made in the past on the precise measurement<br />

<strong>of</strong> the Casimir force [4, 5], a detailed quantitative comparison between theory and<br />

experiments in the nanometer regime is still lacking when speaking about heat transfer.<br />

Here, we report experimental data on the thermal flux spatial dependence. Theory based<br />

on the Derjaguin approximation, is successfully used here for the first time to describe<br />

radiative heat transfer from the far field to the near field regimes. It reproduces the<br />

measured dependence with an agreement better than 4 % for gaps varying between 40 nm<br />

and 5 μm.<br />

[1] Rytov, S.M., Kratsov, Yu.A., Tatarskii, V.I. Principles <strong>of</strong> statistical Radiophysics, vol<br />

3, Springer-Verlag, New-York, (1987) (Chapter 3)<br />

[2] Lifshitz, E. M. The theory <strong>of</strong> molecuar attractive forces between solids. Zh. Eksp.<br />

Teor. Fiz. 29, 94 (1955) [Sov. Phys. JETP 2, 73 (1956)].<br />

[3] A. V. Shchegrov, K. Joulain, R. Carminati, and J.-J. Greffet, Phys. Rev. Lett. 85, 1548 (2000)<br />

[4] U. Mohideen and A. Roy, Physical Review Letters 81, 4549 (1998)<br />

[5] G. Jourdan, A. Lambrecht, F. Comin, and J. Chevrier, EPL 85, 3, 31001 (2009)<br />

37


Near-field radiative transfer measurements and implications for<br />

Casimir force measurements<br />

Arvind Narayanaswamy 1 , Sheng Shen 2 , Ning Gu 3 , and Gang Chen 2<br />

1 Department <strong>of</strong> Mechanical <strong>Engineering</strong>, Columbia University, New York, USA<br />

2 Department <strong>of</strong> Mechanical <strong>Engineering</strong>, Massachusetts Institute <strong>of</strong> Technology, Cambridge, USA<br />

1 Department <strong>of</strong> Electrical <strong>Engineering</strong>, Columbia University, New York, USA<br />

Mo-1520<br />

Near–field force and energy exchange between two objects due to quantum electrodynamic<br />

fluctuations give rise to interesting phenomena such as Casimir and van der Waals forces,<br />

and thermal radiative transfer exceeding Planck’s theory <strong>of</strong> blackbody radiation. Although<br />

significant progress has been made in the past on the precise measurement <strong>of</strong> Casimir force<br />

related to zero-point energy, experimental demonstration <strong>of</strong> near-field enhancement <strong>of</strong><br />

radiative heat transfer is difficult. In this work, we present a sensitive technique <strong>of</strong> measuring<br />

near–field radiative transfer between a microsphere and a substrate using a bi–material<br />

atomic force microscope (AFM) cantilever, resulting in “heat transfer-distance” curves.<br />

Measurements <strong>of</strong> radiative transfer between a sphere and a flat substrate show the presence <strong>of</strong><br />

strong near–field effects resulting in enhancement <strong>of</strong> heat transfer over the predictions <strong>of</strong> the<br />

Planck blackbody radiation theory. We have been able to identify unambiguously the<br />

contribution <strong>of</strong> electromagnetic surface phonon polaritons to near-field radiative transfer. The<br />

implications <strong>of</strong> measurement <strong>of</strong> near-field radiative heat transfer for determining <strong>of</strong> the<br />

magnitude <strong>of</strong> the thermal component <strong>of</strong> the Casimir force will be discussed.<br />

38


Mo-1615<br />

Tricks and facts in a high precision measurement <strong>of</strong> the Casimir force<br />

with transparent conductors<br />

S. de Man 1 , K. Heeck 1 , R. J. Wijngaarden 1 and D. Iannuzzi 1<br />

1 Department <strong>of</strong> Physics and Astronomy, Faculty <strong>of</strong> Sciences, VU University Amsterdam, The Netherlands<br />

The most promising and simple tool to tailor the Casimir force is to alter the behavior <strong>of</strong><br />

quantum fluctuations by properly choosing the materials on the interacting surfaces.<br />

According to the Lifshitz theory, the interaction between two objects depends on their<br />

dielectric functions. Transparent dielectrics, for example, attract less than reflective<br />

mirrors. Unfortunately, transparent dielectrics are prone to charge accumulation. Even a<br />

small amount <strong>of</strong> charges can give rise to electrostatic forces that easily overcome the<br />

Casimir attraction. For this reason, most <strong>of</strong> the Casimir force experiments reported so far<br />

have been limited to the investigation <strong>of</strong> surfaces coated with metal. In that case,<br />

however, there is not much room to tune the interaction strength, because the diversity in<br />

the dielectric functions <strong>of</strong> different metals is simply not large enough.<br />

In this paper we present a precise experiment where we have investigated the Casimir<br />

force between a gold coated sphere and a glass plate coated with either a thick gold layer<br />

or a highly conductive, transparent oxide film [1]. The measurements were performed in<br />

air, and no electrostatic force due to residual charges was observed over several weeks in<br />

either case. The decrease <strong>of</strong> the Casimir force due to the different dielectric properties <strong>of</strong><br />

the reflective gold layer and the transparent oxide film resulted to be as high as 40%−<br />

50% at all explored separations (from 50 to 150 nm), the largest modification <strong>of</strong> the<br />

Casimir force ever observed at ambient conditions.<br />

In my talk, I will review the new experimental technique we have developed to<br />

simultaneously (i) calibrate the absolute separation between the two surfaces and the<br />

force sensitivity <strong>of</strong> the setup, (ii) compensate for and measure the contact potential, and<br />

(iii) measure the Casimir force. I will comment on the behavior <strong>of</strong> the contact potential as<br />

a function <strong>of</strong> both absolute surface separation and time. I will also address the issues<br />

related to mechanical drifts, and how we are able to control those down to a level <strong>of</strong> 1<br />

nm/hour [2]. Finally, I will discuss how our experimental technique allows us to double<br />

check the electrostatic calibration procedure by investigating the hydrodynamic force that<br />

arises from the cushion <strong>of</strong> air between the two interacting surfaces [1].<br />

[1] S. de Man, K. Heeck, R. J. Wijngaarden, and D. Iannuzzi. arXiv:0901.3720 (2009)<br />

[2] S. de Man, K. Heeck, and D. Iannuzzi. Phys Rev A 79, 024102 (2009)<br />

39


Mo-1640<br />

Measurements <strong>of</strong> the Casimir force gradient by AFM for Different<br />

Materials<br />

Gauthier Torricelli 1 , Stuart Thornton 1 , and Chris Binns 1<br />

1 Department <strong>of</strong> Physics and Astronomy, University <strong>of</strong> Leicester, UK.<br />

We present here quantitative measurements <strong>of</strong> the Casimir force gradient in the 50-600<br />

nm range using a commercial <strong>Atomic</strong> <strong>Force</strong> Microscope operating in UHV (VT AFM<br />

Omicron). The measurements were done in the sphere-plate geometry between a Au<br />

sphere and plates consisting <strong>of</strong> three different classes <strong>of</strong> materials, that is a metal (Au), a<br />

semimetal (HOPG) and a quasicrystal (AlPdMn). The variation in the optical properties<br />

<strong>of</strong> the materials produces clearly observed differences in the Casimir force as predicted<br />

by calculations using the Lifshitz formula. We will discuss in detail the method <strong>of</strong> our<br />

measurements, for which some <strong>of</strong> the key points are listed below:<br />

• The calibration <strong>of</strong> the system is performed using the electrostatic interaction<br />

without any contact between the sphere and the surface.<br />

• The electrostatic interaction is also used to verify the stability <strong>of</strong> our<br />

measurements notably if there is any variation <strong>of</strong> the contact potential.<br />

• A feedback loop is used between two acquisitions in order to prevent any<br />

changes in the distance separation which can be induce by thermal drift<br />

• Several measurements are performed in different area <strong>of</strong> the sample in order to<br />

test the reproducibility <strong>of</strong> the measurement.<br />

40


Measuring the topological dependence <strong>of</strong> the Casimir force on<br />

nanostructured silicon surfaces<br />

Mo-1705<br />

Ho Bun Chan 1 , Y. Bao 1 , J. Zou 1 , R. A. Cirelli 2 , F. Klemens 2 , W. M. Mansfield 2 , and C.<br />

S. Pai 2<br />

1 Department <strong>of</strong> Physics, University <strong>of</strong> Florida, Florida, USA<br />

2 Bell Laboratories, Lucent Technologies, New Jersey, USA.<br />

The Casimir force is usually regarded as an extension <strong>of</strong> the van der Waals (vdW)<br />

interaction between molecules in the retarded limit. By dividing two solid plates into<br />

small elementary constituents and summing up the vdW force, it is possible to recover<br />

the distance dependence <strong>of</strong> the Casimir force between them. For smooth curved surfaces,<br />

the Casimir force is <strong>of</strong>ten calculated using the proximity force approximation (PFA)<br />

when the separation is small. However, the PFA breaks down when the deformation is<br />

strong. In fact, one important characteristic <strong>of</strong> the Casimir force is its strong dependence<br />

on geometry. The Casimir energy for a conducting spherical shell or a rectangular box<br />

has been calculated to have opposite sign to parallel plates. Whether such geometries<br />

exhibit repulsive Casimir forces remains a topic <strong>of</strong> current interest.<br />

We performed an experiment to demonstrate the strong geometry dependence <strong>of</strong> the<br />

Casimir force [1]. The interaction between a gold sphere and a silicon surface with an<br />

array <strong>of</strong> nanoscale, rectangular corrugations is measured using a micromechanical<br />

torsional oscillator. Deviation <strong>of</strong> up to 20% from PFA is observed. The data will be<br />

compared to recent theories on strongly deformed surfaces. In particular, the interplay<br />

between finite conductivity corrections and geometry effects complicates the comparison<br />

<strong>of</strong> data with theories. Ongoing efforts on shallow corrugations will also be presented.<br />

a) b) c)<br />

Figure 1: (a) Cross section <strong>of</strong> the nanoscale rectangular trenches on a silicon surface. (b)<br />

Schematic <strong>of</strong> the experimental setup (not to scale) including the micromechanical torsional<br />

oscillator, gold spheres and silicon trench array. (c) Ratio ρ <strong>of</strong> the measured Casimir force<br />

gradient to the force gradient expected from PFA, for two samples with trench arrays with<br />

different periodicity. The lines represent calculations based on perfect conductors<br />

[1] H. B. Chan, Y. Bai, J. Zou, R. A. Cirelli, F. Klemens, W. M. Mansfiled and C. S. Pai, Phys. Rev. Lett.<br />

100, 030401 (2008).<br />

41


Mo-1730<br />

Short range Casimir force measurements under ambient conditions and<br />

liquid environments<br />

P.J. van Zwol 1 , V.B. Svetovoy 2 , G. Palasantzas 1 , J. Th. M. De Hosson 1<br />

1 Materials Innovation Institute and Zernike Institute for Advanced Materials<br />

University <strong>of</strong> Groningen, 9747 AG Groningen, The Netherlands<br />

2 MESA_Research Institute, University <strong>of</strong> Twente, PO 217, 7500 AE Enschede, The Netherlands<br />

We briefly review ellipsometry [1] and our force measurement in the sphere plane<br />

geometry [2,3] on smooth and rough gold surfaces, and focus specifically on the contact<br />

point between rough surfaces, and the optical quality <strong>of</strong> the surfaces in use. Furthermore<br />

we outline inverse colloid probe force measurements [4] in liquid environment (fig 1), for<br />

which we found local charge variation <strong>of</strong> glass spheres resulting in varying strengths <strong>of</strong><br />

the electrical double layer. In addition we show that when the dielectric function <strong>of</strong> the<br />

liquid becomes comparable to that <strong>of</strong> the surfaces, resulting Casimir forces are very<br />

small. However sample dependence and uncertainty in the measured dielectric functions<br />

indicate that the theory is even more uncertain than the measurements, even if measured<br />

dielectric data is available, for all materials, over all relevant frequency ranges. In the<br />

extreme limit this can lead to the weird situation that small variations in the dielectric<br />

data can lead to a complete change <strong>of</strong> shape <strong>of</strong> the force, flipping between atractive and<br />

repulsive, and multiple atractive-repulsive transitions with distance.<br />

Figure 1: Optical image <strong>of</strong> sintered borisilicate glass spheres on glass.<br />

[1] V. B. Svetovoy, P.J. van Zwol, G. Palasantzas, J. Th. M. DeHosson, Phys. Rev. B 77, 035439 (2008)<br />

[2] P.J. van Zwol, G. Palasantzas, J. Th. M. DeHosson, Phys. Rev. B 77, 075412 (2008)<br />

[3] P.J. van Zwol, G. Palasantzas, M. van de Schootbrugge, J. Th. M. De Hosson, Appl. Phys. Lett. 92, 054101 (2008)<br />

[4] P.J. van Zwol, G. Palasantzas, J. Th. M. DeHosson, To appear in Phys. Rev. E (2009).<br />

42


Mo-1755<br />

Contact potential difference (CPD) in a Casimir force measurement:<br />

How do we deal with it?<br />

W. J. Kim 1 , A. Sushkov 1 , D. A. R. Dalvit 2 , S. K. Lamoreaux 1<br />

1 P. O. Box 208120, Department <strong>of</strong> Physics, <strong>Yale</strong> University, New Haven, CT 06510-8120<br />

2 Theoretical Devision, MS B213 , Los Alamos National Laboratory, Los Alamos, NM 87545, USA<br />

In order to achieve the minimum force condition for a pair <strong>of</strong> conducting plates in electric<br />

force microscopy, one must always apply a non-zero value <strong>of</strong> electric potential difference<br />

across the plates. This so-called contact potential difference (CPD) is ubiquitously<br />

present in all force measurements and poses an enormous challenge for those wishing to<br />

accomplish precision quantification <strong>of</strong> a force <strong>of</strong> non-electrostatic origin, such as van der<br />

Waals force and Casimir force.<br />

In this talk, we will briefly review physical origins <strong>of</strong> CPD and how it has been taken into<br />

previous analysis <strong>of</strong> short-range force measurements. We will also discuss CPD in a<br />

broader context by presenting some <strong>of</strong> the experimental studies undertaken in other<br />

subfields in physics that are challenged by the similar surface potential effects including<br />

gravitational wave missions like LIGO and LISA as well as our recent measurement <strong>of</strong><br />

the Casimir force between Ge plates at <strong>Yale</strong>.<br />

43


Oral<br />

Presentations<br />

Tuesday, 11 August<br />

44


3D Scanning <strong>Force</strong> <strong>Microscopy</strong> at Solid/Liquid Interface<br />

Takeshi Fukuma 1,2 and Yasumasa Ueda 1<br />

1 Frontier Science Organization, Kanazawa University, Kanazawa, Japan<br />

2 PRESTO, Japan Science and Technology Agency, Kawaguchi, Japan.<br />

Tu-0920<br />

The solid/liquid interface inherently has a three-dimensional (3D) extent in XYZ<br />

directions. This is particularly evident when solvation layers exist at the interface [1]. In<br />

frequency modulation atomic force microscopy (FM-AFM), a tip is scanned in XY<br />

directions to produce a 2D height (or Δf ) image having no vertical extent in Z direction.<br />

Therefore, the information contained in a 2D FM-AFM image is insufficient for<br />

understanding 3D distribution <strong>of</strong> solvent molecules interacting with the atoms or<br />

molecules constituting the solid surface. In an effort to resolve this issue, 3D force<br />

mapping technique has been proposed, where a number <strong>of</strong> force curves are measured at<br />

arrayed XY positions. This, however, results in a considerable increase <strong>of</strong> imaging time<br />

and hence has been used mainly in vacuum at low temperatures.<br />

In this study, we propose a method referred to as “3D scanning force microscopy (3D-<br />

SFM)”, where a tip is scanned in Z as well as XY directions. While 3D force mapping is<br />

based on 1D force spectroscopy, 3D-SFM is based on 2D imaging technique. This<br />

methodological advancement, together with our recent development <strong>of</strong> high-speed<br />

scanner and frequency detector, has enabled us to obtain 3D-SFM image <strong>of</strong> mica/water<br />

interface in 53 sec (Fig. 1). The atomically-resolved XY and XZ cross-sectional images<br />

obtained from the 3D-SFM image allow us to correlate the lateral positions <strong>of</strong> the surface<br />

atoms with the vertical distribution <strong>of</strong> the force field. The site-specific Δf vs distance<br />

curves are obtained from the Z pr<strong>of</strong>iles <strong>of</strong> the 3D-SFM image, revealing the remarkable<br />

variation <strong>of</strong> the force pr<strong>of</strong>iles depending on the tip positions. The results demonstrate that<br />

3D-SFM provides diverse information that has not been accessible with conventional 2D<br />

imaging techniques.<br />

Figure 1: (a) XY and XZ cross-sectional images and (b) Δf vs distance curves obtained from a<br />

3D-SFM image <strong>of</strong> mica/water interface (53 sec / 3D image, 0.82 sec / XZ image, Raw data).<br />

[1] T. Fukuma, M. J. Higgins and S. P. Jarvis, Biophys. J. 92 (2007) 3603-3609.<br />

45


Tu-0940<br />

Anisotropic Hydration <strong>of</strong> Biological Molecules Visualized by Three-<br />

Dimensional Scanning <strong>Force</strong> <strong>Microscopy</strong><br />

Hitoshi Asakawa 1 , Yasumasa Ueda 1 and Takeshi Fukuma 1,2<br />

1 Frontier Science Organization, Kanazawa University, Japan<br />

2 PRESTO, JST, Japan<br />

Hydration <strong>of</strong> biological molecules has grave impact on important biological processes<br />

such as membrane fusions and protein activities. However, molecular-scale details <strong>of</strong><br />

local hydration structures have remained to be understood due to the lack <strong>of</strong> a method<br />

able to visualize their three-dimensional (3D) distribution at sub-nanometer resolution. In<br />

order to resolve this issue, we have recently developed a novel technique referred to as<br />

3D scanning force microscopy (3D-SFM). Combined with frequency modulation atomic<br />

force microscopy (FM-AFM) in liquid, the method allows us to record 3D volume data <strong>of</strong><br />

frequency shift (Δf ) induced by tip-sample interaction force in less than 1 min. We are<br />

able to slice the obtained 3D-SFM image in any directions to produce 2D cross-sectional<br />

Δf images.<br />

In this study, we have investigated hydration structures formed at the interface between<br />

a model biological membrane and aqueous solution. A dipalmitoylphosphatidylcholine<br />

(DPPC) bilayer was formed on a cleaved mica surface as a model biological membrane in<br />

HEPES buffer solution. The XY cross-sectional Δf image inserted in Fig. 1(a) was<br />

obtained by slicing the 3D-SFM image near the tip position closest to the membrane<br />

surface. The image shows the pseudo-hexagonal packing <strong>of</strong> DPPC molecules. The<br />

periodic Δf variation corresponding to the DPPC molecule is also found in a crosssectional<br />

image obtained by slicing the 3D-SFM image with a plane perpendicular to the<br />

membrane surface (Fig. 1(b)). The image reveals formation <strong>of</strong> multiple hydration layers<br />

on the DPPC bilayer. In addition, we found that the first hydration layer shows<br />

anisotropic distribution around the DPPC headgroups elongated in one direction in a 3D<br />

space as indicated by the white arrows. The result demonstrates the capability <strong>of</strong> 3D-SFM<br />

for visualizing 3D distribution <strong>of</strong> water molecules interacting with a biological molecule.<br />

a) b)<br />

Figure 1: 3D-SFM images <strong>of</strong> the interface between a DPPC bilayer and HEPES/NaCl (10/100<br />

mM, pH 7.4) buffer solution. (a) Schematic illustration <strong>of</strong> a membrane/water interface. Inset: XY<br />

cross-sectional Δf image, (b) Z cross-sectional Δf image.<br />

46


Tu-1000<br />

Experimental and Theoretical Studies on 3D Hydration Structures<br />

on Muscovite Mica Surfaces in Aqueous Solution<br />

N. Oyabu 1 , K. Kimura 2 , S. Ido 1 , K. Suzuki 1 , K. Kobayashi 3 T. Imai 4 , and H. Yamada 1<br />

1 Department <strong>of</strong> Electronic Science & <strong>Engineering</strong>, Kyoto University, Japan<br />

2 Department <strong>of</strong> Chemistry, Kobe University, Japan<br />

3 Innovative Collaboration Center (ICC), Kyoto University, Japan<br />

4 Computational Science Research Program, RIKEN, Japan<br />

Recent progress in the two-dimensional (2D) frequency shift (Δf) mapping method<br />

using the high-resolution FM-AFM in liquids allows us to investigate molecular-scale<br />

hydration structures on various solid surfaces [1]. A precise comparison <strong>of</strong> the<br />

experimental data with theoretical calculations <strong>of</strong> water structures requires the threedimensional<br />

(3D) Δf mapping method, which can clarify more the relationship between<br />

the crystal site on the surface and the hydration structure. However, there are several<br />

difficulties in the 3D-Δf mapping in liquids including a large, linear and nonlinear<br />

thermal drift <strong>of</strong> the tip position relative to the surface.<br />

We have developed a low-thermal-drift FM-AFM working in liquids based on a<br />

commercial AFM. A sufficiently low, lateral thermal drift rate <strong>of</strong> less than 1 nm/min was<br />

achieved in liquids by an accurate temperature control <strong>of</strong> the environment and by a large<br />

reduction <strong>of</strong> the liquid evaporation. We obtained 3D-Δf data on a muscovite mica surface<br />

in a 1M KCl solution. Figure 1 shows three examples <strong>of</strong> the 2D-Δf image taken out <strong>of</strong> the<br />

3D-Δf data. Figures 2(a) and (b) are constant height (Δf distribution) X-Y images<br />

reconstructed from the 3D-Δf data. The vertical position (Z) <strong>of</strong> the plane <strong>of</strong> Fig. 2(a) is<br />

about 0.2 nm closer to the surface compared to that <strong>of</strong> Fig. 2(b). Figures 2(c) and (d) are<br />

X-Y water density distributions calculated using the 3D reference interaction site model<br />

(3D-RISM) theory at the corresponding vertical positions, respectively. At each position<br />

the experimental image agrees well with the calculated density distribution. In addition,<br />

this 3D-Δf mapping method in liquids has been also applied to the study <strong>of</strong> hydration<br />

structures on biomolecules.<br />

Z1 Z0 Z<br />

Y<br />

X<br />

Figure 1: Three examples <strong>of</strong> the 2D<br />

(X-Z)-Δf mapping image taken out <strong>of</strong><br />

the 3D-Δf data (X = 3 nm, Y = 3 nm, Z<br />

= 3 nm).<br />

Figure 2: (a), (b) Constant height X-Y images reconstructed from<br />

the 3D-Δf data at different heights. The Z-position (a) is 0.2 nm<br />

closer to the surface than the position (b). (c), (d) Water density<br />

distributions calculated by the 3D-RISM theory at the<br />

corresponding heights.<br />

[1] K. Kimura et al., The 11 th International conference on Non-Contact <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong>, Sept.<br />

2008, Madrid, Spain, Abstract booklet, p62<br />

47


Theory <strong>of</strong> Tip-Sample Interaction <strong>Force</strong> Mediated by Water<br />

M.Tsukada 1 , M.Harada 2 , and K.Tagami 2<br />

1 WPI-Advanced Institute for Materials Research, Tohoku University, Sendai, Japan<br />

2 Advance S<strong>of</strong>t Corp. Tokyo, Japan<br />

Tu-1020<br />

<strong>Noncontact</strong> AFM(ncAFM) in liquids is a powerful method for observing nano-scale<br />

images <strong>of</strong> s<strong>of</strong>t-materials as protein. We are developing theoretical simulation<br />

methodologies <strong>of</strong> ncAFM in liquid[1,2], and by the simulation, we found various<br />

remarkable features <strong>of</strong> tip-surface interaction force mediated by water[2]. They are<br />

obtained either by the molecular dynamics (MD) calculation or by the 3D-RISM<br />

(Reference Interacting Site Model) calculation.<br />

As a case study <strong>of</strong> the MD method, ncAFM images and 3D force map <strong>of</strong> mica surface<br />

in water were calculated and compared with the experiments by Yamada’s Group. The<br />

oscillatory layer structures <strong>of</strong> water exist around the interface tracing the nano-scale<br />

shapes <strong>of</strong> the sample. The phase/amplitude <strong>of</strong> the force oscillation is changed depending<br />

on the lateral tip position, and the 3D force map shows a complicated network structure<br />

<strong>of</strong> attractive and repulsive regions. The numbers <strong>of</strong> the water layers in the narrow gap<br />

decreased abruptly and finally disappears when the tip approaches to the surface.<br />

The 3D-RISM method was applied to clarify the local charge separation effect on the<br />

interaction force in water. The distribution <strong>of</strong> water molecules around the atomically<br />

charged portion on the substrate is considerably disturbed, and this effect is observed in<br />

the force map from far region. We discuss how drastically different the water mediated<br />

tip-sample force compared with that in vacuum, and how they influence on the images.<br />

a) b)<br />

Figure 1 a) Simulated force map <strong>of</strong> mica in water by the MD method, and b)<br />

The distribution function <strong>of</strong> water molecules near the flat solid surface. The atom shown<br />

with red and blue sphere are ionized positively and negatively, respectively.<br />

[1]M. Tsukada, N. Watanabe, Jpn.J.Appl.Phys., vol.48 No.3 2009<br />

[2] Tsukada, Tagami , J.Phys.Soc.Jpn., submitted.<br />

48


Dynamic <strong>Force</strong> Spectroscopy <strong>of</strong> Single Chain-like Molecules<br />

Daniel Ebeling 1 , Harald Fuchs 1 , Filipp Oesterhelt 2 , and Hendrik Hölscher 3<br />

Tu-1120<br />

1<br />

Center for Nanotechnology (CeNTech), Heisenbergstrasse 11, 48149 Münster, Germany and<br />

Physikalisches Institut, Westfälische Wilhelms-Universität Münster, Wilhelm-Klemm-Strasse 10, 48149<br />

Münster, Germany<br />

2<br />

Institut für Physikalische Chemie II, Heinrich-Heine-Universität Düsseldorf, Universitätsstrasse 1,<br />

40225 Düsseldorf, Germany<br />

3<br />

Institute for Microstructure Technology (IMT), Forschungszentrum Karlsruhe, P.O. box 36 70, 76021<br />

Karlsruhe, Germany<br />

In order to measure forces acting on a single chain-like molecule during a stretching<br />

experiment in liquid, we introduce a dynamic approach based on the frequencymodulation<br />

technique with constant-excitation. In difference to the classical approach<br />

where the force is recorded as a conventional force vs. distance curve in a static<br />

measurement, we are able to detect simultaneously the conservative force as well as the<br />

energy dissipation during the elongation <strong>of</strong> a chain-like molecule.<br />

Therefore the amplitude and the frequency shift <strong>of</strong> an oscillating cantilever are measured<br />

during the retraction from the surface (Fig. 1a and b) [1]. The tip-sample force (Fig. 1c)<br />

and the energy dissipation (Fig. 1d) can be reconstructed from these data sets via an<br />

analytical approach. We apply this technique to dextran monomers and demonstrate the<br />

agreement <strong>of</strong> the experimental force curves with a ``single-click'' model [2].<br />

Figure 1: a) Amplitude and b) frequency shift curves measured during approach to and retraction<br />

from the surface covered with dextran molecules. c) <strong>Force</strong> acting on the dextran molecule as a<br />

function <strong>of</strong> the actual tip position. The experimental result is well described by a ``single-click''<br />

model (solid line) d) Energy dissipation per oscillation cycle.<br />

[1] D. Ebeling, F. Oesterhelt, H. Hölscher (submitted).<br />

[2] R. G. Haverkamp, A. T. Marshall, and M. A. K. Williams, Phys. Rev. E 75, 021907 (2007).<br />

49


Spatial force fields above a single atom defect<br />

André Schirmeisen 1 and Domenique Weiner 2<br />

1 CeNTech (Center for Nanotechnology) and Institute <strong>of</strong> Physics, University <strong>of</strong> Muenster, Germany<br />

2 SPECS Zurich GmbH, Switzerland<br />

Tu-1140<br />

<strong>Atomic</strong> force microscopy under ultrahigh vacuum conditions is a powerful tool to<br />

investigate the atomic structure <strong>of</strong> surfaces. The method <strong>of</strong> 3D force field spectroscopy<br />

[1] allows the spatial analysis <strong>of</strong> vertical and lateral interatomic forces [2], as well as the<br />

potential energy landscape with atomic resolution [3]. In this study we focus on the<br />

analysis <strong>of</strong> surface defects on a NaCl(001) crystal by 3D force field spectroscopy. The<br />

spatial force fields along different crystallographic directions were measured above a<br />

defect, which appeared as a valley <strong>of</strong> molecular dimensions in the surface topography.<br />

We find that the vertical tip-sample force directly above the defect is repulsive. From the<br />

force fields we calculate the atomic scale potential energy landscape, which is compared<br />

to model calculations. This model is based on electrostatic interactions <strong>of</strong> hard spheres<br />

and assumes an ion terminated tip apex. According to this model our experimental<br />

potential energy fields agree best with a situation where a single ion is missing in the<br />

surface. This raises interesting questions about the unexpected stability <strong>of</strong> single charge<br />

defects in an ionic surface.<br />

Figure 1: Left: Topography scan <strong>of</strong> NaCl(001) showing a surface defect. Right: Frequency shift<br />

slice extracted from 3D spectroscopy data along a row <strong>of</strong> ions including the surface defect. The<br />

force and energy pr<strong>of</strong>iles best agree with a model assuming a singular missing ion.<br />

[1] Hölscher, Langkat, Schwarz, Wiesendanger, Appl. Phys. Lett. 81, 4428 (2002)<br />

[2] Ruschmeier, Schirmeisen, H<strong>of</strong>fmann, Phys. Rev. Lett 101, 156102 (2008)<br />

[3] Schirmeisen, Weiner, Fuchs, Phys. Rev. Lett. 97, 136101 (2006)<br />

50


Simultaneous measurement <strong>of</strong> force and tunneling current<br />

Tu-1200<br />

Daisuke Sawada, Yoshiaki Sugimoto, Ken-ichi Morita, Masayuki Abe, and Seizo Morita<br />

Graduate school <strong>of</strong> <strong>Engineering</strong>, Osaka University, Osaka, Japan<br />

We have performed simultaneous STM and AFM measurements in the dynamic mode<br />

using Pt-Ir coated Si cantilevers at room temperature. Frequency shift (Δf) and timeaverage<br />

tunneling current () images were obtained by tip scanning on the Si(111)-<br />

(7×7) surface at constant height mode to prevent a crosstalk between these two channels<br />

[Fig.1 (a)]. To compensate the thermal drift <strong>of</strong> the tip-surface distance, feed forward<br />

technique was applied [1]. Analysis <strong>of</strong> 25 sets <strong>of</strong> AFM/STM images using different tips<br />

shows that when atomic resolution is obtained simultaneously by both AFM and STM,<br />

the tunneling current is much larger than the typical values in conventional STM. We<br />

have also performed simultaneous measurements <strong>of</strong> site-specific force/tunneling<br />

spectroscopy. The Δf and versus tip-surface distance curves were converted into the<br />

short-range force (FSR) and the tunneling current (It) at closest separation between the<br />

sample surface and the oscillating tip [Fig.1 (b)]. We observed the drop in the tunneling<br />

current due to the chemical interaction between the tip apex atom and the surface adatom,<br />

which was found recently [2], and estimated the value <strong>of</strong> the chemical bonding force.<br />

Furthermore, scanning tunneling spectroscopy (STS) was performed on the same site<br />

using the same tip [Fig.1 (c)]. The spectrum is in good agreement with previous STM<br />

results. Our results demonstrate that one can quantitatively measure the local density <strong>of</strong><br />

state and the chemical bonding force above the same atom using the same tip [3].<br />

Figure 1: (a) Δf and images <strong>of</strong> a Si(111)-(7×7) surface simultaneously obtained in the<br />

constant height mode at room temperature. The acquisition parameters are f0=284092.5 Hz,<br />

A=300 Å, k=22.9 N/m, and Vs=-400 mV respectively. (b) It-z and FSR-z curves derived from z<br />

and Δf-z curves obtained above a center adatom in a faulted half unit cell, respectively. (c) STS<br />

spectrum and Δf - Vs curves obtained at z=4.4 Å.<br />

[1] M. Abe, et al., Appl. Phys. Lett. 90, 203103 (2007).<br />

[2] P. Jelinek, M. Svec, P. Pou, R. Perez, and V. Chab, Phys. Rev. Lett. 101, 176101 (2008).<br />

[3] D. Sawada, Y. Sugimoto, K. Morita, M. Abe, and S. Morita, Appl. Phys. Lett. in press.<br />

51


Atom Manipulation on Cu(110)-O Surface with LT-AFM<br />

52<br />

Tu-1220<br />

Yasuhiro Sugawara, Yukinori Kinoshita, Yan Jun Li, Yoshitaka Naitoh, and Masami<br />

Kageshima<br />

Department <strong>of</strong> Applied Physics, Osaka University, Suita, Japan<br />

Manipulation <strong>of</strong> single atoms and molecules is an innovative experimental technique<br />

<strong>of</strong> nanoscience. So far, a scanning tunneling microscopy (STM) has been widely used to<br />

fabricate artificial structures by laterally pushing, pulling or sliding single atoms and<br />

molecules. However, the driving forces involved in manipulation have not been<br />

measured. Recently, an atomic force microscopy (AFM) has been also used to manipulate<br />

single atoms and molecules. Atom manipulation with an AFM is particularly promising,<br />

because it allows the direct measurement <strong>of</strong> the required forces [1].<br />

Recently, we investigated the forces in AFM lateral manipulation for a top single Cu<br />

atom (super Cu atom) on the Cu(110)-O surface. In the case <strong>of</strong> Cu-adsorbed AFM tip, the<br />

super Cu atom on the surface was pulled at a lateral tip position on the adjacent binding<br />

site. In contrast, in the case <strong>of</strong> O-adsorbed AFM tip, the super Cu atom was pushed over<br />

the top <strong>of</strong> the super Cu atom. Thus, we found that the forces (attractive or repulsive<br />

forces) to move an atom laterally on the surface strongly depend on the atom species <strong>of</strong><br />

the AFM tip apex and the surface.<br />

In the present study, in order to further clarify the manipulation process depending on<br />

the chemical nature <strong>of</strong> tip-sample interaction, we investigate the full tip-sample potential<br />

landscape necessary to manipulate atoms.<br />

All experiments were performed by using homebuilt noncontact AFM using frequency<br />

modulation technique operating in ultrahigh vacuum at a temperature <strong>of</strong> about 78 K. The<br />

AFM tip apex was coated with Cu or O atoms in situ by slightly making a tip-sample<br />

mechanical contact on the Cu(110)-O surface prior to the imaging. The tip-sample<br />

potentials are determined form the frequency shift versus distance curves by<br />

mathematical analysis. We found that the tip-sample potentials to move the super Cu<br />

atom laterally on the surface strongly depend on the atom species <strong>of</strong> the AFM tip apex<br />

and the surface. These results strongly suggest that the chemical nature <strong>of</strong> tip-sample<br />

interaction plays an important role in lateral atom manipulation. Furthermore, we discuss<br />

the pathways for moving the super Cu atom.<br />

Reference<br />

[1] M. Ternes, C. P. Lutz, C. F.<br />

Hirjibehedin, F. J. Gissibl and<br />

Andreas J. Heinrich. Science,<br />

319, 66 (2008).<br />

(a) (b)<br />

Figure 1: AFM images (a) before and (b) after atom<br />

manipulation performed on Cu(110)-O surface at 78K.


Water, Ions, Membranes, Real Metals, Finite Temperature:<br />

Is there ever a pure Casimir force?<br />

Adrian Parsegian 1<br />

1 National Institutes <strong>of</strong> Health, Bethesda, MD 20892-0924<br />

Tu-1430<br />

The powerful idea <strong>of</strong> vacuum fluctuations strictly confined between ideal metals<br />

dominates the psychology <strong>of</strong> much thinking about electrodynamic forces. In fact, most<br />

systems reveal interactions that depend on material properties wherein fluctuations<br />

penetrate unto permeate interacting bodies. This talk will attempt to delineate<br />

the conditions under which true Casimir forces can be expected as compared with the<br />

material-based van der Waals forces formulated by Lifshitz, Dzyaloshinskii, and<br />

Pitaevskii, then generalized by their descendents.<br />

53


Short and medium range electrostatic forces analyzed by<br />

Kelvin probe force microscopy<br />

Th. Glatzel, S. Kawai, S. Koch, A. Barat<strong>of</strong>f, and E. Meyer<br />

Department <strong>of</strong> Physics, University <strong>of</strong> Basel, 4056 Basel, Switzerland.<br />

Tu-1510<br />

We discuss the influences <strong>of</strong> short and medium range electrostatic forces on the local<br />

contact potential difference (LCPD) by means <strong>of</strong> amplitude modulated Kelvin probe<br />

force microscopy (AM-KPFM) measured on single crystal KBr(100) surfaces. Bias- and<br />

z-spectroscopic curves at atomic scale showing clear features related to the influence <strong>of</strong><br />

the short range electrostatic interaction [1].<br />

In Fig.1 a typical nc-AFM /KPFM measurement <strong>of</strong> a KBr(100) in UHV is shown. While<br />

the topography (a) was measured at the first resonance frequency <strong>of</strong> the cantilever<br />

(approx 160kHz), the LCPD image (b) was measured by compensating the inphase signal<br />

detected at he second resonance (approx 1MHz). A clear contrast between the different<br />

ionic sides is observed. To clarify the origin <strong>of</strong> this contrast we performed bias- and zspectroscopy<br />

measurements. c) shows a bias-spectroscopy measurement illustrating the<br />

amplitude R, the inphase signal X, and the frequency shift at a maxima <strong>of</strong> the topography.<br />

Comparing these measurements with the ones at the minimas <strong>of</strong> the topography (not<br />

presented here) clearly shows the expected CPD difference.<br />

Figure 1: Typical topography (3x3nm 2 , �z=100pm) and<br />

local contact potential image (�LCPD=300mV) <strong>of</strong> a<br />

KBr(100) surface. In c) the lockin signals amplitude R,<br />

inphase X (second resonance <strong>of</strong> the cantilever) as well as<br />

the frequency shift <strong>of</strong> the first resonance �f are plotted<br />

over the bias voltage applied to the sample. The<br />

measurement was done at a maxima <strong>of</strong> the topography.<br />

[1] F. Bocquet et al., Phys. Rev. B 78, (2008), 035410.<br />

54


Tu-1530<br />

The Effective Quality Factor in Dynamic <strong>Force</strong> Microscopes with<br />

Fabry-Perot Interferometer Detection<br />

Hendrik Hölscher 1 , Peter Milde 2 , Ulrich Zerweck 2 , Lukas M. Eng 2 , Regina H<strong>of</strong>fmann 3<br />

1 Institute for Microstructure Technology (IMT), Forschungszentrum Karlsruhe, Karlsruhe, Germany<br />

2 Institut für Angewandte Photophysik, Technische Universität Dresden, Germany<br />

3 Physikalisches Institut and DFG-Center for Functional Nanostructures, Universität Karlsruhe, Germany<br />

Recently, the self-oscillation <strong>of</strong> micr<strong>of</strong>abricated silicon resonators by photo-induced<br />

forces has become a focus <strong>of</strong> current research in order to reach the quantum ground state<br />

[1]. Having these results in mind, it is interesting to note that the experimental set-up<br />

used in these studies is similar in design to the standard instrumentation <strong>of</strong> lowtemperature<br />

ultra-high vacuum atomic force microscopes (LT-UHV-AFM) using<br />

interferometer detection [2,3].<br />

In order to quantify the relevance <strong>of</strong> photo-induced forces for NC-AFM<br />

experiments, we analyzed the oscillation behavior <strong>of</strong> a micr<strong>of</strong>abricated cantilever in a<br />

LT-UHV-AFM with a Fabry-Perot interferometer. We observed that photo-induced<br />

forces depend on the cantilever-fiber distance and that they lead to a pr<strong>of</strong>ound change <strong>of</strong><br />

the effective quality factor <strong>of</strong> the cantilever. Depending on the slope <strong>of</strong> the interference<br />

signal, the effective quality factor <strong>of</strong> the cantilever is increased or decreased. The overall<br />

effect increases with the laser power <strong>of</strong> the laser diode and decreases for higher<br />

temperatures. The Q-factor <strong>of</strong> the cantilever, however, is an essential parameter when the<br />

energy dissipation between tip and sample is measured with a dynamic force microscope.<br />

Hence, the effect addressed here has to be taken into account whenever the cantilever<br />

oscillation is measured with an interferometer at low temperatures.<br />

a) b) c)<br />

Figure 1: (a) Schematic set-up <strong>of</strong> the interferometer used for the detection <strong>of</strong> the cantilever<br />

vibration in dynamic force microscopy (b) The resulting interference intensity is periodic and has<br />

a positive or negative slope. (c) Depending on the slope and the temperature the resonance curves<br />

change dramatically due to a change <strong>of</strong> the effective Q-factor <strong>of</strong> the cantilever.<br />

[1] C. Höberger-Metzger and K. Karrai, Nature 432, 1002 (2004).<br />

[2] D. Rugar, H. J. Mamin, and P. Guethner, Appl. Phys. Lett. 55, 2588 (1989).<br />

[3] A. Moser et al., Meas. Sci. Technol. 4, 769 (1993).<br />

55


Oral<br />

Presentations<br />

Wednesday, 12 August<br />

56


We-0900<br />

Imaging Single Atoms on Oxide Surfaces – Gold on Alumina/NiAl(110)<br />

Markus Heyde, Georg Hermann Simon, Thomas König, and Hans-Joachim Freund<br />

Fritz-Haber-Institute <strong>of</strong> the Max-Planck-Society, Faradayweg 4-6, D-14195 Berlin, Germany<br />

A logical next step after gaining atomic resolution on complex surfaces <strong>of</strong> wide<br />

band-gap materials with FM-DFM is the determination and characterization <strong>of</strong> ad-atom<br />

adsorption sites on such materials. Imaging metal ad-atoms on complex oxide surfaces<br />

with atomic resolution is <strong>of</strong> a certain value in the study <strong>of</strong> these materials and has not<br />

been shown before. Here we present recent work on adsorbed gold atoms on the ultrathin<br />

alumina on NiAl(110) with its large complex surface unit cell as an example. It can be<br />

shown that an established contrast in FM-DFM [1,2] enables the determination <strong>of</strong><br />

adsorption sites with atomic scale precision. This holds despite the fact that gold is rather<br />

easily removed by the tip and gives confidence for atomic scale characterisation and<br />

manipulation <strong>of</strong> ad-species on oxide surfaces with FM-DFM. Obtained results are<br />

compared to density functional theory calculations for preferred adsorption sites <strong>of</strong> gold<br />

atoms on this alumina film [3].<br />

Figure 1: (a) <strong>Atomic</strong> resolution FM-DFM image <strong>of</strong> the ultrathin alumina on NiAl(110). The<br />

large unit cell (here an A domain) is indicated (scan range 4.2 nm × 4.2 nm ). (b) FM-DFM image<br />

<strong>of</strong> adsorbed gold atoms on the alumina film (scan range 10 nm × 10 nm). (c) Preferential<br />

adsorption sites for gold atoms as determined by density functional theory [3].<br />

[1] G. H. Simon, T. König, M. Nilius, H.-P. Rust, M. Heyde, and H.-J. Freund; Phys. Rev. B 78 (2008)<br />

113401.<br />

[2] G. Kresse, M. Schmid, E. Napetschnig, M. Shishkin, L. Köhler, and P. Varga; Science 308 (2005)<br />

1440.<br />

[3] N. Nilius, M.V. Ganduglia-Pirovano, V. Brázdová, M. Kulawik, J. Sauer, and H.-J. Freund; Phys.<br />

Rev. Lett. 100 (2008) 096802.<br />

57


Site-specific force spectroscopy on TiO2 (110) surface at lowtemperature<br />

A. Yurtsever, A. Pratama, Y. Sugimoto, M. Abe, and S. Morita<br />

Graduate <strong>School</strong> <strong>of</strong> <strong>Engineering</strong>, Osaka University, Osaka, Japan<br />

We-0920<br />

Non-contact AFM (nc-AFM) has opened the way for the measurement <strong>of</strong> the forces<br />

that are generated between different kinds <strong>of</strong> interaction on surfaces. Using the homemade<br />

low temperature nc-AFM, we study the site-specific force spectroscopy on the<br />

rutile TiO2 (110)-(1x1) metal oxide surface. The rutile TiO2(110) surface has been<br />

intensively used as a subject <strong>of</strong> numerous surface science investigations over the years<br />

due to its wide variety <strong>of</strong> technological applications such as catalysis, solar cells, surface<br />

protective coatings, and gas sensing devices for pollution control [1]. In AFM, imaging <strong>of</strong><br />

metal-oxide surfaces, the image contrast strongly depends on the chemical constitution <strong>of</strong><br />

the surface layer and the chemical identity <strong>of</strong> the tip-apex structure [2]. Depending on the<br />

polarity <strong>of</strong> tip-apex charged state, the distinct types <strong>of</strong> image contrast are observed. In<br />

order to clarify the imaging mechanism <strong>of</strong> each contrast modes, we need to perform the<br />

measurement <strong>of</strong> force-displacement curves on this surface. By employing the atomtracking<br />

technique [3], force–distance measurements over three different atomic sites,<br />

using three different tip states, were obtained. We observe that the maximum attractive<br />

interaction force is smallest on OH groups for different tip-apex states. In this<br />

contribution, we will discuss the effect <strong>of</strong> the different tip terminations on interaction<br />

force over surface atomic (e.g., Ob, Ti) and defect (e.g., OH) sites.<br />

Figure 1: Low-temperature atomic force microscopy spectroscopic measurements over various<br />

sites in the TiO2 (110) surface. The force versus distance data obtained with a neutral tip (e.g., Si)<br />

(a) and with a positively terminated tip (b). Inset: Shows the corresponding topographic images <strong>of</strong><br />

each contrast modes at the same surface area.<br />

[1] U. Diebold, Surf. Sci. Rep. 48, 53-229 (2003).<br />

[2] G. H. Enevoldsen et al., Phys. Rev. B 78, 045416 (2008).<br />

[3] M. Abe et al., Appl. Phys. Lett. 90, 203103 (2007).<br />

58


We-0940<br />

Character <strong>of</strong> the short-range interaction between a silicon based tip and<br />

the TiO2(110) surface: a DFT study<br />

Cesar Gonzalez 1,2 , Pavel Jelínek 1 , and Ruben Pérez 3<br />

1<br />

Department <strong>of</strong> Thin Films, Institute <strong>of</strong> Physics <strong>of</strong> the ASCR, Prague, Czech Republic<br />

2<br />

Departamento de Superficies y Recubrimientos, Instituto de Ciencia de Materiales de Madrid del CSIC,<br />

Spain<br />

3 Departamento de Fisica Teorica de la Materia Condensada, Universidad Autonoma de Madrid, Spain<br />

Non-contact atomic force microscopy (NC-AFM) provides a rich variety <strong>of</strong> atomic<br />

contrasts on the rutile TiO2 (110) surface, imaging either the bridging oxygen atoms<br />

(hole mode) or the titanium atoms (protrusion mode) [1]. These contrast modes have been<br />

assigned to purely electrostatic interaction. Another contrast mode that does not ?t into<br />

the scheme <strong>of</strong> purely electrostatic interaction, the so-called neutral mode, has also been<br />

reported[]. Recently it has been demonstrated that NC-AFM is capable <strong>of</strong> imaging both,<br />

the bridging oxygen atoms and the titanium rows simultaneously in a new “all inclusive”<br />

mode [3]. Such diversity <strong>of</strong> contrast modes can be attributed to the complex character <strong>of</strong><br />

the short range interaction between tip and characteristic sites <strong>of</strong> the rutile TiO2 (110)<br />

surface driven by (i) a weak short-range electrostatic interaction [4] depending on atomic<br />

termination <strong>of</strong> tip and its polarization and (ii) the onset <strong>of</strong> chemical bond formed between<br />

a tip and surface [5]. A proper characterization <strong>of</strong> the different regimes in the short-range<br />

interaction regime Si-based tips and this oxide surface is crucial for the interpretation <strong>of</strong><br />

the experimental images and the design <strong>of</strong> protocols for single atom manipulation and<br />

chemical identification.<br />

Here we have employed density-functional theory (DFT) calculations to understand<br />

the character <strong>of</strong> tip-sample interaction between clean and contaminated Si-based tips and<br />

the TiO2 surface. We have performed a detailed analysis <strong>of</strong> electronic structure and the<br />

charge transfer between tip and sample. Our calculations show that the relative<br />

contribution <strong>of</strong> the weak short-range electrostatic interaction and the on-set <strong>of</strong> chemical<br />

bonding between the closest tip and surface atoms is very sensitive to the tip-sample<br />

distance, de?ning di? erent interaction regimes along the tip-sample distance. In<br />

particular, we show the short-range electrostatic interaction in weak interaction regime<br />

can provide a complex atomic contrast such as the experimentally reported neutral and<br />

all-inclusive contrast modes.<br />

[1] J.V. Lauritsen et al., Nanotechnology 17, 3436 (2006).<br />

[2] G.H. Enevoldsen et al., Phys. Rev. B 78, 045416 (2008).<br />

[3] R. Bechstein et al (submitted).<br />

[4] A.S. Foster et al., Phys. Rev. B 68, 195420 (2003).<br />

[5] R. Pérez et al., Phys, Rev. B 58, 10835 (1998).<br />

59


We-1000<br />

True atomic resolution imaging on an application-oriented system:<br />

Understanding photocatalytic reactivity <strong>of</strong> transition-metal doped TiO2<br />

Ralf Bechstein 1,2 , Mitsunori Kitta 3 , Jens Schütte 1 , Hiroshi Onishi 3 , and Angelika Kühnle 1<br />

1 Fachbereich Physik, Universität Osnabrück, Barbarastraße 7, 49076 Osnabrück, Germany<br />

2 present address: iNano, Aarhus University, DK-8000 Aarhus C, Denmark<br />

3 Department <strong>of</strong> Chemistry, Kobe University, Rokko-dai, Nada-ku, Kobe 657-8501, Japan.<br />

The influence <strong>of</strong> transition-metal dopants, namely chromium and antimony, on the<br />

surface structure <strong>of</strong> TiO2(110) was investigated at the atomic level using frequency<br />

modulation NC-AFM. We found the vacancy density created upon doping to depend<br />

strongly on the dopant ratio [1, 2], explaining differences in photocatalytic reactivity [3,<br />

4]. The surface roughness and the arrangement at the atomic scale (see e.g. Fig. 1) gave<br />

further insight into the implications <strong>of</strong> doping <strong>of</strong> titanium dioxide.<br />

This work demonstrates that NC-AFM is capable <strong>of</strong> true atomic resolution, even on<br />

rough and highly corrugated surfaces. Thus, it constitutes a step towards investigating<br />

more realistic, application-oriented systems using NC-AFM, revealing detailed insights<br />

into the implications <strong>of</strong> transition-metal doping on the surface structure <strong>of</strong> the wide band<br />

gap photocatalyst titanium dioxide.<br />

(a) (b)<br />

Figure 1: NC-AFM detuning images <strong>of</strong> TiO2(110) codoped with chromium and antimony.<br />

Enhanced surface roughness is observed (a). At a first glance, the atomic structure <strong>of</strong> the doped<br />

surface resembles the structure <strong>of</strong> pristine TiO2(110) (b). A careful analysis <strong>of</strong> the high-resolution<br />

images unravelled a surface reconstruction that closely resembles a (1 x 4) structure, indicating a<br />

firm integration <strong>of</strong> the dopant atoms into the titanium dioxide crystal.<br />

[1] R. Bechstein et al., J. Phys. Chem. C, 113, 3277 (2009).<br />

[2] R. Bechstein et al., submitted to J. Phys. Chem. C (2009).<br />

[3] H. Kato and A. Kudo, Catal. Today 78, 561 (2003).<br />

[4] T. Ikeda et al., J. Phys. Chem. C 112, 1167 (2008).<br />

60


We-1020<br />

Local surface photovoltage spectroscopy <strong>of</strong> molecular clusters using<br />

Kelvin probe force microscopy<br />

Sascha Sadewasser, Martha Ch. Lux-Steiner<br />

Helmholtz Zentrum Berlin für Materialien und Energie, Glienicker Str. 100, 14109 Berlin, Germany<br />

The high optical absorption properties <strong>of</strong> organic molecules are <strong>of</strong> high interest for<br />

photovoltaic application, as well in purely organic as in dye sensitized solar cells. In the<br />

latter, the electron conductor is typically a porous TiO2, which is used to enhance the<br />

surface and contact areas. Investigation <strong>of</strong> the local optoelectronic properties <strong>of</strong> such<br />

molecules and interfaces is <strong>of</strong> high interest for the understanding <strong>of</strong> charge separation<br />

processes and ultimately the improvement <strong>of</strong> these types <strong>of</strong> solar cells.<br />

With the goal to gain access to optoelectronic properties on the nanometer scale, we<br />

have recently set up a surface photovoltage spectroscopy (SPS) system in combination<br />

with a Kelvin probe force microscope (KPFM) operated in ultra high vacuum (UHV).<br />

Sample illumination is realized by a halogen lamp and monochromator in the visible<br />

spectrum range using an optical fiber introduced into the UHV. Using chopped light and<br />

lock-in detection the SPV signal can be measured with a sensitivity <strong>of</strong> better than 1mV.<br />

For the KPFM operation we use the resonant amplitude modulation technique, which<br />

allows using a low modulation voltage (~0.1V).<br />

Here we present a study <strong>of</strong> various organic molecules used in organic and dye<br />

sensitized solar cells. For the case <strong>of</strong> the dye N3, small amounts <strong>of</strong> molecules were<br />

deposited onto TiO2 single crystals and subsequently measured by SPS in the UHV-<br />

KPFM. Fig. 1 shows the topography and work function image <strong>of</strong> the sample, clearly<br />

showing a higher work function for the molecule clusters. Spectroscopy in various points<br />

on the sample shows clearly distinct spectra on the molecules as compared to the TiO2<br />

surface. The absorption and charge separation <strong>of</strong> the various molecules will be discussed<br />

comparatively.<br />

Figure 1: (a) Topography (800×800 nm 2 ) and (b) corresponding work function image obtained<br />

by KPFM. The symbols indicate the position <strong>of</strong> SPV spectra. (c) Averaged SPV spectra taken on<br />

TiO2 (squares) and on N3 molecule clusters (circles). Clearly the enhanced absorption <strong>of</strong> N3 in<br />

the range below ~ 600 nm can be seen.<br />

61


We-1120<br />

Why is Graphite so Slippery? Gathering Clues from Three-Dimensional<br />

Lateral <strong>Force</strong>s Measurements<br />

Udo D. Schwarz 1 , Mehmet Z. Baykara 1 , Todd C. Schwendemann 1,2 , Boris J. Albers 1 ,<br />

Nicolas Pilet 1 , and Eric I. Altman 2<br />

1<br />

Department <strong>of</strong> Mechanical <strong>Engineering</strong> and Center for Research on Interface Structures and Phenomena<br />

(CRISP), <strong>Yale</strong> University, New Haven, USA<br />

2<br />

Department <strong>of</strong> Chemical <strong>Engineering</strong> and Center for Research on Interface Structures and Phenomena<br />

(CRISP), <strong>Yale</strong> University, New Haven, USA<br />

Conventional lateral force experiments give insufficient insight into the fundamental<br />

reasons for graphite’s outstanding qualities as a solid lubricant due to an averaging effect<br />

caused by the finite contact area <strong>of</strong> the tip with the sample. To overcome this limitation,<br />

we used a noncontact atomic force microscopy-based approach that enables use <strong>of</strong><br />

atomically sharp tips. The new technique [1], performed using a home-built low<br />

temperature, ultrahigh vacuum atomic force microscope [2], allows the measurement <strong>of</strong><br />

normal and lateral surface forces in all three dimensions with picometer and piconewton<br />

resolution.<br />

In this presentation, we analyze the height and lattice site dependence <strong>of</strong> lateral forces,<br />

their dependence on normal load, and the effect <strong>of</strong> tip shape in detail. The lateral forces<br />

are found to be heavily concentrated in the hollow sites <strong>of</strong> the graphite lattice, surrounded<br />

by a matrix <strong>of</strong> vanishingly small lateral forces. It will be argued that this astonishing<br />

localization may be a reason for graphite’s excellent lubrication properties. In addition,<br />

the distance and load dependence <strong>of</strong> the lateral forces experienced along possible “escape<br />

routes” from the hollow sites, which would be followed by a slider that is dragged out <strong>of</strong><br />

them, are studied. Surprisingly, the maximum lateral forces along these escape routes,<br />

which ultimately determine the static friction, are found to depend linearly on normal<br />

load, suggesting the validity <strong>of</strong> Amontons' law in the noncontact regime.<br />

Figure 1: a) Possible escape routes which would be preferred by a nanoslider as it is dragged<br />

away from a hollow site. b) Dependence <strong>of</strong> static friction in direction (I) on normal load.<br />

[1] B. J. Albers et al., Nature Nanotechnology 4, 307 (2009).<br />

[2] B. J. Albers et al., Rev. Sci Instrum. 79, 033704 (2008).<br />

62


We-1140<br />

Theoretical DFT simulations <strong>of</strong> the STM/AFM atomic scale imaging on<br />

graphite.<br />

Vít Rozsíval 1 , Martin Ondráček 1 , Pavel Jelínek 1<br />

1 Department <strong>of</strong> Thin Films, Institute <strong>of</strong> Physics <strong>of</strong> the ASCR, Prague, Czech Republic<br />

Graphite is used as a standard benchmark surface to achieve atomic resolution by<br />

Scanning Probe Methods (SPM). True atomic scale contrast <strong>of</strong> graphite provided by<br />

simultaneous measurement <strong>of</strong> both forces and tunneling current is still source <strong>of</strong><br />

controversy [1-3]. The hexagonal surface unit cell contains to non-equivalent atoms: the<br />

α-site atom located directly above adjacent planes and β-site atom above the hollow site.<br />

While there is a consensus on STM measurements seeing every second, β-site, atom in<br />

the hexagonal unit cell [4], the situation turns more conflicting in the case <strong>of</strong> nc-AFM<br />

experiments. Main doubt arises where the maximum <strong>of</strong> the contrast is located: (i) over<br />

hollow site [2] or one C atom [1]; and (ii) if the α and β atoms can be distinguished by<br />

nc-AFM.<br />

The direct quantitative interpretation <strong>of</strong> SPM measurements in the near-to-contact<br />

mode is a not straightforward task due to several factors playing an important role in<br />

imaging processes. Realistic conductance calculations have to involve properly several<br />

effects: e.g. the tip/surface structural relaxations and multiple electron scattering to<br />

understand properly the relation between forces and currents measured in experiments<br />

[5]. Here we performed state-<strong>of</strong>-the-art theoretical studies based on a combination <strong>of</strong> total<br />

energy DFT calculations with a Green's function approach for the electron STM transport<br />

[5] between silicon or tungsten tip and graphite. We analyzed the electronic structure<br />

evolution along tip-sample distance. In addition, we analyzed the relation between the<br />

chemical short-range force and the conductance from the tunnelling to the contact<br />

regime.<br />

[1] S. Hembacher, F. J. Giessibl, J. Mannhart, and C. F. Quate, PNAS 100, 12539 (2003).<br />

[2] H. Holscher, W. Allers, U.D. Schwarz, A. Schwarz, and R. Wisendanger Phys. Rev. B 62, 6967 (2000)<br />

[3] S, Kawai and H. Kawakatsu, Phys. Rev. B 79, 115440 (2009).<br />

[4] G. Binning et al, Europhys. Lett. 1, 31 (1986).<br />

[5] P. Jelínek, M. Švec, P. Pou, R. Pérez and V. Cháb, Phys.Rev. Lett. 101 (2008).<br />

63


We-1200<br />

<strong>Atomic</strong>-Resolution Damping <strong>Force</strong> Spectroscopy on Nanotube Peapods<br />

with Different Tube Diameters<br />

M. Ashino 1 , R. Wiesendanger 1 , A. N. Khlobystov 2 , S. Berber 3,4 , and D. Tománek 4<br />

1 Inst. <strong>of</strong> Applied Physics & Microstructure Research Center, University <strong>of</strong> Hamburg, Hamburg, Germany<br />

2 <strong>School</strong> <strong>of</strong> Chemistry, University <strong>of</strong> Nottingham, Nottingham, UK<br />

3 Physics Department, Gebze Institute <strong>of</strong> Technology, Gebze, Turkey<br />

4 Physics and Astronomy Department, Michigan State University, East Lansing, USA.<br />

Damping <strong>of</strong> the oscillating cantilever in dynamic atomic force microscopy (AFM)<br />

includes valuable information about the local vibrational structure and elastic compliance<br />

<strong>of</strong> the samples. We present atomically-resolved maps <strong>of</strong> damping in nanotube peapods,<br />

showing their capability <strong>of</strong> identifying the presence and location <strong>of</strong> encapsulated Dy@C82<br />

metall<strong>of</strong>ullerenes as well as their packing structure in the different diameters <strong>of</strong> carbon<br />

nanotubes. In our study, the physical origin <strong>of</strong> damping is elucidated in a microscopic<br />

model, and its relationship to the outer tube diameter is quantitatively interpreted by<br />

calculating the vibrational spectrum and energy dissipation in peapods with different<br />

diameters using ab initio total energy and molecular dynamics calculations [1].<br />

In our experiment we probed sparsely deposited (Dy@C82)@SWNT peapods on insulating<br />

oxide layers <strong>of</strong> a Si substrate by dynamic AFM under constant oscillation amplitude conditions in<br />

ultrahigh vacuum (p≈10 -8 Pa) and at low temperature (T≈13 K). As seen in Fig. 1(a), the<br />

atomically site-specific damping signal is obviously detected over the peapod. Besides,<br />

Figure 1(b) shows that its maximum energy is closely related to the enclosing nanotube<br />

diameter d, determined on the basis <strong>of</strong> the simultaneously observed topographic images.<br />

(a) (b)<br />

Figure 1: (a) Simultaneously obtained dynamic AFM topography (left) and damping (right)<br />

images <strong>of</strong> an empty SWNT (1,3) and a (Dy@C82)@SWNT peapod (2,4) with the same outer<br />

tube diameter d≈1.62 nm. (b) Relationship between outer tube diameter d and the maximum<br />

damping energy ΔEmax.<br />

[1] M. Ashino, R. Wiesendanger, A. N. Khlobystov, S. Berber, and D. Tománek,<br />

submitted for publication.<br />

64


We-1220<br />

Theoretical study <strong>of</strong> the forces and atomic configurations <strong>of</strong> NC-AFM<br />

P. Pou 1 , and R. Perez 1<br />

experiments on low-dimension carbon materials.<br />

1 Dpto. de Física Teórica de la Materia Condensada, Universidad Autónoma de Madrid, Madrid, Spain<br />

During the last years we have been witnessing the rise <strong>of</strong> low-dimension carbon-based<br />

materials. Fullerenes, nanotubes and graphene are nowadays key materials nowadays key<br />

materials for nanotechnology applications due to their promising electronic and<br />

mechanical properties. The outstanding capabilities <strong>of</strong> NCAFM to image, manipulate and<br />

determine the tip-surface forces at the atomic scale make it the technique <strong>of</strong> choice for<br />

the study <strong>of</strong> the basic properties and for the development <strong>of</strong> these materials [1-5].<br />

In this work we present a DFT study <strong>of</strong> the interaction <strong>of</strong> a large set <strong>of</strong> AFM tips<br />

with different low-dimension carbon materials. We have considered several possible tip<br />

terminations: reactive clean Si tip apexes, non-reactive apexes, oxygen contaminated Si<br />

apexes and metallic tips. For each <strong>of</strong> these tips, we have characterized both the<br />

conservative and non-conservative part <strong>of</strong> the tip-sample interaction, determining the<br />

force versus distance curves during approach and retraction and calculating also the<br />

dissipated energy. Our results provide insight into the origin <strong>of</strong> the atomic contrast<br />

observed in recent experiments on both graphite and a SWNT [2-4]. Moreover, we have<br />

studied the atomic origin <strong>of</strong> the particular 3D <strong>Force</strong> mapping obtained in the experiments<br />

on peapods constituted by endo-fullerenes inserted in a SWNT [1] (see Fig. 1).<br />

a) b)<br />

Figure 1: (a) Ball-and-stick model <strong>of</strong> Si tip apex model over (Dy@C82)@SWNT. (b) Electronic<br />

charge density calculated with DFT in a plane parallel to the peapod over the top atoms <strong>of</strong> the<br />

SWNT. A slight modulation produced by the Dy@C82 is observed.<br />

[1] M. Ashino et al. Nature Nanotech. 3, 337 (2008).<br />

[2] B. J. Albers et al. Nature Nanotech. 4, 57 (2009).<br />

[3] S. Hembacher et al. PNAS. 100, 12539-12542 (2003); S. Hembacher et al. PRL 94, 056101 (2005).<br />

[4] M. Ashino et al. PRL 93, 136101 (2004); M. Ashino et al. Nanotechnology 16, S134-S137 (2005).<br />

[5] H. Hölscher et al. Phys. Rev. B 62, 6967-6970 (2000).<br />

65


Anchoring highly polar molecules onto an ionic crystal<br />

We-1430<br />

Jens Schütte 1 , Ralf Bechstein 1,2 , Michael, Rohlfing 1 , Michael Reichling 1 , and Angelika<br />

Kühnle 1<br />

1 Fachbereich Physik, Universität Osnabrück, Barbarastraße 7, 49076 Osnabrück, Germany<br />

2 present address: iNano, Aarhus University, DK-8000 Aarhus C, Denmark<br />

Precise control <strong>of</strong> molecular structure formation on surfaces is most important for<br />

creating functional molecular devices for future molecular (opto)electronics applications.<br />

So far, however, controlled structure formation on dielectric surfaces has <strong>of</strong>ten been<br />

hindered by weak molecule-substrate interactions, frequently leading to clustering at step<br />

edges [1]. Here, we explore the possibility <strong>of</strong> increasing the molecule-substrate<br />

interaction by deposition <strong>of</strong> a molecule having a high dipole moment, cytosine, onto the<br />

surface <strong>of</strong> an ionic crystal, CaF2(111). Our results reveal molecular trimer structures that<br />

are stable at room temperature. Density-functional theory calculations provide insights<br />

into molecular diffusivity and structure formation. Our findings indicate that employing<br />

well-adapted molecules with high dipole moment constitutes a promising strategy for<br />

controlling structure formation on insulating surfaces.<br />

Figure 1: (a) Overview image showing cytosine structures on CaF2(111) at room temperature.<br />

The most dominant structure is a trimer as marked by a box. (b) Zoom into the marked region <strong>of</strong><br />

(a), revealing a molecular structure that can be explained by three cytosine molecules forming a<br />

hydrogen-bonded ring. The model in the lower part illustrates a plausible adsorption position with<br />

respect to the underlying substrate.<br />

[1] T. Kunstmann et al., Phys. Rev. B 71, 121403 (2005); S. Maier et al., Small 4, 1115 (2008).<br />

66


Steering the formation <strong>of</strong> molecular nanowires and compact<br />

nanocrystallites on NaCl(001)<br />

We-1450<br />

Sweetlana Fremy 1 , Alexander Schwarz 1 , Knud Laemmle 1 , Roland Wiesendanger 1 , and<br />

Marc Prosenc 2<br />

1 Institute <strong>of</strong> Applied Physics, University <strong>of</strong> Hamburg, Jungiusstr. 11, 20355 Hamburg, Germany<br />

2 Institute <strong>of</strong> Anorganic and Applied Chemistry, University <strong>of</strong> Hamburg, Martin-Luther-King Platz 6, 20146<br />

Hamburg, Germany<br />

Recently, molecular-based magnetism raised a considerable interest in the scientific<br />

community. It is envisaged that the flexibility <strong>of</strong> organic synthesis might allow in the<br />

future to combine magnetism with other physical properties like transparency or optical<br />

switchability to create multifunctional devices. Investigations <strong>of</strong> such molecules using<br />

local probes have been mostly performed on metallic surfaces. However, due to strong<br />

hybridization the electronic state, and hence the magnetic properties as well, <strong>of</strong> adsorbed<br />

molecules is usually strongly modified.<br />

Fig. 1: (a) Co-Salen molecule. (b) Nanowire networks and compact nanocrystallites on NaCl(001).<br />

(c) Molecular resolution on a compact nanocrystallite reflecting the curved shape <strong>of</strong> Co-Salen.<br />

Therefore, we deposited Co-Salen (see Fig. 1a), a paramagnetic metal-organic Schiff-<br />

Base complex, on NaCl(001), a bulk insulator, by in-situ thermal evaporation. Since the<br />

molecular states are located in the band gap <strong>of</strong> this insulating substrate, hybridization<br />

effects are absent. Co-Salen on NaCl(001) forms either compact nanocrystallites or<br />

networks <strong>of</strong> nanowires or a mixture <strong>of</strong> both (see Fig. 1b). We found that the assembly <strong>of</strong><br />

these different molecular nanostructures can be controlled by adjusting appropriate<br />

growth parameters. Moreover, the arrangement <strong>of</strong> the molecules in both morphologies<br />

could be resolved by performing molecular resolution imaging (see Fig. 1c). It turned out<br />

that the adsorption geometry at step edges, where nucleation takes place, is crucial to<br />

understand initial formation and growth <strong>of</strong> the two different morphologies.<br />

67


2-Dimensional growth <strong>of</strong> phenylenediboronic acid assisted by<br />

H-bonding<br />

We-1510<br />

R. Pawlak, L. Nony, F. Bocquet, M. Sassi, V. Oison, J.-M. Debierre, Ch. Loppacher,<br />

and L. Porte<br />

IM2NP, Aix-Marseille Université, Avenue Escadrille Normandie-Niemen, F-13397 Marseille Cedex 20,<br />

and CNRS, IM2NP (UMR 6242), F-13397 Marseille-Toulon, France<br />

Christian.Loppacher@im2np.fr<br />

Organic molecules are <strong>of</strong>ten synthesized in order to support the formation <strong>of</strong> selfassembled<br />

nanostructures. In such a way, directed non-covalent [1] as well as covalent<br />

interactions [2] have already been used to tune the size as well as the shape <strong>of</strong> molecular<br />

assemblies. 1,4 Phenylenediboronic acids (C6H8B2O4, BDBA) are molecules which are<br />

designed to engage in multiple interactions with neighbors via their –B(OH)2 groups,<br />

either by hydrogen bonding or by covalent bonding (obtained by molecular dehydration).<br />

In our work, we investigate the molecular growth <strong>of</strong> BDBA on the surface <strong>of</strong> singlecrystal<br />

potassium chloride (KCl) as well as on highly oriented pyrolytic graphite (HOPG)<br />

by means <strong>of</strong> noncontact atomic force microscopy (ncAFM). Other than on the surface <strong>of</strong><br />

Ag(111) BDBA does not form the honey-comb like covalent network [3], but rather<br />

forms chains <strong>of</strong> hydrogen-bonded dimers, and the chains then associate by lateral<br />

hydrogen bonding to create sheets [4]. The ncAFM results displayed in the figure below<br />

show that extended and well organized monolayers <strong>of</strong> BDBA are formed on KCl. The<br />

observed structure fits very well to the one observed in crystals where the benzene ring is<br />

tilted by ~ 40° with respect to the mean plane <strong>of</strong> the sheet [4]. Interestingly, this structure<br />

seems to be very stable since it is also observed on the surface <strong>of</strong> HOPG. Experimental<br />

results are discussed in respect with structural models as well as calculations <strong>of</strong> the<br />

reaction path for the formation <strong>of</strong> covalent networks.<br />

Figure 1: 0.5 ML <strong>of</strong> BDBA adsorbed on KCl imaged by ncAFM. Large monolayer islands are<br />

formed which show the formation <strong>of</strong> sheets similar to the ones observed in single crystals.<br />

[1] T. Yokoyama et al., Nature 413, 619 (2001).<br />

[2] L. Grill et al., Nature Nanotechnology 2, 687 (2007)<br />

[3] N. Zwaneveld et al., JACS 190, 6678 (2008)<br />

[4] P. Rodriguez-Cuematzi et al., Acta Cryst. E60, o1315 (2004)<br />

68


Molecular scale dissipation in oligothiophene monolayers<br />

measured by dynamic force microscopy<br />

We-1530<br />

Carlos J. Gómez 1 , Nicolas F. Martínez 1 , Wojciech Kamiński 2,4 , Cristiano Albonetti 3 ,<br />

Fabio Biscarini 3 , Ruben Perez 4 and Ricardo Garcia 1<br />

1 Instituto de Microelectronica de Madrid, CSIC, Isaac Newton 8 28760 Tres Cantos, Madrid, Spain<br />

2 Institute <strong>of</strong> Experimental Physics, University <strong>of</strong> Wrocław, p. Maksa Borna 9, 50-204 Wrocław, Poland<br />

3 CNR- (ISMN) Via P. Gobetti 101, I-40129 Bologna, Italy<br />

4 Dep. de Física Teórica de la Materia Condensada,Universidad Autónoma 28049 Madrid, Spain<br />

Amplitude modulation atomic force microscopy has enabled high resolution imaging <strong>of</strong><br />

s<strong>of</strong>t materials 1 and Phase-imaging has been applied to identify nanoscale dissipation<br />

processes 2 . We perform a combined experimental and theoretical approach to establish<br />

the atomistic origin <strong>of</strong> dissipation occurring while imaging a molecular surface with an<br />

amplitude modulation force microscope 3 . The configuration space sampled by the tip<br />

depends on whether it approaches or withdraws from the surface. The asymmetry arises<br />

because <strong>of</strong> the presence <strong>of</strong> energy barriers among different deformations <strong>of</strong> the molecular<br />

geometry. This is the source <strong>of</strong> the material contrast provided by the images.<br />

[1] R. Garcia, R. Magerle, R. Perez, Nat. Mater. 7, 405 (2007).<br />

[2] R. Garcia, C. J. Gomez, N. F. Martinez, S. Patil, C. Dietz, R. Magerle, Phy. Rev. Lett. 97, 016103<br />

(2006).<br />

[3] Phy. Rev. Lett. (2009) Accepted.<br />

69


Oral<br />

Presentations<br />

Thursday, 13 August<br />

70


Dependence <strong>of</strong> the atomic scale image <strong>of</strong> a Si adatom on the tip apex<br />

termination: a DFT study<br />

Anna Campbellova 1 , Pablo Pou 2 , Ruben Pérez 2 , Petr Klapetek 1 and Pavel Jelínek 3<br />

1 Czech Metrology Institute, Brno, Czech Republic<br />

2 Department <strong>of</strong> Physics, <strong>Yale</strong> University, New Haven, Spain.<br />

3 Department <strong>of</strong> Thin Films, Institute <strong>of</strong> Physics <strong>of</strong> the ASCR, Prague, Czech Republic<br />

Th-0900<br />

High-resolution measurements <strong>of</strong> tip-sample forces became possible with dynamic force<br />

spectroscopy (DFS) [1]. Furthermore this technique was adopted to measure threedimensional<br />

force ?elds with atomic resolution [2,3]. In principle, these measurements<br />

can provide, by separating the short-range contribution from the total tip-sample<br />

interaction, detailed information about e.g. the surface energy landscape, adhesion forces<br />

and inter-atomic forces. Sub-atomic resolution <strong>of</strong> Si adatoms on Si(111)-(7x7) has also<br />

been reported and its interpretation is still source <strong>of</strong> debate.<br />

Here we have performed DFT simulations <strong>of</strong> a 3D scan over a silicon cluster –that<br />

reproduces the local coordination <strong>of</strong> a surface adatom--, using different Sibased tips [5].<br />

Our aim is to reveal the effect <strong>of</strong> apex atomic termination on the resulting atomic scale<br />

imaging <strong>of</strong> individual Si adatom. We have found a strong variation <strong>of</strong> the internal atomic<br />

contrast with the symmetry and chemical composition <strong>of</strong> the tip apex. In particular, the<br />

theoretical atomic image provided by a dimer-like Si tip with an oxygen atom positioned<br />

on apex mimics very well the characteristic subatomic features observed experimentally<br />

[4].<br />

Figure 1: Theoretical DFT images <strong>of</strong> Si adatom provided by three different apex tip structures.<br />

[1] M. A. Lantz et al, Science 291, 2580 (2001).<br />

[2] H. Hölscher, S. M. Langkat, A. Schwarz, and R. Wiesendanger, Appl. Phys. Lett. 81, 4428 (2002).<br />

[3] Y. Sugimoto, T. Namikawa, K. Miki, M. Abe, and S. Morita Phys. Rev. B 77, 195424 (2008).<br />

[4] F. J. Giessibl, S. Hembacher, H. Bielefeldt, and J. Mannhart, Science 289, 422 (2000)<br />

[5] P.Pou et al Nanotechnology accepted (2009)<br />

71


AFM probe tips with a small front atom<br />

Th-0920<br />

T. H<strong>of</strong>mann, 1 J. Welker, 1 M. Ternes, 2 C. P. Lutz, 2 A. J. Heinrich, 2 Franz J. Giessibl 1<br />

1 University <strong>of</strong> Regensburg, Institute <strong>of</strong> Experimental and Applied Physics, D-93040 Regensburg, Germany<br />

2 IBM Research Division, Almaden Research Center, 650 Harry Rd, San Jose, CA 95120, USA<br />

A direct comparison <strong>of</strong> scanning tunneling microscopy and AFM data showed that when<br />

probing a tungsten tip with a graphite surface (a “light-atom probe”), subatomic orbital<br />

structures with a spatial resolution <strong>of</strong> less than one Angstrom can be obtained in the force<br />

map, while a map <strong>of</strong> the tunneling current only shows the known atomic resolution. 1<br />

Optimized subatomic contrast is obtained when recording the higher harmonics <strong>of</strong> the<br />

cantilever motion. 1 The idea <strong>of</strong> the light atom probe was carried further by using an adsorbed<br />

CO molecule to probe the tip atom. It requires quite a large force to move a CO molecule<br />

laterally 2 and this molecule is an excellent probe for the orbital structure <strong>of</strong> the front atom <strong>of</strong><br />

the metal tip. In contrast to previous examples <strong>of</strong> “subatomic” resolution, where a flat sample<br />

with a periodic array <strong>of</strong> sample atoms created multiple tip images, using a single adsorbed CO<br />

molecule as a “tip” allows to isolate the force contribution <strong>of</strong> a single atom-pair contact.<br />

Ideally, one wishes to create a probe with a perfectly perpendicularly oriented CO molecule at<br />

the very end <strong>of</strong> the tip. First results will be discussed where a CO molecule has been picked<br />

up by a metal tip and used to image adsorbed atoms.<br />

In order to allow simultaneous STM and AFM, high electrical conductivity is another<br />

desirable property <strong>of</strong> a light atom probe. Because <strong>of</strong> its small atomic radius, high mechanical<br />

strength and fair electric conductivity Beryllium is a promising choice as a probe material.<br />

However, due to its high reactivity, preparation is a challenge. Results using Be tips prepared<br />

by high-voltage field emission for atomic imaging <strong>of</strong> Si(111)-(7×7) will be presented.<br />

1. S. Hembacher, F. J. Giessibl, J. Mannhart, Science 305, 380 (2004).<br />

2. M. Ternes, C. P. Lutz, C.F. Hirjibehedin, F. J. Giessibl, A. J. Heinrich, Science 319, 1066<br />

(2008).<br />

72


Th-0940<br />

Intramolecular features <strong>of</strong> organic molecules characterized by force<br />

field spectroscopy: The case <strong>of</strong> PTCDA on Cu and Ag<br />

Gernot Langewisch 1 , Daniel-Alexander Braun 1 , Domenique Weiner 2 ,<br />

Bartosz Such 3 , Harald Fuchs 1 , and Andre Schirmeisen 1<br />

1CeNTech (Center for Nanotechnology) and Institute <strong>of</strong> Physics, University <strong>of</strong> Muenster, Germany<br />

2 SPECS Zurich GmbH, Switzerland<br />

3 Marian Smoluchowski Institute <strong>of</strong> Physics, Jagiellonian University Krakow, Poland<br />

Thin films <strong>of</strong> π-conjugated organic molecules, like the organic semiconductor PTCDA,<br />

are <strong>of</strong> high relevance for nanoelectronic applications. We use non-contact atomic force<br />

microscopy in ultrahigh vacuum at room temperature to investigate the forces between<br />

the tip and PTCDA molecules deposited on Cu(111) and Ag(111) surfaces by molecular<br />

beam epitaxy.<br />

Submolecular features <strong>of</strong> the PTCDA layers on both substrates are resolved in the<br />

topography scans. In particular we find that the second layer molecules show an<br />

intramolecular structure with a height corrugation <strong>of</strong> up to 40pm, while molecules in the<br />

first layer above the substrate are depicted as featureless ovals [1]. To study this effect in<br />

detail, 2-dimensional cuts <strong>of</strong> the spatial tip-sample force landscape above individual<br />

molecules were obtained by force field spectroscopy. In the double layer these cuts, each<br />

consisting <strong>of</strong> 40 force spectroscopy curves, reveal an enhanced tip-sample force<br />

interaction at the molecular end groups compared to the centre <strong>of</strong> the molecule [2].<br />

However, for the monolayer molecules this effect is not present. This is interpreted with<br />

respect to different mechanical relaxation processes <strong>of</strong> the molecular functional end<br />

groups as well as variations <strong>of</strong> the internal electron density distributions <strong>of</strong> the molecules.<br />

Figure 1: Left: Topography scans <strong>of</strong> a double layer <strong>of</strong> PTCDA molecules on Cu(111) deposited<br />

by molecular beam epitaxy, showing intramolecular contrast. Right: <strong>Force</strong> field spectroscopy<br />

image along line in left image, visualizing the intramolecular variations <strong>of</strong> the tip-sample force.<br />

[1] B. Such, A. Schirmeisen, D. Weiner, and H. Fuchs, Appl. Phys. Lett. 98, 093104 (2006).<br />

[2] D.-A. Braun, D. Weiner, B. Such, H. Fuchs and A. Schirmeisen, Nanotechnology (2009) submitted.<br />

73


Th-1000<br />

Analysis <strong>of</strong> bimodal and higher mode small-amplitude near-contact<br />

AFM and energy dissipation<br />

S. Kawai, A. Barat<strong>of</strong>f, Th. Glatzel, S. Koch, B. Such, and E. Meyer<br />

Department <strong>of</strong> Physics, University <strong>of</strong> Basel, Klingelbergstr. 82, 4056 Basel Switzerland<br />

Recently enhanced atomic-scale contrast was achieved on KBr(001) in UHV using<br />

resonant self-excitation <strong>of</strong> the lowest two cantilever flexural modes with constant tip<br />

oscillation amplitudes A1 >> A2 by monitoring the interaction induced frequency shift Δf2<br />

while controlling the closest approach distance z via Δf1 [1]. Optimum sensitivity and<br />

contrast in non-contact AFM using the 2 nd mode alone can in principle be achieved for z<br />

below the interatomic separation and A2 close to the range <strong>of</strong> the tip-sample interaction [2<br />

3]. Stable operation under such conditions can, however, be jeopardized by atomic jumps<br />

even if macroscopic jump-to-contact is prevented by using a sufficiently stiff deflection<br />

sensor. Simultaneous excitation <strong>of</strong> the fundamental mode with a large A1 alleviates this<br />

problem for reasons to be discussed.<br />

Under the stated conditions, Δf2 is to a good approximation proportional to the time<br />

average <strong>of</strong> the interaction force gradient over one cycle <strong>of</strong> the fundamental oscillation, as<br />

demonstrated and verified by simulations and measurements [1]. An application <strong>of</strong> the<br />

same argument to atomic-scale Kelvin <strong>Force</strong> <strong>Microscopy</strong> using voltage modulation at f2<br />

shows that the resulting deflection signal at f2 is proportional to the time average <strong>of</strong> the<br />

electrostatic force [4].<br />

Measurements to be reported separately using the 2 nd mode alone with amplitudes as<br />

small as 0.5 nm show smooth force vs. distance curves extracted from Δf2, as well as a<br />

small time-averaged deflection proportional to the time-average <strong>of</strong> the force.<br />

In all three cases, the site-dependent distance dependence <strong>of</strong> the time-averaged force<br />

gradient, electrostatic or total force can be extracted from the measured frequency shift,<br />

deflection signal at f = f2 or f = 0 using slightly modified versions <strong>of</strong> inversion algorithms<br />

proposed for single-mode non-contact AFM [5, 6]<br />

The magnitude, distance and amplitude dependences <strong>of</strong> the interaction-induced energy<br />

dissipation per oscillation cycle simultaneously measured using the 2 nd mode suggest that<br />

the dominant cause <strong>of</strong> dissipation is a force hysteresis loop narrower than twice A, but<br />

significantly smeared by thermal activation. However, some velocity-dependent<br />

damping with a coefficient decaying with increasing distance must also be present.<br />

Possible mechanisms and a hitherto ignored difficulty in extracting meaningful<br />

"dissipative forces" in the presence <strong>of</strong> both mechanisms will be discussed.<br />

[1] S. Kawai et al., submitted.<br />

[2] F. J. Giessibl et al., Appl. Surf. Sci.140, 352 (1999).<br />

[3] S. Kawai et al., Appl. Phys. Lett. 89, 023113 (2006).<br />

[4] S. Kawai et al., unpublished<br />

[5] O. Pfeiffer et al., Phys. Rev. B 65, 161403R (2002)<br />

[6].J. E. Sader and S. P. Jarvis, Appl. Phys. Lett. 84, 1801 (2004)<br />

74


Th-1020<br />

Adhesion-induced energy dissipation and atom-tracked tip changes<br />

S. Kawai, Th. Glatzel, S. Koch, B. Such, A. Barat<strong>of</strong>f, and E. Meyer<br />

Department <strong>of</strong> Physics, University <strong>of</strong> Basel, Klingelbergstr. 82, 4056 Basel Switzerland<br />

We have simultaneously measured the frequency shift Δf and the interaction-induced<br />

dissipated energy per cycle Ets above a maximum in a topographic image <strong>of</strong> KBr(001).<br />

The measurements were performed at room temperature in UHV using the 2nd flexural<br />

mode <strong>of</strong> a silicon cantilever (1039369 Hz) at 11 constant amplitudes A between 12.8 and<br />

0.51 nm. As shown in fig. 1, Ets rises smoothly above the noise level, then almost<br />

saturates as the closest approach distance extends below the attractive force minimum.<br />

Whereas the conservative force vs. distance curves extracted from Δf were smooth and<br />

coincided apart from noise, Ets decreased smoothly by only a factor <strong>of</strong> 2 and became less<br />

noisy as A was reduced from 12.8 to 0.51 nm. This behavior, and the magnitude <strong>of</strong> Ets<br />

suggest that the dominant cause <strong>of</strong> dissipation is a force hysteresis loop narrower than<br />

twice A, but significantly smeared by thermal activation. However, some velocitydependent<br />

damping with a coefficient decaying with increasing distance is also present.<br />

Using A = 0.285 nm, one thousand approach-retraction curves were recorded, separated<br />

by intervals for atom-tracking (Nanonis AT4). In most cases one <strong>of</strong> three distinct Ets vs.<br />

distance curves are obtained, whereas the conservative force curves are almost the same,<br />

as shown in fig. 1 (a-c). This is consistent with the higher sensitivity <strong>of</strong> Ets to the tip apex<br />

configuration. In a few cases we observed jumps between curves corresponding to two <strong>of</strong><br />

those long-lived configurations, accompanied by significant deviations between approach<br />

and retraction. As shown in fig. 2, the tracked 3D motion <strong>of</strong> the sample relative to the tip<br />

at Δf = -150 Hz exhibits strong correlations between more frequent jumps (strongly<br />

reduced by averaging over each tracking interval). This suggests that tip changes occur<br />

over a range <strong>of</strong> time scales. Further implications <strong>of</strong> those results will be discussed.<br />

Figure 1 Figure 2<br />

75


Th-1120<br />

Combined Qplus-AFM and STM imaging <strong>of</strong> the Si(100) surface:<br />

Activating the c(4x2) to p(2x1) transition with subnanometre oscillation<br />

amplitudes<br />

A. Sweetman, S. Gangopadhyay, R. Danza, and P. Moriarty<br />

<strong>School</strong> <strong>of</strong> Physics and Astronomy, University <strong>of</strong> Nottingham,<br />

Nottingham NG7 2RD, UK<br />

We report the first combined Qplus AFM and STM images <strong>of</strong> the Si(100) surface. The<br />

study <strong>of</strong> the Si(100)-(2x1) and c(4x2) reconstructions using scanning probe methods<br />

(both STM and NC-AFM) has led to a variety <strong>of</strong> intriguing and controversial issues<br />

associated with identifying the ground state <strong>of</strong> the system [1]. Foremost among these<br />

have been the determination <strong>of</strong> whether the silicon dimers, which give rise to the (2x)<br />

periodicity at the Si(100) surface, are asymmetric or symmetric, and the extent to which<br />

the measurement technique might influence the energy balance. Previous experimental<br />

NC-AFM measurements (with ~ 10 nm amplitude) combined with theoretical<br />

calculations [2] showed that the AFM probe could strongly influence the Si(100) surface,<br />

stabilising a p(2x1) phase at low temperatures. Our Qplus measurements, taken with<br />

probe oscillation amplitudes ranging from 250 pm to 10 nm, not only confirm that the<br />

strong influence <strong>of</strong> the tip is retained at subnanometre oscillation amplitudes, but, as<br />

shown in Fig. 1, suggest that tip symmetry plays a central role in stabilisation <strong>of</strong> a<br />

particular surface phase. In addition, there are clear and striking differences between<br />

STM and Qplus AFM images <strong>of</strong> the Si(100)-(2x1) phase arising from the residual dimer<br />

buckling associated with probe-induced stabilisation <strong>of</strong> the p(2x1) structure.<br />

[1] Lev Kantorovich and Chris Hobbs, Phys. Rev. B 73 245420 (2006) and references therein<br />

[2] Yan Jun Li et al., Phys. Rev. Lett. 96 106104 (2006)<br />

Fig. 1 Qplus AFM images <strong>of</strong> the Si(100) surface. (a) Regions <strong>of</strong> both (2x1) and c(4x2)<br />

symmetry are visible. (df=-10.9 Hz, A vib=350 pm (700 pm pk-pk), V gap=+0.3V). (b) and (c)<br />

Images <strong>of</strong> the same surface region taken in parallel with identical scan parameters showing the<br />

influence <strong>of</strong> the scan direction on the surface phase. Image (b) was taken while the tip was<br />

moving from left to right. For Image (c) the tip was moving from right to left.<br />

76


Measuring <strong>Atomic</strong> Charge States by nc-AFM<br />

Th-1140<br />

Leo Gross 1 , Fabian Mohn 1 , Peter Liljeroth 1,2 , Jascha Repp 1,3 , Franz J. Giessibl 3 , and<br />

Gerhard Meyer 1<br />

1<br />

IBM Research, Zurich Research Laboratory, 8803 Rüschlikon, Switzerland<br />

2<br />

Debye Institute for Nanomaterials Science, Utrecht University, 3508 TA Utrecht, The Netherlands<br />

3<br />

Institute <strong>of</strong> Experimental and Applied Physics, University <strong>of</strong> Regensburg,<br />

93040 Regensburg, Germany<br />

We investigated the charge state switching <strong>of</strong> individual gold and silver adatoms on<br />

ultrathin NaCl films on Cu(111) using a qPlus tuning fork AFM operated at 5 Kelvin<br />

with oscillation amplitudes in the sub-Ångstrom regime. Charging <strong>of</strong> a gold adatom by<br />

one electron charge increases the force on the AFM tip by a few piconewtons. Moreover,<br />

the local contact potential difference is shifted depending on the sign <strong>of</strong> the charge and<br />

allows the discrimination <strong>of</strong> positively charged, neutral, and negatively charged atoms.<br />

Furthermore, we modified AFM tips by means <strong>of</strong> vertical manipulation techniques, i.e.<br />

deliberately picking up known adsorbates, to study the effect <strong>of</strong> the atomic tip<br />

termination on AFM contrast and sensitivity.<br />

77<br />

Figure 1: (a) Constant-current STM<br />

measurement (V = –50mV, I = 2pA) <strong>of</strong> a<br />

neutral (Au 0 , left) and a negatively charged<br />

gold adatom (Au − , right) adsorbed on<br />

NaCl(2ML)/Cu(111). The line scan is<br />

through the center <strong>of</strong> both adatoms shown<br />

in the inset (image size <strong>of</strong> insets:<br />

55Å × 17Å). (b) Current and (c) frequency<br />

shift recorded simultaneously in a constantheight<br />

measurement <strong>of</strong> the same area<br />

(Δz = 5.0Å, V = –5mV, A = 0.3Å).


Th-1200<br />

Mechanism <strong>of</strong> Dissipative Interaction by Tunneling Single-Electrons<br />

Yoichi Miyahara 1 , L. Cockins 1 , S. D. Bennett 1 , A. A. Clerk 1 , S. A. Studenikin 2 ,<br />

P. Poole 2 , A. Sachrajda 2 and P. Grutter 1<br />

1 Department <strong>of</strong> Physics, McGill University, Montreal, Canada<br />

2 Institute for Microstructural Science, National Research Council <strong>of</strong> Canada, Ottawa, Canada<br />

The frequency modulation mode atomic force microscopy (FM-AFM) can be used as a<br />

sensitive electrometer which can detect the motion <strong>of</strong> a single electron. An oscillating<br />

AFM tip with an applied dc-bias voltage (Vbias) can allow a single electron to tunnel back<br />

and forth between a quantum dot (QD) and a back-electrode by oscillating the chemical<br />

potential <strong>of</strong> the QD around that <strong>of</strong> the back-electrode, owing to Coulomb blockade effect.<br />

The resulting electron motion in turn modulates the electrostatic force acting on the tip<br />

and causes the dissipation as well as the resonance frequency shift (Δf) <strong>of</strong> the cantilever.<br />

Both dissipation and Δf images taken in constant-height mode show characteristic<br />

concentric rings around the QD, each <strong>of</strong> which reflects the above-mentioned single-<br />

electron tunneling. The corresponding features also appear as peaks (dips) in the<br />

dissipation (Δf) versus Vbias spectra taken above the QD. These spectra can be interepreted<br />

as addition energy spectra which enable us to investigate the electronic structure <strong>of</strong> a<br />

single QD.<br />

In this contribution, we present the mechanism <strong>of</strong> the dissipative electrostatic<br />

interaction due to the tunneling single-electrons in detail. In essence, this dissipative<br />

interaction arises from the delayed response <strong>of</strong> a single tunneling electron to the<br />

oscillating chemical potential induced by the oscillating tip. The delay is due to the finite<br />

tunneling rate which is determined by the tunnel barrier.<br />

We developed a theoretical model for this dissipation process and obtained a very<br />

good agreement between the theoretical dissipation versus Vbias curve and the<br />

experimental one (Fig. 1). We also discuss the effect <strong>of</strong> the tip oscillation amplitude and<br />

temperature and the relation between Δf and dissipation signal.<br />

78<br />

Figure 1: Theoretical (solid) and<br />

experimental (dashed) dissipation versus<br />

bias voltage curves. (T = 30 K, A=0.5 nm,<br />

Tip-QD distance = 15 nm)


Th-1220<br />

Controlling electron transfer processes on insulating surfaces with the<br />

NC-AFM<br />

Thomas Trevethan 1,2 and Alexander Shluger 2<br />

1 Department <strong>of</strong> Physics and Astronomy, University College London, London, UK<br />

2 WPI Advanced Institute for Materials Research, Tohoku University, Sendai, Japan.<br />

We present theoretical modeling that predicts how a single electron transfer process can<br />

be induced between two defects on an insulating surface using the non-contact <strong>Atomic</strong><br />

<strong>Force</strong> Microscope. On an insulating substrate one can realize the possibility <strong>of</strong> using the<br />

localized perturbation produced by an AFM tip to modify electronic structure locally on<br />

the surface, which could result in the control <strong>of</strong> electronic processes and transitions. The<br />

field produced by the tip at close approach may modify both the relative energies <strong>of</strong><br />

distinct electronic states as well as the potential energy barrier separating them.<br />

A model but realistic system is employed which consists <strong>of</strong> a neutral oxygen vacancy and<br />

a noble metal (Pt or Pd) adatom on the MgO (001) surface. We show that the ionization<br />

potential <strong>of</strong> the vacancy and the electron affinity <strong>of</strong> the metal adatom can be significantly<br />

modified by the electric field produced by an ionic tip apex at close approach to the<br />

surface. The relative energies <strong>of</strong> the two states are also a function <strong>of</strong> the separation <strong>of</strong> the<br />

two defects. Therefore the transfer <strong>of</strong> an electron from the vacancy to the metal adatom<br />

can be induced either by the field effect <strong>of</strong> the tip or by manipulating the position <strong>of</strong> the<br />

metal adatom on the surface. We also show how the transition <strong>of</strong> an electron can be<br />

observed in the change <strong>of</strong> image contrast. The realization <strong>of</strong> these procedures<br />

experimentally would mean that a single electron transfer process could be directly<br />

observed and controlled, and would herald a new regime <strong>of</strong> control on the atomic scale.<br />

Figure 1: (a) Schematic <strong>of</strong> the system, showing an Mg terminated MgO tip directly above a<br />

metal adatom on the surface, adjacent to an oxygen vacancy. (b) Illustration <strong>of</strong> adiabatic total<br />

energy curves for the two states at two defect separations (A > B), where Q is a structural degree<br />

<strong>of</strong> freedom. (c) Illustration <strong>of</strong> potential energy wells for the two states, depicting how barrier<br />

width increases at larger separation.<br />

79


Th-1240<br />

NC-AFM imaging with atomic resolution in a temperature range<br />

Andreas Bettac and Albrecht Feltz<br />

between 5 K and 1083 K<br />

Omicron NanoTechnology GmbH, Taunusstein, Germany<br />

We present the latest results <strong>of</strong> atomically resolved NC-AFM (QPlus) measurements on a<br />

Si(111) 7x7 surface at high sample temperatures. Starting at room temperature, the<br />

temperature was increased stepwise to 1083 K by direct current heating. Even at these<br />

high temperatures it is possible to achieve atomic resolution <strong>of</strong> the 7x7 reconstruction in<br />

NC-AFM mode (Fig 1a) .<br />

Since the phase transition between the 1x1 and the 7x7 reconstruction starts to occur at<br />

this temperatures, we focused our experiments on the observation <strong>of</strong> the phase transition.<br />

In NC-AFM and dynamic STM mode we observed fluctuating formations <strong>of</strong> step edges<br />

and kinks. Typically, these kinks have a width <strong>of</strong> half a unit cell - most likely due to<br />

breaking the dimer bonds. Both effects are indicators for the phase transition. At the same<br />

high temperature and without changing the sensor we were able to image this<br />

characteristic step movement in STM mode with atomic resolution <strong>of</strong> the 7x7<br />

reconstruction. The size <strong>of</strong> the 1x1 areas at the lower side <strong>of</strong> a step edge permanently<br />

changes - affecting the 7x7 area.<br />

With the developed instrument we could resolve the rest atoms <strong>of</strong> the Si(111) 7x7<br />

reconstruction at a sample temperature <strong>of</strong> 50 K and observed a different brightness for the<br />

rest atoms <strong>of</strong> the faulted and unfaulted half (Fig. 1b).<br />

Furthermore, atomic resolution in NC-AFM mode on metallic Au(111) and Ag(111)<br />

surfaces at 5 K will be presented [1].<br />

a) b) c)<br />

T= 1053 K<br />

T= 50 K T= 5 K<br />

Figure 1: NC-AFM results: a) atomically resolved<br />

Si(111) 7x7 at a sample temperature<br />

<strong>of</strong> 1053 K; b) Si(111) 7x7 at 50K, the atomically resolved image <strong>of</strong> the Si(111) 7x7<br />

surface clearly shows the restatoms and their different brightness for the faulted and<br />

unfaulted half; and c) atomically resolved Au(111) surface at 5K.<br />

[1]<br />

A. Bettac, J. Koeble, K. Winkler, B. Uder, M. Maier, and A. Feltz, Nanotechnology 20, 264009 (2009).<br />

80


Oral<br />

Presentations<br />

Friday, 14 August<br />

81


Th-1240 Fr-0900<br />

<strong>Atomic</strong> scale elasticity mapping <strong>of</strong> Ge(001) surface by multifrequency<br />

FM-AFM<br />

Yoshitaka Naitoh, Zongmin Ma, Yanjun Li, Masami Kageshima and Yasuhiro Sugawara<br />

Department <strong>of</strong> Applied Physics, Osaka University, Suita 565-0871, JAPAN<br />

Surface elasticity is quite important to study atomic scale properties <strong>of</strong> the surface<br />

such as bounding strength <strong>of</strong> surface atoms and surface phonons. The conventional FM-<br />

AFM provides topographic information <strong>of</strong> various surfaces at atomic scale. However, it is<br />

hardly accessible to the elasticity and the adhesion on surfaces. In this study, we propose<br />

a new technique, multifrequency FM-AFM, on Ge(001) surface using a commercial<br />

cantilever in order to investigate the surface elasticity at atomic scale with the<br />

topographic image.<br />

The cantilever <strong>of</strong> the multifrequency FM-AFM was simultaneously excited at the<br />

by two sets <strong>of</strong> automatic gain<br />

1st and the 2nd flexural resonant frequencies (f1st, f2nd)<br />

controller and phase locked loop electronics. Their oscillating amplitudes are set constant<br />

at A1st, A2nd. The cantilever oscillating signal, detected by an optical interferometer<br />

system, is divided into the 1st and the 2nd components through band pass filters. The<br />

surface topography is obtained from feedback signal maintaining the frequency shift <strong>of</strong><br />

the 1st component (Δf1st) constant. From the theoretical consideration, we found that the<br />

surface elasticity is obtained as a mapping <strong>of</strong> the frequency shift <strong>of</strong> the 2nd component<br />

(Δf2nd).<br />

The cleaned Ge(001) surface showing buckled dimer structure was adopted as a<br />

sample with the elastic distribution at atomic scale. Figures 1(a) and (b) show the<br />

simultaneously obtained topography and the Δf2nd mapping <strong>of</strong> the surface. The dimer<br />

structure <strong>of</strong> the surface was clearly resolved in both images. We found Δf2nd <strong>of</strong> the fig.<br />

1(b) was high at the dimer atom position in comparison with the topography. This result<br />

suggests multifrequency FM-AFM is a nicer tool for exploring the surface elasticity at<br />

atomic scale.<br />

Figure 1: (a) Topographic image <strong>of</strong> Ge(001) surface with Δf1st=-70Hz, f1st=160kHz and<br />

A1st=4nm. Scan area is 7x7 nm 2 . (b) Simultaneously obtained Δf2nd mapping <strong>of</strong> the surface area<br />

with Δf2nd=-21Hz, f2nd=1.0MHz and A2nd=0.3nm.<br />

82


Small amplitude atomic resolution NC-AFM imaging and force<br />

spectroscopy experiments using a stiff piezoelectric force sensor<br />

Stefan Torbrügge 1 , Jörg Rychen 2 , Oliver Schaff 1<br />

1 SPECS GmbH, Voltastrasse 5, 13355 Berlin, Germany<br />

2 SPECS Zurich GmbH, Technoparkstrasse 1, 8005 Zurich, Switzerland<br />

Fr-0920<br />

We present the design <strong>of</strong> the recently introduced KolibriSensor TM and its application to<br />

frequency modulation atomic force microscopy. The KolibriSensor TM , schematically<br />

shown in Fig. 1(a), is a new type <strong>of</strong> piezoelectric force sensor based on a quartz length<br />

extension resonator (LER) [1]. Key features <strong>of</strong> the KolibriSensor TM are a high spring<br />

constant (k~540,000 N/m), avoiding snap into contact when operated at small amplitudes,<br />

and a large Q-factor exceeding 10,000 at room temperature. Its high resonance frequency<br />

(fres~1 MHz) enables fast scanning and an excellent signal-to-noise ratio necessary for<br />

stable operation at sub-nanometer oscillation amplitudes. In-situ sputtering enables<br />

repeated sharpening <strong>of</strong> the separately contacted tungsten tip. Switching between AFM,<br />

STM, or combined feedback modes is possible on the fly during measurement.<br />

We demonstrate atomic resolution measurements on Si(111)-(7x7) and insulating<br />

KBr(001), see Fig. 1(b), with oscillation amplitudes down to 30 pm at room temperature<br />

and scanning speeds <strong>of</strong> several lines/s. Furthermore the performance <strong>of</strong> the sensor in<br />

force spectroscopy experiments is evaluated by acquisition <strong>of</strong> two-dimensional force<br />

maps measured on KBr(001) (see Fig. 1 (c)).<br />

Figure 1: (a) Schematic drawing <strong>of</strong> the KolibriSensor TM . (b) Topographic atomic resolution NC-<br />

AFM image <strong>of</strong> the KBr(001) surface recorded at room temperature. (c) Vertical tip-sample force<br />

map F(x,z) derived from 20 Δf(z) curves taken along the line in the insert image in (c).<br />

[1] T. An, et al. Appl. Phys. Lett., 88, 149903 (2006); T. An, et al., Rev. Sci. Instrum., 79, 033703 (2008)<br />

83


Fr-0940<br />

Determination <strong>of</strong> the Optimum Spring Constant and Oscillation<br />

Amplitude for <strong>Atomic</strong>/Molecular-Resolution FM-AFM<br />

Yoshihiro Hosokawa 1 , Kei Kobayashi 2 , Hir<strong>of</strong>umi Yamada 1 and Kazumi Matsushige 1<br />

1 Department <strong>of</strong> Electronic Science and <strong>Engineering</strong>, Kyoto University, Kyoto, Japan<br />

2 Innovative Collaboration Center, Kyoto University, Kyoto, Japan.<br />

Several studies have shown that the lateral resolution <strong>of</strong> FM-AFM on Si(111)-7x7 surface<br />

can be improved by oscillating a force sensor with a very high spring constant (> 1,000<br />

N/m) at a very small amplitude (< 1 nm) [1,2]. However, the resolution <strong>of</strong> FM-AFM on<br />

an organic thin film was not improved by the use <strong>of</strong> a cantilever with a spring constant <strong>of</strong><br />

about 700 N/m [3]. Here we propose a general procedure to determine the optimum<br />

imaging parameters (spring constant and oscillation amplitude) for obtaining atomic or<br />

molecular resolution by FM-AFM.<br />

We first determine the minimum distance between the tip and the sample, which is<br />

limited by the instability caused by jump-into-contact the sample or by the dissipative tipsample<br />

interaction forces. Then we defined effective signal intensity for atom/molecularresolution<br />

FM-AFM as the difference between the frequency shift on top <strong>of</strong> the<br />

atom/molecule and on the gap between the atoms/molecules at the minimum distance.<br />

Figure 1 shows a plot <strong>of</strong> the calculated signal-to-noise ratio on lead-phthalocyanine thin<br />

films on MoS2. The optimum spring constant and oscillation amplitude were 10 N/m and<br />

25 nm, respectively. Figure 2(a) is a preliminary experimental result using 7.4 N/m<br />

cantilever. Line pr<strong>of</strong>iles obtained from Fig. 2(a) (red, bold) and from the image (black,<br />

thin) using 40 N/m cantilever [3] are shown in Fig. 2(b). From Fig.2(b), the corrugation<br />

was larger and consequently signal-to-noise ratio was improved when we used a<br />

cantilever with a smaller spring constant and a larger oscillation amplitude.<br />

Figure 1: Signal-to-noise ratio for<br />

molecular-resolution FM-AFM on PbPc with<br />

various spring constant and oscillation<br />

amplitude. The optimum spring constant and<br />

oscillation amplitude is 10 N/m and 25 nm.<br />

[1] F. J. Giessibl et al., Appl. Surf. Sci., 140 (199) 352.<br />

[2] S. Kawai et al., Appl. Phys. Let., 86 (2005) 193107.<br />

[3] Y. Hosokawa et al., Jpn. J. Appl. Phys., 47(2008) 6125.<br />

84<br />

Figure 2: (a) Topographic<br />

image <strong>of</strong> PbPc<br />

obtained with using small spring constant<br />

cantilever (7.4 N/m) at large amplitude (15<br />

nm). (b) Line pr<strong>of</strong>iles obtained from (a)<br />

(red, bold) and from the image obtained<br />

using 40 N/m cantilever


Fr-1000<br />

Visualization <strong>of</strong> Anisotropic Conductance in Polydiacetylene Crystal<br />

by Two-probe FM-AFM/KFM<br />

Eika Tsunemi 1 , Kei Kobayashi 2 , Kazumi Matsushige 1 , and Hir<strong>of</strong>umi Yamada 1<br />

1 Department <strong>of</strong> Electronic Science & <strong>Engineering</strong>, Kyoto University, Katsura Nishikyo, Kyoto, Japan<br />

2 Innovative Collaboration Center, Kyoto University, Katsura Nishikyo, Kyoto, Japan<br />

<strong>Atomic</strong> force microscopy probe tips serve as a wide variety <strong>of</strong> roles such as<br />

nanometer-scale electrical probes or atom manipulation tools in addition to imaging tools.<br />

However, their simultaneous, multiple use is <strong>of</strong>ten limited because only a single probe is<br />

available in AFM. Implementation <strong>of</strong> two or more probe tips can tremendously expand<br />

the capability <strong>of</strong> AFM. We recently developed a high-resolution two-probe AFM system<br />

using the optical beam deflection method [1]. It allows us to make a multi-probe<br />

electrical measurement or to conduct stimulus-response experiments for various materials.<br />

In this presentation, anisotropic conduction in a polydiacetylene (PDA) single<br />

crystal was studied by the two-probe FM-AFM/KFM. A PDA single crystal shows an<br />

anisotropic conductance due to the quasi-one-dimentional electronic band structure along<br />

the diacetylene main chain. In this experiment, we used a poly-PTS (polydiacetylene<br />

para-toluene sulfonate) single crystal (inset <strong>of</strong> the right <strong>of</strong> Fig. 1). The left <strong>of</strong> Fig. 1<br />

shows an experimental setup. While a bias voltage was locally applied to the surface with<br />

Probe-1, the surface potential was mapped with Probe-2 as a FM-KFM probe to visualize<br />

the injected carrier distribution. An obtained KFM image (the right <strong>of</strong> Fig. 1) shows that<br />

the higher potential region extends linearly along the PDA main chain from the Probe-1<br />

contact point, which means that the carriers injected from Probe-1 travel mainly through<br />

the diacetylene chain. In addition, the effect <strong>of</strong> a defect on the carrier transport was<br />

investigated. We positioned Probe-1 at a point on a line parallel to the main chain through<br />

a line defect which we had found. The right image <strong>of</strong> Fig. 2 shows that the potential<br />

sharply drops near the edge <strong>of</strong> the defect, which indicates a resistance increase at this<br />

point.<br />

Figure 1: AFM (Left) and KFM (Right) images<br />

<strong>of</strong> a poly-PTS single crystal obtained by twoprobe<br />

FM-AFM/KFM. Schematic <strong>of</strong> two-probe<br />

measurement is also shown (Left).<br />

[1] E. Tsunemi et al, Jpn. J. Appl. Phys. 46, 5636 (2007).<br />

85<br />

Figure 2: AFM (Left) and KFM (Right)<br />

images <strong>of</strong> a poly-PTS single crystal surface<br />

with a line defect.


Scattering Scanning Near-Field Optical <strong>Microscopy</strong><br />

performed by NC-AFM<br />

Fr-1020<br />

Ulrich Zerweck 1 , S.C. Schneider 1 , M.T. Wenzel 1 , H.-G. von Ribbeck 1 , S. Grafström 1 ,<br />

R. Jacob 2 , S. Winnerl 2 , M. Helm 2 , and L.M. Eng 1<br />

1 Institute <strong>of</strong> Applied Photophysics, Technische Universität Dresden, Dresden, Germany,<br />

2 Forschungszentrum Dresden Rossendorf, Dresden, Germany<br />

Many samples <strong>of</strong> interest show resonances in the IR to THz range, which are based<br />

e.g. on phonons or on chemical bonding. At these wavelengths, the classical optical<br />

resolution is strongly limited by the diffraction limit. In scattering scanning near-field<br />

optical microscopy (s-SNOM), an AFM tip operated in the non-contact mode is used both<br />

for the electric field concentration and the field enhancement at the tip apex to increase<br />

the optical resolution. The method is independent <strong>of</strong> the wavelength, and hence allows<br />

for high resolution measurements even at very large wavelengths (IR and THz<br />

frequencies).<br />

We combine s-SNOM with the free-electron laser (FEL) at the Forschungszentrum<br />

Dresden-Rossendorf 2 used as a precisely tunable and monochromatic high-power light<br />

source covering a wavelength range from 4 to 200 µm. With this unique setup, we are<br />

able to perform optical imaging <strong>of</strong> a sample as well as spectroscopy by tuning the<br />

wavelength <strong>of</strong> the FEL [1] (see Fig. 1).<br />

We study different types <strong>of</strong> phonon-resonant samples such as ferroelectric LiNbO3<br />

and BaTiO3, semiconductors with different doping concentrations, and Si-SiO2<br />

nanostructures.<br />

Figure 1: a) Sketch <strong>of</strong> the s-SNOM setup inspecting uniaxial ferroelectric BaTiO3 with<br />

orientation <strong>of</strong> its optical axis perpendicular or parallel to the sample surface on different domains;<br />

b) near-field imaging <strong>of</strong> BaTiO3 with contrast reversal at different optical wavelengths, and c)<br />

distance-dependent near-field spectroscopy on BaTiO3.<br />

[1] S.C. Schneider, PhD thesis, TU Dresden, Germany (2007)<br />

86


<strong>Atomic</strong> resolution dynamic lateral force microscopy in liquid<br />

Shuhei Nishida, Dai Kobayashi, Noriyuki Okabe, and Hideki Kawakatsu<br />

Fr-1120<br />

Institute <strong>of</strong> Industrial Science, The University <strong>of</strong> Tokyo, 4-6-1 Komaba, Meguro-ku, Tokyo 153-8505, Japan<br />

snishida@iis.u-tokyo.ac.jp<br />

A key step to achieve atomic resolution imaging in dynamic lateral force<br />

microscopy (DLFM) is to reduce the lateral tip amplitude down to less than atomic lattice<br />

scale <strong>of</strong> the sample. Use <strong>of</strong> high-frequency cantilever vibration modes is effective for<br />

reducing the tip amplitude. In this contribution, we demonstrate atomic resolution DLFM<br />

imaging in liquid using a high-frequency torsional mode.<br />

We performed the DLFM imaging using an optically based method combining<br />

photothermal excitation and laser Doppler velocimetry for utilizing various cantilever<br />

vibration modes [1,2]. We excited and detected the first torsional mode with the<br />

resonance frequency <strong>of</strong> 1.15 MHz, and reduced the lateral tip amplitude down to 1.3 Å<br />

with keeping vibration stability.<br />

Figure 1(a) shows a DLFM image <strong>of</strong> a muscovite mica surface immersed in purified<br />

water. The small tip amplitude around a quarter <strong>of</strong> the mica’s lattice constant allowed us<br />

to achieve atomic resolution imaging. Cross sectional analysis <strong>of</strong> the image at the<br />

meandering features suggests that the force acting on the tip from a tetrahedral silicate<br />

group was overlapped with the force from the neighboring silicate group along the<br />

vibration direction (Fig. 1(b)).<br />

(a) (b)<br />

Figure 1: DLFM imaging <strong>of</strong> a muscovite mica surface immersed in purified water. (a) A<br />

topographic image, drive frequency: 1,158,225 Hz, scan size: 10 x 10 nm 2 . (b) A line pr<strong>of</strong>ile<br />

along the solid line in (a).<br />

[1] S. Nishida, D. Kobayashi, T. Sakurada, T. Nakazawa, Y. Hoshi, and H. Kawakatsu, Rev. Sci. Instrum.<br />

79, 123703 (2008).<br />

[2] S. Nishida, D. Kobayashi, H. Kawakatsu, and Y. Nishimori, J. Vac. Sci. Technol. B 27, 964 (2009).<br />

87


Fr-1140<br />

Molecular-scale Investigations <strong>of</strong> Biomolecules in Liquids by FM-AFM<br />

Shinichiro Ido 1 , Noriaki Oyabu 1,2 , Kei Kobayashi 2,3 , Yoshiki Hirata 4 , Masaru Tsukada 5 ,<br />

Kazumi Matsushige 1 , and Hir<strong>of</strong>umi Yamada 1,2<br />

1 Department <strong>of</strong> Electronic Science and <strong>Engineering</strong>, Kyoto University, Kyoto, Japan<br />

2 Japan Science and Technology Agency /Adv. Meas. & Analysis, Japan<br />

3 Innovative Collaboration Center, Kyoto University, Kyoto, Japan<br />

4 National Institute <strong>of</strong> Advanced Industrial Science and Technology, Tsukuba, Japan<br />

5 Advanced Institute for Materials Research,Tohoku University ,Sendai, Japan<br />

Recent, rapid progress in high-resolution frequency modulation atomic force<br />

microscopy (FM-AFM) in liquid environments has opened a new way to directly<br />

investigate "in vivo" biological processes with molecular resolution [1,2]. In this study,<br />

submolecular-resolution imaging <strong>of</strong> DNA molecules and proteins has been conducted<br />

toward the molecular scale analysis <strong>of</strong> DNA-protein interactions dominating site-specific<br />

binding <strong>of</strong> proteins to DNA. In addition, three-dimensional (3D) hydration structures on<br />

the biomolecules have been visualized for exploring the roles <strong>of</strong> water molecules in the<br />

biological functions.<br />

A low-thermal-drift FM-AFM developed based on a commercial AFM apparatus<br />

with a home-built low-noise electronics system. Figure 1(a) shows an FM-AFM image <strong>of</strong><br />

a pUC18 plasmid DNA (2686 bp) adsorbed onto a mica surface in a buffer solution<br />

containing 50 mM NiCl2. The major grooves (2.2 nm wide) indicated by the gray arrows<br />

and the minor grooves (1.2 nm wide) indicated by the white arrows along the DNA<br />

chains were clearly resolved. Figure 1(b) shows an FM-AFM image <strong>of</strong> bR trimers on the<br />

cytoplasmic side <strong>of</strong> a purple membrane patch on a mica substrate in a phosphate buffer<br />

solution. A two-dimensional (2D) frequency shift (Δf) mapping image (X-Z) on the<br />

membrane was also shown.<br />

a) b)<br />

Figure 1: (a) FM-AFM image <strong>of</strong> pUC18 plasmid DNA on a mica substrate in 50mM NiCl2<br />

solution. Image size: 48.4 nm × 20.0 nm. Inset: simulated structural model <strong>of</strong> DNA taking the tip<br />

size effect into account. Image size: 10.0 nm × 5.5 nm. (b) FM-AFM image (image size: 29.1 nm<br />

× 29.1 nm) and 2D frequency shift mapping image (image size: 29.1 nm × 2.5 nm) <strong>of</strong> a purple<br />

membrane patch in buffer solution.<br />

[1] T. Fukuma, K. Kobayashi, K. Matsushige, and H. Yamada. Appl. Phys Lett. 87, 034101 (2005).<br />

[2] S. Rode, N. Oyabu, K. Kobayashi, H. Yamada, A. Kuhnle. Langmuir 25, 2850 (2009).<br />

88


Bimodal AFM imaging <strong>of</strong> antibodies and chaperonins in liquids<br />

E.T. Herruzo, C. Dietz, J.R. Lozano and R.Garcia<br />

Instituto de Microelectrónica de Madrid (CSIC), Madrid, Spain<br />

Fr-1200<br />

Bimodal AFM is novel probe microscopy method based on the simultaneous excitation <strong>of</strong><br />

the first two flexural modes <strong>of</strong> the cantilever. The instrument opens up additional<br />

channels (amplitude and phase <strong>of</strong> the 2 nd mode) which can be used for imaging with<br />

enhanced lateral resolution and compositional contrast with respect to amplitude<br />

modulation AFM.<br />

Here, we discuss the performance <strong>of</strong> the Bimodal AFM in water and buffer solutions by<br />

imaging and resolving the structural components <strong>of</strong> isolated biomolecules and packed<br />

protein arrays. Bimodal AFM images <strong>of</strong> antibodies reveal the subunits in both monomer<br />

and pentameric forms by applying very low forces. Unprocessed high resolution images<br />

<strong>of</strong> GroEL reveal the seven-fold symmetry <strong>of</strong> the protein.<br />

We also perform theoretical and numerical simulations to characterize Bimodal AFM<br />

operation. Our model allows us to study the material contrast sensitivity <strong>of</strong> the two<br />

additional channels (amplitude and phase <strong>of</strong> the 2 nd mode) that can be used for imaging.<br />

The theoretical approach also allows us to estimate the forces applied on the sample<br />

during bimodal AFM operation. The calculated forces are small, due to the enhanced<br />

sensitivity <strong>of</strong> 2 nd mode phase to detect changes while the cantilever is far away from the<br />

sample.<br />

Figure 1<br />

a) b)<br />

Figure 2<br />

Figure 1. Bimodal (a) and topography (b) image <strong>of</strong> an isolated IgM antibody in water.<br />

Figure 2. Topography image under Bimodal AFM imaging <strong>of</strong> GroEl chaperonins in buffer.<br />

[1] Rodriguez T R and Garcia R 2004 Appl. Phys. Lett. 84 449–51<br />

[2] Martinez N F, Patil S, Lozano J R and Garcia R 2006, Appl. Phys. Lett. 89 153115–3<br />

[3] Patil S, Martinez N F, Lozano J R and Garcia R 2007, J. Mol. Recognit. 20 516–23<br />

[4] Lozano J R and Garcia R 2008 Phys. Rev. Lett,100 076102–4<br />

[5] N F Martínez et al 2008 Nanotechnology 19 384011<br />

89


Fr-1220<br />

Redox-state Dependent Reversible Change <strong>of</strong> Molecular Ensembles in<br />

Water Solution by Electrochemical FM-AFM<br />

Ken-ichi Umeda 1† , Yasuyuki Yokota 2 , and Ken-ichi Fukui 2,3<br />

1 Department <strong>of</strong> Chemistry, Tokyo Institute <strong>of</strong> Technology, Tokyo, Japan.<br />

2 Department <strong>of</strong> Materials <strong>Engineering</strong> Science, Osaka University, Osaka, Japan.<br />

3 PRESTO, Japan Science and Technology Agency, Saitama, Japan.<br />

Electrochemisty can provide wide range <strong>of</strong> science and technology because <strong>of</strong> easy<br />

control <strong>of</strong> electrochemical potentials <strong>of</strong> interfaces and adsorbed molecules, which initiate<br />

electron transfer and chemical reactions hard to achieve in vacuum. Redox-active<br />

molecules fixed on the electrode can reversibly change their charges depending on the<br />

electrode potential, which makes it possible to introduce a point charge at the<br />

electrode/solution interface. Following the recipe in literature to reduce the deflection<br />

noise density in liquid-phase measurements, we have developed a frequency-modulation<br />

AFM applicable in electrochemical environments with controlling the potential <strong>of</strong> the tip<br />

and the sample. By using the EC-FM-AFM, we have studied how the charging <strong>of</strong> the<br />

molecule fixed at the electrode/solution interface affect the local structure.<br />

Ferrocenylundecanethiol (FcC11H22SH) molecules were embedded in an n-decanethiol<br />

(C10H21SH) self-assembled monolayer (SAM) matrix as molecular islands In Figure 1(a),<br />

the Fc-islands are shown as bright protrusions, whose apparent height differ depending<br />

on their lateral size (number <strong>of</strong> involved molecules). By changing the charge <strong>of</strong> Fcterminal<br />

groups from Fc 0 to Fc +1 , the apparent height and the lateral size <strong>of</strong> each Fcisland<br />

has clearly increased as shown in Figure 1(b). The process was reversible against<br />

the potential change, and originated from the change in the local charge. Simultaneously<br />

obtained energy dissipation signal has decreased at the island when the Fc-islands had<br />

positive charge. These results can be reasonably explained by the formation <strong>of</strong> ion<br />

network from Fc + and ClO4 - counter anions supplied from the electrolyte water solution.<br />

(a) Fc (b) 0<br />

Fc +1<br />

Figure 1: In situ EC-FM-AFM images (Δƒ = +528 Hz) <strong>of</strong> FcC11H22SH islands embedded in<br />

C10H21SH SAM (200×160 nm 2 50 nm 50 nm<br />

) at different electrochemical potentials: (a) substrate potential (Es)<br />

= tip potential (Et) = -0.8 V, (b) Es = Et = -0.4 V vs. Au/AuOx, respectively. The amplitude <strong>of</strong><br />

cantilever vibration was maintained in a feedback loop to be 0.5 nm.<br />

†<br />

Present address: Department <strong>of</strong> Electronic Science and <strong>Engineering</strong>, Kyoto University, Japan.<br />

90


Poster<br />

Session I<br />

Tuesday, 11 August<br />

91


P.I-01<br />

<strong>Force</strong> and Tunneling Current Measurements on the Semiconductor<br />

Surface<br />

Daisuke Sawada, Yoshiaki Sugimoto, Ken-ichi Morita, Masayuki Abe, and Seizo Morita<br />

Graduate <strong>School</strong> <strong>of</strong> <strong>Engineering</strong>, Osaka University, 2-1 Yamada-Oka, Suita, Osaka, 565-0871, Japan.<br />

Non-contact atomic force microscopy (NC-AFM) and scanning tunneling microscopy<br />

(STM) enable us to obtain atomically-resolved information, for example, chemical<br />

bonding force or local density <strong>of</strong> state (LDOS). Recently, at the near contact region, the<br />

drop <strong>of</strong> the tunneling current was reported using conventional STM [1]. The drop <strong>of</strong> the<br />

tunneling current was explained by the chemical bonding formation, however, such<br />

correlation has not been measured experimentally. We measured the force and the<br />

tunneling current simultaneously using metal coated Si cantilevers. Here, we report our<br />

results on both the Si(111)-(7×7) surface and the Ge(111)-c(2×8) surface at room<br />

temperature using the feedforward technique in order to compensate the thermal drift [2].<br />

In constant height simultaneous measurement, atomic image contrast between the<br />

AFM image and the STM image are different [Fig.1 a), b)]. We obtained FSR-z and It-z<br />

curves on the adatom <strong>of</strong> the Ge(111)-c(2×8) surface by measuring Δf-z and -z curves<br />

[Fig.1 c)]. The drop <strong>of</strong> It was oabserved at the onset <strong>of</strong> FSR with decreasing the tip-surface<br />

distance. The drops were observed on the Ge(111)-c(2×8) surface as well as the Si(111)-<br />

(7×7) surface [3].<br />

A part <strong>of</strong> this work was supported by Grant-in-Aid for Scientific Research on<br />

Priority Areas “Nano Materials Science for <strong>Atomic</strong> Scale Modification” and Global COE<br />

Program, “Center for Electronic Devices Innovation”.<br />

Figure 1: Δf image a) and image b) were obtained simultaneously in the constant height<br />

mode. The sample bias (Vs) is +500 mV. c) FSR-z and It-z curves measured on the Ge adatom<br />

at Vs= +500 mV.<br />

[1] P. Jelínek, et al., Phys. Rev. Lett. 101, 176101 (2008).<br />

[2] M. Abe, et al., Appl. Phys. Lett. 90, 203103 (2007).<br />

[3] D. Sawada, et al., Appl. Phys. Lett. 94, 173117 (2009).<br />

92


P.I-02<br />

<strong>Force</strong> Map <strong>of</strong> <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong> on Si(111)-(5x5)-DAS Surface<br />

Akira Masago and Masaru Tsukada<br />

WPI-AIMR, Tohoku University, 2-1-1 Katahira, Aoba, Sendai, 980-8577 Japan<br />

Recently, lateral force mapping has been performed by atomic force microscopy (AFM)<br />

on the Si(111)-(7x7)-(Dimer-Adatom-Stacking fault (DAS)) surface, as well as potential<br />

and vertical force mapping. We believe that they provide us with useful information for<br />

atomic identification and manipulation. In this study, we simulated force maps along the<br />

long diagonal direction <strong>of</strong> the unit cell <strong>of</strong> the Si(111)-(5x5)-DAS surface using the<br />

density-functional based tight-binding (DFTB) calculation. As a result, we obtained force<br />

maps shown in Fig. 1, using a Si4H9 cluster tip. In the vertical force map, the small peaks<br />

located at the rest atoms (Rs) as well as the large peaks at the adatoms (Ad). Moreover,<br />

the plateaus also exist at bridge site (Br) <strong>of</strong> the adatoms out <strong>of</strong> the scan trajectory. In the<br />

lateral force maps, we can see paired peaks located at the adatoms and rest atoms. These<br />

characters must be related with the atomic configuration <strong>of</strong> the tip apex. In our<br />

presentation, we will talk about the relation between remarkable characters that present<br />

on the force maps and various configurations <strong>of</strong> the tip apex.<br />

Figure 1: Vertical and lateral force maps when the tip scans along a long diagonal line <strong>of</strong> a unit<br />

cell <strong>of</strong> the Si(001)-(5x5)-DAS surface as well as side and top views <strong>of</strong> the surface. All circles<br />

denote Si atoms, where open and fill circles denote adatoms and rest atoms, respectively. In<br />

particular, open circles written in bold face are adatoms on the scan trajectory.<br />

93


From non-contact to atomic scale contact between a Si tip and a Si<br />

surface analyzed using an nc-AFM and nc-AFS based instrument<br />

P.I-03<br />

Toyoko Arai 1 , Kosei Kiyohara 1 , Taiki Sato 1 , Shugaku Kushida 2 and Masahiko Tomitori 2<br />

1 Graduate <strong>School</strong> <strong>of</strong> Natural science & Technology, Kanazawa University, Kanazawa, Ishikawa, Japan<br />

2 <strong>School</strong> <strong>of</strong> Materials Science, Japan Advanced Institute <strong>of</strong> Science and Technology, Nomi, Ishikawa Japan<br />

The scanning probe microscopy (SPM) is a powerful tool to observe a sample surface<br />

with atomic resolution as well as spectroscopic measurements and atom/molecule<br />

manipulation. Since the advent <strong>of</strong> SPM, we are able to bring an atomically sharpened tip<br />

very close to a sample in a well-controlled manner, in particular, by non-contact atomic<br />

force microscopy (nc-AFM); the force interaction detected by nc-AFM changes so<br />

drastically at atomically close separations that the precise control <strong>of</strong> separation is<br />

performed. The quantum mechanical properties become prominent at those distances<br />

between a tip and a sample; chemical covalent bonding is formed at less than 1 nm, the<br />

tunneling barrier between them collapses, and at closer distances, an atomically necking<br />

can be shaped and quantized conductance comes up through atomically confined<br />

channels or intermediate electronic states in between. The quantized phenomena at the<br />

point contact have attracted much interest, including its formation process from noncontact<br />

through pseudo-contact to contact. Not only the force interaction, but also the<br />

electric conductance properties can be evaluated using an instrument based on nc-AFM<br />

combined with spectroscopic methods at those separations, i.e., bias-voltage non-contact<br />

atomic force spectroscopy (nc-AFS) with the ability to measure electric current with<br />

respect to bias voltage between a tip and a sample [1]. From a viewpoint <strong>of</strong> application,<br />

contact formation between two pieces <strong>of</strong> condensed matter is crucial to fabricate novel<br />

nanoscale devices. Here we focus on Si-Si contact and non-contact states analyzed by an<br />

nc-AFM and nc-AFS based instrument. Although silicon-silicon contacts sounds very<br />

important in industries and various types <strong>of</strong> Si nanowires have been synthesized, there are<br />

few reports on their nanoscale electromechanical properties.<br />

Experiments were conducted using a home-made UHV-AFM/AFS instrument with a<br />

B-doped Si piezoresistive cantilever having a [001]-oriented Si for samples <strong>of</strong> n- and ptype<br />

Si(111). After cleaning the tip and the sample by heating in UHV, we took nc-AFM<br />

images simultaneously with current, averaged over a cantilever oscillation cycle, and<br />

damping with changing bias voltage, while taking spectroscopic curves <strong>of</strong> frequency shift<br />

current and damping versus bias voltage or tip-sample separation. By slowly approaching<br />

and retracting with compensation <strong>of</strong> z-direction thermal drift, the current-separation<br />

curves with a good S/N ratio exhibited features from tunneling regime to saturation<br />

toward tunneling barrier collapse, leading to chemical bond formation. The damping also<br />

exhibited curious features possibly due to charge change around the tip and the sample.<br />

For Si point contacts current-voltage curves showed p-p or p-n junction characteristics<br />

with staircase behaviors. The details and discussion will be presented.<br />

1. [1] T. Arai and M. Tomitori: Phys.<br />

Rev. B 73 (2006) 073307.<br />

94


P.I-04<br />

Improved atomic-scale contrast via bimodal dynamic force microscopy<br />

S. Kawai, Th. Glatzel, S. Koch, B. Such, A. Barat<strong>of</strong>f, and E. Meyer<br />

Department <strong>of</strong> Physics, University <strong>of</strong> Basel, Klingelbergstr. 82, 4056 Basel Switzerland<br />

We extend multi-frequency AFM to atomically resolved frequency-modulation<br />

dynamic force microscopy and demonstrate a further improvement <strong>of</strong> spatial resolution in<br />

ultra-high vacuum [1]. The first and second flexural resonance modes <strong>of</strong> a commercially<br />

available Si cantilever are simultaneously self-excited with given amplitudes, and the<br />

resonance frequency shifts (Δf1st and Δf2nd) are demodulated by two phase-locked loop<br />

circuits (Nanonis: Dual-OC4). The combination <strong>of</strong> sub-angstrom amplitude oscillation<br />

A2nd at the second resonance with the commonly used large amplitude oscillation A1st at<br />

the first resonance enables higher resolution imaging at closer tip-sample distances while<br />

avoiding atomic jump-to-contact instabilities, which are more likely at such distances<br />

during small amplitude operation with a single mode [2-4].<br />

Figure 1 shows a series <strong>of</strong> simultaneously recorded Δf1st and Δf2nd maps <strong>of</strong> KBr(001),<br />

obtained with A1st=16 nm and A2nd=50 pm at decreasing tip-sample distances in the<br />

attractive region, recorded in the quasi-constant height mode. Being more sensitive to the<br />

short-range interaction, Δf2nd exhibits a stronger distance dependence and contrast than<br />

Δf1st at short distances. Although A2nd was much smaller then the width <strong>of</strong> the measured<br />

force minimum, the signal-to-noise in the Δf2nd map was higher than in the Δf1st map at<br />

close tip-sample distances [(e) and (f)]. The enhanced sensitivity to the short-range<br />

interaction could even reveal tip and/or sample deformations induced by the interaction<br />

force.<br />

Figure 1: A series <strong>of</strong> simultaneously recorded Δf1st and Δf2nd maps at decreasing tip-sample<br />

distances (a – f). The first and second resonance frequencies are 154021 Hz and 960874 Hz, and<br />

the Q factors and amplitudes are 31059 and 6246, and A1st=16 nm and A2nd=50 pm, respectively.<br />

[1] S. Kawai et al., submitted.<br />

[2] F. J. Giessibl et al., Science 289, 422 (2000).<br />

[3] S. Kawai et al., Appl. Phys. Lett. 86, 193107 (2005).<br />

[4] S. Kawai and H. Kawakatsu, Appl. Phys. Lett. 88, 133103 (2006).<br />

95


Static cantilever deflection in dynamic force microscopy<br />

S. Kawai, Th. Glatzel, S. Koch, B. Such, A. Barat<strong>of</strong>f, and E. Meyer<br />

Department <strong>of</strong> Physics, University <strong>of</strong> Basel, Klingelbergstr. 82, 4056 Basel Switzerland<br />

P.I-05<br />

So far, the influence <strong>of</strong> the time-averaged cantilever deflection in dynamic force<br />

microscopy (DFM) and spectroscopy [1] has been assumed negligibly small or constant,<br />

or else regarded as the source <strong>of</strong> jump-to-contact instabilities. But in fact, for a given tipsample<br />

distance <strong>of</strong> closest approach, the static deflection increases with decreasing<br />

oscillation amplitude. As a consequence, small-amplitude DFM is able to simultaneously<br />

detect the influence <strong>of</strong> atomic-scale interaction forces on time-averaged, as well as<br />

dynamic measured quantities. The use <strong>of</strong> higher resonances [2,3] is well-suited for this<br />

comparison because a typical static stiffness kc ~ 30 N/m can cause a measurable timeaveraged<br />

deflection. Nevertheless kc is still high enough to avoid cantilever jump-tocontact<br />

instabilities when a surface with a low reactivity is scanned with a sharp Si tip [4]<br />

We theoretically and experimentally studied the effects <strong>of</strong> the time-averaged<br />

deflection in DFM. Figure 1(a) and 1(b) shows a series <strong>of</strong> frequency shift and static<br />

deflection vs. distance curves measured using different amplitudes A2nd <strong>of</strong> the second<br />

flexural mode above a maximum <strong>of</strong> the topographic image <strong>of</strong> KBr(001). Figure 1(c)<br />

shows force vs. distance curves with and without compensation <strong>of</strong> the static deflection. It<br />

is found that the static deflection affects not only the Z-distance scale,<br />

but also the<br />

converted force. <strong>Force</strong> curves extracted from Δf2nd(Z) measured with different A2nd<br />

coincide if shifted along Z and exhibit a wiggle which is likely due to a sideways<br />

displacement <strong>of</strong> the tip apex towards a nearby counterion on the sample surface [1].<br />

Figure 1: A series <strong>of</strong> (a) frequency shift and (b) static deflection vs. distance curves. (c) <strong>Force</strong> vs.<br />

distance curves with (Zc) and without (Z’c) compensation <strong>of</strong> the static deflection.<br />

[1] A. Schirmeisen et al., Phys. Rev. Lett. 97, 136001 (2007).<br />

[2] S. Kawai et al., Appl. Phys. Lett. 86, 193107 (2005).<br />

[3] S. Kawai and H. Kawakatsu, Appl. Phys. Lett. 88, 133103 (2006)<br />

[4] S. Kawai and H. Kawakatsu, Phys. Rev. B 79, 115440 (2009).<br />

96


P.I-06<br />

Resonance frequency shift due to tip-sample interaction in the thermal<br />

oscillations regime<br />

Giovanna Malegori, Gabriele Ferrini<br />

Dipartimento di Matematica e Fisica, Università Cattolica, Brescia, Italy<br />

We present results on the interaction <strong>of</strong> a cantilever in the thermal oscillations regime<br />

with the force gradients near various kind <strong>of</strong> surfaces. As is well known, the power<br />

density spectrum <strong>of</strong> a cantilever freely oscillating in the absence <strong>of</strong> external modulation<br />

shows the thermal amplitude and resonance frequency <strong>of</strong> its flexural modes. As the tip<br />

approaches a surface, a frequency shift in the amplitude and resonance frequency <strong>of</strong> the<br />

thermally excited flexural modes is measured as a function <strong>of</strong> the tip-sample separation,<br />

until jump-to-contact sets in. After jump-to-contact, the power density spectrum <strong>of</strong> the<br />

cantilever fluctuations changes due to the fact it has a supported end. A comparison with<br />

available models permits to identify the flexural modes in the two situations (we detect<br />

four flexural modes for the free end cantilever and two for the supported end cantilever)<br />

and to extract information about the interaction force gradients. The mean vibration<br />

amplitude in the various flexural modes is in the range <strong>of</strong> 10-100 pm at room<br />

temperature, according to the stiffness <strong>of</strong> the cantilever. Finally, we describe the<br />

acquisition techniques and background requirements for the instrumentation.<br />

97


Influence <strong>of</strong> thermal noise on measurements <strong>of</strong><br />

chemical bonds in UHV-AFM<br />

Peter M. H<strong>of</strong>fmann<br />

Department <strong>of</strong> Physics and Astronomy, Wayne State University, Detroit, MI 48201<br />

P.I-07<br />

Ultra-high vacuum based AFM has been repeatedly shown to provide high resolution<br />

measurements <strong>of</strong> forces and force gradients associated with chemical bonds 1-5 . However,<br />

the bond parameters, such as bond length and energies, have varied between<br />

measurements. One theory put forward for these discrepancies has been the influence <strong>of</strong><br />

thermal noise on measurements performed at room temperature. In FM-AFM it was<br />

originally found that bond force curves seem broadened when measured at room<br />

temperature 1 , and theoretically expected bond lengths were only recovered when<br />

measurements were obtained at low temperature 2 . On the other hand, AM-AFM<br />

performed at small amplitudes found no such broadening at room temperature 3 . More<br />

recent measurements using FM AFM at room temperature, however, also seem to show<br />

shorter interaction lengths 4,5 . What, if any, is the influence <strong>of</strong> cantilever thermal noise on<br />

measured interaction lengths? We present a simulation study <strong>of</strong> the influence <strong>of</strong> thermal<br />

noise on measurements <strong>of</strong> interaction forces using FM or AM AFM.<br />

[1] M. Guggisberg, et al., Phys. Rev. B 61, 11151 (2000).<br />

[2] M. A. Lantz, et al., Science 291, 2580 (2001).<br />

[3] P.M. H<strong>of</strong>fmann, et al.. Proc. Royal Soc. Lond. 457, 1161-1174 (2001).<br />

[4] Y. Sugimoto et al., Phys. Rev. B 77, 195424 (2008).<br />

[5] K. Ruschmeier et al., Phys. Rev. Lett. 101, 156102 (2008)<br />

98


P.I-08<br />

<strong>Atomic</strong> force microscope cantilever resonance frequency shift based<br />

thermal metrology<br />

Arvind Narayanaswamy 1 , Carlo Canetta 1 , and Ning Gu 2<br />

1 Department <strong>of</strong> Mechanical <strong>Engineering</strong>, Columbia University, New York, USA.<br />

2 Department <strong>of</strong> Electrical <strong>Engineering</strong>, Columbia University, New York, USA.<br />

The atomic force microscope (AFM) is now an indispensable tool not only in obtaining<br />

topographical images <strong>of</strong> the surfaces <strong>of</strong> physical objects but also measuring the forces,<br />

such as van der Waals and Casimir, between two objects, one <strong>of</strong> which is the tip <strong>of</strong> the<br />

AFM cantilever or a microsphere attached to the cantilever, and the other is a substrate<br />

mounted on a translation stage. We have used a bi-material AFM cantilever to measure<br />

the radiative transfer, instead <strong>of</strong> force, between a microsphere attached to a cantilever and<br />

a substrate, thereby extending the capabilities <strong>of</strong> a conventional AFM with minor<br />

modifications. Unlike force measurement, which can use the static deflection <strong>of</strong> the<br />

cantilever or the frequency shift <strong>of</strong> a resonance mode, the heat transfer can be measured<br />

only through the static deflection <strong>of</strong> the cantilever. However, with cantilevers to which<br />

microspheres are attached to the tip, the flexural as well as torsional modes <strong>of</strong> the<br />

cantilever exhibit a frequency shift due to the temperature changes <strong>of</strong> the cantilever that<br />

is independent <strong>of</strong> the variation <strong>of</strong> material properties with temperature. The frequency<br />

shift is due to the change in curvature <strong>of</strong> the cantilever, in this case due to temperature<br />

changes, and is more pronounced for higher order modes compared to the fundamental<br />

flexural resonance mode. Though not as sensitive a technique as measuring temperature<br />

changes via static deflection, the resonance frequency based technique is more robust and<br />

can be calibrated to measure the absolute temperature as opposed to temperature shifts<br />

alone. While the frequency shift <strong>of</strong> the first flexural mode is small enough not to affect<br />

current non-contact force measurement techniques, it could affect force detection based<br />

on higher order modes.<br />

99


P.I-09<br />

Contact potential difference on the atomic-scale probed by<br />

Kelvin Probe <strong>Force</strong> <strong>Microscopy</strong>: an imaging scenario<br />

Laurent Nony 1 , Adam Foster 2 , Franck Bocquet 1 , and Christian Loppacher 1<br />

1<br />

Aix-Marseille Université, IM2NP, Av. Escadrille Normandie-Niemen, F-13397 Marseille and CNRS, IM2NP (UMR<br />

6242), Marseille-Toulon, France<br />

2<br />

Department <strong>of</strong> Physics, Tampere University <strong>of</strong> Technology,<br />

P.O. Box 692 FIN-33101 Tampere, Finland<br />

A fully numerical analysis <strong>of</strong> the origin <strong>of</strong> the atomic-scale contrast in Kelvin probe force<br />

microscopy (KPFM) is presented. The numerical implementation mimics recent experimental<br />

results on the (001) surface <strong>of</strong> a bulk alkali halide single crystal for which a simultaneous<br />

topographical and Kelvin, so-called Contact Potential Difference (CPD), atomic-scale contrast<br />

had been reported [1]. In this work, we have combined atomistic simulations <strong>of</strong> the tip-sample<br />

force field with our non-contact AFM/KPFM simulator [2] to compute topographical and CPD<br />

images. The force field notably includes the required short-range bias voltage dependence to<br />

account for the CPD atomic-scale contrast. The sample is a NaCl(001) single crystal. The<br />

atomistic simulations have been performed by means <strong>of</strong> the code SCIFI [3]. The simulator has<br />

been used in the Frequency-Modulation KPFM operating mode. The tip consists <strong>of</strong> a spherical<br />

metallic cap within which a 4x4x4 NaCl cluster with the [111] direction pointing towards the<br />

surface is partly embedded. The sample consists <strong>of</strong> a 10x10x4 NaCl slab embedded within a<br />

semi-infinite continuous medium merely described by its dielectric constant. The sample, with a<br />

macroscopic height (several millimeters), lies on a metallic counter-electrode that is biased with<br />

respect to the tip. Beyond the short-range chemical and electrostatic forces computed with SCIFI,<br />

we have included a usual Van der Waals long-range term and a less usual, although fundamental,<br />

long-range electrostatic interaction. This term mimics the weak capacitive interaction between tip<br />

and sample holders and is necessary to describe the behavior <strong>of</strong> the CPD with the distance.<br />

Figs.1a and b show that the simulator is able to account for the simultaneous topographical and<br />

CPD atomic-scale contrast, respectively. The occurrence <strong>of</strong> such a contrast is intricately<br />

connected to the dynamic polarization <strong>of</strong> the ions <strong>of</strong> the slab triggered by the modulation <strong>of</strong> the<br />

bias voltage, unless otherwise no Kelvin contrast is measured. The nc-AFM/KPFM simulator is a<br />

tool which helps in interpreting the experimental data quantitatively.<br />

References :<br />

[1]- F. Bocquet et al., Phys. Rev. B<br />

78, 035410 (2008)<br />

[2]- L. Kantorovich et al., Surf. Sci.<br />

445, 283 (2000)<br />

[3]- L. Nony et al., Phys. Rev. B 74,<br />

235439 (2006); L. Nony et al.,<br />

accepted in Nanotechnology<br />

Fig.1a- Topography channel<br />

measured at 0.45nm to the<br />

surface: the vertical contrast is<br />

40pm<br />

100<br />

Fig.1b- simultaneously<br />

acquired CPD channel: The<br />

vertical contrast is 0.6V


Self-assembled Boronitride Nanomesh on Rh(111)<br />

Investigated by Means <strong>of</strong> Kelvin Probe <strong>Force</strong> <strong>Microscopy</strong><br />

P.I-10<br />

Sascha Koch 1 , Markus Langer 1 , Jorge Lobo-Checa 1,3 , Thomas Brugger 2 , Shigeki<br />

Kawai 1 , Bartosz Such 1 , Ernst Meyer 1 and Thilo Glatzel 1<br />

1 Department <strong>of</strong> Physics, University <strong>of</strong> Basel, 4056 Basel, Switzerland<br />

2 Department <strong>of</strong> Physics, University <strong>of</strong> Zurich, 8006 Zurich, Switzerland<br />

3 Centre d'Investigaciò en Nanociència i Nanotecnologia (CIN2), Campus Universitat Autònoma de<br />

Barcelona, Spain<br />

By high temperature exposure <strong>of</strong> borazine [(HBNH)3], a self-assembled hexagonal<br />

nanomesh with a periodicity <strong>of</strong> about 3nm and a hole size <strong>of</strong> 2nm is formed on Rh(111)<br />

under UHV conditions [1]. The stable and periodic structure is perfectly suitable for<br />

trapping single molecules at room temperature [2]. The determination <strong>of</strong> local work<br />

function variations is a major requirement to understand the process <strong>of</strong> molecular<br />

adsorption. NC-AFM combined with Kelvin probe force microscopy (KPFM) enables<br />

imaging <strong>of</strong> the surface and simultaneous detection <strong>of</strong> the work function map with subnanometer<br />

resolution [3].<br />

Figure 1(a) shows the topography signal <strong>of</strong> the nanomesh which is in good agreement<br />

with previous STM results [1,2,4]. The Rh(111) surface is found to be completely<br />

covered with the mesh. Figure 1(b) shows the simultaneously obtained LCPD signal <strong>of</strong><br />

this superstructure clearly resolving a work function difference <strong>of</strong> the structure. Here a<br />

variation <strong>of</strong> the LCPD <strong>of</strong> approximately 600mV was detected.<br />

Figure 1: (a) AFM image <strong>of</strong> the bn-nanomesh (20x20nm 2 ). Parameters:<br />

Δf=-35Hz, A=4nm, f0=153kHz; Δz=650pm. (b) Simultaneously measured<br />

LCPD image. ΔLCPD=600mV, VAC=500mV, f1≈fAC=954kHz.<br />

[1] M.Corso, W. Auwärter, M. Muntwiler, A. Tamai, T. Greber, J. Osterwalder, Science 303, 217 (2004)<br />

[2] H. Dil, J. Lobo-Checa, R. Laskowski, et al., Science 319, 1824 (2008)<br />

[3] Th. Glatzel, L. Zimmerli, S. Koch, S. Kawai, E. Meyer; Appl. Phys. Lett., 94, 3 (2009)<br />

[4] S. Berner, M. Corso, R. Widmer et al., Ang. Chem. Int. Ed. 46, 5115 (2007)<br />

101


Kelvin probe force microscopy in application to organic thin films:<br />

P.I-11<br />

frequency modulation, amplitude modulation, and hover mode KPFM<br />

Brad Moores 1 , Francis Hane 2 , Lukas Eng 3 , and Zoya Leonenko 1,2<br />

1 Department <strong>of</strong> Physics and Astronomy, University <strong>of</strong> Waterloo, Waterloo, Canada<br />

2 Department <strong>of</strong> Biology, University <strong>of</strong> Waterloo, Waterloo, Canada<br />

3 Institute <strong>of</strong> Applied Photophysics, Technical University <strong>of</strong> Dresden, Dresden, Germany<br />

We applied Kelvin probe force microscopy (KPFM) to visualize the lateral surface<br />

potential distribution in thiol self-assembled monolayers and lipid-protein films. We have<br />

shown earlier (Leonenko, et al. Biophys. J. 2007) that function <strong>of</strong> Bovine Lipid Extract<br />

Surfactant (BLES) is related to the specific molecular architecture <strong>of</strong> surfactant films.<br />

Defined molecular arrangement <strong>of</strong> the lipids and proteins <strong>of</strong> the surfactant film give rise<br />

to a local highly variable electrical surface potential <strong>of</strong> the interface. In this work the<br />

resolution <strong>of</strong> frequency modulation (FM-KPFM), amplitude modulation (AM-KPFM)<br />

and hover (HM-KPFM) modes were compared. At larger scale and larger surface<br />

potential deviations all modes give high resolution images. At smaller scans and smaller<br />

differences in surface potential FM-KPFM mode gives superior resolution, and therefore<br />

is preferable for imaging lipid- and lipid-protein films. The response and sensitivity <strong>of</strong><br />

KPFM modes was addressed with the help <strong>of</strong> force measurements.<br />

102


P.I-12<br />

Resolution enhanced multifrequency electrostatic force microscopy<br />

under ambient conditions<br />

X. D. Ding 1 , J. B. Xu 2 , and J. X. Zhang 1<br />

1 State Key Laboratory <strong>of</strong> Optoelectronic Materials and Technologies, and <strong>School</strong> <strong>of</strong> Physics Science &<br />

<strong>Engineering</strong>, Sun Yat-sen University, Guangzhou 510275, China<br />

2 Department <strong>of</strong> Electronic <strong>Engineering</strong>, and Materials Science and Technology Research Center, The<br />

Chinese University <strong>of</strong> Hong Kong, Shatin, New Territories, Hong Kong, China.<br />

Electrostatic force microscopy (EFM) and Kelvin probe force microscopy (KPFM)<br />

are widely used to study the electrical and electrochemical characteristics <strong>of</strong> a variety <strong>of</strong><br />

material. The lateral resolution for EFM ia better than 10~20nm under vacuum conditions<br />

In contrast, the lateral resolution reported in literature is only 100 nm or so under ambient<br />

conditions. Though suffering from strong mechanical non-linearity <strong>of</strong> repulsive tip–<br />

sample contact, multifrequency method has been introduced to EFM recently[1].<br />

Here, we report an extension <strong>of</strong> multifrequency AFM with enhanced resolution to<br />

measure electrostatic force under ambient conditions[2]. The first eigenmode <strong>of</strong> a<br />

cantilever is used for topography imaging, while the third eigenmode is resonantly<br />

excitated with a sinusoidal modulation voltage applied on tip to measure electrostatic<br />

force in lift mode. Figure 1 shows the images for a thermally evaporated aluminum (Al)<br />

film on Si(111). Due to the surface potential variation <strong>of</strong> the sample, the EFM image in<br />

figure 1(b) shows a well reproduced contrast different from its corresponding topographic<br />

image in figure 1(a). The lateral resolution is better than 15nm determined from the line<br />

pr<strong>of</strong>ile in figure 1(c). The enhancement <strong>of</strong> resolution is ascribed to the suppress <strong>of</strong> the<br />

cantilever-sample interactions and the increase <strong>of</strong> the tip-sample interactions due to the<br />

use <strong>of</strong> the third eigenmode <strong>of</strong> the probe.<br />

80nm<br />

30 50 70<br />

Position, nm<br />

(a) (b) (c)<br />

Figure 1: Experimental results <strong>of</strong> the multifrequency EFM for Al film on Si(111) under ambient<br />

conditions. (a) Topography image. (b) EFM Image, and (c) The line pr<strong>of</strong>ile along the arrow in (b)<br />

[1] R. W. Stark, N. Naujoks, and A. Stemmer, Nanotechnology 18, 065502 (2007).<br />

[2] X. D. Ding, C. Li, R. Y. Zeng, J. An, and J. B. Xu, Appl. Phys. Lett. (To be published)<br />

(X. D. Ding, Category 7-Kelvin probe microscopy is fitted and an Oral presentation is preferred)<br />

103<br />

Electrostatic <strong>Force</strong>, a.u.<br />

1.5<br />

1.0<br />

0.5<br />

10-15nm


P.I-13<br />

Deconvolution and Tip Geometry Effects in <strong>Atomic</strong>- and Nanoscale Kelvin probe<br />

<strong>Force</strong> <strong>Microscopy</strong><br />

George Elias 1 , Yossi Rosenwaks 1 , Amir Boag 1 , Ernst Meyer 2 , and Thilo Glatzel 2<br />

1 Dept. <strong>of</strong> Physical Electronics, Tel-Aviv University, Tel Aviv 69978, Israel<br />

2 Dept. <strong>of</strong> Physics, University <strong>of</strong> Basel, Klingelbergstr. 82, 4056 Basel, Switzerland<br />

In Kelvin probe force microscopy (KPFM) the long range electrostatic forces between the<br />

tip and the surface prevents quantitative measurement <strong>of</strong> nanostructures; however this<br />

can become feasible by developing appropriate deconvolution algorithms that restore the<br />

actual sample work function. We present such novel algorithms and methods that enable<br />

us to restore measurements conducted at tip-sample distances below 1 nm, while taking<br />

into account also the measured topography. Fitting the restored images with KPFM<br />

measurements <strong>of</strong> samples with well defined work function values has allowed us to<br />

extract the measuring tip geometry and validate our deconvolution methods.<br />

The three dimensional potential <strong>of</strong> the tip-sample system is calculated using an<br />

integral equation based boundary element method, combined with modeling the sample<br />

by an equivalent dipole-layer and image-charge model. The tip is modeled using MSC<br />

Patran finite element mesh, to create a high density mesh on the tip apex and a lower<br />

density mesh far from the sample as shown in Figure 1 (left). The middle figure shows a<br />

measured UHV-KPFM image <strong>of</strong> NaCl thin film grown on Cu (111) and a comparison<br />

with convolutions performed with 3 different tip apex radii (right). The excellent<br />

agreement obtained for a tip apex <strong>of</strong> 10nm allows to estimate the tip apex geometry, and<br />

helps to validate our convolution algorithms. Implications for real tip shapes (nanotips)<br />

and atomic resolution imaging is demonstrated and discussed<br />

Figure 1: (left) Full tip geometry mesh and tip apex mesh ; (middle) a KPFM measurement <strong>of</strong><br />

NaCl thin films grown on Cu(111); (right) a comparison between the measured KPFM line in the<br />

middle image and convolutions performed for 3 tip apex radii (half opening angle <strong>of</strong> 17.5º) <strong>of</strong> 3,<br />

6, and 10 nm.<br />

104


P.I-14<br />

Kelvin <strong>Force</strong> <strong>Microscopy</strong> Dynamic Behavior and Noise Propagation<br />

Heinrich Diesinger, Dominique Deresmes, Jean-Philippe Nys, and Thierry Mélin<br />

Institut d’Electronique, Microelectronique et Nanotechnologie, CNRS UMR 8520, Department<br />

ISEN,Avenue Henri Poincaré, 59652 Villeneuve d’Ascq, France<br />

The bandwidth <strong>of</strong> scanning probe control loops limits the sampling rate in data<br />

acquisition. In this work, the dynamic behavior <strong>of</strong> amplitude detection (AM) and<br />

frequency detection (FM) Kelvin force microscopy (KFM) setups is analyzed and<br />

optimized. Since enhanced bandwidth alone would increase speed at the cost <strong>of</strong> tolerating<br />

more noise, the origin <strong>of</strong> noise and its propagation within the control loops are studied in<br />

parallel.<br />

Laser Diode Photodetector<br />

F N = 4γk B T<br />

CPD<br />

V PD<br />

V PD, N = 10 μV/sqrt(Hz)<br />

In<br />

Lock-in<br />

Osc<br />

V<br />

bias, AC<br />

fres2 X<br />

PI Amplifier<br />

Kelvin Controller<br />

+<br />

Σ =<br />

+<br />

V Kelvin<br />

open closed<br />

Figure 1: Setup <strong>of</strong> the Kelvin control loop <strong>of</strong> an AM-KFM, consisting <strong>of</strong> a lock-in amplifier<br />

exciting the probe electrostatically at its second resonance, and a PI error amplifier to close the<br />

feedback loop and to apply the Kelvin voltage to the probe. The loop can be opened. Main noise<br />

sources are the thermal probe excitation force and white noise at the output <strong>of</strong> the photodetector.<br />

Fig. 1 shows the Kelvin control loop <strong>of</strong> an AM-KFM in ultrahigh vacuum, using the<br />

second cantilever resonance for KFM while distance control is based on the first, similar<br />

to the setup demonstrated by Kikukawa et al.[1]. The noise power spectral density <strong>of</strong> the<br />

Kelvin output signal can be modeled after the transfer functions <strong>of</strong> the different stages<br />

have been measured in open loop configuration. A benchmark criterium for comparing<br />

hardware with respect to noise is issued. An FM KFM close to the configuration<br />

suggested by Kitamura [2] is also analyzed. Advantages and drawbacks <strong>of</strong> both methods<br />

in terms <strong>of</strong> bandwidth and signal to noise are discussed.<br />

[1] A. Kikukawa, S. Hosaka, and R. Imura. Appl. Phys. Lett. 66, 3510 (1995).<br />

[2] S. Kitamura and M. Iwatsuki. Appl. Phys. Lett. 72, 3154 (1998)<br />

105


Charge transfer from doped silicon nanocrystals<br />

P.I-15<br />

Łukasz BOROWIK 1 , Thuat NGUYEN-TRAN 2 , Pere ROCA i CABARROCAS 2 , Koku<br />

KUSIAKU 1 , Didier THERON 1 , Heinrich DIESINGER 1 , Dominique DEREMES 1 , and<br />

Thierry MELIN 1<br />

1<br />

Institut d’Electronique, de Microélectronique et de Nanotechnologie, CNRS-UMR 8520, Av . Poincaré,<br />

BP 60069, 59652 Villeneuve d’Ascq, France<br />

2<br />

Laboratoire de Physique des Interfaces et des Couches Minces, CNRS-UMR 7647, Ecole Polytechnique<br />

91128 Palaiseau, France<br />

We present ultra-high vacuum (UHV) atomic force experiments performed on doped<br />

silicon nanocrystals fabricated by plasma enhanced chemical vapour deposition. The aim<br />

<strong>of</strong> this work is to study the doping properties and charge transfers from doped<br />

nanocrystals using UHV amplitude modulation Kelvin force microscopy (AM-KFM).<br />

The doping properties <strong>of</strong> hydrogen passivated nanocrystals are studied by monitoring the<br />

nanocrystal surface potential VS as a function <strong>of</strong> the nanocrystal and substrate doping,<br />

and also as a function <strong>of</strong> the nanocrystal size. The case <strong>of</strong> intrinsic nanocrystals is<br />

illustrated in Figure 1, in which the surface potential VS <strong>of</strong> the nanocrystals is plotted as a<br />

function <strong>of</strong> their height, showing - in average - positive or negative charge transfer from<br />

the p-doped and n-doped substrate, together with strong potential fluctuations attributed<br />

to the variation <strong>of</strong> the nanocrystal surface states. This situation is then compared to the<br />

case <strong>of</strong> n-doped nanocrystals, for which much lower potential fluctuations can be<br />

observed, and for which the nanocrystal surface potential VS is found almost independent<br />

<strong>of</strong> the doping level. These two effects are understood as stemming from the doping <strong>of</strong> the<br />

nanocrystal, which provides the necessary charge to compensate for the nanocrystal<br />

surface states, and induce a charge transfer to the substrate. The interpretation <strong>of</strong> the<br />

charge transfer equilibrium will be detailed quantitatively[1], using numerical simulations<br />

<strong>of</strong> the KFM signals taking into account capacitance averaging effects known to occur in<br />

KFM [2]. It will be shown that the charge transfer is enhanced due to the quantum<br />

confinement in nanocrystals with size


Open source scanning probe microscopy control and data analysis<br />

Percy Zahl 1<br />

s<strong>of</strong>tware package Gxsm.<br />

1 Brookhaven National Laboratory, Center for Functional Nanomaterials, Upton NY 11973, USA<br />

P.I-16<br />

Gxsm 1 is a full featured and modern scanning probe microscopy (SPM) s<strong>of</strong>tware. It can<br />

be used both in stand alone mode for powerful image processing and analysis and<br />

connected to an instrument operating many different flavors <strong>of</strong> SPM, e.g., scanning<br />

tunneling microscopy (STM) and atomic force microscopy (AFM) or in general twodimensional<br />

multi-channel data acquisition instruments. The Gxsm core can handle<br />

different data types, e.g., integer and floating point numbers. An easily extendable plugin<br />

architecture provides many image analysis and manipulation functions. A powerful<br />

digital signal processor (DSP) subsystem runs the feedback loop, does optional adaptive<br />

real-time signal filtering, generates the scanning signals and acquires all the data during<br />

SPM measurements. We will show lastest results and demonstrate the performance <strong>of</strong> the<br />

new SignalRanger Mark2 and Analog A810 now supported by GXSM. The<br />

programmable Gxsm vector probe engine (real-time, DSP based) performs virtually any<br />

thinkable spectroscopy and manipulation task, such as scanning tunneling spectroscopy<br />

(STS) or tip formation. The Gxsm s<strong>of</strong>tware is released under the GNU general public<br />

license (GPL) and can be obtained via the Internet.<br />

Visit and follow the Gxsm home page for further information:<br />

1 Homepage location: http://gxsm.sourceforge.net.<br />

107


2 nd generation Dynamic Nanostencil AFM/DFM/STM<br />

for in-situ (UHV) resistless patterning <strong>of</strong> nanostructures<br />

Percy Zahl 1 , Peter Sutter 1<br />

1 Brookhaven National Laboratory, Center for Functional Nanomaterials, Upton NY 11973, USA<br />

P.I-17<br />

The development and construction <strong>of</strong> the 2 nd generation Nanostencil, which will be an<br />

upgrade <strong>of</strong> the original and worldwide unique 1 st all-in-one Nanostencil as build at IBM<br />

Zurich Research Laboratory 1,2 . This Nanostencil instrument combines the shadow mask<br />

stencil method with an versatile scanning probe microscope and allows in-situ creation<br />

and inspection <strong>of</strong> even arbitrary single nanostructures down to 30nm as demonstrated at<br />

IBM.<br />

With the new design <strong>of</strong> the 2 nd generation Stencil a set <strong>of</strong> optimizations and partial<br />

redesigns will be incorporated. Nevertheless, this new CFN Nanostencil will be user<br />

accessible.<br />

The Nanostencil technique integrates a system for in-situ nano structuring and<br />

<strong>Atomic</strong>/Dynamic <strong>Force</strong> <strong>Microscopy</strong> (AFM/DFM) and Scanning Tunneling <strong>Microscopy</strong><br />

(STM) capability for in-situ structure analysis with atomic resolution capability but also<br />

wide (80µm for the 2 nd generation) scan range. In contrast to lithography methods, this is<br />

a direct writing or deposition method and no resist removal or etching steps are involved,<br />

allowing to make structures from any material which can be evaporated in UHV. Thus<br />

advanced sensitive or functional materials <strong>of</strong> an wide range (including metals, insulators,<br />

magnetic materials or even organic materials) can be used for patterning.<br />

The Scanning Probe <strong>Microscopy</strong> capabilities allow imaging and optional manipulation <strong>of</strong><br />

the deposited nanostructures/atoms and thus extends structure tune-ups down to the<br />

atomic scale.<br />

The 2 nd generation Nanostencil is currently in development at the CFN. It is expected to<br />

deliver similar or better stencil performance as the IBM-Stencil and will add more<br />

versality and a calibrated high precision scanning table with up to 80µm scan range. It<br />

will also provide a high resolution scanner (piezo tube style) for reaching atomic<br />

resolution. Optimized optical access for coarse alignments (Long Distance Microscope<br />

with 3µm resolution) while providing simultaneous access for deposition is planned.<br />

1<br />

All-in-one static and dynamic nanostencil atomic force microscopy/scanning tunneling microscopy<br />

system, P. Zahl, M. Bammerlin, G. Meyer and R.R. Schlittler, Rev. Sci. Instrum. 76, 0230707 (2005)<br />

2<br />

Fabrication <strong>of</strong> ultrathin magnetic structures by nanostencil lithography in dynamic mode, L. Gross, R.R.<br />

Schlittler, G. Meyer, A. Vanhaverbeke and R. Allenspach, Appl. Phys. Lett. 90, 093121 (2007)<br />

108


P.I-18<br />

Design <strong>of</strong> a Variable Temperature Variable Magnetic Field <strong>Noncontact</strong><br />

Scanning <strong>Force</strong> Microscope for the Characterization <strong>of</strong> Nanoscale<br />

Electronic and Magnetic Phenomena<br />

Peter Staffier 1 , Marcus Liebmann 1 , Jens Falter 1 , Nicolas Pilet 1 , Charles Ahn 2 , and Udo<br />

D. Schwarz 1<br />

1<br />

Department <strong>of</strong> Mechanical <strong>Engineering</strong> and Center for Research on Interface Structures and Phenomena<br />

(CRISP), <strong>Yale</strong> University, New Haven, USA<br />

2<br />

Department <strong>of</strong> Applied Physics and Center for Research on Interface Structures and Phenomena<br />

(CRISP), <strong>Yale</strong> University, New Haven, USA<br />

In thin film manganese oxides variations in the electric charge carrier density can result<br />

in increased conductivity, magnetic domain formation, and lattice distortions due to<br />

strong electron correlation effects. We intend to characterize the length scales at which<br />

the electronic, magnetic, and structural properties <strong>of</strong> thin films <strong>of</strong> La1-xSrxMnO3 remain<br />

correlated by using localized electrostatic fields to modulate the electric charge density<br />

while applying noncontact scanning force microscopy methods to observe the resulting<br />

phase changes. Since manganese oxides are most sensitive to electrostatic perturbation<br />

near their phase transition temperatures Tc, a critical part <strong>of</strong> the planned experiments will<br />

be to measure samples at various temperatures.<br />

For that purpose, we have developed a new ultrahigh vacuum variable temperature<br />

noncontact scanning force microscope (NC-AFM). The requirement to run experiments<br />

at distinct, stable temperatures shortly below and above Tc has been addressed by<br />

choosing a low vibration flow cryostat for cooling. Thermal gradients are minimized by<br />

cooling the entire microscope, in contrast to most commercially available variable<br />

temperature NC-AFMs. In addition, an electromagnet mounted to the vacuum system<br />

supplies magnetic fields up to 180 mT for in-field magnetic force measurements. A<br />

quartz microbalance and two evaporators allow for the preparation <strong>of</strong> multilayer<br />

magnetic force sensors in-situ. Both cantilever and sample stages permit the exchange <strong>of</strong><br />

force sensors and samples in-situ with the microscope at low temperatures. An x-y<br />

translation stage provides up to 4 mm × 4 mm <strong>of</strong> course motion in the horizontal plane so<br />

that we may measure at various positions on the sample. We will present the<br />

microscope’s design, initial measurements in topographical and magnetic operational<br />

modes, and details on some <strong>of</strong> the further planned experiments.<br />

109


An Active Q Control System in Scanning <strong>Force</strong> <strong>Microscopy</strong><br />

Jeehoon Kim 1 , Martin Zech 1 , and J. E. H<strong>of</strong>fman 1<br />

1 Department <strong>of</strong> Physics, Harvard University, Cambridge MA, USA<br />

P.I-19<br />

We have developed a high vacuum, cryogenic scanning force microscope with a fiber<br />

optic interferometer detection system. The microscope has several features: (1) flexibility<br />

to employ either a horizontal cantilever or a vertical cantilever; (2) rigidity <strong>of</strong> design to<br />

allow DC cantilever deflection detection as small as 1 picometer; (3) an x-y walker to<br />

position the tip above interesting features within a 3 mm × 3 mm area.<br />

We have also developed an active quality factor (Q) control system by employing a<br />

phase shifter using capacitive coupling. The active Q control system allows Q reduction<br />

by a factor <strong>of</strong> 60 without sacrificing signal to noise ratio. This active Q control system is<br />

essential for non-contact force microscopy with a vertical cantilever in order to increase<br />

both a force resolution and spatial resolution. The Q control system could also be used in<br />

either vertical or horizontal cantilever mode to increase scanning speed, or to employ<br />

amplitude modulation mode scanning in a vacuum or low temperature environment.<br />

a) b)<br />

Figure 1: (a) The home-built high vacuum, cryogenic magnetic force microscope and (b)<br />

Schematic <strong>of</strong> the active Q system, showing phase shifter and capacitive coupling.<br />

110


P.I-20<br />

Besocke style quartz tuning fork FM-AFM/STM for use in UHV and<br />

low temperatures<br />

Shawn M. Huston 1 , Rachel T. Port 1 , Katie M. Andrews 1 , and Thomas P. Pearl 1<br />

1 Department <strong>of</strong> Physics, North Carolina State University, Raleigh, USA<br />

The design and operation <strong>of</strong> a Besocke-style scanning probe microscope with both<br />

frequency modulated atomic force and scanning tunneling microscopy capabilities will be<br />

presented. This instrument has been optimized for use at low temperatures, either with<br />

helium or nitrogen cooling, in ultrahigh vacuum. FM-AFM is performed via a quartz<br />

crystal tuning fork in the qPlus sensor configuration. Cooling <strong>of</strong> the microscope is<br />

achieved through the use <strong>of</strong> a top-loaded bath cryostat that cools a pair <strong>of</strong> radiation<br />

shields which house the spring-suspended microscope assembly. This Besocke-style<br />

microscope is designed to record tunneling current and resonance frequency shift for a<br />

conductive tip mounted on the free prong <strong>of</strong> the tuning fork simultaneously Single<br />

molecule images <strong>of</strong> pentacene on Ag(111) at ~5 K recorded in both NC-AFM and STM<br />

modes will be pr<strong>of</strong>fered as a testament to the microscope’s effectiveness.<br />

111


P.I-21<br />

Design <strong>of</strong> a Low Temperature <strong>Noncontact</strong> <strong>Atomic</strong> <strong>Force</strong> Microscope<br />

Combined with a Field Ion Microscope<br />

J. Falter 1 , D.-A. Braun 1 , H. Hölscher 2 , U. D. Schwarz 3 , A. Schirmeisen 1 , and H. Fuchs 1<br />

1 CeNTech Center for Nanotechnology and Institute <strong>of</strong> Physics, University <strong>of</strong> Münster, Germany<br />

2 IMT, Forschungszentrum Karlsruhe, Germany<br />

3 Department <strong>of</strong> Mechanical <strong>Engineering</strong>, <strong>Yale</strong> University, New Haven, USA<br />

Non-contact atomic force microscopy has become a standard tool for imaging surfaces<br />

with atomic resolution. The underlying contrast mechanism is influenced by both<br />

interaction partners, the sample as well as the tip. Theoretical simulations <strong>of</strong> the force<br />

interaction allow a deeper insight in the contrast mechanisms but require information<br />

about the exact position <strong>of</strong> all atoms <strong>of</strong> sample and the tip. A method that is able to<br />

determine the configuration <strong>of</strong> the probing tip with atomic precision is the field ion<br />

microscope (FIM).<br />

We present a home-built low temperature ultra high vacuum (UHV) system that<br />

combines these two microscopy techniques. The design <strong>of</strong> the AFM has been proven to<br />

operate in UHV at low temperature and features a high mechanical stability [1]. The<br />

original design was modified to use piezoelectric tuning forks as force sensors, with a<br />

glued-on tungsten wire, prepared by the double lamella etching technique [2] (Figure 1a).<br />

For stable oscillation, the tuning fork is excited electronically [3] in the frequency<br />

modulation mode. Our force sensors have been successfully employed in the AFM as<br />

well as the FIM. Figure 1b shows a FIM image <strong>of</strong> a tuning fork tungsten tip allowing the<br />

atom-by-atom reconstruction <strong>of</strong> the tip apex demonstrated in Figure 1c.<br />

a) b) c)<br />

Figure 1: a) Image <strong>of</strong> the tuning fork force sensor with glued-on tungsten wire. b) FIM image <strong>of</strong><br />

a tuning fork tungsten tip, indicating the different crystallographic axis. A bright spot in the FIM<br />

image corresponds to a tip apex atom. c) Atom-by-atom reconstruction <strong>of</strong> the whole tip apex.<br />

[1] B. Albers et al, Rev. Sci. Instrum. 79 033704 (2008).<br />

[2] M. Kulawik et al, Rev. Sci. Instrum. 74, 1027 (2002)<br />

[3] J. Jersch et al, Rev. Sci. Instrum. 77, 083701 (2006)<br />

112


P.I-22<br />

A homebuilt low-temperature STM / tuning fork AFM combination<br />

Manfred Lange 1 , Johannes Schaffert 1 , Nikolai Wintjes 1 and Rolf Möller 1<br />

1 Department <strong>of</strong> Physics, University <strong>of</strong> Duisburg-Essen, Germany<br />

We present details on a homebuilt, compact scanning probe microscope (Fig. 1) which<br />

can be switched between tuning fork AFM and STM operation without breaking vacuum.<br />

The geometry <strong>of</strong> the microscope resembles a cylinder with a height <strong>of</strong> only 15 cm and a<br />

diameter <strong>of</strong> 4 cm. It is attached to a commercially available continuous flow cryostat<br />

which allows cooling to about 5-7 K. Because <strong>of</strong> the very compact design, low helium<br />

consumption <strong>of</strong> only 1 liter/hour is achieved. Drift rates in lateral direction are


Development <strong>of</strong> quartz force sensors for noncontact atomic force<br />

microscopy/spectroscopy<br />

Kenichirou Hori 1 , Toyoko Arai 1 and Masahiko Tomitori 2<br />

(a) (b)<br />

Amplitude[nm]<br />

Flexural oscillation amplitude and phase<br />

450<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

Amplitude[nm]<br />

Phase[degree]<br />

5370 5380 5390 5400 5410 5420 5430<br />

Frequency[Hz]<br />

P.I-23<br />

1 Graduate <strong>School</strong> <strong>of</strong> Natural science & Technology, Kanazawa University, Kanazawa, Ishikawa, Japan<br />

2 <strong>School</strong> <strong>of</strong> Materials Science, Japan Advanced Institute <strong>of</strong> Science and Technology, Nomi, Ishikawa Japan<br />

In order to extend the application <strong>of</strong> noncontact atomic force microscopy (nc-AFM),<br />

force sensors have still held the key to supplement the functions <strong>of</strong> nc-AFM, in particular,<br />

for force spectroscopy. For example, bias-voltage noncontact atomic force spectroscopy<br />

(nc-AFS) [1] developed on the basis <strong>of</strong> nc-AFM, invokes probes suitable to electric<br />

conductance measurements with high force sensitivity to characterize the relationship<br />

between binding state and electronic state. One possibility is to use quartz, one <strong>of</strong><br />

piezoelectric materials with a high Q, e.g., successfully demonstrated by Giessibl [2], on<br />

which a metal or a Si needle is attached as a probe. Though a Si probe seems suitable to<br />

examine the relationship between electronic states and chemical bonding force for Si<br />

samples from point contact to non-contact through pseudo-contact regime, a metallic<br />

needle sounds as a good conductor tip. The tip material is also desired to be easily<br />

exchangeable. Moreover, the reduction <strong>of</strong> cantilever oscillation amplitude takes<br />

advantage for conductance measurement [3], which were also realized using quartz force<br />

sensors. Here we report the development <strong>of</strong> a quartz force sensor using lithographic<br />

techniques from a quartz wafer, which has electrodes suitable to simultaneous<br />

conductance measurement with high sensitivity <strong>of</strong> force at a small oscillation amplitude<br />

with feasibility <strong>of</strong> probe material change.<br />

We fabricated a quartz sensor with gold-plated multi-electrodes, one <strong>of</strong> which is to<br />

fix the potential <strong>of</strong> a probe for conductance measurement. Figure 1 shows a schematic <strong>of</strong><br />

the sensor and mechanical characteristics <strong>of</strong> a flexure mode in air near its resonance<br />

frequency measured using a heterodyne optical interferometer. This sensor exhibits a<br />

longitudinal mode as well,<br />

analyzed them by a finite<br />

element method. To avoid Q<br />

degradation, bonding <strong>of</strong> Si tip<br />

onto quartz without adhesive<br />

agent is adopted.<br />

[1] T. Arai and M. Tomitori: Phys.<br />

Rev. B 73 (2006) 073307.<br />

[2] F.J. Giessibl: Appl. Phys. Lett. 76<br />

(2000) 1470.<br />

[3] T. Arai and M. Tomitori: Jpn. J.<br />

Appl. Phys. 39 (2000) 3753.<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

-20<br />

-40<br />

-60<br />

-80<br />

-100<br />

Fig.1(a) Schematic <strong>of</strong> a fabricated sensor. (b) Resonance<br />

curves <strong>of</strong> the flexural mode <strong>of</strong> the sensor without a probe.<br />

114<br />

Phase[degree]


P.I-24<br />

High-Speed Frequency Modulation <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong> using<br />

Wideband Digital Phase-Locked Loop Detector<br />

Takeshi Fukuma 1,2 and Yuji Mitani 1<br />

1 Frontier Science Organization, Kanazawa University, Kanazawa, Japan<br />

2 PRESTO, Japan Science and Technology Agency, Kawaguchi, Japan<br />

Recent advancement in the instrumentation <strong>of</strong> frequency modulation atomic force<br />

microscopy (FM-AFM) has enabled to operate FM-AFM in liquid with true atomic<br />

resolution. To date, liquid-environment FM-AFM has been used for several biological<br />

applications, which has demonstrated its excellent force sensitivity and high spatial<br />

resolution. However, unlike applications in vacuum, biological applications require highspeed<br />

operation <strong>of</strong> FM-AFM due to the mobile nature, high complexity and large<br />

dimension <strong>of</strong> biological systems. In this study, we develop a high-speed FM-AFM using<br />

FPGA-based digital signal processing circuit in order to improve the applicability <strong>of</strong> this<br />

method to practical biological studies.<br />

Development <strong>of</strong> high-speed FM-AFM requires enhancing the bandwidth and resonance<br />

frequencies <strong>of</strong> all the components involved in the tip-sample distance feedback loop. In<br />

particular, a frequency detector has been one <strong>of</strong> the major speed limiting factors in the<br />

feedback loop. A conventional frequency detector using a digital phase-locked loop<br />

(PLL) typically utilizes multiplication-based phase comparator, which has limited its<br />

bandwidth to less than 10 kHz. In this study, we have developed a wideband digital PLL<br />

using a subtraction-based phase comparator (Fig. 1(a)). The developed PLL detector has<br />

a bandwidth <strong>of</strong> wider than 1 MHz (Fig. 1(b)). A high-speed FPGA circuit has been used<br />

for implementing the PLL as well as the cantilever excitation circuit, the distance and<br />

amplitude feedback control circuits. Combined with our recently developed high-speed<br />

scanner and wideband photo-thermal excitation system, we aim at high-speed operation<br />

<strong>of</strong> FM-AFM in liquid.<br />

Figure 1: (a) Block diagram and (b) frequency response <strong>of</strong> the developed PLL using a<br />

subtraction-based phase comparator (fc : cut<strong>of</strong>f frequency <strong>of</strong> a low-pass filter at the output).<br />

115


Recent Advances in Multi-Spectral <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong><br />

S. Jesse, N. Balke, P. Maksymovych, O. Ovchinnikov, A.P. Baddorf, S.V. Kalinin<br />

Oak Ridge National Laboratory, Oak Ridge, TN 37831<br />

P.I-25<br />

Piezoresponse force microscopy has become the primary method for the<br />

characterization <strong>of</strong> electromechanical activity on the nanometer scale. The recent<br />

introduction <strong>of</strong> the band excitation mode has allowed resonance enhancement in PFM<br />

through the detection <strong>of</strong> amplitude-frequency curve at each pixel. Similarly, switching<br />

spectroscopy PFM has allowed measurements <strong>of</strong> polarization dynamics. Both these<br />

techniques give rise to 3D data arrays. Combining both techniques into BE SSPFM yields<br />

4D data, while implementation <strong>of</strong> first order reversal curve measurements yields 5D data<br />

sets. In the absence <strong>of</strong> exact and reliable analytical models, data analysis <strong>of</strong> high<br />

dimensional data sets to elucidate relevant aspects <strong>of</strong> materials behavior and link them to<br />

theory represents a complex problem.<br />

Here we discuss an approach for the analysis <strong>of</strong> multi-dimensional, spectroscopicimaging<br />

data based on principal component analysis (PCA) coupled with neural network<br />

recognition algorithms. PCA selects and ranks relevant response components based on<br />

variance within the data. It is shown that for examples with small relative variations<br />

between spectra, the first few PCA components coincide with results obtained using<br />

model fitting, and achieves this at rates approximately 4 orders <strong>of</strong> magnitude faster. For<br />

cases with strong response variations, PCA allows an effective approach to rapidly<br />

process, de-noise, and compress data. The combination <strong>of</strong> PCA with correlation analysis<br />

is demonstrated as a means to delineate the principal components containing information<br />

from those dominated by noise. In ferroelectric data, PCA allows for the selection <strong>of</strong><br />

minute effects related to polarization orientation changes during switching. Here, the<br />

PCA based approach is used to study domain wall dynamics in ferroelectric and<br />

multiferroic materials and polarization behavior in ferroelectric capacitors and relaxors.<br />

Further prospects for other multidimensional SPM methods are discussed.<br />

Research was supported by the U.S. Department <strong>of</strong> Energy Office <strong>of</strong> Basic<br />

Energy Sciences Division <strong>of</strong> Scientific User Instruments and was performed at Oak<br />

Ridge National Laboratory which is operated by UT-Battelle, LLC.<br />

Figure 1: The first 4 PCA components <strong>of</strong> band excitation magnetic force microscopy image <strong>of</strong><br />

yttrium-iron garnet (data acquired in collaboration with R. Proksch, Asylum Research). PCA is<br />

able to decompose the data into 2 conservative, 1 background, and 1 dissipation component.<br />

116


P.I-26<br />

Deciphering Nanoscale Interactions: Artificial Neural Networks and<br />

Scanning Probe <strong>Microscopy</strong><br />

Maxim Nikiforov, Stephen Jesse, Oleg Ovchinnikov, and Sergei V. Kalinin<br />

Oak Ridge National Laboratory, Oak Ridge, TN 37831<br />

Scanning Probe <strong>Microscopy</strong> techniques provide a wealth <strong>of</strong> information on<br />

nanoscale interactions. The rapid emergence <strong>of</strong> spectroscopic imaging techniques in<br />

which the response to local force, bias, or temperature is measured at each spatial<br />

location necessitates the development <strong>of</strong> data interpretation and visualization techniques<br />

for 3- or higher dimensional data sets.<br />

In this presentation, we summarize recent advances in the application <strong>of</strong> neural<br />

network based artificial intelligence methods to scanning probe microscopy. The<br />

examples will include biological identification based on the dynamics <strong>of</strong> the<br />

electromechanical response, direct mapping <strong>of</strong> dynamic disorder in ferroelectric relaxors,<br />

and reconstruction <strong>of</strong> random bond-random field Ising model parameters in ferroelectric<br />

capacitors. The future prospects for smart, multispectral SPMs are discussed.<br />

Research was supported by the U.S. Department <strong>of</strong> Energy Office <strong>of</strong> Basic<br />

Energy Sciences Division <strong>of</strong> Scientific User Facilities and was performed at Oak Ridge<br />

National Laboratory which is operated by UT-Battelle, LLC.<br />

Figure 1: Results <strong>of</strong> neural network identification <strong>of</strong> bacteria (M. Lysodeikticus (red), P.<br />

Fluorescens (green)) are overlaid on the topographic image. Training area for neural network is<br />

outlined by black rectangle (in collaboration with A. Vertegel and V. Reukov, Clemson<br />

University).<br />

117


P.I-27<br />

NC-AFM study <strong>of</strong> a cleaved InAs (110) surface using modified Si probes<br />

under ambient atmospheric pressure<br />

Yonkil Jeong 1,2 , Masato Hirade 2,3 , Ryohei Kokawa 2,3 , Hir<strong>of</strong>umi Yamada 2,4 ,<br />

Kei Kobayashi 2,4 , Noriaki Oyabu 2,4 , Hiroshi Yamatani 2,5 , Toyoko Arai 2,5 ,<br />

Akira Sasahara 1,2 , and Masahiko Tomitori 1,2,*<br />

1 <strong>School</strong> <strong>of</strong> Materials Science, Japan Advanced Institute <strong>of</strong> Science and Technology, Ishikawa, Japan,<br />

2 JST Advanced Measurement and Analysis, 3 Shimadzu Corp., Kyoto, Japan,<br />

4 Kyoto University, Kyoto, Japan, 5 Kanazawa University, Ishikawa, Japan<br />

Characterization <strong>of</strong> crystalline defects in thin films with interfaces <strong>of</strong> lattice-mismatched<br />

III-V compound semiconductors such as GaAs and InP is an important issue for practical<br />

device application. We pursue a simple and quick method to characterize them on an<br />

atomic scale using NC-AFM by cleaving devices in a controlled environment without a<br />

UHV system. In this work, we demonstrate the possibilities <strong>of</strong> high-resolution imaging <strong>of</strong><br />

a cleaved InAs (110) surface using the NC-AFM under atmospheric pressure <strong>of</strong> air or<br />

pure Ar, and <strong>of</strong> acquiring charge state information with specially fabricated probes,<br />

which are designed to emphasize electronic interaction between the tip and the sample.<br />

High aspect ratio (HAR) Si and Ge/Si probes were fabricated from commercial<br />

AFM Si probes by an FIB and a Ge deposition system. Fig. 1(a) shows the changes in ? f<br />

due to the reduction <strong>of</strong> capacitance between probes and a sample. A high resolution<br />

image with a periodic structure was successfully obtained with a fabricated probe, in Fig.<br />

1(c); possibly residual H2O molecules were adsorbed on a cleaved InAs (110) surface.<br />

(a (b<br />

Δf = -43Hz, A = 0.4nm<br />

Figure 1: (a) Bias-Δf curves between probes and a sample surface. (b) SEM images <strong>of</strong><br />

FIB fabricated probes. (c) NC-AFM image <strong>of</strong> a cleaved InAs (110) with HAR-Si in Ar.<br />

*corresponding author e-mail: tomitori@jaist.ac.jp<br />

118<br />

(c)


Dual-Frequency-Modulation AFM Spectroscopy<br />

Gaurav Chawla, C. Alan Wright, and Santiago D. Solares<br />

Department <strong>of</strong> Mechanical <strong>Engineering</strong>, University <strong>of</strong> Maryland at College Park, USA<br />

P.I-28<br />

<strong>Force</strong> spectroscopy is an important application <strong>of</strong> atomic force microscopy (AFM),<br />

whereby tip-sample interaction forces are measured with probes <strong>of</strong> varying geometry and<br />

chemistry. Most methods rely on measuring either the instantaneous cantilever deflection<br />

in static mode or the frequency shift in low-amplitude frequency-modulation mode. In<br />

the first case the force is calculated directly from the cantilever deflection. In the second<br />

case the frequency shift is used to compute the tip-sample force gradient, from which the<br />

tip-sample force curve can be reconstructed after scanning a region for which a boundary<br />

condition is known. In both cases, acquiring a 3D representation <strong>of</strong> the tip-sample forces<br />

requires fine-grid scanning <strong>of</strong> a volume above the surface. Using computational<br />

simulations, we have developed a dual-frequency-modulation AFM spectroscopy method<br />

[1,2] in which the cantilever is simultaneously excited at the fundamental frequency and<br />

at a higher-order eigenfrequency (with a much smaller amplitude), such that<br />

topographical imaging is performed using the fundamental frequency response and force<br />

spectroscopy is performed using the higher-order response. Our results suggest that this<br />

method could enable measurement <strong>of</strong> the 3D tip-sample forces above the surface with a<br />

single 2D scan. The procedure was originally developed for ambient air, but has been<br />

extended to vacuum conditions, where we are evaluating the 3D imaging <strong>of</strong> atomic<br />

orbitals by integrating quantum mechanics and cantilever dynamics calculations. We are<br />

especially interested in the effect <strong>of</strong> tip apex geometry on the ability to image sub-atomic<br />

features. The method’s experimental implementation in air is in progress.<br />

1.85<br />

a b<br />

Dual-frequency tip response<br />

ν 1 band<br />

ν 2 band<br />

imaging<br />

spectroscopy<br />

Effective frequency, MHz<br />

1.75<br />

1.65<br />

1.55<br />

1.45<br />

0 20 40 60 80 100<br />

Time, ms<br />

Tip approach<br />

Tip retract<br />

Figure 1: (a) Illustration <strong>of</strong> the dual-frequency-modulation AFM concept, where the cantilever<br />

vibrates at two eigenfrequencies such that the fundamental response is used to perform imaging<br />

and the high-frequency response is used to perform force spectroscopy; (b) illustration <strong>of</strong> the<br />

instantaneous frequency <strong>of</strong> the self-excited high-frequency response for one full oscillation <strong>of</strong> the<br />

low-frequency response, under positive frequency shift for the fundamental frequency. Since the<br />

frequency shift follows the tip-sample force gradient, it is in principle possible under ideal<br />

conditions to extract the tip-sample force curve every low-frequency oscillation.<br />

[1] G. Chawla and Solares, S.D., Measurement Science and Technology 20, No. 015501 (2009).<br />

[2] S.D. Solares and G. Chawla, Measurement Science and Technology 19, No. 055502 (2008).<br />

119


Theory <strong>of</strong> Multifrequency Method in FM-AFM<br />

Zongmin Ma, Yoshitaka Naitoh, Yanjun Li, Masami Kageshima<br />

and Yasuhiro Sugawara<br />

Department <strong>of</strong> Applied Physics, Graduate <strong>School</strong> <strong>of</strong> <strong>Engineering</strong>, Osaka University, 2-1 Yamada-oka,<br />

Suita, Osaka 565-0871, Japan<br />

P.I-29<br />

FM-AFM has been succeeded in measuring topography and energy dissipation<br />

with atomic resolution on the surface. Furthermore, force spectroscopy is<br />

demonstrated to identify the atomic species. The elasticity information <strong>of</strong> the surface<br />

is important physical information <strong>of</strong> the sample materials. Actually, it has been<br />

measured by using multifrequency method in AM-AFM recently, which was proposed<br />

by Rodriguez and Garcia and successfully demonstrated its usefulness theoretically<br />

and experimentally [1-2].<br />

In this paper, for the first time, we<br />

proposed usage <strong>of</strong> the multifrequency<br />

f 2<br />

BPF PLL Δf 2<br />

method in FM-AFM to measure the<br />

AGC<br />

sample properties such as elasticity.<br />

Figure 1 shows the block diagram <strong>of</strong><br />

×<br />

multifrequency method in FM-AFM.<br />

f 1<br />

BPF PLL<br />

The cantilever is excited simultaneously<br />

at the first and the second resonances.<br />

Band pass filters (BPFs) are used to<br />

separate the vibration signals <strong>of</strong> the<br />

Optical<br />

interferometer<br />

+<br />

+<br />

AGC<br />

×<br />

cantilever. For each signal, the cantilever<br />

Cantilever<br />

Δf 1<br />

is vibrated by self-oscillation at the first<br />

Feedback<br />

and the second resonances. Topography<br />

Tube Scanner<br />

Topography<br />

<strong>of</strong> the sample is obtained by the<br />

feedback signal to keep the frequency<br />

Figure 1. Block diagram <strong>of</strong> multifrequency<br />

shift (△ f1) constant. The elasticity information <strong>of</strong> the sample surface is measured by<br />

frequency shift (△ f2) <strong>of</strong> the second resonance, which is proportional to the<br />

conservative force between tip and sample. Here, we show the theory for the<br />

multifrequency method in FM-AFM.<br />

[1] Rodriguez and R. Garcia, Appl. Phys. Lett. 84, 449 (2004).<br />

[2] Jose R. Lozano and Ricardo Garcia. Phy Rev B 79, 014110 (2009).<br />

120


P.I-30<br />

Internal Resonances and Spatio-Temporal Instabilities in Nonlinear<br />

Multi-mode NC-AFM Dynamics<br />

Oded Gottlieb, Sharon Hornstein, Wei Wu and Arthur Shavit<br />

Department <strong>of</strong> Mechanical <strong>Engineering</strong>, Technion - Israel Institute <strong>of</strong> Technology, Haifa, Israel<br />

The use <strong>of</strong> multi-mode excitation in atomic force microscopy has been proposed and<br />

validated in the past decade. Examples include the combined excitation <strong>of</strong> the first two<br />

bending modes where the latter has been found to be sensitive to low surface force<br />

variations [1], and excitation <strong>of</strong> a third bending mode in the vicinity <strong>of</strong> a torsion mode to<br />

map nanomechanical changes <strong>of</strong> a polymer near its glass transition [2]. While small<br />

amplitude excitation results in a single-valued frequency response, finite amplitude<br />

dynamics can yield nonlinear coexisting bi-stable solutions and complex aperiodic<br />

dynamics that reveal nonstationary energy transfer between multiple spatial modes. The<br />

accuracy <strong>of</strong> force estimation from measured data crucially depends on the quality <strong>of</strong> the<br />

mathematical model in use. Thus, as lumped mass models cannot describe multi-mode<br />

dynamics, a continuum mechanics description is required to consistently incorporate<br />

nonlinear atomic interaction and coupled elastic and internal damping forces that govern<br />

noncontact operation in ultrahigh vacuum. Thus, the objectives <strong>of</strong> this paper include<br />

theoretical derivation and analysis <strong>of</strong> a continuum spatio-temporal model for the<br />

vibrating NC-AFM cantilever that consistently incorporates both nonlinear atomic<br />

interaction and coupled thermo-visco-elastic damping. We derive a nonlinear initialboundary-value<br />

problem using the extended Hamilton's principle [3] for the coupled<br />

viscoelastic and temperature fields for torsion, transverse and out-<strong>of</strong>-plane bending. The<br />

continuum system is then reduced via a Galerkin procedure to a multi-mode dynamical<br />

system which is analyzed numerically. System response reveals that the dynamic jumpto-contact<br />

bifurcation threshold is not sensitive to the damping level but includes lengthy<br />

chaotic transients for low damping. Furthermore, both subharmonic and quasiperioic<br />

solutions are found when combination and internal resonances are excited. The latter<br />

reveal energy transfer between the 3rd and 2ond microbeam modes due to a strong 3:1<br />

internal resonance [Fig.1] and complex chaotic like spatio-temporal instabilities when<br />

this internal resonance is coupled with either its out-<strong>of</strong>-plane or torsion resonances.<br />

Figure 1: Quasiperiodic dynamics <strong>of</strong> a 3:1 internal resonance between the 3 rd and 2 nd bending<br />

modes in UHV: time series (left), power spectra(center), and Poincare' map (right).<br />

[1] T.R. Rodriguez and R. Garcia, APL, 83, 449 (2004).<br />

[2] O. Sahin et al. Nature Nanotechnology 2, 507 (2007).<br />

[3] S. Hornstein and O. Gottlieb, Nonlinear Dynamics 54, 93 (2008).<br />

121


P.I-31<br />

Frequency Noise in Frequency Modulation <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong><br />

Kei Kobayashi 1 , Hir<strong>of</strong>umi Yamada 2 , and Kazumi Matsushige 2<br />

1 Innovative Collaboration Center, Kyoto University, Kyoto, Japan.<br />

2 Department <strong>of</strong> Electronic Science and <strong>Engineering</strong>, Kyoto University, Kyoto, Japan.<br />

<strong>Atomic</strong> force microscopy (AFM) using the frequency modulation (FM) detection method<br />

has been widely used for atomic/molecular-scale investigations <strong>of</strong> various materials.<br />

Recently, it has been shown that high-resolution imaging in liquids by the FM-AFM is<br />

also possible by reducing the noise-equivalent displacement in the cantilever<br />

displacement sensor and by oscillating the cantilever at a small amplitude, even with the<br />

extremely reduced Q-factor due to the hydrodynamic interaction between the cantilever<br />

and the liquid. However, it has not been clarified how the noise reduction <strong>of</strong> the<br />

displacement sensor contributes to the reduction <strong>of</strong> the frequency noise in the FM-AFM<br />

in low-Q environments. In this presentation, the contribution <strong>of</strong> the displacement sensor<br />

noise to the frequency noise in the FM-AFM is analyzed in detail to show how it is<br />

important to reduce the noise-equivalent displacement in the displacement sensor<br />

especially in low-Q environments.<br />

As a general equation for the frequency noise density <strong>of</strong> the oscillator, we have to<br />

consider the contribution <strong>of</strong> the displacement sensor noise to the oscillator noise in<br />

addition to the frequency noise <strong>of</strong> the high-Q cantilevers,<br />

where N ds is the noise-equivalent displacement sensor noise density.<br />

Figure 1: Schematics <strong>of</strong> the evolution <strong>of</strong> the displacement noise into the frequency noise without<br />

and with the displacement sensor noise. The displacement noise spectrum <strong>of</strong> the cantilever around<br />

the resonance frequency without the displacement sensor noise (a) and the corresponding<br />

oscillator frequency noise (b). The total frequency noise considering the measurement noise<br />

(dotted area in (b)) becomes constant as shown in (c). If there is non-zero displacement sensor<br />

noise, it brings additional oscillator frequency noise and measurement noise as shown in (d).<br />

Dark gray and black areas represent two levels (small and large) <strong>of</strong> additional displacement<br />

sensor noise.<br />

[1] K. Kobayashi, H. Yamada, and K. Matsushige, Rev. Sci. Instrum. accepted for publication (2009).<br />

122<br />

,


Relation between lateral forces and dissipation in FM-AFM<br />

Michael Klocke, Dietrich E. Wolf<br />

Department <strong>of</strong> Physics, University <strong>of</strong> Duisburg-Essen, Germany.<br />

P.I-32<br />

We study the coupling <strong>of</strong> torsional and normal cantilever oscillations and their effect on<br />

the imaging process <strong>of</strong> a frequency-modulated atomic force microscope by means <strong>of</strong><br />

molecular dynamics simulations. We show that the bending and torsional modes <strong>of</strong> the<br />

cantilever are coupled if the tip is near the surface and connect this coupling to the<br />

damping <strong>of</strong> the cantilever oscillation. The strength <strong>of</strong> the coupling is determined roughly<br />

by the strength <strong>of</strong> the lateral forces on the closest approach <strong>of</strong> the tip. Energy is<br />

transferred from the normal to the torsional excitation which can be detected as damping<br />

<strong>of</strong> the cantilever oscillation. Energy is actually dissipated by the usually uncontrolled<br />

mechanical damping <strong>of</strong> the torsional excitation. For high Q factors, the transferred energy<br />

is not completely dissipated during one cycle. The question, what happens to the<br />

remaining energy <strong>of</strong> the lateral degree <strong>of</strong> freedom in the long run, is addressed by<br />

studying a simplified two-dimensional point-mass model. We show that in succeeding<br />

cycles, energy is transferred back into the normal degree <strong>of</strong> freedom. The observation <strong>of</strong><br />

the energy swapping process (amplitude and frequency) can therefore give additional<br />

information <strong>of</strong> the surface structure, especially on lateral forces.<br />

123


P.I-33<br />

Experimental Study <strong>of</strong> Dissipation Mechanisms in AFM Cantilevers<br />

Fredy Zypman<br />

Department <strong>of</strong> Physics, Yeshiva University, New York, USA.<br />

The presence <strong>of</strong> energy dissipation in an oscillating AFM cantilever is a fact that must be<br />

included in any force reconstruction algorithm. Although commonly low quality factor<br />

(Q) cantilevers get on the way <strong>of</strong> high accuracy non-contact measurements, some times<br />

they can be used purposely. A case in point is the study <strong>of</strong> the measurement <strong>of</strong> fluids<br />

viscosity by monitoring changes in Q [1]. In this work we study theoretically the<br />

mechanisms <strong>of</strong> energy dissipation in AFM, and validate the corresponding models<br />

experimentally. This understanding <strong>of</strong> the origins <strong>of</strong> energy dissipation is not only<br />

interesting for its own sake, but it is also relevant with an eye on practical applications.<br />

We have done extensive work on cantilevers <strong>of</strong> various shapes but, to fix ideas we<br />

consider a rectangular cantilever in this abstract. In that case, the beam model is<br />

commonly used to obtain the corresponding frequency spectrum via the steady state<br />

4<br />

2<br />

EI ∂ u(<br />

x,<br />

t)<br />

∂ u(<br />

x,<br />

t)<br />

solution <strong>of</strong> + = 0 , where u( x,<br />

t)<br />

is the deflection <strong>of</strong> the cantilever<br />

4<br />

2<br />

Aρ<br />

∂x<br />

∂t<br />

at location x and time t , E is the Young's modulus, I the cross sectional moment <strong>of</strong><br />

inertia, A the cross sectional area, and ρ the mass density. In the beam model, a<br />

∂u dissipation term proportional to is usually introduced. The proportionality factor is a<br />

∂t<br />

difficult problem in itself that involves the interaction <strong>of</strong> the cantilever with the<br />

surrounding fluid [2]. The complete equation describes quite well the central frequency<br />

and width <strong>of</strong> multiple peaks. However, a detailed analysis shows that the peak shape is<br />

not in full agreement with experiment if fine details are included. We will show that this<br />

is improved by introducing explicit dissipation terms based on basic physical principles<br />

that represent cantilever-fluid interaction and internal viscoelasticity. The model is<br />

solved and compared with experimental data obtained on a Veeco AFM. We will also<br />

argue that this improvement is necessary to produce reconstruction algorithms consistent<br />

with the resolutions necessary today to measure objects at the subnanometer scale, like<br />

charges in biological molecules.<br />

[1] A. Schilowitz, D. Yablon, E. Lansey, and F. Zypman, Measurement 41, 1169 (2008).<br />

[2] J.E. Sader, J. Appl. Phys. 84, 64 (1998)<br />

124


M. Teresa Cuberes<br />

Ultrasonic Nanolithography on Hard Substrates<br />

Laboratory <strong>of</strong> Nanotechnology, University <strong>of</strong> Castilla-La Mancha, Almadén, Spain.<br />

P.I-34<br />

Ultrasonic- AFM techniques provide a means to monitor ultrasonic vibration at the<br />

nanoscale and open up novel opportunities to improve nan<strong>of</strong>abrication technologies [1].<br />

Ultrasonic <strong>Force</strong> <strong>Microscopy</strong> (UFM) relies on the mechanical-diode cantilever response<br />

when ultrasonic vibration is excited at the tip-sample contact [2]. Strictly, UFM is a<br />

dynamic AFM (<strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong>) implementation in which the tip-sample<br />

contact is periodically broken at ultrasonic frequencies. In the presence <strong>of</strong> ultrasonic<br />

vibration, the tip <strong>of</strong> a s<strong>of</strong>t cantilever can dynamically indent hard samples due to its<br />

inertia. Fig. 1 (a) demonstrates scratching <strong>of</strong> a silicon sample in the presence <strong>of</strong> ultrasonic<br />

vibration, using a cantilever with nominal stiffness <strong>of</strong> 0.11 Nm -1 and a SiN tip. In the<br />

absence <strong>of</strong> ultrasound, it was not possible to scratch the Si surface using such a s<strong>of</strong>t<br />

cantilever. Interestingly, no debris is found in the proximity <strong>of</strong> lithographed areas.<br />

Fig. 1 (b) displays ultrasonic curves -UFM response for increasing ultrasonic<br />

excitation amplitude- recorded during UFM imaging (curve (a)) and during scratching <strong>of</strong><br />

silicon (curves c-d). UFM imaging is carried out without substrate modification by using<br />

a lower tip-sample load. When increasing the load, the silicon substrate is scratched<br />

provided a sufficiently high ultrasonic amplitude is excited. By monitoring the ultrasonicinduced<br />

cantilever responses during manipulation, information about the substrate<br />

modification can be obtained. As the substrate is being scratched, the UFM responses<br />

change. Apparently, for a fixed load, a critical ultrasonic amplitude is needed to induce<br />

modification (see Fig 1 (b)). In the poster, the advantages <strong>of</strong> ultrasonic nanolithography,<br />

and the shape <strong>of</strong> UFM curves recorded during manipulation will be discussed in detail.<br />

a) b)<br />

Figure 1. AFM scratching <strong>of</strong> Si(111) in the presence <strong>of</strong> normal surface ultrasonic vibration <strong>of</strong> ~<br />

5 MHz using a pyramidal SiN cantilever tip with nominal stiffness 0.11 Nm -1 (a) Nanotrenches<br />

formed in 50, 70 and 100 cycles respectively, at a load <strong>of</strong> ~ 40 nN and maximum ultrasonic<br />

amplitude Am = 500 mV (b) UFM response for a triangular modulation <strong>of</strong> the ultrasonic<br />

excitation (see lowest curve). Load: (i) ~ 26 nN. (ii-iv) ~ 40 nN. Am=500 mV.<br />

[1] M. T. Cuberes, J. <strong>of</strong> Phys.: Conf. Ser. 61 (2007) 219.<br />

[2] M. T. Cuberes in “Applied Scanning Probe Methods”, B. Bhushan and H. Fuchs (ed.), Springer 2009.<br />

125


P.I-35<br />

Silicon nanowire transistors with a channel width <strong>of</strong> 4 nm fabricated by<br />

atomic force microscope nanolithography<br />

Javier Martinez, Ramses V. Martinez, and Ricardo Garcia<br />

Instituto de Microelectronica deMadrid - CSIC, Isaac Newton 8, 28760 Tres Cantos, Madrid, Spain<br />

The emergence <strong>of</strong> an ultrasensitive sensor technology based on silicon nanowires<br />

requires both the fabrication <strong>of</strong> nanoscale diameter wires and the integration with<br />

microelectronic processes [1]. Here we demonstrate an atomic force microscopy<br />

lithography that enables the reproducible fabrication <strong>of</strong> complex single-crystalline silicon<br />

nanowire field-effect transistors with a high electrical performance. The nanowires have<br />

been carved from a silicon-on-insulator wafer by a combination <strong>of</strong> local oxidation<br />

processes [2] with a force microscope and etching steps. We have fabricated and<br />

measured the electrical properties <strong>of</strong> a silicon nanowire transistor with a channel width <strong>of</strong><br />

4 nm. The flexibility <strong>of</strong> the nan<strong>of</strong>abrication process is illustrated by showing the<br />

electrical performance <strong>of</strong> two nanowire circuits with different geometries [3]. The<br />

fabrication method is compatible with standard Si CMOS processing technologies and,<br />

therefore, can be used to develop a wide range <strong>of</strong> architectures and new microelectronic<br />

devices.<br />

(a) (b)<br />

1 µm<br />

Figure 1: (a) AFM image <strong>of</strong> a silicon nanowire transistor that combines linear and circular<br />

regions and (b) output characteristics <strong>of</strong> the silicon nanowire transistor.<br />

[1] Y. Ciu, C. M. Lieber. Science 291, 851 (2001).<br />

[2] R. Garcia, R. V. Martinez, J. Martinez. Chemical Society Reviews 35, 29. (2006).<br />

[3] J. Martinez, R. V. Martinez, R. Garcia. Nano Letters 8, 3636, (2008).<br />

Contact author: Ricardo García, rgarcia@imm.cnm.csic.es<br />

126


Contacting self-ordered molecular wires by<br />

nanostencil lithography<br />

L. Gross, R.R. Schlittler, and G. Meyer<br />

IBM Research, Zurich Research Laboratory, 8803 Rüschlikon, Switzerland<br />

Th. Glatzel, S. Kawai, S. Koch, and E. Meyer<br />

Department <strong>of</strong> Physics, University <strong>of</strong> Basel, 4056 Basel, Switzerland.<br />

P.I-36<br />

We grew self-ordered cyanoporphyrin molecular wires [1] on thin epitaxial NaCl(100)<br />

layers on top <strong>of</strong> GaAs substrates under UHV conditions. This molecular assemblies form<br />

one- and two-dimensional wires with a length <strong>of</strong> several 10nm depending on the substrate<br />

conditions. Using a shadow masking technique [2], a nanostencil combined with a room<br />

temperature nc-AFM in UHV, we additionally evaporate Au and Cr electrodes with a<br />

thickness <strong>of</strong> around 3nm in situ. The resulting combined molecular and metal structures<br />

are investigated by means <strong>of</strong> nc-AFM and KPFM (Kelvin probe force microscopy).<br />

While nc-AFM enabled us to control the tip sample distance on the very complex and<br />

partly insulating surface, KPFM has been used to determine and compensate changes <strong>of</strong><br />

the local contact potential difference (LCPD).<br />

Figure 1: (a) nc-AFM image (size 1.7um x 1.7um, LCPD compensated) <strong>of</strong> NaCl/GaAs covered<br />

with cyanoporphyrin molecular wires and clusters. A Cr electrode was evaporated through a<br />

stencil mask in order to contact molecular wires. b) Simultaneously measured LCPD <strong>of</strong> the<br />

structure.<br />

[1] Th. Glatzel, L. Zimmerli, S. Koch, S. Kawai, and E. Meyer, Appl. Phys. Lett. 94, 063303 (2009).<br />

[2] P. Zahl, M. Bammerlin, G. Meyer, and R. R. Schlittler, Rev. Sci. Instrum. 76, 023707 (2005).<br />

127


Poster<br />

Session II<br />

Wednesday, 12 August<br />

128


Molecular Structures <strong>of</strong> Organic Single Crystals Investigated by<br />

Frequency Modulation <strong>Atomic</strong> <strong>Force</strong> Microscopes<br />

Taketoshi Minato 1 , Hiroto Aoki 2 , Thorsten Wagner 3 2, 4<br />

, and Kingo Itaya<br />

129<br />

P.II-01<br />

1<br />

Institute for International Advanced Interdisciplinary Research (IIAIR), Tohoku Univ., 6-6-07 Aoba,<br />

Aramaki, Aoba-ku, Sendai, Miyagi 980-8579, Japan<br />

2<br />

Department <strong>of</strong> Applied Chemistry, Tohoku Univ., 6-6-07 Aoba, Aramaki, Aoba-ku, Sendai, Miyagi 980-<br />

8579, Japan<br />

3<br />

University at Duisburg-Essen, Fachbereich Physics, Lotharstr. 1-21, 47057 Duisburg, Germany<br />

4<br />

World Premier International Research Center (WPI), Tohoku Univ., 6-6-07 Aoba, Aramaki, Aoba-ku,<br />

Sen-dai, Miyagi 980-8579, Japan<br />

Recently, single crystals <strong>of</strong> organic semiconductors such as rubrene [1],<br />

pentacene [2] and PTCDA [3] are attracting great interest as materials to be used for<br />

organic field effect transistors (OFETs). It is important to characterize geometric and<br />

electronic structures <strong>of</strong> organic layers in molecular levels in order to achieve higher<br />

carrier mobility. Here, we describe applications <strong>of</strong> Frequency Modulation <strong>Atomic</strong> <strong>Force</strong><br />

<strong>Microscopy</strong> (FM-AFM) for single crystal to investigate the molecular structure [4, 5].<br />

First, we measured FM-AFM images on<br />

prepared pentance single crystals in UHV. The<br />

obtained molecular resolved FM-AFM images<br />

showed very wide terrace (2 – 3 um) and high<br />

crystallity <strong>of</strong> prepared single crystal without<br />

molecular defects [2]. Also, we have recently<br />

achieved molecular resolution on a rubrene single<br />

crystal in UHV [5]. Fig. 1 shows a FM-AFM<br />

image in 5.0 nm × 5.0 nm. This also showed high<br />

crystallity <strong>of</strong> prepared single crystal without<br />

molecular defects as same as pentacene single<br />

crystals. Such results strongly suggest that the<br />

prepared organic single crytals are applicable to<br />

OFETs.<br />

Figure 1 A FM-AFM image<br />

(5.0 nm × 5.0 nm) <strong>of</strong> rubrene<br />

single crystal. Tsample = 298 K.<br />

References<br />

[1] R. W. I. de Boer, M. E. Gershenson, A. F.<br />

Morpurgo, V. Podzorov, Phys. Status Solidi A, 201, 1302 (2004).<br />

[2] V. Y. Butko, X. Chi, D. V. Lang, A. P. Ramirez, Appl. Phys. Lett. 83, 4773 (2003).<br />

[3] T. Wagner, A. Bannani, C. Bobisch, H. Karacuban, M. Stohr, M. Gabriel, R. Moller,<br />

Org. Elect., 5, 35 (2004).<br />

[4] K. Sato, T. Sawaguchi, M. Sakata, and K. Itaya, Langmuir, 23, 12788 (2007).<br />

[5] T. Minato, H. Aoki, T. Wagner, H. Fukidome and K. Itaya, J. Am. Chem. Soc.,<br />

submitted (2009).


Imaging <strong>of</strong> aromatic molecules by tuning-fork based LT-NC-AFM<br />

P.II-02<br />

Bartosz Such 1 , Thilo Glatzel 1 , Shigeki Kawai 1 , Sacha Koch 1 , Alexis Barat<strong>of</strong>f 1 , Ernst<br />

Meyer 1 , Catelijne H. M. Amijs 2 , Paula de Mendoza 2 , and Antonio M. Echavarren 2<br />

1 Department <strong>of</strong> Physics, University <strong>of</strong> Basel, Kilngelbergstrasse 82, 4056 Basel, Switzerland<br />

2 Institute <strong>of</strong> Chemical Research <strong>of</strong> Catalonia (ICIQ), Av. Països Catalans 16, 43007 Tarragona, Spain.<br />

Utilizing tuning forks [1,2] as sensors for NC-AFM allowed for extending<br />

capability <strong>of</strong> the technique to cryogenic temperatures.<br />

In the presentation, high resolution imaging <strong>of</strong> large aromatic molecules at 5K will<br />

be presented. Functionalized truxeses (10,15-dihydro-5H-diindeno[1,2-a;1',2'c]fluorine)<br />

with three 4-cyanobenzyl groups added in positions 5, 10, and 15 in a syn relative<br />

configuration have been evaporated onto a clean Cu(111) surface. The molecules adopt<br />

planar configuration and form a triangular network (see fig. 1). However, the<br />

cyanobenzyl groups are flexible and can adopt various configurations for neighbouring<br />

molecules. Supposedly the cyanobenzyl groups standing in the upright position appear in<br />

frequency shift images as dark spots (areas <strong>of</strong> larger interaction) due to enhanced<br />

electrostatic interaction with the tip, proving NC-AFM capability <strong>of</strong> imaging internal<br />

structures <strong>of</strong> aromatic molecules.<br />

Figure 1: 7.8x7.8 nm constant height image <strong>of</strong> truxene layer on Cu(111) surface. a) current (tip<br />

bias = -1.8 V, colour scale form 173 pA to 503 pA); b) frequency shift, Ap-p = 260 pm f0 = 26417,<br />

Q = 47000, colour scale from -56.5 Hz to -34.8 Hz.<br />

[1] F.J. Giessibl, Appl. Phys. Lett. 76, 1470 (2000).<br />

[2] M. Ternes et al. Science 319, 5866 (2008).<br />

130


One and Two Dimensional Structure <strong>of</strong> Water on Cu(110) and<br />

O/Cu(110)-(2x1) Surface<br />

Byoung Y. Choi 1 , Yu Shi 1,2 , Thomas Duden 3 , and Miquel Salmeron 1,2<br />

1 Material Science Division, Lawrence Berkeley National Laboratory, Berkeley, USA<br />

2 Department <strong>of</strong> Materials Science and <strong>Engineering</strong>, University <strong>of</strong> California, Berkeley, USA<br />

3 National Center for Electron <strong>Microscopy</strong>, Lawrence Berkeley National Laboratory, Berkeley, USA.<br />

P.II-03<br />

The interaction <strong>of</strong> water on metal surfaces like Ru, Pt, or Pd has been studied for a long<br />

time to understand its role in catalysis, electrochemistry and environmental science.<br />

Especially, the structure <strong>of</strong> water on atomically clean surfaces has been an important<br />

subject because <strong>of</strong> its complexity. Water molecules can easily diffuse and aggregate on<br />

the surface at over ~40K to form a two dimensional (2D) hexagonal structure on closepacked<br />

metallic surfaces [1] .<br />

For water on Cu, it has been reported that it forms 1D-like chains at low coverage on<br />

the (110) surface while it forms 2D hexagonal networks at higher coverage. It has been<br />

proposed that this 1D wire is built from hydrogen bonded hexagons as building blocks<br />

[2]. The 1D chain structure was recently suggested to be a combination <strong>of</strong> pentagons<br />

based on experiment and theory [3]. An important question is how the pentagon based<br />

chains evolve into hexagon based 2D networks. To determine this, more careful<br />

investigation at the atomic scale is required.<br />

<strong>Atomic</strong> force microscopy (AFM) in non-contact (NC) mode can be used to study the<br />

molecules on metals and insulators with atomic resolution. We built a tuning fork base<br />

NC-AFM which works in both 4K and 77K with a FET type current amplifier on the low<br />

temperature side to reduce a noise coupling through the signal lines. Our instrument can<br />

work in either the NC-AFM or STM mode, which provides complementary information<br />

and is useful for reference and calibration.<br />

With this instrument we investigated the evolution <strong>of</strong> water from 1D chain to 2D<br />

network on clean and oxygen precovered Cu(110) surfaces at low temperature, which<br />

reveals the growth mechanism <strong>of</strong> water at various coverages and growth temperatures.<br />

By dosing water on partially covered O/Cu(110)-(2x1) surface, we were able to show that<br />

the oxygen atoms at the edges <strong>of</strong> CuO rows play an important role in forming well<br />

ordered 2D hexagonal networks even at low coverage. Finally we suggest that water<br />

hexamer can be a base structure <strong>of</strong> both 1D and 2D islands when it grows adjacent to<br />

CuO row.<br />

[1] Angelos Michaelides and Karina Morgenstern, Nature Mater. 6, 597 (2007).<br />

[2] T. Yamada, S. Tamamori, H. Okuyama, and T. Aruga, Phys. Rev. Lett. 96, 036105 (2006)<br />

[3] Javier Carrasco, Angelos Michaelides, Matthew Forster, Sam Haq, Rasmita Raval and Andrew<br />

Hodgson, Nature Mater. doi:10.1038/nmat2403 (2009).<br />

131


P.II-04<br />

Dynamical simulations <strong>of</strong> truxene molecules adsorbed on the KBr (001)<br />

surface<br />

Thomas Trevethan 1,2 and Alexander Shluger 2<br />

1 Department <strong>of</strong> Physics and Astronomy, University College London, London, UK<br />

2 WPI Advanced Institute for Materials Research, Tohoku University, Sendai, Japan.<br />

We present the results <strong>of</strong> calculations performed to model the dynamical behavior <strong>of</strong><br />

truxene derivative molecules adsorbed on the KBr (001) surface, which have been<br />

imaged at room temperature with the NC-AFM with both molecular and atomic<br />

resolution. An atomistic potential model <strong>of</strong> the system was built from extensive quantum<br />

chemical calculations and used to investigate the room temperature mobility <strong>of</strong> the<br />

molecules on different surface sites at room temperature using molecular dynamics<br />

simulations. We find a hierarchy <strong>of</strong> rates <strong>of</strong> diffusion on different surface structures: the<br />

molecule is highly mobile on the perfect terrace and is bound to, but mobile along, the<br />

perfect monolayer step edge. The molecule is more strongly bound to double layer kinks,<br />

as observed by their immobilization at these sites in the experimental images; however<br />

the binding energies <strong>of</strong> the molecule on the two different polarity kinks are very similar.<br />

We show using dynamical simulations how one polarity kink results in a much higher<br />

residence time than the other, suggesting the identity <strong>of</strong> the kinks in the experimental<br />

images and hence <strong>of</strong> the identity <strong>of</strong> the sublattice.<br />

132


Temperature-dependent growth <strong>of</strong> C60 on CaF2(111)<br />

Felix Loske, Philipp Maaß, Jens Schütte, and Angelika Kühnle<br />

Department <strong>of</strong> Physics, Universität Osnarück, Barbarastraße 7, 49076 Osnabrück ,Germany<br />

E-Mail: floske@uos.de<br />

P.II-05<br />

Non-contact atomic force microscopy was employed to study the system <strong>of</strong> C60<br />

molecules on the CaF2(111) surface in-situ at various temperatures. The molecules were<br />

observed to be very mobile on the surface at room temperature (RT), as they nucleate at<br />

step edges and form large islands on the terraces. Both, regular triangular islands as well<br />

as branched structures were observed to coexist (Fig. 1b and 1c). The branched structures<br />

were seen to change shape during measuring. S. Burke et al. have previously reported<br />

about such branched structures on alkali halides [1].<br />

The compact islands <strong>of</strong> C60 on CaF2 are at least two layer high, and at higher coverages<br />

C60 molecules grow in a dendritic manner onto the these compact C60 islands (Fig. 1d). In<br />

contrast, at low temperatures (LT) hexagonal islands are predominant, which are initially<br />

only one layer high. Here, already the second layer is growing in a dendritic manner (Fig.<br />

1a).<br />

These temperature-dependent measurements allow to gain insight into the molecular<br />

dynamics <strong>of</strong> this system and to specify the diffusion barrier. Microscopic growth models<br />

are shortly addressed to explain the measured island morphologies.<br />

Figure 1: Topographic NC-AFM images. (a) C60 island with dendritic second layer at LT. (b)<br />

and (c) represent the coexistence <strong>of</strong> compact triangular and branched islands at RT, with dendritic<br />

growth starting at the second layer (d).<br />

[1] S.A. Burke, J.M. Mativetsky, S. Fostner, P. Grütter, Phys. Rev. B 76, 035419 (2007)<br />

133


134<br />

P.II-06<br />

Creating 1D nanostructures: Heptahelicene-carboxylic acid on Calcite<br />

Philipp Rahe 1 , Markus Nimmrich 1 , Jens Schütte 1 , Irena G. Stara 2 and Angelika Kühnle 1<br />

1 Fachbereich Physik, Universität Osnabrück, Barbarastraße 7, 49076 Osnabrück, Germany<br />

2 Institute <strong>of</strong> Organic Chemistry and Biochemistry, Flemingovo nám. 2, 166 10 Prague 6, Czech Republic<br />

Creating nanostructures on a molecular level has received an exceptional level by<br />

depositing molecules on metal surfaces. Structures with two-, one- or quasi zerodimensionality<br />

have successfully been created [1,2]. When deposited onto insulating<br />

substrates, one-dimensional nanostructures may act as wires in future molecular<br />

electronic devices. Recently, first results <strong>of</strong> such structures on insulating surfaces have<br />

been presented, binding molecules to step edges [3] or forming wires consisting <strong>of</strong><br />

several molecules in width [4].<br />

Here, we present a study <strong>of</strong> heptahelicene-carboxylic acid adsorbed on the calcite<br />

(10-14) surface. The molecules are sublimated in-situ from a glass crucible. Both,<br />

molecule deposition and NC-AFM imaging are performed in ultra-high vacuum. As<br />

presented in Fig 1, we observe single and double rows, which align along the [010]<br />

substrate direction. These structures are not bound to step edges. Rows are formed at<br />

room temperature and stabilize presumably by hydrogen bonds between the carboxylic<br />

end groups forming pairs as well as by π-π stacking along the molecular row.<br />

[1] J. Barth et al. Annu Rev Phys Chem 58, 375 (2007).<br />

[2] A. Kühnle. Curr Op Coll Interf Sci 14, 157 (2009).<br />

[3] S. Maier et al. Small 4, 1115 (2007).<br />

[4] M. Fendrich et al. Appl Phys Lett 91, 023101 (2007).<br />

Figure 1: Frequency shift image taken at<br />

room temperature. Rows along the [010]<br />

substrate direction are presented: Single<br />

rows (solid circle) and double rows (dashed<br />

circle). Adjacent to the double rows, further<br />

single rows may attach as marked by a white<br />

triangle.


P.II-07<br />

<strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong> Study <strong>of</strong> Cross-Linked C32H66 Monolayer by<br />

Low-Energy (10eV) Hyperthermal Bombardment<br />

Y. Liu 1 , H.Y. Nie 2 , D.Q. Yang 2 , M.W. Lau 2 and J. Yang 1<br />

1 Department <strong>of</strong> Mechanical and Materials <strong>Engineering</strong>, University <strong>of</strong> Western Ontario, Ontario, Canada<br />

2 Surface Science Western, University <strong>of</strong> Western Ontario, Ontario, Canada<br />

Email: yliu452@uwo.ca<br />

Low-energy (10eV) hydrogen projectiles generated in a special production prototype<br />

reactor are being developed in our study to only break the C-H bonds with other bonds<br />

intact on C32H66 monolayer, as spin cast on a silicon wafer. The generation <strong>of</strong> free carbon<br />

radicals leads to neighboring cross-link and form covalent C-C bonds [1]. The newly<br />

formed surface is characterized by a combination <strong>of</strong> XPS, optical contact angle<br />

measurement and dynamic atomic force microscopy (AFM). Especially, AFM study is<br />

focused on to correlate with the results from other two techniques. Roughness AFM<br />

results demonstrate a temporal behavior <strong>of</strong> C32H66 monolayer surface modification,<br />

corresponding to bombardment time which is proportional to the influence during<br />

bombardment process. The transiting roughness measurements can also interpret the<br />

directional properties <strong>of</strong> the formed covalent C-C bonds.<br />

The controllability <strong>of</strong> the surface modification is additionally demonstrated by the<br />

evolution <strong>of</strong> hydrophibilicity <strong>of</strong> the monolayer as measured by the aid <strong>of</strong> AFM and<br />

contact angle measurements. Phase image by tapping mode AFM provides necessary<br />

contrast on other mechanical properties and/or adhesion energy to evaluate the untreated<br />

and treated areas on an incompletely bombarded monolayer surface [2]. The results are<br />

further associated with the measurements from force modulation and torsion resonance<br />

modes to provide a critical bombardment time essential to carry on a complete<br />

bombardment and guarantee the surface homogeneity. All <strong>of</strong> the results confirm crosslinking<br />

C-C bonds as formed after bombardment and explain the enhanced surface and<br />

mechanical properties through the developed hyperthermal bombardment.<br />

[1] Z. Zheng, X.D. Xu, X.L. Fan, W.M. Lau, and R.W.M. Kwok, J. Am. Chem. Soc. 126, 12336 (2004)<br />

[2] J. 1. Tamayo, J. and R. Garcia, Appl. Phys. Lett., 71, 2394 (1997).<br />

135


P.II-08<br />

Towards a molecule-based Ferroelectric-OFET: surface modification <strong>of</strong><br />

PZT mediated through functionalized thiophene derivates<br />

Peter Milde 1 , Kinga Haubner 2 , Evelyn Jaehne 2 , Denny Köhler 1 , Ulrich Zerweck 1 , and<br />

Lukas M. Eng 1<br />

1 Department <strong>of</strong> Applied Photophysics, TU Dresden, Dresden, Germany<br />

2 Institute <strong>of</strong> Macromolecular Chemistry and Textile Chemistry, TU Dresden, Dresden, Germany<br />

Organic field effect transistors (OFETs) and their application have become a field <strong>of</strong><br />

intense research due to significant advances in molecular design and integration [1]. For<br />

an OFET-design whith a gate “electrode” that is made out <strong>of</strong> a ferroelectric (FE), we may<br />

expect even more functionality due to the strong and remanent electric field arising from<br />

bound surface charges at the FE/molecule interface [2]. In order to achieve excellent<br />

electric transport properties, a high degree <strong>of</strong> intermolecular ordering is inevitable [3].<br />

In our approach, lead zirconate titanate (PZT) is used as material <strong>of</strong> choice for designing<br />

an ultrathin ferroelectric gate electrode in a Ferroelectric-OFET. The focus <strong>of</strong> the present<br />

work is on the film formation process <strong>of</strong> the molecularly thin organic conduction layer<br />

based on α,ω-dicyano-β,β*-dibutylquaterthiophene (DCNDBQT). Film formation is<br />

effectively promoted through specifically designed, bifunctional self-assembling<br />

molecules (CNBTPA: 5-cyano-2-(butyl-4-phosphonic acid)-3butylthiophene) which act<br />

as template layer (see Fig. 1). We report on nc-AFM and KPFM [4] investigation <strong>of</strong> the<br />

template layer's structural and electronical properties.<br />

D<br />

Org.<br />

DCNDBQT<br />

Ferroelectric Gate<br />

Conductive Gate<br />

Figure 1: Schematic <strong>of</strong> the proposed molecular OFET design. The CNBTPA template layer<br />

promotes both bonding <strong>of</strong> the organic semiconductor onto the PZT substrate and the self<br />

assembling <strong>of</strong> the DCNDBQT layer.<br />

[1] see for instance: phys. stat. solidi A 205(3) (2008)<br />

[2] S. Gemming, R, Luschtinetz, W. Alsheimer, G. Seifert, Ch. Loppacher, and L.M. Eng, Journal <strong>of</strong><br />

Computer-Aided Materials Design 14(S1), 211 (2007)<br />

[3] K. Haubner, E. Jähne, H.-J.P. Adler, D. Köhler, Ch. Loppacher, L.M. Eng, J. Grenzer, A.<br />

Herasimovich, and S. Scheinert, phys. stat. solidi A 205(3), 430 (2008)<br />

[4] U. Zerweck, Ch. Loppacher, T. Otto, S. Grafström, and L.M. Eng, Phys. Rev. B 71, 125424 (2005)<br />

S<br />

136


Imaging and Detection <strong>of</strong> Single Molecule Recognition Events on<br />

Organic Semiconductor Surfaces<br />

P.II-09<br />

N. S. Losilla 1 ,J. Preiner 2 , , A. Ebner 2 , P. Annibale 3 , F. Biscarini 3 , R. Garcia 1 and P.<br />

Hinterdorfer 2<br />

1 Instituto de Microelectrónica de Madrid, CSIC, Tres Cantos, Spain<br />

2 Johannes Kepler University <strong>of</strong> Linz, Austria<br />

3 CNR-Istituto per lo Studio dei Materiali Nanostrutturati (ISMN),Italy<br />

The combination <strong>of</strong> organic thin film transistors and biological molecules could open<br />

new approaches for the detection and measurement <strong>of</strong> properties <strong>of</strong> biological entities. To<br />

generate specific addressable binding sites on such substrates, it is necessary to determine<br />

how single biological molecules, capable <strong>of</strong> serving as such binding sites behave upon<br />

attachment to semiconductor surfaces. Here, we use a combination <strong>of</strong> high-resolution<br />

atomic force microscopy topographical imaging and single molecule force spectroscopy<br />

(TREC), to study the functionality <strong>of</strong> antibiotin antibodies upon adsorption on pentacene<br />

islands, using biotin-functionalized, magnetically coated AFM tips. The antibodies could<br />

be stably adsorbed on the pentacene, preserving their functionality <strong>of</strong> recognizing biotin<br />

over the whole observation time <strong>of</strong> more than one hour. We have resolved individual<br />

antigen binding sites on single antibodies for the first time. This highlights the resolution<br />

capacity <strong>of</strong> the technique.<br />

Figure 1: (a) Scheme <strong>of</strong> the simultaneous topography and recognition AFM imaging <strong>of</strong> antibiotin<br />

antibodies adsorbed on pentacene islands. A magnetically driven cantilever oscillates across the<br />

surface. The oscillation signal is split into two parts <strong>of</strong> which the lower part is used to generate<br />

the topography (topography signal, red) and the upper part is used for the recognition image<br />

(recognition signal, green).(b) Antibody molecule on pentacene with its expected size and shape.<br />

[1] J. Preiner et al. Nano Lett., , 9, 571( 2009).<br />

a) b)<br />

137<br />

10 nm<br />

100 nm


P.II-10<br />

Transverse conductance image <strong>of</strong> DNA probed by current-feedback<br />

noncontact AFM<br />

T. Matsumoto 1, Y. Maeda 2, and T Kawai 1<br />

1 Institute <strong>of</strong> Scientific and Industrial Research (ISIR), Osaka Universit,Osaka, Japan<br />

2 National Institute <strong>of</strong> Advanced Industrial Science and Technology (AIST), Osaka Japan<br />

Scanning probe microscopy (SPM), which provides means to access to individual<br />

molecule, has a special significance for investigations <strong>of</strong> molecular-scale electronics. In<br />

particular, scanning tunneling microscopy (STM) has been used to evaluate molecular<br />

conductivity. However, STM measurement is effective only for height-foreknown system<br />

such as the molecules embedded in self-assembled monolayer. For the molecules having<br />

conformational variations, STM measurement is unable to separate the information <strong>of</strong><br />

transverse conduction through the molecules from the unknown height variation.<br />

We demonstrated that simultaneous measurement <strong>of</strong> tunnel current and frequency<br />

shift realized to separate electrical properties from topographic variations. Figure 1(a)<br />

shows simultaneously obtained STM topography (constant-current mode) and frequency<br />

shift image <strong>of</strong> DNA on Cu (111) surface. These images are totally similar to each other<br />

but the contrast distribution <strong>of</strong> these images shows differences. For the quantitative study<br />

<strong>of</strong> these variances, the correlation between STM height and frequency shift is analyzed by<br />

plotting each pixel as shown in Figure 2(a). The negative frequency shift overall increase<br />

as increasing STM height. This indicates that the tip close to the molecule at the high<br />

region <strong>of</strong> STM topography. Assuming the transverse electron tunneling through DNA<br />

molecules, frequency shift and STM height show linear relationship as a function <strong>of</strong><br />

transconductance G0 at molecule/surface interface and attenuation decay constant β. The<br />

line represented in Figure 2(a) means the data sets <strong>of</strong> STM height and frequency shift<br />

satisfying a parameter set <strong>of</strong> G0=0.27 and β= 0.26 . The pixel data picked from the plots<br />

around this line is used for reconstructing an image which shows a slice indicating the<br />

conduction with G0=0.27 and β= 0.26. As shown in Figure 2(b), the conductance image<br />

differ from either STM topography and frequency shift image in that it reflects the<br />

conductance variations affected by molecular conformation and/or molecules/substrate<br />

interface.<br />

STM topography Frequency shift image<br />

Fig. 1. Simultaneously obtained STM<br />

topography and frequency shift image<br />

<strong>of</strong> DNA on Cu (111).<br />

(a) Correlation analysis (b) Conductance image<br />

Frequency shift (Hz)<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

(G 0 ,β ) = (0.26, 0.27)<br />

-8<br />

-6 -4 -2 0 2 4 6 8<br />

STM height (nm)<br />

Fig. 2. (a) Correlation analysis between STM height<br />

and frequency. (b) Conductance image reconstructed<br />

from the data around the line <strong>of</strong> G0=0.27, β= 0.26.<br />

138


P.II-11<br />

Imaging Schwann Cell NGF Receptors using <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong><br />

Ryan Williamson and Cheryl Miller<br />

Department <strong>of</strong> Biomedical <strong>Engineering</strong>, Saint Louis University, St Louis, MO, USA.<br />

Nerve growth factor (NGF) is a necessary neurotrophic agent that promotes neural<br />

survival and proliferation. Production <strong>of</strong> NGF by Schwann cells is essential for<br />

successful nerve regeneration. During neural axotomy and the resulting Wallerian<br />

degeneration, Schwann cells increase proliferation while axons and their myelin sheaths<br />

are degraded. The resulting formation, the band <strong>of</strong> Bungner, is crucial for guidance <strong>of</strong><br />

axon sprouts which form during regeneration. Schwann cells in the distal axon will<br />

express the high affinity NGF receptor tyrosine kinase A (TrkA) selectively in the bands<br />

<strong>of</strong> Bungner as well as the low affinity receptor p75. Using force measurements taken<br />

with a modified atomic force microscopy (AFM) tip to detect binding events, NGF<br />

receptor locations were identified. Using AFM with Schwann cells, we investigated the<br />

expression and conformation <strong>of</strong> NGF receptors. Receptor location and change during<br />

axon-Schwann cell contact could explain Schwann cell role during regeneration and<br />

possible clinical solutions.<br />

139


Comparative Studies on Water Structures on Hydrophilic and<br />

Hydrophobic Surfaces by FM-AFM<br />

Kazuhiro Suzuki 1 , Noriaki Oyabu 1,2 , Kei Kobayashi 3 , Kazumi Matsushige 1 ,<br />

and Hir<strong>of</strong>umi Yamada 1 .<br />

1 Department <strong>of</strong> Electronic Science and <strong>Engineering</strong>, Kyoto University, Kyoto, Japan<br />

2 Japan Science and Technology Agency /Adv. Meas. and Analysis, Japan<br />

3 Innovative Collaboration Center, Kyoto University, Kyoto, Japan<br />

P.II-12<br />

Water structures play essential roles in various biochemical functions such as selfassembly<br />

<strong>of</strong> proteins and molecular recognition between ligands and receptors. Recently<br />

we succeeded in three-dimensional molecular-scale visualization <strong>of</strong> water structures on<br />

hydrophilic surfaces including a mica surface and a purple membranes using frequency<br />

modulation atomic force microscopy (FM-AFM) [1].<br />

In this study water structures on a hydrophobic highly oriented pyrolytic graphite<br />

(HOPG) surface were investigated using FM-AFM. The results were compared with<br />

those obtained on the hydrophilic mica surfaces. Figure 1 shows frequency shift-distance<br />

curves measured on the HOPG surface (a) and on the mica surface (b). The oscillatory<br />

behavior, originating from the water structure on each surface, can be observed in each<br />

curve. However, there is a clear difference in the number <strong>of</strong> the water layers between the<br />

two curves, which reflects the difference in the hydrophilicities <strong>of</strong> the two surfaces. In<br />

addition, a possible effect <strong>of</strong> the hydration structure on the AFM tip surface was<br />

investigated by using a chemically modified tip with a self-assembled monolayer film.<br />

Figure 1: Frequency shift-distance curves measured on a hydrophobic HOPG surface (a) and on<br />

a hydrophilic mica surface (b). Arrows indicate the peaks produced by the water layers.<br />

[1] K. Kimura et al, The 11 th International Conference on Non-Contact <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong>,<br />

Madrid 2008, Oral Presentation Th-1200.<br />

140


Molecular Resolution Investigation <strong>of</strong> Lysozyme Crystal<br />

in Liquid by Frequency-Modulation AFM<br />

P.II-13<br />

K. Nagashima 1,2 , M. Abe 1,2 , S. Morita 1,2 , N. Oyabu 2,3 , K. Kobayashi 2,3,4 , H. Yamada 2,3 ,<br />

R. Murai 1 , H. Adachi 1,5 , K. Takano 1,5 , H. Matsumura 1,5 , S. Murakami 5,6 , T. Inoue 1,5 , Y.<br />

Mori 1,5 , M. Ohta 2,7 , R. Kokawa 2,7<br />

1Grad. <strong>School</strong> <strong>of</strong> Eng., Osaka Univ.; 2 SENTAN, JST; 3 Electronic Sci. & Eng., Kyoto Univ.; 4 Innovative<br />

Collaboration Center, Kyoto Univ.; 5 SOSHO Inc.; 6 Grad. <strong>School</strong> <strong>of</strong> Biosci. and Biotech., Tokyo Insti. <strong>of</strong><br />

Tech.; 7 Shimadzu Corporation<br />

Recently, Frequency-Modulation AFM (FM-AFM) in liquids has opened a new<br />

way to direct visualization with atomic or molecular resolution [1,2]. We have developed<br />

a high-resolution FM-AFM working in liquids based on a commercially available AFM<br />

(Shimadzu: SPM-9600), which obtained atomic or molecular resolution in liquid [3].<br />

We have tried to observe protein crystals in liquid. To make high quality protein<br />

crystals is very important for determining its molecular structure by X-ray diffraction<br />

analysis. We observed (110) face <strong>of</strong> tetragonal lysozyme crystal in saturated solution. A<br />

surface unit cell (11.2 x 3.8 nm) consists <strong>of</strong> 4 molecules with different orientations from<br />

one another (Fig. 1a). We observed it for the first time at molecular resolution (Fig. 1b),<br />

which was not achieved in contact mode AFM [4]. In addition, we also succeeded to<br />

observe changes <strong>of</strong> the terrace with admolecule and the point defect when the terrace was<br />

covered with a single unit layer (height = 5.6 nm) in supersaturated solution. Such<br />

observations are considered to be fruitful technique for investigating growth mechanism<br />

at molecular level.<br />

a) b)<br />

c c<br />

5 nm<br />

Figure 1: (a) The molecular-packing arrangement <strong>of</strong> the tetragonal lysozyme (110) face. The<br />

surface unit cell shows black rectangle. (b) FM-AFM image <strong>of</strong> tetragonal lysozyme (110) face.<br />

[1] T. Fukuma, K. Kobayashi, K. Matsushige, H. Yamada, Appl. Phys. Lett. 86 (2005) 193108.<br />

[2] T. Fukuma, K. Kobayashi, K. Matsushige, H. Yamada, Appl. Phys. Lett. 87 (2005) 34101.<br />

[3] S. Rode, N. Oyabu, K. Kobayashi, H. Yamada, A. Kuhnle, Langmuir 25 (2009) 2850.<br />

[4] H. Li, M. A. Perozzo, J. H. Konnert, A. Nadarajaha, M. L. Pusey, Acta Cryst. D55 (1999) 1023.<br />

141


<strong>Noncontact</strong> <strong>Atomic</strong> <strong>Force</strong> Microscope Observation<br />

<strong>of</strong> TiO2(110) Surface in Pure Water<br />

Akira Sasahara 1 , Yonkil Jeong 1 , and Masahiko Tomitori 1<br />

1 <strong>School</strong> <strong>of</strong> Materials Science, Japan Advanced Institute <strong>of</strong> Science and Technology, Ishikawa, Japan<br />

P.II-14<br />

Titanum dioxide (TiO2) has a wide range <strong>of</strong> industrial applications such as pigments,<br />

optical devices, catalyst supports, photocatalysts, and photoelectrodes. A rutile<br />

TiO2(110) surface, which exhibits a simple truncation <strong>of</strong> the bulk structure (Fig. 1(a)) by<br />

simple cleaning procedures, has been most frequently employed in ultra high vacuum<br />

(UHV) studies to understand the origin <strong>of</strong> its properties. In many <strong>of</strong> the applications,<br />

TiO2 demonstrates the excellent properties in the presence <strong>of</strong> aqueous media. Nanoscale<br />

investigation <strong>of</strong> a well-defined rutile TiO2(110) surface in a liquid possibly provides<br />

further insight into the surface chemical properties <strong>of</strong> TiO2. We attempted noncontact<br />

atomic force microscope (NC-AFM) observation <strong>of</strong> TiO2(110) surface in pure water.<br />

The experiments were performed by using an intermittent contact mode atomic force<br />

microscope (5500 AFM/SPM, Agilent Technologies). The microscope was operated as<br />

an NC-AFM by using a phase locked loop detector (easy PLL plus, Nanosurf AG). The<br />

microscope was placed in a glass chamber, and imaging was performed under a flow <strong>of</strong><br />

Ar. The TiO2(110) surface was cleaned by cycles <strong>of</strong> Ar + sputtering and annealing in<br />

UHV, removed from the vacuum chamber, and immersed in Ar-purged Milli-Q water.<br />

Figure 1(b) shows an NC-AFM image <strong>of</strong> the surface in water. Strings elongated to the<br />

[001] direction were observed. Along cross section 1 in Fig. 1(c), subnanometer<br />

corrugation was observed perpendicular to the rows. The distance between the lowest<br />

rows, indicated by arrowheads, was approximately 0.65 nm. On some <strong>of</strong> the rows,<br />

periodic corrugation was observed as shown in the cross sections 2 and 3. We attributed<br />

the lowest rows to the oxygen atom rows, and the corrugation along the rows to H2O<br />

clusters.<br />

Figure 1: (a) Ball model <strong>of</strong> a rutile TiO2(110)-(1×1) surface. Size <strong>of</strong> the unit cell is 0.65<br />

nm×0.30 nm. (b) NC-AFM image <strong>of</strong> the TiO2(110) surface in pure water (15×15 nm 2 ).<br />

Frequency shift = +150 Hz, sample bias voltage = 0 V, peak-to-peak amplitude <strong>of</strong> the cantilever<br />

oscillation = 2 nm. (c) Cross sections obtained along the lines in the image.<br />

142


Development <strong>of</strong> Multifrequency High-speed NC-AFM in Liquid<br />

Y. J. Li 1,2 , K. Takahashi 1 , N. Kobayashi 1 , Y. Naitoh 1,2 , M. Kageshima 1,2<br />

and Y. Sugawara 1,2<br />

1 Department <strong>of</strong> Applied Physics, Osaka University and 2 CREST, Japan<br />

P.II-15<br />

AFM is a power tool to observe the solid-liquid interface, and it has succeeded in true<br />

atomic-resolution imaging on the nano-scale. Especially, the high-speed imaging is useful<br />

for understanding the dynamic behavior <strong>of</strong> biomolecules and the clarification <strong>of</strong> the<br />

physical phenomena that happen in the solid-liquid interface. In order to clarify unknown<br />

physical phenomena on the surface, not only the topography image and energy<br />

dissipation image but also elasticity including the information <strong>of</strong> surface physical<br />

properties is also necessary.<br />

Previously, we have succeeded in simultaneous imaging topography and energy<br />

dissipation image by our developed high-speed phase-modulation AFM (PM-AFM) in<br />

constant-amplitude (CA) mode [1-2]. Recently, Garcia et al have developed a theory <strong>of</strong><br />

the multifrequency method in AM-AFM [3]. So far, the multifrequency method for<br />

measuring elasticity by high-speed PM-AFM in CA mode has not been developed yet.<br />

In this research, we develop the multifrequency high-speed PM-AFM in CA mode with<br />

the capability <strong>of</strong> simultaneous imaging elasticity in liquid. From the theory and the<br />

experiment, we discuss the multifrequency method in high-speed PM-AFM in CA mode.<br />

The high-speed elasticity image can be measured by phase shift Ф2from high-speed<br />

phase detector with the bandwidth <strong>of</strong> 3MHz (the second resonance mode). We<br />

demonstrate the multifunctional high-speed images <strong>of</strong> polymer film with the scan<br />

speed <strong>of</strong> 10 frames /s in water.<br />

(a) (b) (c)<br />

Figure 2 (a) topographic, (b) energy dissipation and (c) elasticity images <strong>of</strong> polymer film<br />

with 10frames/s. 300x300nm 2 . f1 =600 kHz, f2 =3MHz.<br />

[1] N. Kobayashi et, al., Jpn. J. Appl. Phys., 45 (2006) L793.<br />

[2] Y. Sugawara, et, al., Appl. Phys. Lett., 90 (2007) 194104.<br />

[3] J. R. Lozano and R. Garcia, Phys. Rev. Lett., 100 (2008) 076102.<br />

143


P.II-16<br />

Frequency-Domain and Time-Domain Analyses <strong>of</strong> S<strong>of</strong>t-Matter<br />

Dynamics Using Wide-Band Magnetic Excitation AFM<br />

Masami Kageshima, Tatsuya Ogawa, Shinkichi Kurachi, Yoshitaka Naitoh, Yan Jun Li,<br />

and Yasuhiro Sugawara<br />

Department <strong>of</strong> Applied Physics, Osaka University, Suita, Osaka, Japan.<br />

Dynamics analysis <strong>of</strong> nanometer-scale s<strong>of</strong>t matter systems is essentical in<br />

understanding mechanical properties <strong>of</strong> polymer systems, function <strong>of</strong> biological<br />

molecules, interaction <strong>of</strong> biological systems with water molecules, etc. There are two<br />

major experimentally accessible quantities to represent viscoelastic properties <strong>of</strong> a system,<br />

a frequency response function and a step response function, each <strong>of</strong> which can be<br />

understood as a frequency-domain and a time-domain measurement, respectively.<br />

Bothapproaches can be implemented with AFM by applying a sinusoidal force with a<br />

varied frequency or a step force onto the AFM cantilever, given that the excitation device<br />

posesses a sufficient bandwidth. Here the magnetic excitation technique was extended so<br />

that a supurious-free and well-characterized cantilever excitation force can be applied<br />

onto the cantilever beyond 1 MHz [1]. As an example <strong>of</strong> a frequency-domain<br />

measurement, frequency-dependent viscoelasticity <strong>of</strong> a single dextran chain was already<br />

presented [1]. Alternatively, a step force can be applied to the cantilever using the same<br />

excitation system. By driving a coreless electromagnet with a wide-band constant-current<br />

circuit (Fig. 1), a current step <strong>of</strong> 1 A can be settled in ca. 500 ns (Fig 2(a)). Since the<br />

cantilever deflection makes a ringing motion mainly due to its fundamental resonance<br />

even in liquid, an active feedback (Q-control) circuit is employed to suppress the<br />

resonance (Fig. 2(b)). Results <strong>of</strong> analysis <strong>of</strong> a single polymer chain and water molecules<br />

interacting with a hydrophilic surface will be presented.<br />

Input<br />

Sinusoidal<br />

Step<br />

+<br />

OP amp.<br />

-<br />

HV amp.<br />

-<br />

Electromagnet<br />

Differential amp.<br />

Fig. 1: Constant-current circuit and electromagnet.<br />

[1] M. Kageshima, T. Chikamoto, T. Ogawa, Y.<br />

Hirata, T. Inoue, Y. Naitoh, Y. J. Li, and Y.<br />

Sugawara, Rev. Sci. Instrum. 80 (2009) 023705.<br />

+<br />

Cantilever<br />

Magnet<br />

144<br />

Iuput (V)<br />

(a)<br />

0.08<br />

0.03<br />

-0.02<br />

-1 0 1<br />

(b) Time (microsec.)<br />

input<br />

deflection 3 nm<br />

0.4<br />

-0.1<br />

-0.6<br />

Current (A)<br />

0 1 2 3 4 5<br />

Time (msec.)<br />

Fig.2: (a) Current pr<strong>of</strong>ile and (b) Qsuppressed<br />

cantilever deflection in water.


<strong>Noncontact</strong> observation in liquid with van der Pol-type FM-AFM<br />

M. Kuroda 1 , H. Yabuno 2 , T. Someya 3 , R. Kokawa 4 , and M. Ohta 4<br />

1 National Institute <strong>of</strong> Advanced Industrial Science and Technology (AIST), Tsukuba, Japan<br />

2 Department <strong>of</strong> Mechanical <strong>Engineering</strong>, Keio University, Yokohama, Japan<br />

3 Mitsubishi Heavy Industries Ltd., Hiroshima, Japan<br />

4 Analytical & Measuring Instruments Division, Shimadzu Corp., Kyoto, Japan<br />

<strong>Atomic</strong> force microscopy (AFM) is crucial for nanobiotechnology<br />

studies. Observing bio-related samples<br />

in liquids using AFM is important, but deformable,<br />

uneven, and easily damaged surfaces <strong>of</strong> such<br />

specimens require non-contact AFM observation.<br />

For probe-cantilever excitation, the eigenfrequency<br />

must be estimated based on frequency response<br />

characteristics for the external excitation method. It<br />

cannot estimate the probe cantilever’s eigenfrequency<br />

precisely because <strong>of</strong> many spurious peaks (Fig. 1). But,<br />

the self-excitation method needs fine adjustment <strong>of</strong> the<br />

linear feedback gain near the oscillation limit by an<br />

automatic gain controller (AGC) to prevent the probe<br />

cantilever from touching the sample surface:<br />

oscillation can easily halt, disabling the observation.<br />

In solving those problems, van der Pol-type (vdP)<br />

self-excited oscillation represents a positive use <strong>of</strong><br />

nonlinearity [1–2]. Along with conventional positive<br />

linear velocity feedback for self-excited oscillation,<br />

vdP self-excited oscillation is realized by adding<br />

nonlinear feedback proportional to the squared<br />

deflection times the velocity. With the eigenfrequency<br />

component alone (Fig. 2), vdP self-excited oscillation<br />

guarantees existence <strong>of</strong> a stable stationary amplitude.<br />

Increased nonlinear feedback gain reduces the response<br />

amplitude but retains the high gain linear feedback.<br />

A 19-nm-high SiN4 surface pattern with 3 μm pitch<br />

was observed in pure water using vdP-AFM (Fig. 3).<br />

The smaller probe-cantilever vibration amplitude than<br />

the probe-cantilever – sample gap verifies noncontact<br />

observation. The vdP-AFM takes sample surface<br />

images with equal accuracy to that <strong>of</strong> contact-mode<br />

AFM. Even in liquid, oscillation continues.<br />

[1] H. Yabuno et al. Nonlinear Dynamics 54, 137 (2008).<br />

[2] M. Kuroda et al. Journal <strong>of</strong> System Design and Dynamics 2-3, 886 (2008).<br />

145<br />

P.II-17<br />

Figure 1: Frequency response<br />

curve in pure water (External<br />

excitation method)<br />

f=54.4 kHz<br />

Figure 2: Power spectrum in pure<br />

water (vdP self-excited oscillation<br />

method)<br />

Figure 3: Sample observed in pure<br />

water by vdP FM-AFM


P.II-18<br />

Development <strong>of</strong> a NC – AFM for Ambient and Liquid Environments<br />

Haider I. Rasool 1 , Shivani Sharma 1 , James K. Gimzewski 1,2,3<br />

1 Department <strong>of</strong> Chemistry and Biochemistry, University <strong>of</strong> California – Los Angeles, USA<br />

2 California NanoSystems Institute, 570 Westwood Plaza, Los Angeles, USA.<br />

3 International Center for Materials Nanoarchitectonics (MANA), Tsukuba, Japan<br />

High resolution Non – Contact <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong> imaging has been<br />

achieved by various groups in both ambient and liquid environments [1] – [3]. The recent<br />

success <strong>of</strong> NC – AFM in these “low – Q” environments has been attributable to the<br />

development <strong>of</strong> low noise detection schemes for measuring small amplitude deflections.<br />

We describe current research on the development <strong>of</strong> a robust home – built scanning probe<br />

microscope with an incorporated low noise all fiber interferometer deflection sensor for<br />

imaging in both ambient and liquid environments. The current design <strong>of</strong> the scanning<br />

probe microscope in its AFM configuration is depicted in Figure 1 (a).<br />

The developing instrument has been used to image various biological species in<br />

both ambient and liquid environments in different imaging modes. In particular, our<br />

group has been successful in imaging isolated human saliva exosomes fixed to a mica<br />

substrate in both air and buffer solution. Both amplitude and phase modulation imaging<br />

have been accomplished on these samples under ambient conditions. Using constant<br />

excitation phase modulated imaging, a simultaneous mapping <strong>of</strong> topography and energy<br />

dissipation has been possible at a constant average tip sample force. Figure 1(b) shows a<br />

typical image <strong>of</strong> the topography <strong>of</strong> an isolated exosome.<br />

a) b)<br />

Figure 1: The above images show (a) the design <strong>of</strong> the home built microscope and (b) a PM –<br />

AFM image <strong>of</strong> a human saliva exosome.<br />

[1] T. Fukuma, M. Kimura, K. Kobayashi, K. Matsushige, and H. Yamada. Rev. Sci. Instrum. 76, 053704<br />

(2005).<br />

[2] B.W. Hoogenboom, H.J. Hug, Y. Pellmont, S. Martin, P.L.T.M. Frederix, D. Fotiadis,and A. Engel.<br />

Appl. Phys. Lett. 88, 193109 (2006).<br />

[3] T. Fukuma, and S. Jarvis. Rev. Sci. Instrum. 77, 043701 (2006).<br />

146


P.II-19<br />

Cantilever Holder Design for Spurious-Free Cantilever Excitation in<br />

Hitoshi Asakawa 1 and Takeshi Fukuma 1,2<br />

Liquid by Piezoactuator<br />

1 Frontier Science Organization, Kanazawa University, Japan<br />

2 PRESTO, JST, Japan<br />

Recent advancement in frequency modulation atomic force microscopy (FM-AFM)<br />

has enabled us to obtain atomic and molecular resolution images in liquid, which has<br />

encouraged us to explore the possibilities <strong>of</strong> FM-AFM application in biological research.<br />

In liquid-environment FM-AFM, a method for cantilever excitation is particularly<br />

important for ensuring the stability and accuracy in imaging and quantitative force<br />

measurements. This is because FM-AFM utilizes a cantilever deflection signal not only<br />

for obtaining a distance feedback signal (i.e., frequency shift signal) but also for<br />

producing a cantilever excitation signal. Although the cantilever excitation with a<br />

piezoactuator is most commonly used, the method typically results in an excitation <strong>of</strong><br />

spurious resonances around the natural resonance frequency <strong>of</strong> a cantilever due to the<br />

uncontrolled propagation <strong>of</strong> acoustic waves from a piezoactuator to the cantilever (Fig.<br />

1(a)).<br />

In order to suppress the influence <strong>of</strong> the spurious resonances, we propose a cantilever<br />

holder design, where the propagation <strong>of</strong> an acoustic wave is restricted the surrounding<br />

boundaries between two materials having significantly different specific acoustic<br />

resistances (ρ·c [kg/m 2 ·s]). In our design, an acoustic wave is greatly reduced by the<br />

reflection at the boundary between a piezoactuator (PZT, ρ·c=35×10 6 ) and its supporting<br />

material (PEEK, ρ·c=3.3×10 6 ). Then the remaining acoustic wave is further reduced by<br />

the reflection at the boundaries between the PEEK part and other components such as a<br />

cantilever support (stainless steel 316, ρ·c=36×10 6 ) and an optical window (glass,<br />

ρ·c=11×10 6 ). Figure 1 demonstrates the remarkable effect <strong>of</strong> the acoustic boundaries for<br />

suppressing spurious vibrations induced by the acoustic waves.<br />

a) b)<br />

Figure 1: Frequency responses <strong>of</strong> cantilever oscillation amplitude induced by a piezoactuator.<br />

(Excitation amplitude, : 0.5 V, : 1 V, : 1.5 V). Supporting materials for a piezoactuator: (a)<br />

stainless steel 316 and (b) PEEK.<br />

147


The different faces <strong>of</strong> the calcite (10-14) surface<br />

Jens Schütte 1 , Lutz Tröger 1 , Philipp Rahe 1 , Ralf Bechstein 1,2 and Angelika Kühnle 1<br />

1 Fachbereich Physik, Universität Osnabrück, Barbarastraße 7, 49076 Osnabrück, Germany<br />

2 present address: iNano, Aarhus University, DK-8000 Aarhus C, Denmark<br />

P.II-20<br />

Although atomic resolution has been achieved on calcite (10-14) using static atomic force<br />

microscopy (AFM) in liquid environment since 1992 [1], it is still unclear whether the<br />

surface reconstructs or not. Most experiments carried out so far using static AFM in<br />

liquids show a reconstruction <strong>of</strong> the surface in the [-4-21] direction, called “row pairing”<br />

[2,3]. Besides the row pairing, several results exist, showing a (2 x 1) reconstruction [4].<br />

This reconstruction has been discussed rather controversial and was explained by the<br />

adsorption <strong>of</strong> water.<br />

Here, we present an NC-AFM study carried out in ultra-high vacuum <strong>of</strong> the in-situ<br />

cleaved calcite (10-14) surface showing real atomic resolution (see Fig. 1 (a)) and a<br />

detailed research <strong>of</strong> the many “faces” <strong>of</strong> the surface. It is shown, that the (2 x 1)<br />

reconstruction (see Fig. 1 (b)) <strong>of</strong> the surface is not induced by water adsorption, but<br />

imaging the surface with a (2 x 1) reconstruction is strongly influenced by the tip-sample<br />

distance.<br />

a) b)<br />

Figure 1: NC-AFM frequency shift images <strong>of</strong> calcite (10-14). In (a) the surface shows “row<br />

pairing” in [-4-21] direction, however, no (2 x 1) reconstruction is seen. The average detuning is -<br />

9.3 Hz. In (b) a (2 x 1) reconstruction is clearly visible. The average detuning is -12.1Hz.<br />

[1] P. E. Hillner et al., Ultramicroscopy 42-44, 1387 (1992).<br />

[2] S. L. S. Stipp et al., Geochim. Cosmochim. Ac. 58, 3023 (1994).<br />

[3] F. Ohnesorge and G. Binnig, Science 260, 1451 (1993).<br />

148


P.II-21<br />

Manipulation Mechanism <strong>of</strong> Single Cu Atoms on Cu(110)-O Surface<br />

with Low Temperature Non-Contact AFM<br />

Y. Kinoshita, T. Satoh, Y. J. Li, Y. Naitoh, M. Kageshima and Y. Sugawara<br />

Department <strong>of</strong> Applied Physics, Graduate <strong>School</strong> <strong>of</strong> <strong>Engineering</strong>, Osaka University Osaka, Japan<br />

Atom manipulation is a fascinating technique for artificially fabricating nanometer<br />

scale structures. Recently using non-contact atomic force microscopy (NC-AFM), the<br />

well-controlled manipulations <strong>of</strong> individual atoms [1] or molecules on surfaces were<br />

demonstrated. Furthermore, using the force spectroscopy techniques, atomic force-vector<br />

field [2] or driving forces involved in a single atom manipulation [3] were determined.<br />

More recently our group have successfully manipulated single topmost Cu atoms<br />

laterally on Cu(110)-O surfaces and identified manipulation modes (pulling or pushing)<br />

from the AFM feedback signals. At the same time, we found that the atom species (Cu or<br />

O) <strong>of</strong> the tip apex changes not only the AFM topographic images drastically but also the<br />

mode for lateral manipulation <strong>of</strong> a Cu atom on the surface. However the difference in tipsample<br />

potential energy from the species <strong>of</strong> the tip apex has still been unclear. In this<br />

study, we investigate the potential distributions depending on the species <strong>of</strong> the topmost<br />

atom on the tip and clarify the different mechanisms for lateral manipulation <strong>of</strong> Cu atoms<br />

All measurements were performed in ultrahigh vacuum at 78 K. By using the image<br />

tracking scheme, the precise positioning <strong>of</strong> the tip with small drift velocity


P.II-22<br />

Simultaneous NC-AFM/STM Imaging <strong>of</strong> the Surface Oxide Layer on<br />

Cu(100) and Identification <strong>of</strong> Lattice Sites<br />

Mehmet Z. Baykara 1 , Todd C. Schwendemann 1,2 , Eric I. Altman 2 , Udo D. Schwarz 1<br />

1<br />

Department <strong>of</strong> Mechanical <strong>Engineering</strong> and Center for Research on Interface Structures and Phenomena<br />

(CRISP), <strong>Yale</strong> University, New Haven, USA<br />

2<br />

Department <strong>of</strong> Chemical <strong>Engineering</strong> and Center for Research on Interface Structures and Phenomena<br />

(CRISP), <strong>Yale</strong> University, New Haven, USA<br />

Exposure <strong>of</strong> Cu(100) to molecular oxygen at elevated temperatures leads to the (2√2 ×<br />

√2) R45 o missing row reconstruction pictured in Fig. 1a, where oxygen atoms are found<br />

to sit nearly co-planar with the Cu atoms in the outermost layer [1]. Since O and Cu<br />

atoms occupy sublattices with different symmetries, this surface represents an ideal<br />

model where atoms responsible for contrast formation in NC-AFM and STM imaging<br />

modes can be identified by symmetry alone. Using our homebuilt low temperature,<br />

ultrahigh vacuum atomic force microscope [2], we have obtained simultaneously<br />

recorded NC-AFM and tunneling current images on the oxidized Cu(100) surface.<br />

Comparing the visual distinctions between the two images, bright features in the images<br />

are assigned to specific locations on the surface. In particular, metal tips are found to<br />

interact strongly with bridge sites between two under-coordinated oxygen atoms, leading<br />

to a rectangular unit cell contrast in high-resolution NC-AFM images (Fig. 1b), whereas<br />

simultaneously collected tunneling current data exhibit an elongated hexagonal unit cell,<br />

which is assigned to surface Cu atoms (Fig. 1c). These measurements demonstrate the<br />

ability <strong>of</strong> multi-dimensional, atomic-resolution scanning probe microscopy to identify<br />

adsorption sites on oxide surfaces.<br />

Figure 1: a) Model <strong>of</strong> the (2√2 ×√2) R45 o O-induced reconstruction <strong>of</strong> Cu(100). The grey balls<br />

represent O, the light orange surface Cu, and the darker orange second layer Cu. b) NC-AFM<br />

image acquired with a frequency shift <strong>of</strong> -0.95 Hz at T = 6 K and c) simultaneously acquired<br />

tunneling current image collected at a sample bias <strong>of</strong> -0.4 V.<br />

[1] M. C. Asensio et al, Surf. Sci. 236, 1 (1990).<br />

[2] B. J. Albers et al, Rev. Sci Instrum. 79, 033704 (2008).<br />

150


nc-AFM Investigations <strong>of</strong> Metal Nanoclusters on α-alumina<br />

P.II-23<br />

K Venkataramani 1 , M C R Jensen 1 , M Reichling 2 , F Besenbacher 1 , and J V Lauritsen 1<br />

1 Interdisciplinary Nanoscience Center (iNANO), Department <strong>of</strong> Physics, Aarhus University, Denmark<br />

2 Department <strong>of</strong> Physics, University <strong>of</strong> Osnabrück, Germany<br />

AFM operated in the non-contact mode has proved very successful in imaging insulating<br />

surfaces with atomic resolution and recently has also proved the capability to image<br />

supported nanoclusters. Insight from studies <strong>of</strong> such systems may be important for a<br />

better and much needed understanding <strong>of</strong> heterogeneous catalysis, since most catalysts<br />

consists <strong>of</strong> nanoclusters dispersed on insulating metal oxide substrate. It is well known<br />

that while imaging nanoclusters with nc-AFM, the tip apex is one <strong>of</strong> the main limiting<br />

factors in the topography imaging mode i.e. constant detuning mode (Z) but a<br />

complementary scanning mode, the constant height mode (df), has surprisingly been<br />

shown to give a more precise estimation <strong>of</strong> lateral cluster shape [1]. In this study, using<br />

nc-AFM, we present high resolution topography images <strong>of</strong> Cu nanoclusters supported on<br />

clean (Fig. 1(a)) and pre-hydroxylated α-Al2O3(0001) surfaces and can thereby report in<br />

great detail on cluster growth, thermal stability and morphology. Further, by scanning in<br />

constant height mode, we obtained high quality images <strong>of</strong> the top (111) facet <strong>of</strong> Cu<br />

nanoclusters revealing a clear hexagonal shape (Fig. 1(b)). This equilibrium shape<br />

determined solely from AFM can be related to the adhesion energy (Eadh) by the equation<br />

derived from the Wulff-Kaichew reconstruction (Fig. 1(c)) [2], which gives an<br />

understanding <strong>of</strong> the strong cluster-substrate binding. Further, this study was extended to<br />

investigate the growth <strong>of</strong> Ni nanoclusters and our initial results revealed nanoscale<br />

ordering <strong>of</strong> Ni clusters at room temperature.<br />

Figure 1: (a) Topography image <strong>of</strong> supported Cu nanoclusters on α-Al2O3(0001) surface. (b) Cu<br />

clusters imaged in constant height mode. (c) Schematic representation <strong>of</strong> Wulff-Kaichew<br />

reconstruction and the equation relating free energy <strong>of</strong> cluster facet (γmetal(i)) with Eadh.<br />

[1] C. Barth, O. H. Pakarinen, et al. Nanotechnology 17, S128-S132 (2006).<br />

[2] C. R. Henry. Surface Science Reports 31, 231-325 (1998).<br />

151


Atom-resolved AFM studies <strong>of</strong> the polar MgAl2O4 (001) surface<br />

Morten K. Rasmussen, Jeppe V. Lauritsen, and Flemming Besenbacher<br />

Interdisciplinary Nanoscience Center (iNANO), University <strong>of</strong> Aarhus, Denmark<br />

P.II-24<br />

Magnesium-aluminate spinel (MgAl2O4) has many interesting properties and<br />

applications, among other things the use as a support for catalysts. In particular, MgAl2O4<br />

in a porous form is a support material for the important Ni-based catalyst used for steam<br />

reforming <strong>of</strong> methane, which today is the most common method <strong>of</strong> producing bulk<br />

hydrogen. A study <strong>of</strong> the clean spinel is thus <strong>of</strong> very great importance to understand the<br />

role <strong>of</strong> the support, and still almost no surface science studies on the MgAl2O4 surface<br />

have been presented due to the insulating nature <strong>of</strong> the material. When constructing a<br />

surface model one must consider the fact that the spinel (001) surface is polar and the<br />

surface consequently has to be modified somehow to cancel out or compensate for this.<br />

Additionally, it is well known that the AFM tip can pick up material from the surface,<br />

and this material then can change the polarity <strong>of</strong> the nano-apex which leads to variations<br />

in the AFM contrast seen in atom-resolved images and reveals the individual sub lattices<br />

[1, 2]. In this study we observe from atom-resolved AFM images that the MgAl2O4 (001)<br />

surface may be imaged in two distinctly different contrast modes revealing the anion or<br />

cation sub-lattice. Tentative structural characterizations from the AFM studies suggest a<br />

stoichiometric bulk like termination. Stabilization mechanism may then include<br />

relaxations in the crystal structure below the topmost layer which cannot be probed by<br />

AFM, hence surface x-ray diffraction is used to complement the AFM data.<br />

a) b)<br />

Figure 1: (a) Atom resolved non-contact AFM image (4nm×4nm) <strong>of</strong> the clean MgAl2O4 (001)<br />

surface imaged under UHV conditions together with (b) a ball model <strong>of</strong> the surface. The surface<br />

appears to be Mg-terminated and near stoichiometric with a low atomic defect density less than<br />

0.1ML.<br />

[1] Lauritsen, J.V., et al., Chemical identification <strong>of</strong> point defects and adsorbates on a metal oxide<br />

surface by atomic force microscopy. Nanotechnology, 2006. 17(14): p. 3436.<br />

[2] Enevoldsen, G.H., et al., <strong>Noncontact</strong> atomic force microscopy studies <strong>of</strong> vacancies and hydroxyls<br />

<strong>of</strong> TiO[sub 2](110): Experiments and atomistic simulations. Physical Review B (Condensed<br />

Matter and Materials Physics), 2007. 76(20): p. 205415-14.<br />

152


P.II-25<br />

Contrast formation on cross-linked (1x2) reconstructed titania (110)<br />

Hans H. Pieper 1 , Stefan Torbrügge 1,2 , Stephan Bahr 1,2 , Krithika Venkataramani 1,3 ,<br />

Angelika Kühnle 1 and Michael Reichling 1<br />

1 Fachbereich Physik, Universität Osnabrück, Barbarastraße 7, 49076 Osnabrück, Germany<br />

2 present address: SPECS GmbH, Voltastrasse 5, 13355 Berlin, Germnay<br />

3 present address: iNano, Aarhus University, DK-8000 Aarhus C, Denmark<br />

In NC-AFM imaging, the tip structure plays a prominent role in atomic contrast<br />

formation. It is, for instance, possible to image the cationic or the anionic sub-lattice <strong>of</strong><br />

CaF2(111) by changing the tip termination [1]. For the unreconstructed rutile TiO2 (110)<br />

at least a third mode is known [2].<br />

Here, we investigate the cross-linked (1x2) surface reconstruction <strong>of</strong> rutile<br />

TiO2(110), which has been studied both with STM and NC-AFM, however, the atomic<br />

structure is still discussed controversially. We present NC-AFM results, which will be<br />

interpreted with respect to tip termination and tip-surface distance. A comparison with<br />

existing models provides strong evidence for one <strong>of</strong> these models.<br />

Figure 1: Three consecutive images <strong>of</strong> the cross-linked (1x2) titania surface taken with different<br />

tip terminations. The left image is taken with a positive tip termination. A contact with the sample<br />

leads to a negative tip termination (middle image). The tip in the right picture is unidentified.<br />

Figure 2: High resolution measurements are in good agreement with the surface model suggested<br />

by Bennett [3] as shown to the right.<br />

[1] C. Barth et al. J. Phys.: Condens. Matter 13, 2061 (2001)<br />

[2] G.H. Enevoldsen et al. Phys. Rev. B 76, 205415 (2007)<br />

[3] R.A. Bennett et al. Phys. Rev. Lett. 82, 3831 (1999)<br />

153


P.II-26<br />

Nano volcanoes – the surface structure <strong>of</strong> antimony-doped TiO2(110)<br />

Ralf Bechstein 1,2 , Mitsunori Kitta 3 , Jens Schütte 1 , Hiroshi Onishi 3 , and Angelika Kühnle 1<br />

1 Fachbereich Physik, Universität Osnabrück, Barbarastraße 7, 49076 Osnabrück, Germany<br />

2 present address: iNano, Aarhus University, DK-8000 Aarhus C, Denmark<br />

3 Department <strong>of</strong> Chemistry, Kobe University, Rokko-dai, Nada-ku, Kobe 657-8501, Japan.<br />

Titanium dioxide constitutes a very prominent and widely used example <strong>of</strong> a wide band<br />

gap photocatalyst [1]. Upon codoping with chromium and antimony, rutile titania has<br />

shown photocatalytic activity under irradiation with visible-light. The reactivity has been<br />

found to be optimized in slight excess <strong>of</strong> antimony. Even larger antimony excess has<br />

been observed to again result in reduced photocatalytic activity [2].<br />

Here, we address the role <strong>of</strong> excess antimony in chromium and antimony codoped titania<br />

by studying the influence <strong>of</strong> antimony on the surface structure <strong>of</strong> rutile TiO2(110) using<br />

high-resolution NC-AFM. All findings draw a picture <strong>of</strong> weakly integrated antimony in<br />

antimony-doped titania indicating why an excess <strong>of</strong> antimony beyond the optimum ratio<br />

is unfavourable for photocatalytic applications. Moreover, the presented study revealed<br />

another interesting phenomenon: It seems to be possible to create nano-sized holes with<br />

depths ranging from a few up to more than hundred monatomic steps in a controlled way<br />

by tuning the annealing parameters <strong>of</strong> antimony-doped TiO2(110) [3]. This might be an<br />

appropriate strategy for creating nano-patterned surfaces with macroscopic techniques.<br />

(a) (b)<br />

Figure 1: NC-AFM topography images <strong>of</strong> antimony-doped TiO2(110) obtained after annealing at<br />

1100 K. Holes with a diameter <strong>of</strong> about 50 nm and a depth <strong>of</strong> more than 30 nm are created (a).<br />

Typically, the holes are surrounded by small islands (b). This results in an appearance that<br />

resembles a volcano-like shape.<br />

[1] A. Fujishima, X. Zhang, and D. A. Tryk, Surf. Sci. Rep. 63, 515 (2008).<br />

[2] H. Kato and A. Kudo, Catal. Today 78, 561 (2003).<br />

[3] R. Bechstein et al., submitted to Nanotechnology (2009).<br />

154


P.II-27<br />

Non-contact scanning nonlinear dielectric microscopy imaging <strong>of</strong> TiO2<br />

Nobuhiro Kin and Yasuo Cho<br />

(110) surfaces<br />

Research Institute <strong>of</strong> Electrical Communication, Tohoku University, Sendai, Japan<br />

Scanning nonlinear dielectric microscopy (SNDM) is a pure electrical method that is<br />

used for detecting the local anisotropy <strong>of</strong> dielectric materials and measuring ferroelectric<br />

polarization with sub-nanometer resolution [1]. Recently, we have developed non-contact<br />

SNDM (NC-SNDM) with a height control technique that makes use <strong>of</strong> the detection <strong>of</strong> a<br />

higher-order nonlinear dielectric constant (ε (4) signal). We have already succeeded in<br />

observing the typical Si (111) 7 × 7 atomic structure [2]. An interesting potential future<br />

application <strong>of</strong> NC-SNDM is in the imaging <strong>of</strong> insulating materials such as TiO2, SrTiO3<br />

(STO), Al2O3, SiO2, and MgO. It is well known that TiO2 becomes an n-type<br />

semiconductor after high-temperature annealing under ultra-high vacuum conditions. Due<br />

to its simplicity, the atomic structure <strong>of</strong> TiO2 can be examined easily as compared to that<br />

<strong>of</strong> other materials such as STO, which comprises three types <strong>of</strong> atoms. Moreover, TiO2<br />

has been studied extensively because it is well known and widely used as a superior<br />

substrate for catalysts. Therefore, in this study, we selected TiO2 (110) as our initial<br />

specimen for investigating insulating materials. We have demonstrated that atomresolved<br />

images <strong>of</strong> a TiO2 (110) surface can be obtained by NC-SNDM. Considering<br />

both the topography and the electrical dipole moment (ε (3) image) [2], the parallel bright<br />

stripes in the [001] direction should correspond to Ti 4+ . The protrusion on the bright<br />

stripes corresponds to Ti 3+ , and the other protrusion on the dark rows corresponds to H +<br />

due to water. It should be noted that there exist other possibilities; these will be analyzed<br />

in detail in the future.<br />

a) b)<br />

Figure 1: (a) NC-SNDM image <strong>of</strong> TiO2 (110) recorded using the ε (4) signal as feedback. The<br />

scan resolution is 30 nm × 30 nm. (b) STM image <strong>of</strong> TiO2 (110) having the same scan resolution.<br />

A positive voltage <strong>of</strong> 1.2 V is applied to the bottom <strong>of</strong> the sample.<br />

[1] Y. Cho, A. Kitahara, and T. Saeki. Rev. Sci. Instrum. 67, 2297 (1996).<br />

[2] Y. Cho and R. Hirose. Phys. Rev. Lett. 99, 186101 (2007).<br />

155


156<br />

P.II-28<br />

Characterizations <strong>of</strong> Carbon Material by Non-contact Scanning<br />

Non-linear Dielectric <strong>Microscopy</strong><br />

Shin-ichiro Kobayashi and Yasuo Cho<br />

Research Institute <strong>of</strong> Electrical Communication, Tohoku University, Sendai , Japa<br />

Recently a new and uniquely promising technique, the "non-contact scanning<br />

non-linear dielectric microscopy" (NC-SNDM) was introduced [1,2]. Since this technique has<br />

high sensitivity to variations in the capacitance, we were able to use it to visualize the domain<br />

structure in ferroelectric or carrier in flash memory. The graphite or fullerene (C60) are paied to<br />

attention because <strong>of</strong> the unique properties and the candidate <strong>of</strong> high performance device. In this<br />

study, we present SNDM images <strong>of</strong> the graphite (HOPG) and C60 on Si(111)-7x7 (7x7) surface<br />

by the NC-SNDM to understand the properties <strong>of</strong> carbon materials with new viewpoint. C60<br />

thin-film on 7x7 surfaces were fabricated by the deposition method under the ultra high vacuum<br />

Feedback signal to observe topography or SNDM images was utilized by 2ω amplitude signal.<br />

Figure 1 shows ω amplitude image <strong>of</strong> HOPG. The contrast is inverted to make easily<br />

to understand the image. The clear honeycomb structure is observed. In addition, the three<br />

saddle points in the single hexisagonal ring are also recognized. This means that the NC-SNDM<br />

can detect the A or B site originated from the difference <strong>of</strong> the band structure. These results are<br />

similar to those <strong>of</strong> the STM. This indicates that the origin <strong>of</strong> the SNDM signal for the HOPG is<br />

similar to that <strong>of</strong> the STM. Figure 2 shows the topography <strong>of</strong> C60 for 1ML on 7x7 surface. The<br />

insert is the expanded image <strong>of</strong> Fig. 2. As<br />

is seen from the insert, the internal<br />

structures in some C60 molecules by the<br />

NC-SNDM are observed. This structure is<br />

similar to the charge distribution for the<br />

LUMO orbital. Furthermore, investigating Figure 1 The ω amplitude image for HOPG the<br />

phase reversal in ω phase image for<br />

various coverage <strong>of</strong> C60 on 7x7 surface, we<br />

could guess the characterization <strong>of</strong> the<br />

bonding between C60 and 7x7 surface. The<br />

details <strong>of</strong> the observed SNDM images are<br />

described in the presentation.<br />

[1]R. Hirose, K.Ohara and Y. Cho, Figure 2 Topography <strong>of</strong> C60 on 7x7 surface for 1ML<br />

Nanotechnology. 18 (2007), p084014. [2] Y. Cho and R. Hirose, PRL 99 (2007), p186101.


P.II-29<br />

Local Dielectric Spectroscopy <strong>of</strong> Nanocomposite Materials Interfaces<br />

M. Labardi 1 , D. Prevosto 1 , S. Capaccioli 1,2 , M. Lucchesi 1,2 and P.A. Rolla 1,2<br />

1 INFM-CNR, LR polyLab, Largo Pontecorvo 3, 56127 Pisa, Italy<br />

2 Dipartimento di Fisica, Università di Pisa, Largo Pontecorvo 3, 56127 Pisa, Italy.<br />

Dielectric spectroscopy (DS) has revealed especially successful to disclose structural<br />

properties <strong>of</strong> polymeric materials, like e.g. their glass transition and structural relaxation<br />

phenomena. Local probe DS [1] combines the high spatial resolution <strong>of</strong> frequency<br />

modulation electrostatic force microscopy (FM-EFM) with the physical insight typical <strong>of</strong><br />

DS, being firstly demonstrated on homogeneous poly-vinyl-acetate (PVAc) films under<br />

vacuum [1]. In essence, a sinusoidal potential at frequency fmod is applied to a conductive<br />

probe and the induced modulation <strong>of</strong> the resonant frequency shift Δfres is measured by<br />

dual-phase lock-in technique, referenced at 2fmod. Phase delay indicates dielectric loss due<br />

to polarization relaxation occurring with characteristic time τR ~ 1/ fmod [1].<br />

We apply local DS to address the influence <strong>of</strong> inorganic inclusions on structural<br />

relaxation <strong>of</strong> polymers. A PVAc thin film (30 nm) with a dispersion <strong>of</strong> a layered silicate<br />

(montmorillonite, MMT) was spin-coated onto Au, and partially scratched away to<br />

expose the metal, to obtain a reference surface (Fig. 1a). Presence <strong>of</strong> inorganic layers on<br />

the surface is evident from the TappingMode TM phase image (Fig. 1b). Conventional FM-<br />

EFM (Veeco Lift mode TM in controlled atmosphere [2] and temperature) shows materials<br />

contrast (Fig. 1c), independent <strong>of</strong> T. In Fig. 1d-f, dielectric loss maps at T = 50, 60, 70°C<br />

show highest contrast at T = 60°C for the used fmod = 130 Hz, corresponding to τR ~ 7-8<br />

ms. Dielectric contrast at inclusions can also be studied (Fig. 1e, arrow).<br />

Financial support from INFM-CNR “Seed 2007” project is acknowledged.<br />

Figure 1: a) Topography <strong>of</strong> a PVAc/MMT thin film (left) on Au (right). b) Phase image showing<br />

the location <strong>of</strong> inorganic layers (arrow). c) Δfres map. d)-f) Dielectric loss maps at T = 50,60,70°C.<br />

[1] P.S. Crider, M.R. Majewski, J. Zhang, H. Oukris, N.E. Israel<strong>of</strong>f, Appl. Phys. Lett. 91, 013102 (2007).<br />

[2] M. Lucchesi, G. Privitera, M. Labardi, D. Prevosto, S. Capaccioli, P. Pingue, J. Appl. Phys. 105,<br />

054301 (2009).<br />

157


P.II-30<br />

Probing Local Bias-Induced Phase Transitions on the Single Defect<br />

Level: from Imaging to Deterministic Mechanisms<br />

N. Balke 1 , S. Jesse 1 , P. Maksymovych 1 , Y.H. Chu 2 , R. Ramesh 2 , S. Choudhury 3 , L.Q.<br />

Chen 3 , and S.V. Kalinin 1<br />

1 Oak Ridge National Laboratory, Oak Ridge, TN 37831<br />

2 Dept. <strong>of</strong> Physics and Dept. <strong>of</strong> Mat. Sci. and Eng., University <strong>of</strong> California, Berkeley, CA<br />

3 Dept. <strong>of</strong> Mat. Sci. and Eng., Pennsylvania State University, University Park, PA 16802<br />

<strong>Force</strong>-induced processes such a molecular unfolding spectroscopy and NC-AFM<br />

atomic manipulation now underpin many areas <strong>of</strong> physics and biology. Here, I discuss<br />

recent progress on probing local bias-induced phase transitions in ferroelectrics and<br />

electrochemical systems. Bias-induced phase transitions are controlled by structural<br />

defects that act both as the local nucleation centers and the pinning centers for moving<br />

domain walls. Future progress necessitates understanding <strong>of</strong> polarization switching<br />

mechanisms on the single structural or morphological defect level. In this talk, I will<br />

present results on local studies <strong>of</strong> polarization reversal mechanisms in ferroelectrics [1-2].<br />

The direct imaging <strong>of</strong> a single nucleation center on the sub-100 nanometer level is<br />

demonstrated [3]. By using switching spectroscopy Piezoresponse <strong>Force</strong> <strong>Microscopy</strong> on<br />

systems with engineered defect structures and phase-field modeling, we demonstrate that<br />

deterministic that mesoscopic polarization switching mechanisms on single 1D and 2D<br />

defects can be determined. We develop a framework to link the modeling results to<br />

experimental observations using neural-network based recognition. The future potential<br />

for atomistic studies is discussed. Research was supported by the U.S. Department <strong>of</strong><br />

Energy Office <strong>of</strong> Basic Energy Sciences Division <strong>of</strong> Scientific User Instruments and was<br />

performed at Oak Ridge National Laboratory which is operated by UT-Battelle, LLC.<br />

Figure 1: (a) Surface topography and (b) SS-PFM nucleation bias map <strong>of</strong> BiFeO3 24deg (100)<br />

bicrystal grain boundary (GB). Each point in (b) is a hysteresis loop. (c) Ferroelectric switching<br />

properties across the interface.<br />

[1] S.V. Kalinin, B.J. Rodriguez, S. Jesse, Y.H. Chu, T. Zhao, R. Ramesh, E.A. Eliseev, and A.N.<br />

Morozovska, PNAS 104, 20204 (2007).<br />

[2] S. Jesse, B.J. Rodriguez, A.P. Baddorf, I. Vrejoiu, D. Hesse, M. Alexe, E.A. Eliseev, A.N. Morozovska,<br />

and S.V. Kalinin, Nature Materials 7, 209 (2008).<br />

[3] S.V. Kalinin, S. Jesse, B.J. Rodriguez, Y.H. Chu, R. Ramesh, E.A. Eliseev and A.N. Morozovska, Phys.<br />

Rev. Lett. 100, 155703 (2008)<br />

158


P.II-31<br />

Polarization-dependent electron tunneling into ferroelectric surfaces<br />

Peter Maksymovych 1 , Stephen Jesse 1 , Pu Yu 2 , Ramamoorthy Ramesh 2 , Arthur P.<br />

Baddorf 1 , Sergei V. Kalinin 1<br />

1 Center for Nanophase Materials Sciences, Oak Ridge National Laboratory<br />

2 Department <strong>of</strong> Physics and Dept. <strong>of</strong> Materials Science, University <strong>of</strong> California, Berkeley<br />

Electron tunneling underlies the operation <strong>of</strong> numerous devices relevant to<br />

information technology and has been proposed in futuristic applications for energy<br />

harvesting and quantum computing. Replacing a conventional insulator in the tunnel<br />

junction with electronically correlated materials can yield new types <strong>of</strong> electronic<br />

functionality. In one such concept, dubbed ferroelectric tunneling, the tunneling barrier<br />

height is controlled by the polarization <strong>of</strong> a ferroelectric oxide, enabling non-volatile<br />

conduction states that can be switched with electric field. Although ferroelectric<br />

tunneling has been a recent focus <strong>of</strong> a number <strong>of</strong> theoretical studies, aiming to unveil the<br />

interfacial behaviors <strong>of</strong> the ferroelectric order parameter, a convincing experimental<br />

demonstration <strong>of</strong> this phenomenon has been lacking. The key challenges are to find a<br />

material system that simultaneously satisfies the dimensional constraints for tunneling<br />

and ferroelectricity, as well as to assure that the conductance is not dominated by<br />

extrinsic effects due to charge injection and filamentary conduction, which is ubiquitous<br />

in complex oxides.<br />

In this talk we will demonstrate a highly reproducible polarization control <strong>of</strong> local<br />

electron transport through thin Pb(Zr0.2Ti0.8)O3 film. Despite being 30 nm thick,<br />

conductive atomic force microscopy revealed that the film possessed spatially and<br />

temporally reproducible local conductivity. Fowler-Nordheim electron tunneling was<br />

identified as the conductance mechanism, likely enabled by strong electric field in the<br />

sub-surface region (excess <strong>of</strong> 10 6 V/cm) created by the relatively sharp metal tip. Local<br />

conductance exhibited strong hysteresis depending on the probed bias range. Using a<br />

newly-developed combination <strong>of</strong> piezoresponse force and conductive atomic force<br />

microscopy, we have, for the first time, directly correlated local events <strong>of</strong> ferroelectric<br />

and resistive switching [1]. The large polarization <strong>of</strong> the studied film resulted in as high<br />

as 500-fold enhancement <strong>of</strong> the FN-tunneling conductance upon ferroelectric switching,<br />

sufficient to demonstrate a local non-volatile memory function.<br />

The physical mechanism <strong>of</strong> the observed effect was traced to the polarizationdependence<br />

<strong>of</strong> the height and possibly width <strong>of</strong> the metal-ferroelectric Schottky barrier.<br />

Our measurements have thus extended a well-known polarization dependence <strong>of</strong> surfacepotential<br />

on ferroelectric materials from Kevlin probe force microscopy to demonstrate<br />

local control <strong>of</strong> electron transport through such materials. Variable-temperature<br />

measurements and local effects due to dielectric non-linearities will also be discussed.<br />

[1] P. Maksymovych, S. Jesse, P. Yu, R. Ramesh, A. P. Baddorf, S. V. Kalinin, Science (2009) in review.<br />

159


Local ferroelectric and magnetic investigations on<br />

multiferroic thin films<br />

P.II-32<br />

Ulrich Zerweck 1 , D. Köhler 1 , P. Milde 1 , Ch. Loppacher 2 , S. Geprägs 3 , S.T.B. Goennenwein<br />

3 , R. Gross 3 , L.M. Eng 1<br />

1 Institute <strong>of</strong> Applied Photophysics / Technische Universität Dresden, 01062 Dresden, Germany<br />

2 L2MP, Université Paul Cézanne, 13397 Marseille , France<br />

3 Walther-Meißner-Institut / Bayerische Akademie der Wissenschaften, 85748 Garching, Germany.<br />

Multiferroic materials, which combine ferroelectric and ferromagnetic properties, are<br />

promising candidates for novel nanoscale switching and memory devices, especially if<br />

there is a coupling between those properties.<br />

Here, we present low-temperature noncontact atomic force microscopy (nc-AFM) - in<br />

particular magnetic force microscopy (MFM) and Kelvin probe force microscopy<br />

(KPFM) - investigations <strong>of</strong> multiferroic BiCrO3 and BiFeO3 thin films prepared by<br />

pulsed laser deposition (PLD) as described in [1]. In addition to the nc-AFM<br />

investigations, piezoresponse force microscopy (PFM) was used to study the ferroelectric<br />

switching behaviour <strong>of</strong> the thin films.<br />

Although the separation <strong>of</strong> magnetic and ferroelectric properties is nontrivial due to the<br />

long range nature <strong>of</strong> both <strong>of</strong> the related forces, we were able to clearly separate the two<br />

different domain types by simultaneously running KPFM and MFM.<br />

[1] S. Geprägs et al., Phil. Mag. Lett. 87: 141 (2007).<br />

160


Magnetic Resonance <strong>Force</strong> <strong>Microscopy</strong> in Anisotropic Systems<br />

Thomas Fan and Vladimir I. Tsifrinovich 1<br />

1 Department <strong>of</strong> Physics, Polytechnic Institute <strong>of</strong> NYU, Brooklyn, USA<br />

161<br />

P.II-33<br />

We have developed theory for the detection <strong>of</strong> a single spin S≥1 using magnetic<br />

resonance force microscopy (MRFM), a subset <strong>of</strong> atomic force microscopy. The<br />

anisotropy caused by the crystalline field is taken into account. The MRFM signal (i.e.<br />

the frequency shift <strong>of</strong> the cantilever vibrations) in the oscillating cantilever-driven<br />

adiabatic reversals technique is computed using a semi-classical approach: the spin is<br />

treated quantum mechanically, and the cantilever vibrations classically. We have shown<br />

that the MRFM signal for spin S≥1 is the same as for the spin S=1/2. We have obtained<br />

the analytical estimate for the half-width <strong>of</strong> the MRFM signal.


P.II-34<br />

Sub-10 nm resolution in Magnetic <strong>Force</strong> <strong>Microscopy</strong> (MFM) at ambient conditions<br />

Ö. Karcı 1,2 , H. Atalan 1 , M. Dede 1 , Ü. Çelik 3 , A.Oral 4<br />

1 NanoMagnetics Instruments Ltd., 266 Banbury Road, Oxford OX2 7DL, UK.<br />

2 Hacettepe University, Department <strong>of</strong> Nanoscience & Nanomedicine, Beytepe, Ankara,<br />

Turkey<br />

3 Istanbul Technical University, Department <strong>of</strong> Material Science, Istanbul, Turkey<br />

4 Sabanci University, Faculty <strong>of</strong> <strong>Engineering</strong> & Natural Sciences, 34956 Istanbul, Turkey<br />

aoral@sabanciuniv.edu<br />

We describe designs <strong>of</strong> high resolution Magnetic <strong>Force</strong> Microscopes, which can achieve<br />

better than 10 nm resolution in magnetic imaging even in ambient conditions. We<br />

developed two different MFMs, which can both achieve better than 10 nm lateral<br />

resolution, using commercially available, <strong>of</strong>f the shelf, magnetically coated cantilevers.<br />

Both MFMs have been optimized to have low noise, using RF modulation <strong>of</strong> the laser<br />

diode. One <strong>of</strong> the instruments is designed for low temperature use, 300 mK-300 K using<br />

low noise fiber optic interferometer & alignment free cantilevers. The other MFM is<br />

designed to operate in ambient conditions and employ a beam deflection pickup with<br />

optimized noise performance. Both MFMs can also be operated in bimodal operation,<br />

using first and second resonance <strong>of</strong> the magnetically coated cantilever for topography and<br />

magnetic contrast. One <strong>of</strong> the typical images obtained at ambient conditions with the<br />

MFMs on Seagate’s 394 Gbpsi harddisk is given below.<br />

162


P.II-35<br />

Enhancement <strong>of</strong> the Exchange-bias Effect based on Quantitative<br />

Magnetic <strong>Force</strong> <strong>Microscopy</strong> Results.<br />

N. Pilet 1 , M.A. Marioni 1 , S. Romer 1 , N. Joshi 2 , S. Özer 2 and H.J. Hug, 1,2<br />

1 Empa, Swiss Federal Institute for Materials Testing and Research, CH-8600 Duebendorf, Switzerland,<br />

2 Department <strong>of</strong> Physics, University <strong>of</strong> Basel, Klingelbergstrasse 82, CH-4056 Basel, Switzerland<br />

Exchange biasing (EB) causes a lateral shift <strong>of</strong> the magnetization vs. field curve in<br />

antiferromagnetic-ferromagnetic (AF-F) thin-film structures. It is largely believed that<br />

EB arises from linking the magnetization <strong>of</strong> the F to that <strong>of</strong> (pinned) uncompensated<br />

spins (UCS) in the AF. In the present work we are able to measure the local density <strong>of</strong><br />

UCS on the scale <strong>of</strong> 20 nm, using quantitative magnetic force microscopy (qMFM). With<br />

quantitative results at this lateral resolution, it becomes possible to resolve and measure<br />

the UCS underneath each F domain at scales comparable to the grain size. The average<br />

density <strong>of</strong> UCS around 20% <strong>of</strong> that expected for a full monolayer <strong>of</strong> spins is considerably<br />

higher than expected from the measured exchange field and local values may reach<br />

100%. We found that UCSs coupling antiparallel to the F magnetization stabilize its<br />

orientation, i.e. they are biasing [1,2]. Locally, UCS oriented parallel to the F<br />

magnetization were found. These have an antibiasing effect and are hence expected to<br />

weaken the exchange field. We hypothesized that such anti-biasing arises from<br />

conflicting exchange interaction between the F-layer and AF-grains, and between AFgrains<br />

themselves. Exchange-bias systems grown with suppressed inter-granular coupling<br />

in the AF-layer proved our hypothesis to be correct. These systems showed an exchangebias<br />

effect enhanced by 20%. High resolution MFM revealed a more homogenous pattern<br />

<strong>of</strong> UCS with a correspondingly higher density. The evolution <strong>of</strong> the F-domains in an<br />

applied field was strikingly different.<br />

Figure 1: . (A) UCS density calculated from measured MFM data and the instrument calibration<br />

function. Positive UCS (i.e. parallel to the cooling field) are colored yellow, negative UCS red.<br />

(B) UCS with overlayed contours <strong>of</strong> the F domains at H = 0; (C) Idem for H = 200mT. With few<br />

exceptions (such as pointed out by the arrows) the UCS align antiparallel to the domain<br />

magnetization at H=0.<br />

[1] I. Schmid, P. Kappenberger, O. Hellwig, M. Carey, E. Fullerton and H. J. Hug, EPL 81 (2008) 17001<br />

[2] I. Schmid, M. Marioni et al. submitted.<br />

163


Author<br />

Index<br />

164


Abe M. 51,58,92,141 Clerk A. A. 78<br />

Adachi H. 141 Cockins L. 78<br />

Ahn C. 109 Cuberes M. T. 125<br />

Albers B. J. 62 Dalvit D. 33<br />

Albonetti C. 69 Dalvit D. A. R. 43<br />

Altman E. I. 62,150 Danza, R. 76<br />

Amijs C. H. 130 de Boer M. P. 35<br />

Andrews K. M. 111 de Hosson J. Th. M. 42<br />

Annibale P. 137 de Man S. 39<br />

Aoki H. 129 Debierre J. M. 68<br />

Arai T. 94,114,118 Dede M. 162<br />

Asakawa H. 46,147 DelRio F. W. 35<br />

Ashino M. 64 Deresmes D. 105,106<br />

Atalan H. 162 Diesinger H. 105,106<br />

Baddorf A. P. 116,159 Dietz C. 89<br />

Bahr S. 153 Ding X. D. 103<br />

Balke N. 116,158 Duden T. 131<br />

Bao Y. 41 Ebeling D. 49<br />

Barat<strong>of</strong>f A. 54,74,75,95,96,130 Ebner A. 137<br />

Baykara M. Z. 62,150 Echavarren A. M. 130<br />

Bechstein R. 60,66,148,154 Elias G. 104<br />

Bennett S. D. 78 Eng L. M. 55,86,102,136,160<br />

Berber S. 64 Esquivel-Sirvent R. 32<br />

Besenbacher F. 151,152 Falter J. 109.112<br />

Bettac A. 80 Fan T. 161<br />

Binns C. 40 Feltz A. 80<br />

Biscarini F. 69,137 Ferrini G. 97<br />

Boag A. 104 Foster A. 100<br />

Bocquet F. 68,100 Fremy S. 67<br />

Borowik L. 106 French R. H. 29<br />

Braun D. A. 73,112 Freund H. J. 57<br />

Brugger T. 101 Fuchs H. 49,73,112<br />

Campbellova A. 71 Fukui K. 90<br />

Canetta C. 99 Fukuma T. 45,46,115,147<br />

Capaccioli S. 157 Gangopadhyay, S. 76<br />

Cavero-Pelaez, I. 31 Garcia R. 69,89,126,137<br />

Celik U. 162 Gepraegs S. 160<br />

Chan H. B. 41 Giessibl F. J. 72,77<br />

Chawla G. 119 Gimzewski J. K. 146<br />

Chen L. Q. 158 Glatzel Th. 54,74,75,95,96,101,104,127,130<br />

Chen, G. 38 Goennenwein S. T. B. 160<br />

Chevrier J. 37 Gomez C. J. 69<br />

Ching W. Y. 29 Gonzalez C. 59<br />

Cho Y. 155,156 Gottlieb O. 121<br />

Choi B. Y. 131 Grafstroem S. 86<br />

Choudhury S. 158 Graham N. 30<br />

Chu Y. H. 158 Greffet J-J. 37<br />

Cirelli R. A. 41 Gross L. 77,127<br />

165


Gross R. 160 Kitta M. 60,154<br />

Grutter P. 78 Kiyohara K. 94<br />

Gu N. 38,99 Klapetek P. 71<br />

Hane F. 102 Klemens F. 41<br />

Harada M. 48 Klocke M. 123<br />

Haubner K. 136 Kobayashi D. 87<br />

Heeck K. 39 Kobayashi K. 47,84,85,88,118,122,140,141<br />

Heinrich A. J. 72 Kobayashi N. 143<br />

Helm M. 86 Kobayashi S. 156<br />

Herruzo E. T. 89 Koch S. 54,74,75,95,96,101,127,130<br />

Heyde M. 57 Kokawa R. 118,141,145<br />

Hinterdorfer P. 137 Köhler D. 136,160<br />

Hirade M. 118 König T. 57<br />

Hirata Y. 88 Kusiaku K. 106<br />

Hosokawa Y. 84 Kurachi S. 144<br />

H<strong>of</strong>fman J. E. 110 Kuroda M. 145<br />

H<strong>of</strong>fman P. M. 98 Kushida S. 94<br />

H<strong>of</strong>fman R. 55 Kühnle A. 60,66,133,134,148,153,154<br />

H<strong>of</strong>mann T. 72 Labardi M. 157<br />

Hori K. 114 Laemmle K. 67<br />

Hornstein S. 121 Lamoreaux S. K. 43<br />

Hölscher H. 49,55,112 Lange M. 113<br />

Hug H. J. 163 Langer M. 101<br />

Huston S. M. 111 Langewisch G. 73<br />

Iannuzzi D. 39 Lau M. W. 135<br />

Ido S. 47,88 Lauritsen J. V. 151,152<br />

Imai T. 47 Leonenko Z. 102<br />

Inoue T. 141 Li Y. 82,120<br />

Itaya K. 129 Li Y. J. 52,143,144,149<br />

Jacob R. 86 Liebmann M. 109<br />

Jaehne E. 136 Liljeroth P. 77<br />

Jelinek P. 59,63,71 Liu Y. 135<br />

Jensen M. C. R. 151 Lobo-Checa J. 101<br />

Jeong Y. 118,142 Loppacher Ch. 68,100,160<br />

Jesse S. 116,117,158,159 Losilla N. S. 137<br />

Joshi N. 163 Loske F. 133<br />

Kageshima M. 52,82,120,143,144,149 Lozano J. R. 89<br />

Kalinin S. V. 116,117,158,159 Lucchesi M. 157<br />

Karci O. 162 Lutz C. P. 72<br />

Kawai S. 54,74,75,95,96,101,127,130 Lux-Steiner M. Ch. 61<br />

Kawai T. 138 Ma Z. 82,120<br />

Kawakatsu H. 87 Maass P. 133<br />

Khlobystov A. N. 64 Maeda Y. 138<br />

Kim J. 110 Maksymovych P. 116,158,159<br />

Kim W. J. 43 Malegori G. 97<br />

Kimura K. 47 Mansfield W. M. 41<br />

Kin N. 155 Marioni M. A. 163<br />

Kinoshita Y. 52,149 Martinez J. 126<br />

166


Martinez N. F. 69 Parashar P. 31<br />

Martinez R. V. 126 Parsegian A. 29,53<br />

Masago A. 93 Pawlak R. 68<br />

Matsamura H. 141 Pearl T. P. 111<br />

Matsumoto T. 138 Perez R. 59,65,69,71<br />

Matsushige K. 84,85,88,122,140 Pieper H. H. 153<br />

Melin T. 105,106 Pilet N. 62,109,163<br />

Mendoza P. 130 Podgornik R. 29<br />

Meyer E. 54,74,75,95,96,101,104,127,130 Poole P. 78<br />

Meyer G. 77,127 Port R. T. 111<br />

Milde P. 55,136,160 Porte L. 68<br />

Miller C. 139 Pou P. 65,71<br />

Milton K. A. 31 Pratama A. 58<br />

Minato T. 129 Preiner J. 137<br />

Mitani Y. 115 Prevosto D. 157<br />

Miyahara Y. 78 Prosenc M. 67<br />

Morita S. 51,58,92,141 Rahe P. 134,148<br />

Mohn F. 77 Rajter R. 29,34<br />

Moores B. 102 Ramesh R. 158,159<br />

Mori Y. 141 Rasmussen M. K. 152<br />

Moriarty P. 76 Rasool H. I. 146<br />

Morita K. 51,92 Reichling M. 66,151,153<br />

Möller R. 113 Repp J. 77<br />

Murai R. 141 Ribbeck H.-G. 86<br />

Murakami S. 141 Roca i Cabarrocas P. 106<br />

Nagashima K. 141 Rohlfing M. 66<br />

Naitoh Y. 52,82,120,143,144,149 Rolla P. A. 157<br />

Narayanaswami A. 99,38 Romer S. 163<br />

Nguyen-Tran T. 106 Rosenwalks Y. 104<br />

Nie H. Y. 135 Rousseau E. 37<br />

Nikiforov M. 117 Rozsival V. 63<br />

Nimmrich M. 134 Rychen J. 83<br />

Nishida S. 87 Sachrajda A. 78<br />

Nony L. 68,100 Sadewasser S. 61<br />

Nys J-P. 105 Salmeron M. 131<br />

Oesterhelt F. 49 Sasahara A. 118,142<br />

Ogawa T. 144 Sassi M. 68<br />

Ohta M. 141,145 Sato T. 94<br />

Oison V. 68 Satoh T. 149<br />

Okabe N. 87 Sawada D. 51,92<br />

Ondracek M. 63 Schaff O. 83<br />

Onishi H. 60,154 Schaffert J. 113<br />

Oral A. 162 Schirmeisen A. 50,73,112<br />

Ovchinnikov O. 116,117 Schlittler R. R. 127<br />

Oyabu N. 47,88,118,140,141 Schneider S. C. 86<br />

Ozer S. 163 Schütte J. 60,66,133,134,148,154<br />

Pai C. S. 41 Schwarz A. 67<br />

Palasantzas G. 42 Schwarz U. D. 62,109,112,150<br />

167


Schwendemann T. C. 62,150 Wagner T. 129<br />

Shajesh K. V. 31 Weiner D. 50,73<br />

Sharma S. 146 Welker J. 72<br />

Shavit A. 121 Wenzel M. T. 86<br />

Shen S. 38 Wiesendanger R. 64,67<br />

Shi Y. 131 Wijngaarden R. J. 39<br />

Shluger A. 79,132 Williamson R. 139<br />

Siber A. 29 Winnerl S. 86<br />

Simon G. H. 57 Wintjes N. 113<br />

Siria A. 37 Wojciech K. 69<br />

Solares S. D. 119 Wolf D. E. 123<br />

Someya T. 145 Wright C. A. 119<br />

Staffier P. 109 Wu W. 121<br />

Stara I. G. 134 Xu J. B. 103<br />

Studenikin S. A. 78 Yabuno H. 145<br />

Such B. 73,74,75,95,96,101,130 Yamada H. 47,84,85,88,118,122,140,141<br />

Sugawara Y. 82,120,143,144,149 Yamatani H. 118<br />

Sugimoto Y. 51,58,92 Yang D. Q. 135<br />

Sushkov A. 43 Yang J. 135<br />

Sutter P. 108 Yokota Y. 90<br />

Suzuki K. 47,140 Yu P. 159<br />

Svetovoy V. B. 42 Yurtsever A. 58<br />

Sweetman A. 76 Zahl P. 107,108<br />

Tagami K. 48 Zech M. 110<br />

Takahashi K. 143 Zerweck U. 55,86,136,160<br />

Takano K. 141 Zhang J. X. 103<br />

Tang H. X. 36 Zou J. 41<br />

Ternes M. 72 Zypman F. 124<br />

Theron D. 106<br />

Thornton S. 40<br />

Tomanek D. 64<br />

Tomitori M. 94,114,118,142<br />

Torbruegge S. 83,153<br />

Torricelli G. 40<br />

Trevethan T. 79,132<br />

Troeger L. 148<br />

Tsifrinovich V. I. 161<br />

Tsukada M. 48,88,93<br />

Tsunemi E. 85<br />

Ueda Y. 45,46<br />

Umeda K. 90<br />

van Zwol P. J. 42<br />

Venkataramani K. 151,153<br />

Wagner J. 31<br />

168


List <strong>of</strong><br />

Participants<br />

169


LastName FirstName Institution Email<br />

Altman Eric <strong>Yale</strong> University eric.altman@yale.edu<br />

Arai Toyoko Kanazawa University arai@staff.kanazawa-u.ac.jp<br />

Asakawa Hitoshi Kanazawa University hi_asa@staff.kanazawa-u.ac.jp<br />

Barat<strong>of</strong>f Alexis University <strong>of</strong> Basel alexis.barat<strong>of</strong>f@unibas.ch<br />

Baykara Mehmet <strong>Yale</strong> University mehmet.baykara@yale.edu<br />

Bechstein Ralf Aarhus University ralf@inano.dk<br />

Bernard Laetitia EMPA laetitia.bernard@empa.ch<br />

Bettac Andreas Omicon NanoTechnology a.bettac@omicron.de<br />

Borowik Lukasz Institut d'Electronique<br />

Microelectronique Nanotechnologie<br />

lbor@tlen.pl<br />

Braendlin Dominik Nanosurf AG braendlin@nanosurf.com<br />

Cannara Rachel National Institute <strong>of</strong> Standard and<br />

Technology<br />

rachel.cannara@nist.gov<br />

Celik Umit Istanbul Technical University umitcelik@itu.edu.tr<br />

Chan Ho Bun University <strong>of</strong> Florida hochan@phys.ufl.edu<br />

Chawla Gaurav University <strong>of</strong> Maryland College Park gchawla@umd.edu<br />

Cho Yasuo Tohoku University yasuocho@riec.tohoku.ac.jp<br />

Dalvit Diego LANL dalvit@lanl.gov<br />

de Boer Maarten Sandia National Labs mpdebo@sandia.gov<br />

de Man Sven VU University Amsterdam sdeman@nat.vu.nl<br />

Diesinger Heinrich Institut Elelectronique<br />

Microelectronique Nanotechnologie<br />

170<br />

heinrich.diesinger@isen.iemn.univlille1.fr<br />

Ding Xidong Sun Yat-sen University dingxd@mailsysu.edu.cn<br />

Ebeling Daniel University <strong>of</strong> Muenster Daniel.Ebeling@uni-muenster.de<br />

Esquivel-Sirvent Raul Universidad Nacional Autonoma de<br />

Mexico<br />

raul@fisica.unam.mx<br />

Faddis David Nanosurf Inc. faddis@nanosurf.com<br />

Falter Jens Westfallische Wilhelms-Universitaet<br />

Muenster<br />

JensFalter@Uni-Muenster.de<br />

Ferrini Gabriele University Cattolica gabriele@dmf.unicatt.it<br />

Fukui Ken-ichi Osaka University kfukui@chem.es.osaka-u.ac.jp<br />

Fukuma Takeshi Kanazawa University fukuma@staff.kanazawa-u.ac.jp<br />

Giessibl Franz University <strong>of</strong> Regensburg franz.giessibl@physik.uni-r.de<br />

Glatzel Thilo University <strong>of</strong> Basel Thilo.Glatzel@unibas.ch<br />

Gomez Carlos IMM - CSIC carlos@imm.cnm.csic.es<br />

Graham Noah Middlebury College ngraham@middlebury.edu<br />

Gross Leo IBM Research lgr@zurich.ibm.com<br />

Grutter Peter McGil University grutter@physics.mcgill.ca<br />

Hafizovic Sadik Zurich Instruments sadik.hafizovic@zhinst.com<br />

Heer Flavio Zurich Instruments flavio.heer@zhinst.com<br />

Heyde Markus Fritz-Haber-Institute <strong>of</strong> the Max-<br />

Planck-Society<br />

heyde@fhi-berlin.mpg.de<br />

H<strong>of</strong>fmann Peter Wayne State University h<strong>of</strong>fmann@wayne.edu<br />

Hori Kenichirou Kanazawa University horiken@stu.kanazawa-u.ac.jp<br />

Hosokawa Yoshihiro Kyoto University hosokawa@piezo.kuee.kyoto-u.ac.jp<br />

Hölscher Hendrik Forschungszentrum Karlsruhe hendrik.hoelscher@imt.fzk.de<br />

Huston Shawn North Carolina State University smhuston@ncsu.edu


Ido Shinichiro Kyoto University ido@piezo.kuee.kyoto-u.ac.jp<br />

Jelinek Pavel Institute <strong>of</strong> Physics <strong>of</strong> the ACSR,<br />

v.v.i.<br />

jelinekp@fzu.cz<br />

Jesse Stephen Oak Ridge National Laboratory sjesse@ornl.gov<br />

Jhe Wonho Seoul National University whjhe@snu.ac.kr<br />

Kawai Shigeki University <strong>of</strong> Basel shigeki.kawai@unibas.ch<br />

Kim Woo-Joong <strong>Yale</strong> University Woo-Joong.Kim@yale.edu<br />

Kim Jeehoon Harvard University jeehoonkim@gmail.com<br />

Kin Nobuhiro Tohoku University kin@riec.tohoku.ac.jp<br />

Kinoshita Yukinori Graduate <strong>School</strong> <strong>of</strong> <strong>Engineering</strong>,<br />

Osaka Univ<br />

kinoshita@ap.eng.osaka-u.ac.jp<br />

Kiyohara Kosei Kanazawa University kosei@stu.kanazawa-u.ac.jp<br />

Klocke Michael University <strong>of</strong> Duisburg-Essen m.klocke@uni-duisburg.de<br />

Kobayashi Kei Kyoto University keicoba@iic.kyoto-u.ac.jp<br />

Kobayashi Shin-ichiro Research Institute <strong>of</strong> Electrical<br />

Communication<br />

kshin@atom.che.tohoku.ac.jp<br />

Koch Sascha University <strong>of</strong> Basel sascha.koch@unibas.ch<br />

Kose Rickmer SPECS USA Corp. rkose@specsus.com<br />

Kuroda Masaharu National Institute <strong>of</strong> Advanced<br />

Industrial Science and Technology<br />

(AIST)<br />

m-kuroda@aist.go.jp<br />

Kühnle Angelika University <strong>of</strong> Osnabrueck kuehnle@uos.de<br />

Labardi Massimiliano CNR-INFM labardi@df.unipi.it<br />

Lamoreaux Steve <strong>Yale</strong> University steve.lamoreaux@yale.edu<br />

Lange Manfred University <strong>of</strong> Duisburg-Essen manfred.lange@stud.uni-due.de<br />

Langewisch Gernot University <strong>of</strong> Muenster g.langewisch@uni-muenster.de<br />

Lengel George SPECS USA Corp. lengel@nanonis.com<br />

Li Yanjun Osaka Univarsity liyanjun@ap.eng.osaka-u.ac.jp<br />

Loppacher Christian Aix Marseille University Christian.Loppacher@im2np.fr<br />

Loske Felix University <strong>of</strong> Osnabrueck floske@uos.de<br />

Maksymovych Petro Oak Ridge National Laboratory maksymovychp@ornl.gov<br />

Masago Akira Tohoku University masago@wpi-aimr.tohoku.ac.jp<br />

Mavrokefalos Anastassios Massachusetts Institute <strong>of</strong><br />

Technology<br />

anastass@mit.edu<br />

Milde Peter TU Dresden peter.milde@iapp.de<br />

Milton Kimball University <strong>of</strong> Oklahoma milton@nhn.ou.edu<br />

Minato Taketoshi International Advanced Research<br />

and Education Organization<br />

minato@atom.che.tohoku.ac.jp<br />

Miyahara Yoichi McGill University miyahara@physics.mcgill.ca<br />

Moenig Harry <strong>Yale</strong> University harry.moenig@yale.edu<br />

Moon Christopher Agilent Laboratories chris_moon@agilent.com<br />

Moores Brad University <strong>of</strong> Waterloo brad.moores@gmail.com<br />

Moriarty Philip University <strong>of</strong> Nottingham philip.moriarty@nottingham.ac.uk<br />

Morita Seizo Osaka University smorita@eei.eng.osaka-u.ac.jp<br />

Möller Rolf University <strong>of</strong> Duisburg-Essen rolf.b.moeller@uni-due.de<br />

Nagashima Ken Osaka University nagasima@eei.eng.osaka-u.ac.jp<br />

Naitoh Yoshitaka Osaka University naitoh@ap.eng.osaka-u.ac.jp<br />

Narayanaswamy Arvind Columbia University arvind.narayanaswamy@columbia.edu<br />

171


Nishida Shuhei Institute <strong>of</strong> Industrial Science,<br />

University <strong>of</strong> Tokyo<br />

snishida@iis.u-tokyo.ac.jp<br />

Onishi Hiroshi Kobe University oni@kobe-u.ac.jp<br />

Oral Ahmet Sabanci University aoral@sabanciuniv.edu<br />

Oyabu Noriaki Kyoto University oyabu@piezo.kuee.kyoto-u.ac.jp<br />

Palasantzas Georgios University <strong>of</strong> Groningen g.palasantzas@rug.nl<br />

Parsegian Adrian NIH parsegia@mail.nih.gov<br />

Perez Ruben Universidad Autonoma de Madrid ruben.perez@uam.es<br />

Pieper Hans Hermann University <strong>of</strong> Osnabrueck hapieper@uos.de<br />

Pilet Nicolas EMPA nicolas.pilet@empa.ch<br />

Podgornik Rudolf NIH podgornr@mail.nih.gov<br />

Porthun Steffen RHK Technology, Inc porthun@rhk-tech.com<br />

Rahe Philipp University <strong>of</strong> Osnabrueck prahe@uos.de<br />

Rasmussen Morten Aarhus University mkr@inano.dk<br />

Rasool Haider UCLA haider@chem.ucla.edu<br />

Reichling Michael University <strong>of</strong> Osnabrueck reichling@uos.de<br />

Sadewasser Sascha Helmholtz Zentrum Berlin sadewasser@helmholtz-berlin.de<br />

Sasahara Akira Japan Advanced Institute <strong>of</strong> Science<br />

and Technology<br />

sasahara@jaist.ac.jp<br />

Sawada Daisuke Graduate <strong>School</strong> <strong>of</strong> <strong>Engineering</strong>,<br />

Osaka University<br />

d-sawada@afm.eei.eng.osaka-u.ac.jp<br />

Schirmeisen Andre University <strong>of</strong> Muenster schirmeisen@uni-muenster.de<br />

Schütte Jens University <strong>of</strong> Osnabrueck jens.schuette@uos.de<br />

Schwarz Udo <strong>Yale</strong> University udo.schwarz@yale.edu<br />

Schwarz Alexander University <strong>of</strong> Hamburg aschwarz@physnet.uni-hamburg.de<br />

Schwendemann Todd <strong>Yale</strong> University todd.schwendemann@yale.edu<br />

Shen Sheng MIT sshen1@mit.edu<br />

Siria Alessandro CNRS alessandro.siria@grenoble.cnrs.fr<br />

Staffier Peter <strong>Yale</strong> University peter.staffier@yale.edu<br />

Sugawara Yasuhiro Osaka University sugawara@ap.eng.osaka-u.ac.jp<br />

Sugimoto Yoshiaki Graduate <strong>School</strong> <strong>of</strong> <strong>Engineering</strong> ysugimoto@afm.eei.eng.osaka-u.ac.jp<br />

Sushkov Alexander <strong>Yale</strong> University alex.sushkov@yale.edu<br />

Suzuki Kazuhiro Kyoto University suzuki@piezo.kuee.kyoto-u.ac.jp<br />

Sweetman Adam University <strong>of</strong> Nottingham ppxams@nottingham.ac.uk<br />

Tang Hong <strong>Yale</strong> University hong.tang@yale.edu<br />

Tomas Herruzo Elena IMM-CSIC elena.tomas@imm.cnm.csic.es<br />

Tomitori Masahiko Japan Advanced Institute <strong>of</strong> Science<br />

and Technology<br />

tomitori@jaist.ac.jp<br />

Torricelli Gauthier University <strong>of</strong> Leicester gt47@le.ac.uk<br />

Trevethan Thomas University College London t.trevethan@ucl.ac.uk<br />

Tsukada Masaru Tohoku University tsukada@wpi-aimr.tohoku.ac.jp<br />

Tsunemi Eika Kyoto University eika@piezo.kuee.kyoto-u.ac.jp<br />

van Zwol Peter Groningen petervanzwol@gmail.com<br />

Venkataramani Krithika Aarhus University krithika@inano.dk<br />

Werle Christoph University <strong>of</strong> Basel christop.werle@unibas.ch<br />

Williamson Ryan Saint Louis University williamson.ryan@gmail.com<br />

Wintjes Nikolai University <strong>of</strong> Duisburg-Essen nikolai.wintjes@uni-due.de<br />

Workman Richard Agilent Laboratories richard_workman@agilent.com<br />

172


Wright Alan University <strong>of</strong> Maryland College Park cawright@umd.edu<br />

Yurtsever Ayhan Osaka University ayhan@afm.eei.eng.osaka-u.ac.jp<br />

Zahl Percy Brookhaven National Laboratory pzahl@bnl.gov<br />

Zerweck Ulrich University <strong>of</strong> Technology Dresden ulrich.zerweck@iapp.de<br />

173


Advertisement<br />

174


Ultra-high precision spatial positioning <strong>of</strong> objects is <strong>of</strong> prime importance in the emerging field <strong>of</strong><br />

nanotechnology. attocube systems, a German company located in Munich, manufactures and provides<br />

ultra-high precision spatial positioning systems and complete probing tools which are particularly suitable<br />

for extreme environmental conditions such as cryogenic temperatures (10 mK – 300 K), high magnetic<br />

fields (+31 T) and ultra high vacuum environments (5 x 10-11 mbar).<br />

Two principal product lines which meet today’s market demands are currently <strong>of</strong>fered:<br />

First, ultra-compact nano-precise positioning devices provide linear and rotational movement <strong>of</strong> samples<br />

or probes. They are <strong>of</strong>fered in different sizes and out <strong>of</strong> a variety <strong>of</strong> materials, and feature an<br />

unprecedented variety <strong>of</strong> applications particularly suitable for extreme environmental conditions. This<br />

represents a revolutionary advancement for the positioning market, leading to new research in many<br />

areas. Applications <strong>of</strong> these outstanding nanopositioning modules, well-known in many labs around the<br />

world, include scanning probe techniques such as scanning electron microscopy, confocal microscopy,<br />

scanning force microscopy, scanning tunneling microscopy and near-field optical microscopy, to name just<br />

a few. Furthermore, they are, for instance, suitable for general beam manipulation applications involving<br />

optical fibers and solid state waveguides.<br />

The second product line contains easy-to-use, highly flexible low temperature Scanning Probe<br />

Microscopes like LT-AFM, LT-CFM, LT-SNOM, or LT-STM. These systems are based on these reliable<br />

positioning devices. Thus, the microscopist can perform in-situ coarse and fine positioning, smooth<br />

scanning or automatically focusing any samples in respect to any probes at temperatures down to the<br />

Millikelvin range, at high magnetic fields or under high vacuum conditions.<br />

The product range is completed by innovative and highly flexible control systems for multiple SPM modes.<br />

Various SPM hardware and s<strong>of</strong>tware modules make image acquiring a simple task. 2D and 3D s<strong>of</strong>tware<br />

allow image processing for visually appealing, pr<strong>of</strong>essional and publishable results.<br />

attocube systems AG is focused on three markets. First, they provide nano-positioning systems for the<br />

R&D sector in the field <strong>of</strong> low temperature science. Second, the company produces microscopes capable<br />

for low-temperature applications in R&D. Finally attocube systems AG captures the markets <strong>of</strong> hightech<br />

industries that look for the qualities <strong>of</strong> these positioning systems, especially for the UHV-compatibility <strong>of</strong><br />

attocube systems’ products.<br />

For further technical information concerning attocube systems’ products, please visit the website<br />

www.attocube.com.


Nanosurf ® Saphyr<br />

Your Ultimate “Laboratory-in-a-Box” Signal Analyzer/Controller<br />

The Nanosurf Saphyr is a completely modular digital signal-processing unit that can be used<br />

as an SPM sensor controller for a wide range <strong>of</strong> applications. Its signal modules can be set to<br />

work independently or to use their analysis power on the same microscopy signals. High-end<br />

analog preconditioning stages and digital inter-module communication provide an as <strong>of</strong> yet<br />

unseen usable signal range. Controlled by an easy-to-use s<strong>of</strong>tware interface, the Saphyr is your<br />

all-in-one signal-processing solution.<br />

Go beyond the limits <strong>of</strong> your current measurement capabilities:<br />

• Generate sub-Angstrom amplitudes with the Saphyr PLL Modules and SignalCompensator<br />

• Detect even the smallest forces with the KPFM Controller Module and SignalSuppressor<br />

• Track harmonic frequencies with up to four PLL or Lock-In Modules and SignalSync<br />

• Unify control over multiple cantilevers or scan heads in a single powerful system<br />

Get the most out <strong>of</strong> your investment:<br />

• Combine the Saphyr with any home-built or commercial scan controller and scan head<br />

• Immediately update system characteristics and features through a simple file download<br />

• Extend your system’s capabilities with cost-effective modules to match your growing needs<br />

www.nanosurf.com/saphyr


AD VT-QPlus_ONUS / 09-May<br />

Nanotechnology is our Pr<strong>of</strong>ession!<br />

Z<br />

dF<br />

D<br />

The Si(111) 7x7 surface with<br />

QPlus<br />

Beam<br />

AFM in an Omicron VT at RT<br />

Deflection AFM<br />

(raw data).<br />

with QPlus!<br />

- Expanding the capabilities <strong>of</strong> your Omicron VT<br />

- Ideal sensor for the detection <strong>of</strong> short range forces<br />

- QPlus replaces needle sensor technology<br />

- Provides 2 orders <strong>of</strong> magnitude better force sensitivity<br />

Height [pm]<br />

30<br />

20<br />

10<br />

0<br />

-10<br />

0.5<br />

1 1.5 2 2.5 3<br />

Position [nm]<br />

Imaging <strong>of</strong> rest atoms (raw data) <strong>of</strong> the Si(111) 7x7 surface with QPlus AFM in an Omicron<br />

VT at 50 K (Left image). Line pr<strong>of</strong>ile showing the height difference <strong>of</strong> the rest atoms (above).<br />

Measurement: A. Bettac, Omicron NanoTechnology<br />

Z<br />

dF<br />

Germany: Omicron NanoTechnology GmbH<br />

Tel: +49(0) 6128 / 987 - 0<br />

e-mail: info@omicron.de<br />

Web: www.omicron.de<br />

USA: Omicron NanoTechnology USA<br />

Tel: 952 345 5240<br />

e-mail: info@omicronUS.com<br />

D<br />

The Si(111) 7x7 surface with<br />

QPlus AFM in an Omicron VT at RT<br />

(raw data).


Imaging the Future <strong>of</strong> Nanoscience<br />

World leading technology in UHV Scanning Probe Microscopes, SPM Control Systems, and<br />

Complete Surface Analysis Systems. RHK Technology sets the standard for powerful, flexible<br />

solutions for researchers pushing the limits <strong>of</strong> fundamental science at the atomic scale<br />

PLLPro 2 AFM Control System<br />

RHK’s next generation <strong>of</strong> AFM control.<br />

Ultra high performance and KFM capability in a one box solution.<br />

IDEAL INTERFACE IHDL TM<br />

RHK’s unique Iconic Hardware Description Language unites<br />

flexibility, function, and simplicity for ideal experimental<br />

configuration. Fast, intuitive “drag-and-drop” setup uses icons<br />

to represent hardware blocks. Simply connect icons to design<br />

the experiment. Devices automatically configure and connect<br />

when the template is implemented.<br />

PURPOSE BUILT HARDWARE<br />

ONE BOX INTEGRATION<br />

integrated surface science solutions<br />

Air-Cleaved Mica NC-AFM<br />

Large regularly shaped islands on air-cleaved mica surfaces after<br />

degassing in UHV at 473 K for 14 h. NC-AFM images recorded in<br />

constant detuning mode. F. Ostendorf, C. Schmitz, S. Hirth, A. Kuhnle,<br />

Kolodziej, M. Reichling<br />

Tuning Fork AFM on Silicon<br />

RHK UHV 325 with RHK SPM 1000 and PLLPro Controllers.<br />

True constant height image with current and ∆F(Hz) acquisition<br />

amplitude <strong>of</strong> oscillation, 3Å peak to peak. S. Pryadkin, RHK.<br />

RHK Technology, Inc. | 1050 East Maple, Troy, MI 48083 | 248.577.5426 | www.rhk-tech.com | Visit our booth at NC-AFM 2009


Surface Analysis<br />

Technology<br />

Aarhus STM<br />

State-<strong>of</strong>-the-Art STM Technology<br />

for Advanced SPM Research<br />

Key Features:<br />

� Quartz oscillator based design: sensor (tungsten tip)<br />

separated from oscillator: force and tunneling<br />

signals simultaneously and indepentently<br />

� Imaging <strong>of</strong> conductive and non-conductive samples<br />

in non-contact mode<br />

Vacuum<br />

Components<br />

� <strong>Atomic</strong>ally resolved imaging on a daily basis<br />

JT-STM<br />

Ultra-low temperature Joule Thomson-STM<br />

for ultimate STM spectroscopy and manipulation<br />

Key Features:<br />

� Designed at Aarhus University, group <strong>of</strong><br />

Pr<strong>of</strong>. F. Besenbacher<br />

� Extreme stability<br />

� Highest productivity<br />

� Variable temperature: 115 K to 400 K<br />

(for STM optional: 1300 K)<br />

� One-tip-suits-all: in situ tip sputtering<br />

Key Features:<br />

� Cryostat designed by Pr<strong>of</strong>. W. Wulfhekel<br />

� Base temperature: < 1 K (optional < 500 mK)<br />

� LHe hold time: > 4 days<br />

� NANONIS control package<br />

� Magnetic field up to 3 Tesla<br />

Applications:<br />

Surface Analysis<br />

System S<strong>of</strong>tware<br />

AFM Option<br />

Combining KolibriSensor TM and<br />

NANONIS Control Package<br />

� Inelastic electron tunneling spectroscopy<br />

� Investigation <strong>of</strong> superconductors<br />

� Quantum structures<br />

� Spin polarized tunneling<br />

� <strong>Atomic</strong> manipulation<br />

� Molecular electronics<br />

Computer<br />

Technology


First Announcement:<br />

13th International Conference on<br />

<strong>Noncontact</strong> <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong><br />

Kanazawa, Japan<br />

13th International Conference on<br />

<strong>Noncontact</strong> <strong>Atomic</strong> <strong>Force</strong> <strong>Microscopy</strong> (ncAFM2010)<br />

and<br />

Satellite Workshops on<br />

bio-imaging,<br />

solid-liquid interface,<br />

SPM standardization<br />

Venue:<br />

Ishikawa Ongakudo, Kanazawa, JAPAN<br />

Date:<br />

July 31 - August 1, 2010 (Satellite Workshops)<br />

August 2 - 5, 2010 (ncAFM2010)<br />

Conference Chair:<br />

Toyoko Arai (Kanazawa Univ.)<br />

Program committee:<br />

Masayuki Abe (Osaka Univ.)<br />

Proceedings committee:<br />

Satoshi Watanabe (Univ. <strong>of</strong> Tokyo)<br />

Web page:<br />

http://www.afm.eei.eng.osaka-u.ac.jp/ncafm2010/<br />

Kanazawa information:<br />

http://www.kanazawa-tourism.com/<br />

Co-organizers (tentative):<br />

167th Committee on Nano-Probe Technology <strong>of</strong> JSPS<br />

The Surface Science Society <strong>of</strong> Japan<br />

Osaka University Global Center <strong>of</strong> Excellence Program<br />

Center for Electronic Device Innovation


Notes


Program Summary <strong>of</strong> the NC-AFM 2009 Satellite Workshop on<br />

Casimir <strong>Force</strong>s and Their Measurement<br />

Monday sessions only; Tuesday sessions are joint with the NC-AFM Conference (see the main<br />

conference program for listing). Note in particular that the Casimir workshop’s second invited talk,<br />

given by Adrian Parsegian (NIH, USA), will take place Tuesday 14:30-15:10.<br />

Monday, August 10<br />

7:45 Registration<br />

8:50 Opening Remarks<br />

9:00 Rudi Porgornik (University <strong>of</strong> Ljubljana, Slovenia), invited<br />

9:30 Noah Graham (Middlebury College, USA)<br />

9:55 Kimball Milton (University Oklahoma, USA)<br />

10:20<br />

C<strong>of</strong>fee Break<br />

10:50 Raul Esquivel-Sirvent (Universidad Nacional Autónoma, Mexico)<br />

11:15 Diego Dalvit (Los Alamos National Laboratory, USA)<br />

11:40 Rick Rajter (MIT, USA)<br />

12:05 Maarten de Boer (Sandia National Laboratory, USA)<br />

12:30<br />

Lunch (12:30-14:30)<br />

14:30 Hong Tang (<strong>Yale</strong> University, USA)<br />

14:55 Alessandro Siria (CNRS, France)<br />

15:20 Arvind Narayanaswamy (Columbia University, USA)<br />

15:45<br />

C<strong>of</strong>fee Break<br />

16:15 Sven de Man (VU University Amsterdam, Netherlands)<br />

16:40 Gauthier Torricelli (University <strong>of</strong> Leicester, UK)<br />

17:05 Ho Bun Chan (University <strong>of</strong> Florida, USA)<br />

17:30 Peter van Zwol (University <strong>of</strong> Groningen, Netherlands)<br />

17:55 Woo-Joong Kim (<strong>Yale</strong> University, USA)<br />

18:30-20:30<br />

Welcome Reception with Dinner


NC-AFM 2009<br />

Sessions<br />

Monday<br />

August 10<br />

�� Liquids I<br />

Tuesday 9:20-10:40<br />

�� <strong>Force</strong> Spectroscopy I<br />

Tuesday 11:20-12:40<br />

�� Electronic, photonic, & Casimir forces<br />

Tuesday 14:30-15:50<br />

�� Oxides<br />

Wednesday 9:00-10:40<br />

�� Carbon-based materials<br />

Wednesday 11:20-12:40<br />

Tuesday<br />

August 11<br />

Wednesday<br />

August 12<br />

�� Molecules on insulators<br />

Wednesday 14:30-15:50<br />

�� <strong>Force</strong> spectroscopy II<br />

Thursday 9:00-10:40<br />

�� <strong>Force</strong>s & charges<br />

Thursday 11:20-13:00<br />

�� Method development<br />

Friday 9:00-10:40<br />

�� Liquids II<br />

Friday 11:20-12:40<br />

Thursday<br />

August 13<br />

Friday<br />

August 14<br />

8:30 Registration Registration Registration Registration<br />

9:00 Opening Remarks M. Heyde A. Campbellova Y. Naitoh<br />

9:20 T. Fukuma A. Yurtsever F. J. Giessibl S. Torbrügge<br />

9:40 H. Asakawa R. Pérez G. Langewisch Y. Hosokawa<br />

10:00 N. Oyabu R. Bechstein A. Barat<strong>of</strong>f E. Tsunemi<br />

10:20 M. Tsukada S. Sadewasser S. Kawai U. Zerweck<br />

10:40 C<strong>of</strong>fee Break C<strong>of</strong>fee Break C<strong>of</strong>fee Break C<strong>of</strong>fee Break<br />

11:20 D. Ebeling U. D. Schwarz A. Sweetman S. Nishida<br />

11:40 A. Schirmeisen P. Jelínek L. Gross S. Ido<br />

12:00 Y. Sugimoto M. Ashino Y. Miyahara E. T. Herruzo<br />

12:20 Y. Sugawara P. Pou T. Trevethan K. Fukui<br />

12:40 A. Bettac Closing Remarks<br />

13:00<br />

Lunch<br />

110 min.<br />

(12:40-14:30)<br />

Lunch<br />

110 min.<br />

(12:40-14:30)<br />

14:30<br />

A. Parsegian<br />

A. Kühnle<br />

14:50<br />

(invited)<br />

A. Schwarz<br />

15:10 Th. Glatzel Ch. Loppacher<br />

15:30 H. Hölscher C. J. Gómez<br />

15:50<br />

|<br />

18:00<br />

C<br />

A<br />

S<br />

I<br />

M<br />

I<br />

R<br />

W<br />

O<br />

R<br />

K<br />

S<br />

H<br />

O<br />

P<br />

18:30-20:30<br />

Welcome<br />

Reception/Dinner<br />

Poster Session I Poster Session II<br />

Lunch<br />

90 min.<br />

(13:00-14:30)<br />

18:30-22:00<br />

Conference<br />

Banquet

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!