0712.3792/tm.tex
1: \def\deg      {{^\circ \/}}
2: \def\H        {{$^1$H \/}}
3: \def\F        {{$^{19}$F \/}}
4: \def\P        {{$^{31}$P \/}}
5: \def\CaF      {{CaF$_2$ \/}}
6: \def\etal     {{\it et al. \/}}
7: \def\abinitio {{\it ab initio \/}}
8: \def\eg       {{\it e.g. \/}}
9: \def\ie       {{\it i.e. \/}}
10: \def\first    {{1$^{\rm st}$ \/}}
11: \def\second   {{2$^{\rm nd}$ \/}}
12: \def\third    {{3$^{\rm rd}$ \/}}
13: \def\th       {{$^{\rm th}$ \/}}
14: 
15: \newcommand{\ee}[1]{\cdot10^{#1}}
16: \newcommand{\mr}[1]{\mathrm{#1}}
17: \newcommand{\unit}[1]{\,\mathrm{#1}}
18: \newcommand{\um}{\,\mu{\rm m}}
19: \newcommand{\us}{\,\mu{\rm s}}
20: \newcommand{\kT}{k_{\rm B}T}
21: \newcommand{\kB}{k_{\rm B}}
22: 
23: %\newcommand{\wc}{\omega_c}
24: %\newcommand{\wci}{\omega_{c,i}}
25: \newcommand{\fci}{f_{i}}
26: \newcommand{\kci}{k_{i}}
27: \newcommand{\Qci}{Q_{i}}
28: \newcommand{\tm}{\tau_\mr{m}}
29: \newcommand{\itm}{\tau_\mr{m}^{-1}}
30: \newcommand{\wrf}{\omega_\mr{rf}}
31: \newcommand{\weff}{\omega_\mr{eff}}
32: \newcommand{\theff}{\theta_\mr{eff}}
33: \newcommand{\wo}{\omega_0}
34: \newcommand{\wone}{\omega_1}
35: \newcommand{\wmod}{\omega_\mr{mod}}
36: \newcommand{\Dw}{\Delta\omega}
37: \newcommand{\Gone}{\Gamma_1}
38: \newcommand{\Gr}{\Gamma_{1\rho}}
39: \newcommand{\SB}{S_B}
40: \newcommand{\DE}{\Delta E}
41: \newcommand{\Tr}{T_{1\rho}}
42: \newcommand{\dB}{\nabla|\bm B|}
43: \newcommand{\gradB}{\vec{\nabla}|\vec{B}|}
44: \newcommand{\dxB}{\partial \vec{B}/\partial x}
45: %\newcommand{\dxBx}{\partial B_x/\partial x}
46: %\newcommand{\dxBy}{\partial B_y/\partial x}
47: \newcommand{\dxBz}{\partial B_z/\partial x}
48: \newcommand{\dxBu}{\frac{\partial {\bm B}}{\partial x}}
49: %\newcommand{\dxBxu}{\frac{\partial B_x}{\partial x}}
50: %\newcommand{\dxByu}{\frac{\partial B_y}{\partial x}}
51: \newcommand{\dxBzu}{\frac{\partial B_z}{\partial x}}
52: \newcommand{\dxBzm}{\overline{\partial B_z/\partial x}}
53: \newcommand{\dxBzum}{\overline{\frac{\partial B_z}{\partial x}}}
54: %\newcommand{\dxBzum}{\overline{\left(\frac{\partial B_z}{\partial x}\right)}}
55: \newcommand{\SBz}{S_{Bz}(\wone)}
56: \newcommand{\Geff}{G_\mr{eff}}
57: \newcommand{\Btip}{{\bm B}_\mr{tip}}
58: \newcommand{\Btipz}{B_\mr{tip,z}}
59: \newcommand{\Bext}{B_\mr{ext}}
60: \newcommand{\vBtip}{\vec{B}_\mr{tip}}
61: \newcommand{\vBext}{\vec{B}_0}
62: \newcommand{\vecr}{{\bm r}}
63: \newcommand{\xrms}{x_\mr{rms}}
64: \newcommand{\Tmode}{T_\mr{mode}}
65: \newcommand{\Tmodet}{T_\mr{mode,3}}
66: \newcommand{\ez}{{\bm e}_z}
67: 
68: 
69: %\documentclass[preprint,prl,aps,showpacs]{revtex4}% Phys Rev Lett
70: \documentclass[prl,aps,twocolumn,showpacs]{revtex4}% Phys Rev Lett
71: 
72: %\usepackage[dvips]{graphicx}% Include figure files
73: \usepackage{graphicx}
74: %\usepackage{dcolumn}% Align table columns on decimal point
75: \usepackage{bm}% bold math
76: %\usepackage{epsfig}
77: \usepackage{amsmath}
78: \usepackage{amssymb}
79: %\usepackage{footnpag}
80: \usepackage[colorlinks=true, pdfstartview=FitV, linkcolor=blue, citecolor=blue, urlcolor=blue]{hyperref} % enable links
81: 
82: \bibliographystyle{apsrev.bst}
83: 
84: \begin{document}
85: 
86: % formatting
87: \global\emergencystretch = .1\hsize % adjust the line-breaking to avoid overfull \hbox (standard .15\hsize)
88: 
89: \title{Nuclear spin relaxation induced by a mechanical resonator}
90: 
91: \author{C. L. Degen$^{1}$}
92:   \email{degenc@gmail.com} 
93: \author{M. Poggio$^{1,2}$, H. J. Mamin$^1$, and D. Rugar$^1$}
94:   \affiliation{
95:    $^1$IBM Research Division, Almaden Research Center, 650 Harry Road, San Jose CA, 95120, USA. \\
96:    $^2$Center for Probing the Nanoscale, Stanford University, 476 Lomita Hall, Stanford CA, 94305, USA.}
97: \date{\today}
98: 
99: \begin{abstract}
100: We report on measurements of the spin lifetime of nuclear spins strongly coupled to a micromechanical cantilever as used in magnetic resonance force microscopy. We find that the rotating-frame correlation time of the statistical nuclear polarization is set by the magneto-mechanical noise originating from the thermal motion of the cantilever. Evidence is based on the effect of three parameters: (1) the magnetic field gradient (the coupling strength), (2) the Rabi frequency of the spins (the transition energy), and (3) the temperature of the low-frequency mechanical modes.
101: Experimental results are compared to relaxation rates calculated from the spectral density of the magneto-mechanical noise.
102: \end{abstract}
103: 
104: \pacs{76.60.-k, 85.85.+j, 05.40.Jc}
105: 
106: \maketitle
107: 
108: Sensitive detection of nuclear spin signals requires a sensor that can couple strongly to the weak nuclear magnetic moment.
109: For inductively detected nuclear magnetic resonance (NMR), where a resonant coil picks up the rf signal of the precessing spins,
110: this is achieved by scaling down the coil size so as to improve the current-per-flux ratio \cite{hoult76}.
111: Optimized inductive detectors are currently able to observe ensembles containing roughly $10^{12}$ proton spins,
112: equivalent to about $10^8$ net magnetic moments \cite{seeber01}.
113: Stronger couplings enabling higher spin sensitivity have been shown with magnetic force sensors,
114: such as used in magnetic resonance force microscopy (MRFM)  \cite{sidles95,rugar04}, which recently
115: measured the net moment of about $10^3$ nuclei \cite{mamin07}.
116: 
117: In order to detect spins with still higher sensitivity, the detector must be coupled even more tightly to the nuclear magnetic moment.
118: For a mechanical detector sensitive to forces in the $x$-direction, this is achieved by increasing the magnetic field gradient $\dxBu$,
119: which generates the magnetic force $F_x = {\bm \mu}\cdot\dxBu$,
120: where ${\bm \mu}$ is the total magnetic moment of the spin ensemble.
121: Present MRFM technology employs gradients up to $10^6\unit{T/m}$ \cite{mamin07}.
122: If the gradient can be successfully pushed into the range of $10^8\unit{T/m}$,
123: force detection of single nuclear spins may become feasible.
124: 
125: The strong interaction between spins and sensor, however, also increases the ``back action'' of the sensor on the spins and makes
126: them more susceptible to detector noise. Here we are concerned with magnetic noise generated by the thermally vibrating
127: cantilever, but similar effects can be expected in any real-world detector, as for example in an inductively-coupled rf circuit.
128: The interaction between a mechanical resonator and spins has been the subject of a number of theoretical studies,
129: and is predicted to lead to a host of intriguing effects. These range from shortening of spin lifetimes \cite{mozyrsky03,berman03},
130: to spin alignment by specific mechanical modes either at the Larmor frequency or in the rotating frame \cite{magusin00,butler05},
131: to resonant amplification of mechanical oscillations \cite{bargatin03}.
132: 
133: In this letter, we report direct experimental evidence for accelerated nuclear spin relaxation induced by a single, low-frequency mechanical mode.
134: The observed relaxation has its origin in the random magnetic field ${\bm B}(t) = \dxBu x(t)$
135: created by the thermal (Brownian) motion $x(t)$ of the cantilever in a large field gradient.
136: The nuclear spin and mechanical oscillator degrees of freedom are well decoupled in strong static fields,
137: since the spin precession frequency is orders of magnitude higher than typical cantilever frequencies.
138: However, when an on-resonance transverse rf magnetic field $B_1$ is present,
139: kHz frequency noise that overlaps with the Rabi frequency of the spin will induce spin relaxation.
140: The rate of nuclear spin transitions may be described by the rotating-frame relaxation time $\Tr$,
141: %
142: \begin{equation}
143: \Tr^{-1} \approx \frac{\gamma^2}{2} \SBz,
144: \label{eq_tr}
145: \end{equation}
146: %
147: where $\wone = \gamma B_1$ is the Rabi frequency of the spin, $\gamma$ is the gyromagnetic ratio,
148: and $\SBz$ is the (double-sided) power spectral density of the $z$-component of ${\bm B}(t)$ \cite{slichter_specdens,mozyrsky03}.
149: 
150: For a transversely oscillating cantilever as shown in Fig. \ref{fig_schematics},
151: $\SBz$ is related to the spectrum of cantilever tip motion by $\SBz = \left(\dxBzu\right)^2 S_x(\wone)$,
152: and will --- if $S_x(\wone)$ has thermal origin --- be proportional to $\kT$.
153: Hence, in a strong coupling regime where the spin lifetime is dominated by noise from the mechanical resonator,
154: we expect an explicit dependence on the field gradient, the Rabi frequency, the temperature,
155: and the mechanical mode spectrum.
156: 
157: %
158: \begin{figure}
159:       \begin{center}
160:       \includegraphics[width=0.45\textwidth]{fig1}
161:       \end{center}
162:       \caption{(a) Experimental arrangement (setup {\bf A}). A \CaF single crystal containing \F nuclei
163:       is attached to the end of the cantilever and placed in a fixed position $\sim100\unit{nm}$ above a nanoscale FeCo magnetic tip.
164:       An rf magnetic field induces magnetic resonance in a thin ``resonant slice'' of spins where the Larmor resonance condition is fulfilled.
165:       Raising or lowering the external field $\Bext$ shifts the resonant slice position up or down,
166:       allowing spins further from or closer to the tip to be selectively addressed, respectively.
167:       }
168:       \label{fig_schematics}
169: \end{figure}
170: %
171: We find evidence for mechanically induced spin relaxation while measuring nuclear spin correlation times for small ensembles
172: of statistically polarized \F spins in a \CaF single-crystal sample \cite{bloch46,degen07}.
173: The setup for these experiments, shown in Fig. \ref{fig_schematics}(a), combines three components: 
174: (1) an ultra-sensitive cantilever, (2) a nanoscale ferromagnetic tip, and (3) a micron-scale rf circuit for rf field generation.
175: Experiments are carried out with two different arrangements: setup {\bf A} uses a 90-$\um$-long single-crystal Si cantilever \cite{chui03}
176: together with a FeCo thin-film conical tip and an external rf microcoil \cite{mamin07}.
177: Setup {\bf B} employs a slightly longer cantilever ($120\unit{\um}$)
178: combined with a FeCo cylindrical pillar integrated onto a lithographically patterned rf microwire \cite{poggio07}.
179: The \CaF samples, a few $\um^3$ in size, are glued to the end of the cantilevers.
180: 
181: For spin detection we rely on the ``rf frequency sweep'' method to drive adiabatic spin inversions,
182: thereby modulating the $z$-component of the nuclear polarization at the fundamental cantilever frequency \cite{madsen04,poggio07}.
183: Here, the rf center frequency $\wrf/2\pi$ is $114\unit{MHz}$ with peak frequency deviation of the rf sweep $\Dw/2\pi$ in the range of $400$ to $1000\unit{kHz}$.
184: We measure the correlation time $\tm$ of the cantilever tip oscillation amplitude,
185: which reflects the correlation time of the nuclear polarization \cite{degen07}.
186: $\tm$ can be determined in several ways, for example by calculating the autocorrelation function \cite{degen07},
187: by measuring the linewidth of the associated power spectral density, or by using a bank of filters of different noise bandwidths \cite{rugar04}.
188: $\tm$ is very closely related to the rotating frame relaxation time $\Tr$, as will be discussed below.
189: 
190: In a first set of experiments we observe that spin relaxation depends strongly on the distance
191: between spins and magnetic tip. We attribute this observation
192: to an associated variation of the magnetic field gradient.
193: Figure \ref{fig_gradient}(a) plots the relaxation rate $\itm$ as a function of applied external field $\Bext$.
194: Since the external field sets the region of space where the resonance condition is met ---
195: indicated by the ``resonant slice'' in Fig. \ref{fig_schematics} ---
196: changing $\Bext$ is equivalent to changing the region in the sample where we probe the spins.
197: Spins located closer to the tip require less external field to satisfy the Larmor condition,
198: $\gamma |\Bext \hat{\bm z} +\Btip| = \wrf$.
199: For these spins, the resonance appears on the low end of the $\Bext$ scan [Fig. \ref{fig_gradient}(a)].
200: Likewise, spins far away from the tip experience the smallest $\Btip$ and require the highest $\Bext$.
201: 
202: In order to relate $\itm$ to the magnetic field gradient, we first calculate the tip field $\Btip(\vecr)$ as a function of position $\vecr$.
203: A good model for the tip can be obtained by inferring the tip size from a scanning electron micrograph,
204: and combining it with MRFM data for the tip moment \cite{mamin07}.
205: To estimate the effective lateral field gradient $\Geff(\Bext) \propto \dxBzu$ for each resonant slice,
206: we then derive the gradient from $\Btip(\vecr)$ and average it over the resonant slice volume \cite{Geff}.
207: As a result, we can relate the gradient to the external field and use this knowledge to plot $\itm$ as a function of $\Geff$,
208: shown in Fig. \ref{fig_gradient}(b) and (c).
209: 
210: %
211: \begin{figure}[t]
212:       \begin{center}
213:       \includegraphics[width=0.45\textwidth]{fig2}
214:       \end{center}
215:       \caption{
216:       (a) Spin correlation time $\tm$ as a function of external field $\Bext$.
217:       Data points (dots) represent experimental values, measured with setup {\bf A}.
218:       Spins are resonant at a total field of $|\Bext\hat{\bm z}+\Btip|=2.85\unit{T}$.
219:       (b) Effective field gradient $\Geff(\Bext)$ as a function of $\Bext$ calculated from tip model.
220:       (c) Relaxation rate $\itm$ as a function of field gradient $\Geff$ obtained by combining (a) and (b).
221:       Solid line is a best fit with the lowest point excluded. Dashed line is a guide to the eye.
222:       }
223:       \label{fig_gradient}
224: \end{figure}
225: %
226: We find that spin relaxation increases with the gradient as $\itm\propto \Geff^{1.23\pm0.16}$.
227: Since the gradient is the main parameter describing the coupling between spins and oscillator,
228: a stronger gradient is equivalent to a stronger coupling --- hence the same mechanical noise creates
229: more magnetic noise. While gradient-induced spin relaxation is the dominant relaxation mechanism
230: in Fig. \ref{fig_gradient}, other (intrinsic) processes will eventually dominate in the limit of small gradient.
231: In Fig. \ref{fig_gradient}, this may be the case for gradients $\Geff<0.3\unit{G/nm}$.
232: 
233: In a second set of experiments we investigate the spin relaxation rate as a function of rf field magnitude $\wone = \gamma B_1$.
234: Here we are able to measure $\itm$ for $\wone/2\pi$ between $60$ and $170\unit{kHz}$.
235: The rf field magnitude is calibrated by a spin nutation experiment \cite{poggio07}.
236: For small $\wone$, the spin correlation times are so short that
237: the measurement is dominated by the response time of the actively damped cantilever ($\sim 20\unit{ms}$).
238: The upper limit on $\wone$ is set by heating restrictions of the rf microwire.
239: 
240: The measured relaxation rates are shown in Fig. \ref{fig_driving}(a). 
241: We find that the spin relaxation rate decreases rapidly until reaching a distinct knee at $\wone/2\pi\approx125\unit{kHz}$.
242: We believe that this feature is directly related to the thermal vibration of the third cantilever mode,
243: having a resonance frequency of $f_3=122\unit{kHz}$ [see Fig. \ref{fig_modes}].
244: 
245: %
246: \begin{figure}[t]
247:       \begin{center}
248:       \includegraphics[width=0.45\textwidth]{fig3}
249:       \end{center}
250:       \caption{(a) Spin relaxation rate $\itm$ as a function of rf field magnitude $\wone=\gamma B_1$, measured with setup {\bf B}.
251:       Each time $\wone$ falls below a cantilever mode, indicated by dotted vertical lines, a new dissipation channel is added.
252:       At low fields $\tm$ is presumably limited by the adiabatic condition (dashed line, Ref. \cite{adiabaticcondition}).
253:       %(Data reproduced from Ref. \cite{poggio07}).
254:       (b) Spin relaxation rate $\itm$ as a function of mode temperature $\Tmode$, here for the third cantilever mode.
255:       The thermal motion of the mode is found to affect spin relaxation when $\wone/2\pi<f_3$ (upper figure),
256:       but not when $\wone/2\pi>f_3$ (lower figure). Solid line is a best fit.}
257:       \label{fig_driving}
258: \end{figure}
259: %
260: To further explore the role of the third mode we can modify its effective temperature, denoted by $\Tmode$.
261: For that purpose we excite the mode with bandwidth-limited noise using a piezoelectric actuator mechanically coupled to the cantilever.
262: At the same time as we measure $\tm$, we also monitor the mean-squared motion $\xrms^2$ of the cantilever tip in that mode.
263: We can assign an equivalent mode temperature $\Tmode=\kci\xrms^2/\kB$, where $\kci$ is the effective spring constant of mode $i$ and $\kB$ is Boltzmann's constant.
264: Note that while $\Tmode$ can become very large for strong mechanical actuation, the actual (or bath) temperature of the cantilever
265: remains at $T=4.5\unit{K}$, as the heat capacity of the single mode is much smaller than
266: the heat capacity of the cantilever's phonon bath and the modes are only very weakly coupled.
267: 
268: %
269: \begin{figure}[t]
270:       \begin{center}
271:       \includegraphics[width=0.45\textwidth]{fig4}
272:       \end{center}
273:       \caption{Power spectral density of the thermomechanical noise at $T=4.5\unit{K}$, showing the first three vibrational modes
274:       of cantilever {\bf B}. Sketches illustrate the bending of the cantilever beam.
275:       Frequency, spring constant and quality factor of the lowest mode are $f_{1}=2.57\unit{kHz}$, $k_{1}=86\unit{\mu N/m}$ and $Q_{1}=17,000$.
276:       }
277:       \label{fig_modes}
278: \end{figure}
279: %
280: We investigate the dependence of $\itm$ on $\Tmode$ at several different $B_1$ fields and find
281: the two different behaviors shown in Fig. \ref{fig_driving}(b):
282: If $\wone/2\pi$ lies above $f_3$, no change in spin relaxation is observed even for strong actuation.
283: On the other hand, for $\wone/2\pi$ below $f_3$, relaxation is greatly enhanced.
284: We can fit the dependence of the relaxation rate on the mode temperature and find $\itm \propto (\Tmode)^{0.43\pm0.07}$.
285: Because the slope does not level off even for low $\Tmode$, the spin relaxation rate is set by the thermal fluctuations of the third cantilever mode
286: over the entire investigated mode temperature range, in particular at thermal equilibrium where $\Tmode = T$.
287: At $\wone/2\pi=98\unit{kHz}$, we can also try to enhance spin relaxation by actuating the second mode.
288: No influence is observed, as expected, because $\wone/2\pi>f_2$.
289: 
290: In order to better understand the dependence of the spin correlation time on field gradient, rf field magnitude, and temperature, it is worthwhile to connect $\tm$ to the cantilever's mode spectrum. We calculate the transition rate following the analysis of Mozyrsky \etal \cite{mozyrsky03}.
291: To obtain an expression for $\tm$ similar to Eq. (\ref{eq_tr}),
292: taking into account the time-dependence of the rf field frequency $\wrf(t)$ during cyclic spin inversion,
293: we assume that $\tm$ can be described as the average relaxation rate over a frequency sweep,
294: %
295: \begin{equation}
296: \itm = \frac{1}{T_c} \int_{0}^{T_c/2} dt\, \frac{\wone^2}{\weff^2(t)}\, \gamma^2 S_{Bz}(\weff(t)).
297: \label{eq_itm_int}
298: \end{equation}
299: %
300: (See also Ref. \cite{mozyrsky03}, Eq. (7)). Here, $T_c=1/f_1$ is the oscillation period of the fundamental cantilever mode,
301: $\weff(t) = \{[\wrf(t)-\wo]^2+\wone^2\}^{1/2}$ is the effective magnetic field \cite{slichter_adiabaticsweeps},
302: and $\wrf(t)$ is modulated from $\wo-\Dw$ to $\wo+\Dw$. Note that because $\Dw\gg\wone$,
303: the Rabi frequency $\weff$ traverses a broad range of frequencies during the sweep.
304: A key difference between $\tm$ and $\Tr$ is therefore that $\tm$ is sensitive to noise
305: in a frequency band set by $\wone\leq\weff(t) \underset{\sim}{<} \Dw$, while $\Tr$ is influenced by noise in the vicinity of $\wone$ only.
306: 
307: The magnetic noise spectrum $S_{Bz}(\omega)$ is given by
308: %
309: \begin{equation}
310: S_{Bz}(\omega) = \Geff^2 \sum_{i=1}^n
311: \frac{\kT}{\pi \kci \fci \Qci} \ \frac{(2\pi\fci)^4}{((2\pi\fci)^2-\omega^2)^2+(2\pi\fci \omega/\Qci)^2},
312: \label{eq_modes}
313: \end{equation}
314: %
315: where $n$ is the number of modes with a significant noise contribution. Because of the high quality factors $Q_i$, $S_{Bz}(\omega)$ exhibits
316: a discrete set of sharp peaks at the mode frequencies $f_i$. We can facilitate the analysis of 
317: Eq. (\ref{eq_itm_int}) by treating the peaks as $\delta$-functions, and find that
318: %
319: \begin{equation}
320: \label{eq_itm}
321: \itm \approx \frac{\gamma^2\Geff^2\kT}{2\Dw} \sum_{i=n'}^{n} \frac{\wone^2}{2\pi\fci\kci \sqrt{(2\pi\fci)^2-\wone^2}} ,
322: \end{equation}
323: %
324: where $n'$ is the lowest mode whose resonance frequency $f_{n'}$ is above $\wone/2\pi$.
325: In other words, when $\wone$ is reduced a new relaxation channel is added each time
326: $\wone/2\pi$ becomes less than a cantilever resonance $\fci$. While $n$ can be large,
327: the lowest mode $n'$ will dominate (\ref{eq_itm}) because it has the lowest effective spring constant
328: and hence the largest thermal vibration amplitude.
329: 
330: As a result, we find that Eq. (\ref{eq_itm}) describes both the kink in Fig. \ref{fig_driving}(a) at the frequency of the third mode and
331: produces the correct trends in $\Geff$ and $\Tmode$. Eq. (\ref{eq_itm}) and Refs. \cite{sidles92,mozyrsky03}, however, predict
332: that $\itm \propto \Geff^2\Tmode$ --- a significantly stronger dependence on $\Geff$ and $\Tmode$ than our experimental observation,
333: $\itm \propto \Geff^{1.23\pm0.16}\Tmode^{0.43\pm0.07}$.
334: We are aware of two possible reasons for this discrepancy: First, the analysis of the spin transition rate [Eq. (\ref{eq_itm_int})]
335: is based on the Bloch-Redfield approximation, where the magnetic noise is assumed to be uncorrelated \cite{slichter_blochredfield}.
336: This assumption may be violated because the correlation time of the noise, $\tau_{c,i} = \Qci/\pi\fci$, can be significant, possibly even longer than $\tm$.
337: Furthermore, we assume that spins are non-interacting, which will not necessarily be valid, especially at low rf fields where local fields can easily exceed $B_1$.
338: 
339: In conclusion, we find that nuclear spin relaxation can be induced by a single, low-frequency mode of a micromechanical resonator.
340: The spin relaxation rate is observed to increase both with the gradient and the effective mode temperature.
341: We also find that long spin lifetimes are recovered when increasing the magnitude of the rf field to raise the Rabi frequency
342: above the lowest mechanical mode frequencies.
343: 
344: We thank C. Rettner, M. Hart and M. Farinelli for fabrication of microwire and magnetic tip,
345: B. W. Chui for cantilever fabrication, D. Pearson and
346: B. Melior for technical support, and M. Ernst for helpful discussions.
347: We acknowledge support from the DARPA QUIST
348: program administered through the US Army Research Office,
349: and the Stanford-IBM Center for Probing the Nanoscale, a NSF Nanoscale
350: Science and Engineering Center. C. L. D. acknowledges funding from
351: the Swiss National Science Foundation.
352: 
353: \noindent
354: \begin{thebibliography}{}
355: 
356: %%% conventional mri
357: \bibitem{hoult76}
358: D. I. Hoult, and R. E. Richards, J. Magn. Res. {\bf 24}, 71 (1976).
359: 
360: %%% conventional MRI
361: \bibitem{seeber01} %the 10^12 protons paper
362: D. A. Seeber, R. L. Cooper, L. Ciobanu, and C. H. Pennington,
363: %D. A. Seeber \etal,
364: %Rev. Sci. Instr. {\bf 72}, 2171 (2001).
365: %10.1063/1.1359190
366: \href{http://dx.doi.org/10.1063/1.1359190}{Rev. Sci. Instr. {\bf 72}, 2171 (2001)}.
367: 
368: %%% MRFM
369: \bibitem{sidles95}
370: J. A. Sidles, J. L. Garbini, K. J. Bruland, D. Rugar, O. Z\"uger, S. Hoen, and C. S. Yannoni,
371: %J. A. Sidles \etal,
372: \href{http://dx.doi.org/10.1103/RevModPhys.67.249}{Rev. Mod. Phys. {\bf 67}, 249 (1995)}.
373: 
374: \bibitem{rugar04}
375: %D. Rugar \etal,
376: D. Rugar, R. Budakian, H. J. Mamin, and B. W. Chui,
377: \href{http://dx.doi.org/10.1038/nature02658}{Nature {\bf 430}, 329 (2004)}.
378: 
379: %%% present day MRFM
380: \bibitem{mamin07}
381: %H. J. Mamin \etal,
382: H. J. Mamin, M. Poggio, C. L. Degen, and D. Rugar,
383: \href{http://dx.doi.org/10.1038/nnano.2007.105}{Nature Nanotechnology {\bf 2}, 301 (2007)}.
384: %\href{http://arxiv.org/abs/cond-mat/0702664}{cond-mat/0702664}.
385: 
386: %%% theory spin-mechanical oscillator interaction
387: \bibitem{mozyrsky03}
388: D. Mozyrsky, I. Martin, D. Pelekhov and P. C. Hammel,
389: %D. Mozyrsky \etal,
390: \href{http://dx.doi.org/10.1063/1.1554769}{Appl. Phys. Lett. {\bf 82}, 1278 (2003)}.
391: 
392: \bibitem{berman03}
393: G. P. Berman, V. N. Gorshkov, D. Rugar, and V. I. Tsifrinovich,
394: %10.1103/PhysRevB.68.094402
395: \href{http://dx.doi.org/10.1103/PhysRevB.68.094402}{Phys. Rev. B {\bf 68}, 094402 (2003)}.
396: 
397: %%% spin cooling
398: \bibitem{magusin00}
399: P. C. M. M. Magusin, and W. S. Veeman,
400: \href{http://dx.doi.org/10.1006/jmre.1999.1989}{J. Magn. Res. {\bf 143}, 243 (2000)}.
401: 
402: \bibitem{butler05}
403: M. C. Butler, V. A. Norton, and D. P. Weitekamp,
404: Abstracts of papers of the American Chemical Society {\bf 229}, 
405: U735-U735 217-PHYS, 2005; D. P. Weitekamp, Patent US 6,841,995 B2, 2005.
406: 
407: % cantilaser
408: \bibitem{bargatin03}
409: I. Bargatin, and M. L. Roukes,
410: %Phys. Rev. Lett. {\bf 91}, 138302 (2003).
411: %10.1103/PhysRevLett.91.138302 
412: \href{http://dx.doi.org/10.1103/PhysRevLett.91.138302}{Phys. Rev. Lett. {\bf 91}, 138302 (2003)}.
413: 
414: %%% spectral density / theory relaxation
415: 
416: \bibitem{slichter_specdens}
417: C. P. Slichter,
418: {\sl Principles of Magnetic Resonance}, $3^\mr{rd}$ edition, (Springer, Berlin, 1990), p. 206 ff and p. 242 ff.
419: 
420: %%% statistical polarization
421: 
422: \bibitem{bloch46}
423: F. Bloch, 
424: \href{http://dx.doi.org/10.1103/PhysRev.70.460}{Phys. Rev. {\bf 70}, 460 (1946)}.
425: 
426: \bibitem{degen07}
427: %C. L. Degen  \etal
428: C. L. Degen, M. Poggio, H. J. Mamin, and D. Rugar,
429: \href{http://dx.doi.org/10.1103/PhysRevLett.99.250601}{Phys. Rev. Lett. {\bf 99}, 250601 (2007)}.
430: %\href{http://arxiv.org/abs/0710.2323}{arXiv:0710.2323}.
431: 
432: %%% experimental setup, protocol
433: 
434: \bibitem{chui03}
435: B. W. Chui {\it et al.,}
436: {\sl Technical Digest of the 12$^{th}$ International Conference on Solid-State Sensors and Actuators (Transducers '03)},
437: (IEEE Boston MA, USA, 2003), p. 1120.
438: 
439: \bibitem{poggio07}
440: M. Poggio, C. L. Degen, H. J. Mamin, and D. Rugar,
441: %M. Poggio \etal,
442: \href{http://dx.doi.org/10.1063/1.2752536}{Appl. Phys. Lett. {\bf 90}, 263111 (2007)}.
443: 
444: \bibitem{madsen04}
445: L. A. Madsen, G. M. Leskowitz, and D. P. Weitekamp,
446: \href{http://dx.doi.org/10.1073/pnas.0405232101}{Proc. Nat. Acad. Sci. {\bf 101}, 12804 (2004)}.
447: 
448: %%% gradient
449: 
450: \bibitem{Geff}
451: The introduction of an effective gradient $\Geff$ is necessary because not all spins in the
452: resonant slice experience the same $\dxBzu$. We believe that an appropriate definition for $\Geff$ is
453: the weighted-rms value
454: %$\Geff(\Bext) = \left[ \sum_{k=1}^{n}  \dxBzu({\bm r}_k)^2 w_k \right]^{\frac{1}{2}}$,
455: $\Geff(\Bext) = \left\{ \sum_{k=1}^{n}  \left[\dxBzu({\bm r}_k)\right]^4 / \sum_{k=1}^{n} \left[\dxBzu({\bm r}_k)\right]^2 \right\}^{\frac{1}{2}}$,
456: because it gives more weight to spins contributing large magnetic forces $F_k^2 = \mu^2 \left[\dxBzu({\bm r}_k)\right]^2$ to the signal.
457: $\{\vecr_k\}_{k=1...n}$ are the positions of the spins for which $|\gamma\Btip(\vecr_k)-\wo|<\Dw$.
458: 
459: %%% analysis
460: 
461: \bibitem{slichter_adiabaticsweeps}%adiabatic condition AND rotating frame
462: Ref. \cite{slichter_specdens}, p. 20-24.
463: 
464: \bibitem{sidles92}
465: J. A. Sidles, J. L. Garbini, and G. P. Drobny,
466: %Rev. Sci. Instr. {\bf 63}, 3881 (1992).
467: %10.1063/1.1143837
468: \href{http://dx.doi.org/10.1063/1.1143837}{Rev. Sci. Instr. {\bf 63}, 3881 (1992)}.
469: 
470: \bibitem{slichter_blochredfield}%bloch-redfield approximation
471: Ref. \cite{slichter_specdens}, p. 199.
472: 
473: %%% captions
474: 
475: \bibitem{adiabaticcondition}
476: By simulations of the Bloch equations one can show that the adiabaticity parameter $A$
477: [see J. Baum, R. Tycko, and A. Pines,
478: \href{http://dx.doi.org/10.1103/PhysRevA.32.3435}{Phys. Rev. A {\bf 32}, 3435 (1985)}]
479: relates to the maximum possible $\tm$ as $\itm \approx 2f 10^{-0.68 A}$, where $f$ is the cantilever frequency.
480: 
481: \end{thebibliography}
482: 
483: \end{document}
484: