• Keine Ergebnisse gefunden

Physical interpretation of frequency-modulation atomic force microscopy

N/A
N/A
Protected

Academic year: 2022

Aktie "Physical interpretation of frequency-modulation atomic force microscopy"

Copied!
4
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Physical interpretation of frequency-modulation atomic force microscopy

Franz J. Giessibl and Hartmut Bielefeldt

Universita¨t Augsburg, Institut fu¨r Physik, Experimentalphysik 6, D-86135 Augsburg, Germany 共Received 24 March 1999兲

Frequency modulation atomic force microscopy is a method for imaging the surface of metals, semiconduc- tors and insulators in ultrahigh vacuum with true atomic resolution. The imaging signal in this technique is the frequency shift ⌬f of an oscillating cantilever with eigenfrequency f0, spring constant k and amplitude A, which is subject to tip-sample forces Fts. Here, we present analytical results of⌬f ( f0,k,A) for several basic classes of Fts. With these results, a method to calculate images is derived and demonstrated with an example.

The scanning tunneling microscope 共STM兲 has provided spectacular images of conducting surfaces on the atomic scale.1 After the invention of atomic force microscopy 共AFM兲by Binnig, Quate, and Gerber in 1986,2the possibil- ity of imaging insulating surfaces with atomic resolution in real space seemed to be very close. True atomic resolution with static AFM on inert samples has been reported in 1993.3 However, imaging reactive surfaces like Si共111兲in ultrahigh vacuum by static AFM has proven to be impractical because of chemical bonding between tip and sample and wear on the atomic scale.4 The application of frequency modulation AFM 共FM-AFM兲 共Ref. 5兲with large amplitudes in noncon- tact mode has helped to overcome problems with tip-sample bonding and finally, the 7⫻7 reconstruction of Si 共111兲 could be resolved also by AFM.6–7 The initial experiments have been refined and the quality of FM-AFM images reaches that of the STM.8–10 Other semiconductors,10 ionic crystals,11 metal oxides,12 metals,13,14organic monolayers,15 a film of xenon physisorbed on graphite,16 and deliberately created defects on CaF2共111兲 共Ref. 17兲 have been imaged with atomic resolution. The first and second international workshop on noncontact AFM共NC-AFM98 in Osaka, Japan and NC-AFM99 in Pontresina, Switzerland兲 have helped to establish FM-AFM as a powerful experimental technique.

While the number of surfaces which are imaged by FM- AFM with atomic resolution is growing rapidly, open ques- tions in the quantitative interpretation of the FM-AFM re- main. The imaging signal in FM-AFM is the shift in resonance frequency ⌬f of an oscillating cantilever 共CL兲. Here, we derive for the first time analytical solutions of⌬f for three classes of tip-sample potential Vts: inverse power-, power- and exponential potentials. With this relations, we establish a powerful method for the calculation of FM-AFM images.

In FM-AFM, a CL with a very high-Q factor (⬇105), eigenfrequency f0 and spring constant k is subject to con- trolled positive feedback such that it oscillates with a con- stant amplitude A. When this CL is brought close to a sample, its frequency changes from f0 to ff0⫹⌬f . This frequency change ⌬f is used to create an image zCL– b(x,y ,f ) by scanning the CL in the xy plane while adjusting the z position of the base of the CL zCL– b such that

f stays constant 共see Fig. 1兲. Typical parameters are ⌬f

⬇⫺100 Hz k20 N/m, f0200 kHz and A⬇10 nm—see Table I in Ref. 18 for an overview.

The frequency shift ⌬f as a function of Fts⫽⫺⳵Vts/⳵z has been calculated by first order perturbation theory using the Hamilton-Jacobi approach19

fxCL– b, f0,k,A兲⫽⫺ f02 kA2

0

1/f0

Ftsxtipq

tdt 共1兲

with the CL deflection q

(t)A cos(2f0t) and xtip

⫽关xCL– bconst.,yCL– b,zCL– bq

(t)T. This result has been confirmed by other approaches.20–25First-order pertur- bation theory yields excellent results because max兩Vts兩 ⱗ10 eV, while E0.5kA2⬇5 keV. A Fourier approach q

(t)m0amcos(2␲mft) 共Refs. 20–22兲 shows that both the dc deflection (a0Af / f0) and the higher orders are very small 关兩am兩⭐2A/(1m2)兩⌬f / f0for m⭓2 with

f / f0ⱗ103compared to the fundamental amplitude a1

A. At the lower turnaround point of the tip xtipltpthe precise value of the deflection of the CL is given by the condition of constant energy E0.5kqltp

2Vts(xtipltp)⫽0.5kA2, thus qltp

AAVts(xtipltp)/2E. SinceVtsis small compared to E, the FM-AFM image zCL– b(x, y ,f ) is approximately equal to the map described by xtipltp.

The scaling properties of ⌬f are such that it is useful to introduce a ‘‘normalized frequency shift’’19

␥共z,A兲ªkA3/2

f0fz, f0,k,A兲. 共2兲 Substituting z

A关1⫹cos(2␲f0t)兴in Eq.共1兲we find

FIG. 1. Schematic of an oscillating cantilever共CL兲with tip and front atom共fa兲at its lower turnaround point共ltp兲close to a sample.

PHYSICAL REVIEW B VOLUME 61, NUMBER 15 15 APRIL 2000-I

PRB 61

0163-1829/2000/61共15兲/9968共4兲/$15.00 9968 ©2000 The American Physical Society

(2)

␥共xtipltp,A兲⫽ 1

&␲

0

2AFtsxtipltpz

ez

z

1⫺z

1A2Az

dz

. 共3兲

In typical experiments, A is very large compared to the range of Ftsand the second factor in the integral is close to unity in regions where Fts is nonvanishing. Therefore, the ‘‘large- amplitude’’ normalized frequency shift

lAxtipltp兲⫽

0

Ftsxtipltpz

ez

2z

dz

4 is a good approximation for ␥ for any class of tip-sample forces, provided that the range of the forces is small com- pared to A.

The interaction of a macroscopic tip of an AFM with a sample is a complicated many-body problem, which can be solved with molecular dynamics calculations.26 For a quali- tative analysis, quite realistic model forces can be con- structed from a chemical contribution modeling the interac- tion of the front atom 共fa兲with the sample Ffaplus the van der Waals 共vdW兲 contribution of the rest of the tip Fbackground.19Since␥is linear in Ftsit is practical to expand Fts in terms of basic types: 共a兲 inverse-power forces, 共b兲 power forces, and共c兲exponential forces and superimpose the individual solutions for ␥.

共a兲 Fts(z)C/zn 共Lennard-Jones potential, vdW forces, electrostatic and magnetic forces27兲. Insertion in Eq.共3兲and comparing the result with the integral representation of the hypergeometric function Fca,b(␵) 共Ref. 28兲yields:

␥共z,A兲⫽C

A

zn

F1n,0.5

z2A

F2n,1.5

z2A

冊册

. 5

For large amplitudes (A/zⰇ1) the transformation formula Fca,b(␵)⫽(1⫺␵)bFcb,ca(␵/(␵⫺1)) is useful 共15.5.9. in Ref. 28兲and

lAz兲⫽

21

n12

⌫共nC

zn1/2 共6兲 where⌫(n) is the Gamma function.28

共b兲 Fts(z)C(z)m for z0 and Fts(z)0 for z⬎0 共Hertzian contact forces with m⫽3/2 for a spherical tip on a flat surface and adhesion forces with m⫽0兲 共Ref. 27兲

␥⫽Czm1/2⌫共m⫹1兲

2

F⌫共m0.5,0.5m1.5

3/22Az

2Az F⌫共m0.5,1.5m2.5

5/22Az

.

共7兲 For large amplitudes

lAz兲⫽ ⌫共m⫹1兲

2␲⌫共m⫹3/2兲C共⫺zm1/2 for z⬍0. 共8兲 共c兲Fts(z)F0e⫺␬z共Morse potential兲

␥⫽F0e⫺␬z

AM10.5共⫺2␬A兲⫺M21.5共⫺2␬A兲兴, 共9兲 where Mba(␵) is the Kummer function.28 For large ampli- tudes (␬AⰇ1) we can use the asymptotic expression of Mba(␵) 共13.5.1. in Ref. 28兲and find

lAz兲⫽F0ez

2␲␬ . 10 On conductive samples, simultaneous STM and FM-AFM operation is possible. Since the bandwidth of the tunneling current (It) preamplifier is i.g. much smaller than f0 of the CL, the measured It is a time-average. With It(z)I0e⫺␬tz we find

Itz,A兲典I0e⫺␬tzM11/2共⫺2␬tA兲. 共11兲 When ␬tAⰇ1, 具It(z,A)/It(z,0)⬇1/

2␲␬tA yielding a re- duction to 1/35 for typical parameters 共A⫽10 nm and ␬t

⬇20 nm1for metals兲.

Figure 2 shows the ratio ␥(z,A)/lA(z) as a function of A/z and A, respectively. In a typical experiment, A/z and A␬ are between 10 and 1000 during imaging where␥lA(z) and ␥(z,A) are almost identical. As A/z becomes smaller, the sensitivity to chemical 共short range兲 forces increases. In addition to a better signal-to-noise ratio18 the use of smaller amplitudes makes sharp tips less important since the sensi- tivity to the macroscopic vdW forces decreases.

Equation 共4兲 shows that␥lA increases with the range of Ftsand␥has the dimension of N

m. Table I shows␥lAand Fts

␭ with range ␭for three basic types of forces laws. For exponential forces Fts(z)F0e⫺␬z the range ␭ defined by

FIG. 2. Ratio between normalized frequency shiftand large-amplitude approximationlAas a function of A/z and A, respectively.

TABLE I. Three basic types of tip-sample potentials and corre- sponding normalized frequency shift␥.

Basic type Vts(z) FtsVts/FtslA

Inv.-power n1, z⬎0

1 n⫺1

C zn⫺1

1

n⫺1 C zn0.5

n12

2␲⌫共nC zn0.5 Power

m0, z⬍0

C共⫺zm1 m⫹1

C共⫺zm0.5

m⫹1

⌫共m⫹1兲C共⫺zm0.5

2␲⌫共m⫹1.5兲 Exponential F0e⫺␬z

F0e⫺␬z

F0e⫺␬z

2␲␬

PRB 61 BRIEF REPORTS 9969

(3)

V(z⫹␭)/V(z)1/e is independent of z with ␭⫽V(z)/F(z)

⫽1/␬. For inverse power and power forces, an analog ex- pression␭powerªV(z)/F(z) depends on z, but it shows ana- log scaling properties, i.e., V(z⫹␭power)/V(z)1/e for ex- ponents n⬎1 共⫽1/e for n→⬁兲. Table I shows ␥lA and Fts

Vts/Fts. It is interesting to note that Fts

Vts/Fts/␥lA

2forn兩ⲏ2 and Fts

Vts/Fts/␥lA

2␲ for exponen- tial forces and power forces withn兩→⬁. Since the fre- quency shift is a linear function of Fts关Eq.共1兲兴, the total␥is a linear combination of the contributions of the basic types Ftsi

lAxtipltp兲⬇ 1

2

i Ftsi xtipltp

Vtsi xtipltp/Ftsi xtipltp, 12

where Vtsi is a basic type. Typical chemical forces are⬇⫺1 nN with a range of 0.1 nm and vdW forces are⬇⫺1 nN with a ‘‘range’’ of 1 nm, thus experimental␥

s are expected to be in the order of⫺10 fN

m.

To illustrate how Eq. 共12兲 can be used to calculate the FM-AFM image z(x,y ,␥), we consider a relatively simple system: an adsorbed layer of xenon. Xenon forms a closed packed film with next-neighbor distance ␴Xe⬇0.43 nm on graphite, thus the unit vectors of the xenon surface lattice are given by a1⫽(␴,0,0)T and a2⫽(␴/2,␴)/2,0)T. Allers et al.16have succeeded in imaging such a layer by FM-AFM:

the image shows the expected closed packed structure with a corrugation of 25 pm and atoms appearing high in the image, i.e. no contrast inversion. The following imaging parameters have been used:⌬f⫽⫺92 Hz, k35 N/m, f0⫽160 kHz and A⫽9.4 nm, i.e., at ␥⫽⫺18 fN

m.

The interaction of two xenon atoms can be modeled by a Lennard Jones potential

LJd兲⫽⫺Ebond

2

d

6

d

12

13

with Ebond⫽0.02 eV and ␴⫽0.433 nm.29 Since the tip was made of silicon, we assume that the front atom of the tip is

FIG. 3.Color兲 共aNormalized frequency shiftlA(x0,y ,z) from y⫽⫺2to 2(0.43 nm). Xenon atoms are situated at y0,⫾). The contour lines of constantare cross sections of the corresponding FM-AFM image z(x,y ,). Maximal corrugation is obtained for⬇⫺18 fNmgreen area.b Arrangement of Xenon atoms on the surface,ctop view oflAalong red track inbshowing a maximal corrugation of5 pm.

9970 BRIEF REPORTS PRB 61

(4)

also silicon. Krupp has shown, that the Hamaker constant of two materials is given by the geometric average between the individual Hamaker constants.30With further assuming, that the equilibrium distance in a silicon-xenon bond is given by the mean value between the bulk nearest neighbor distance in Si (␴Si⫽0.235 nm) and Xe, we can create a LJ model poten- tial for interaction of Si and Xe ␾LJ

SiXe

with EbondSiXe

⫽0.047 eV and␴SiXe0.33 nm. Fts is approximated by:

Ftsxfaltp兲⫽⫺ C zfaltp⫹␴Si

⫹ ⳵

zn,m

⫽⫺⬁

LJ SiXe

共兩xfaltpamn兩兲, 共14兲 where C is a constant 共the tip is assumed to be conical or pyramidal, thus the long-range component is given by

C/z, see Refs. 19 and 31and amnna1ma2. For the calculation of ␥lA(x,y ,z), the attractive partLJattd6 and repulsive part ␾LJrepd12 have to be treated separately.32 Figure 3共a兲 shows ␥lA(x0,y ,z) from y

⫽⫺2␴to y⫽2␴. The contour lines correspond to cross sec- tions of the image z(x,y⫽0,␥⫽const.). Stable operation of the microscope is only possible in a z range where ⳵␥/z

⬍0. The maximal corrugation occurs at ␥optimal

⬇⫺18 fN

m and is ⬇5 pm 关Fig. 3共c兲兴. It is noted that the absolute value of ␥optimal depends strongly on Chere C

⫽8.6⫻1019J兲 which is a function of the macroscopic tip shape. However, the maximal corrugation varies little with C and is quite insensitive to the parameters we have calculated for the front atom-sample potential ␾LJ

SiXe. The deviation between the corrugations in theory and experiment is prob- ably due to elastic sample deformation. We assume that the tip is rigid 共i.e., xfaxtipis constant—the force constant be-

tween next neighbors in Si is 170 N/m兲, but the vdW bonds in the Xe are weak (⳵2LJ/⳵d2⬇1 N/m). We believe that the corrugation is strongly enhanced because the Xe atoms are pulled out of the surface. A similar effect has been ob- served in STM where the theoretical corrugation for Al共111兲 was ⬃1 pm while the experimental values were⬃50 pm.33

In summary, we have found a physical interpretation of large amplitude FM-AFM: the images are a map z(x, y ,

⫽const.) where ␥(x,y ,z)⫽兰0Fts(x,y ,zz

)/

22z

dz

. When Fts is known in terms of basic 共i.e., power, inverse power, exponential兲 force types Ftsi(z),

⫽兺iFtsi (z)

i/2␲ whereiVtsi (z)/Ftsi (z). Our analytic results for various basic types of tip-sample forces establish the validity of the large-amplitude approximation for the in- terpretation of images and allow a quantitative analysis of

f (z) curves for larger z values 共i.e., 0⬍A/z⬍100兲. We have further found an analytic result for the dependence of the tunneling current as a function of amplitude. Calculations with molecular dynamics, which currently yield Fts(x, y ,z) 共Ref. 26兲 can be extended to compute the observables in FM-AFM, namely the experimental FM-AFM images z(x,y ,) with Eq.共4兲and␥(x,y ,z,A) curves with Eq.共3兲on specific spots x,y on the sample 共e.g. adatom sites and cor- nerhole centers兲. Comparing the experimental results with these calculations, the force vs distance characteristics of chemical bonds between front atom and samples can be di- rectly derived.

We thank A. Baratoff, P. van Dongen, U. Du¨rig, S. Hem- bacher, U. Mair, and J. Mannhart for discussions and W.

Allers et al. for the preprint of Ref. 16. This work is sup- ported by the BMBF共Project No. 13N6918/1兲.

1G. Binnig et al., Phys. Rev. Lett. 49, 57共1982兲; 50, 120共1983兲.

2G. Binnig et al., Phys. Rev. Lett. 56, 930共1986兲.

3F. Ohnesorge and G. Binnig, Science 260, 1451共1993兲.

4L. Howald et al., Z. Phys. B: Condens. Matter 93, 267共1994兲.

5T. R. Albrecht et al., J. Appl. Phys. 69, 668共1991兲.

6F. J. Giessibl, Science 267, 68共1995兲; S. Kitamura and M. Iwat- suki, Jpn. J. Appl. Phys., Part 2 34, L145共1995兲.

7P. Gu¨thner, J. Vac. Sci. Technol. B 14, 2428共1996兲.

8R. Lu¨thi et al., Z. Phys. B: Condens. Matter 100, 165共1996兲.

9R. Erlandsson et al., Phys. Rev. B 54, R8309共1996兲.

10Y. Sugawara et al., Science 270, 1648共1995兲.

11M. Bammerlin et al., Probe Microsc. 1, 3共1997兲.

12K. Fukui et al., Phys. Rev. Lett. 79, 4202共1997兲; H. Raza et al., ibid. 82, 5265共1999兲.

13Ch. Loppacher et al., Appl. Surf. Sci. 140, 287共1999兲.

14S. Orisaka et al., Appl. Surf. Sci. 140, 243共1999兲.

15B. Gotsmann et al., Eur. Phys. J. B 4, 267共1998兲.

16W. Allers et al., Europhys. Lett. 48, 276共1999兲.

17M. Reichling and C. Barth, Phys. Rev. Lett. 83, 768共1999兲.

18F. J. Giessibl et al., Appl. Surf. Sci. 140, 352共1999兲.

19F. J. Giessibl, Phys. Rev. B 56, 16 010共1997兲.

20A. Baratoff共unpublished兲.

21A. I. Livshits et al., Phys. Rev. B 59, 2436共1999兲.

22U. Du¨rig, Surf. Interface Anal. 27, 467共1999兲.

23H. Ho¨lscher et al., Appl. Surf. Sci. 140, 344共1999兲.

24N. Sasaki and M. Tsukada, Jpn. J. Appl. Phys., Part 2 37, L533 共1998兲; Appl. Surf. Sci. 140, 339共1999兲.

25L. Wang, Appl. Phys. Lett. 73, 3781共1998兲.

26R. Perez et al., Phys. Rev. Lett. 78, 10 835共1997兲.

27J. Israelachvili, Intermolecular and Surface Forces, 2nd ed.共Aca- demic, London, 1991兲.

28M. Abramowitz and I. Stegun, Handbook of Mathematical Func- tions, 9th ed.共Dover, New York, 1970兲.

29N. Ashcroft and N. D. Mervin, Solid State Physics 共Saunders College, Philadelphia, 1981兲, p. 398.

30H. Krupp, Adv. Colloid Interface Sci. 1, 113共1967兲.

31U. Hartmann, J. Vac. Sci. Technol. B 9, 465共1991兲.

32For␾(d)C/dn with d⫽冑x2y2z2, ⳵␾/⳵z⫽⫺NC/dn1 if x,y0 but the deviation is negligible for zx,y .

33J. Winterlin et al., Phys. Rev. Lett. 62, 59共1989兲.

PRB 61 BRIEF REPORTS 9971

Referenzen

ÄHNLICHE DOKUMENTE

I will focus on exploring the adhesion strength of as-synthesized calcium fluoride nanoparticles adsorbed on mica and on tooth enamel in liquid with Amplitude Modulation AFM

Figure 6.3: Correlation of the pairwise Euclidean distances based on vectors consisting either of all of the sorted elements of the overlap matrix (a) or eigenvalues of this matrix

Since the contact resonance frequency of the normal and torsional mode oscillations are tracked simultaneously to the lateral force, a contact resonance map is generated in addition

Atomic Force Microscopy Amplitude Modulation Atomic Force Microscopy Frequency Modulation Atomic Force Microscopy non-contact Atomic Force Microscopy Current Imaging

In summary, a method to obtain the tip–sample potential from frequency shift data has been introduced and demon- strated, and an intuitive connection between frequency shift

True atomic resolution of conductors and insulators is now routinely obtained in vacuum by frequency modulation atomic force microscopy. So far, the imaging parameters

Published frequency versus distance data are used to show that the apex of tips providing atomic resolution is faceted and not rounded!. Further, an extended jump-to-contact

There are three externally adjustable pa- rameters that affect the imaging process: the oscillation amplitude, the set point of the frequency shift, and the bias between tip