Non-equilibrium magnetization dynamics in the Fe 8 single-molecule magnet induced by high-intensity microwave radiation

Similar documents
Photon-induced magnetization changes in single-molecule magnets invited

Radiation- and phonon-bottleneck induced tunneling in the Fe 8 single-molecule magnet

Magnetization relaxation in the single-molecule magnet Ni 4 under continuous microwave irradiation

NYU Spin Dynamics in Single Molecule Magnets. Andrew D. Kent

Andrea Morello. Nuclear spin dynamics in quantum regime of a single-molecule. magnet. UBC Physics & Astronomy

High Frequency Electron Paramagnetic Resonance Studies of Mn 12 Wheels

Quantum characterization of Ni4 magnetic clusters using electron paramagnetic resonance

Quantum Tunneling of Magnetization in Molecular Magnets. Department of Physics, New York University. Tutorial T2: Molecular Magnets, March 12, 2006

Final Report. Superconducting Qubits for Quantum Computation Contract MDA C-A821/0000

arxiv:cond-mat/ v2 10 Dec 1998

Magnetic quantum tunnelling in subsets of

arxiv: v1 [cond-mat.mes-hall] 29 Oct 2015

Spins Dynamics in Nanomagnets. Andrew D. Kent

New example of Jahn-Teller isomerism in [Mn 12 O 12 (O 2 CR) 16 (H 2 O) 4 ] complexes

Quantum tunneling of magnetization in lanthanide single-molecule. magnets, bis(phthalocyaninato)terbium and bis(phthalocyaninato)-

The effect of uniaxial pressure on the magnetic anisotropy of the Mn 12 -Ac single-molecule magnet

Intermolecular interactions (dipolar and exchange)

MIT Department of Nuclear Science & Engineering

magnet Mn 12 -t-butylacetate thermally-activated down to 400 mk

Quantum step heights in hysteresis loops of molecular magnets

Room Temperature Quantum Coherence and Rabi Oscillations in Vanadyl Phthalocyanine: Toward Multifunctional Molecular Spin Qubits

Supplementary information for Quantum delayed-choice experiment with a beam splitter in a quantum superposition

Decoherence in molecular magnets: Fe 8 and Mn 12


Quantum dynamics in Single-Molecule Magnets

Γ43 γ. Pump Γ31 Γ32 Γ42 Γ41

High-frequency ESR and frequency domain magnetic resonance spectroscopic studies of single molecule magnets in frozen solution

The First Cobalt Single-Molecule Magnet

10.5 Circuit quantum electrodynamics

arxiv: v1 [cond-mat.mes-hall] 22 Aug 2014

Single-molecule magnets: Jahn Teller isomerism and the origin of two magnetization relaxation processes in Mn 12 complexes

12. Spectral diffusion

Linear and nonlinear spectroscopy

Supercondcting Qubits

Supplementary Information

Landau-Zener tunneling in the presence of weak intermolecular interactions in a crystal of Mn 4 single-molecule magnets

Time Resolved Faraday Rotation Measurements of Spin Polarized Currents in Quantum Wells

Disordered Solids. real crystals spin glass. glasses. Grenoble

PHYSICAL REVIEW LETTERS

NMR Studies of 3 He Impurities in 4 He in the Proposed Supersolid Phase

Coherent Control of a Single Electron Spin with Electric Fields

Theory of selective excitation in stimulated Raman scattering

Likewise, any operator, including the most generic Hamiltonian, can be written in this basis as H11 H

MAGNETIC QUBITS AS HARDWARE FOR QUANTUM COMPUTERS.

Chapter 8 Magnetic Resonance

Optical Pumping in 85 Rb and 87 Rb

Supplementary Figure 1: Spin noise spectra of 55 Mn in bulk sample at BL =10.5 mt, before subtraction of the zero-frequency line. a, Contour plot of

9 Atomic Coherence in Three-Level Atoms

QUANTUM THEORY OF LIGHT EECS 638/PHYS 542/AP609 FINAL EXAMINATION

Non-linear driving and Entanglement of a quantum bit with a quantum readout

Axion dark matter search using the storage ring EDM method

doi: /PhysRevLett

Unsolved Mysteries of the Universe: Looking for Clues in Surprising Places

Quantum technologies based on nitrogen-vacancy centers in diamond: towards applications in (quantum) biology

Synthesizing arbitrary photon states in a superconducting resonator

Introduction to Relaxation Theory James Keeler

Resonant Magnetization Tunneling in Molecular Magnets

ELECTRON PARAMAGNETIC RESONANCE

Magnetic Dynamics of Nanoscale Magnets: From Classical to Quantum

Magnetic qubits as hardware for quantum computers

A Hands on Introduction to NMR Lecture #1 Nuclear Spin and Magnetic Resonance

Chem 325 NMR Intro. The Electromagnetic Spectrum. Physical properties, chemical properties, formulas Shedding real light on molecular structure:

Simulations of spectra and spin relaxation

Radiation-Induced Magnetoresistance Oscillations in a 2D Electron Gas

NANOSCALE SCIENCE & TECHNOLOGY

Entangled Macroscopic Quantum States in Two Superconducting Qubits

.O. Demokritov niversität Münster, Germany

Measuring Spin-Lattice Relaxation Time

arxiv: v2 [cond-mat.mes-hall] 4 Aug 2010 Jonathan R. Friedman

Multiphoton antiresonance in large-spin systems

A proof-of-principle experiment of EIT with gamma radiation in FePSe 3 single crystal

Switching of magnetic domains reveals spatially inhomogeneous superconductivity

NMR Spectroscopy Laboratory Experiment Introduction. 2. Theory

arxiv: v1 [cond-mat.mes-hall] 20 Sep 2013

Beyond the Giant Spin Approximation: The view from EPR

Electron spins in nonmagnetic semiconductors

Molecular prototypes for spin-based CNOT and SWAP quantum logic gates

CONTENTS. 2 CLASSICAL DESCRIPTION 2.1 The resonance phenomenon 2.2 The vector picture for pulse EPR experiments 2.3 Relaxation and the Bloch equations

Conclusion. 109m Ag isomer showed that there is no such broadening. Because one can hardly

Spin electric coupling and coherent quantum control of molecular nanomagnets

Model Answer (Paper code: AR-7112) M. Sc. (Physics) IV Semester Paper I: Laser Physics and Spectroscopy

October Entrance Examination: Condensed Matter Multiple choice quizzes

Schematic for resistivity measurement

Survey on Laser Spectroscopic Techniques for Condensed Matter

Superconducting qubits (Phase qubit) Quantum informatics (FKA 172)

B 2 P 2, which implies that g B should be

Quantum Computation with Neutral Atoms Lectures 14-15

Deterministic Coherent Writing and Control of the Dark Exciton Spin using Short Single Optical Pulses

PINNING MODES OF THE STRIPE PHASES OF 2D ELECTRON SYSTEMS IN HIGHER LANDAU LEVELS

HYPERFINE STRUCTURE CONSTANTS IN THE 102D3/2 AND 112D 3/2 STATES OF 85Rb M. GLOW

Circuit Quantum Electrodynamics. Mark David Jenkins Martes cúantico, February 25th, 2014

Exploring the quantum dynamics of atoms and photons in cavities. Serge Haroche, ENS and Collège de France, Paris

Driving Qubit Transitions in J-C Hamiltonian

arxiv:cond-mat/ v1 1 Dec 1999

arxiv:cond-mat/ Oct 2001

Superconducting Qubits. Nathan Kurz PHYS January 2007

Superconducting Qubits Lecture 4

Nonlinear Electrodynamics and Optics of Graphene

Physikalische Chemie IV (Magnetische Resonanz) HS Solution Set 2. Hand out: Hand in:

A Double Quantum Dot as an Artificial Two-Level System

Transcription:

EUROPHYSICS LETTERS 1 July 2005 Europhys. Lett., 71 (1), pp. 110 116 (2005) DOI: 10.1209/epl/i2005-10069-3 Non-equilibrium magnetization dynamics in the Fe 8 single-molecule magnet induced by high-intensity microwave radiation M. Bal 1, Jonathan R. Friedman 1, Y. Suzuki 1,2,E.M.Rumberger 3, D. N. Hendrickson 3, N. Avraham 4,Y.Myasoedov 4, H. Shtrikman 4 and E. Zeldov 4 1 Department of Physics, Amherst College - Amherst, MA 01002-5000, USA 2 Physics Department, City College of the City University of New York New York, NY 10031, USA 3 Department of Chemistry and Biochemistry, University of California at San Diego La Jolla, CA 92093, USA 4 Department of Condensed Matter Physics, The Weizmann Institute of Science Rehovot 76100, Israel received 22 February 2005; accepted in final form10 May 2005 published online 1 June 2005 PACS. 75.50.Xx Molecular magnets. PACS. 76.30.-v Electron paramagnetic resonance and relaxation. PACS. 75.45.+j Macroscopic quantum phenomena in magnetic systems. Abstract. Resonant microwave radiation applied to a single crystal of the molecular magnet Fe 8 induces dramatic changes in the sample s magnetization. s between excited states are found even though at the nominal system temperature these levels have negligible population. We find evidence that the sample heats significantly when the resonance condition is met. In addition, heating is observed after a short pulse of intense radiation has been turned off, indicating that the spin system is out of equilibrium with the lattice. Introduction. Unlike classical magnets, single-molecule magnets, with spin values on the order of 10, have a discrete energy-level structure. This allows for the observation of resonant tunneling between up and down orientations when energy levels are brought into resonance by an external magnetic field [1 4]. The discrete level structure also permits radiation to drive transitions between states. Recent experiments show that microwave radiation can be used to enhance the rate for magnetization reversal [5, 6] and change the equilibrium magnetization [7 11]. Much of this research is aimed at studying the spin dynamics of singlemolecule magnets in the presence of radiation with the goal of observing coherent dynamics such as Rabi oscillations and measuring the characteristic spin relaxation time T 1 and decoherence time T 2. Measurements of these parameters are essential for determining whether single-molecule magnets are viable as potential qubits [12]. c EDP Sciences Article published by EDP Sciences and available at http://www.edpsciences.org/epl or http://dx.doi.org/10.1209/epl/i2005-10069-3

M. Bal et al.: Non-equilibrium magnetization dynamics etc. 111 m = -9 m = -10 Tunneling m = 9 m = 10 Thermal activation Photon-induced transition Fig. 1 Double-well potential and energy levels for the Fe 8 magnet. The potential is drawn as a function of the angle θ between the spin and the easy (z) axis. The left well corresponds to the spin pointing down and the right corresponds to it pointing up. The different processes involved in spin relaxation are illustrated schematically by the arrows. Resonant microwave radiation can drive spins between neighboring states in a well (wavy arrow). Phonons are responsible for thermal activation over the barrier. Resonant tunneling can occur when the field tunes levels in opposite wells to align (horizontal arrow). For clarity, only one photon resonance and one tunneling resonances are shown. In this letter, we show that high-amplitude radiation can induce dramatic changes in the equilibrium magnetization of the molecular magnet Fe 8 when the radiation frequency matches the energy difference between various energy levels in the system. This allows us to performwhat is essentially a variant of electron-spin resonance spectroscopy on this system. In addition, we find evidence that the radiation produces significant sample heating when the resonance condition is fulfilled. Using short radiation pulses, we observe spin heating effects even after the radiation has been turned off, an effect that indicates that the spin systemand the lattice have been driven out of equilibrium. These results imply that experiments seeking to study coherent spin dynamics in a radiation field need to be done very fast in order to circumvent heating effects. We study the single-molecule magnet Fe 8 O 2 (OH) 12 (tacn) 6 (hereafter called Fe 8 ), which at low temperatures behaves as a single spin-10 system with the effective Hamiltonian H = DS 2 z + E ( S 2 x S 2 y) + C ( S 4 + + S 4 ) gµb S H, (1) where the anisotropy constants D, E, andc are 0.292 K, 0.046 K, and 2.9 10 5 K, respectively, and g = 2 [13 16]. The first term in the Hamiltonian makes the z-axis the preferred axis of orientation, the so-called easy axis. The second and third terms break the rotational symmetry of the Hamiltonian and result in tunneling between the otherwise unperturbed states. The spin s energy can be described using a double-well potential, shown in fig. 1, where the bottomof the left (right) well corresponds to the spin pointing along (antiparallel to) the easy axis. A 25 K barrier [4] separates the up and down orientations and the 2S +1=21 energy levels correspond to different orientations of the magnetic moment. A magnetic field applied along the easy axis tilts the potential. In the absence of radiation, Fe 8 (similar to other systems of single-molecule magnets) relaxes by a combination of thermal activation and tunneling above 0.4 K and by pure quantum tunneling below this temperature [4]. In the higher-temperature regime, the relaxation rate Γ has an Arrhenius-law dependence on temperature: Γ = ω 0 e U eff /k BT, where ω 0 10 7 rad/s, and U eff is the effective barrier height, which can be modulated by tunneling. Experimental details. In previous work [9], we used a low-power radiation source and found small changes (less than 1%) in the magnetization whenever the radiation frequency

112 EUROPHYSICS LETTERS 2.5 Magnetization (a. u.) 2 1.5 1 0.5-9 to -8-8 to -7-7 to -6 116.06 GHz 117.89 GHz 121.67 GHz No Rad, 4 K -10to-9 0 0 0.5 1 1.5 Fig. 2 Magnetization (a. u.) 3.48 3.44 3.4 3.36 127.62 GHz 129.62 GHz 131.69 GHz 133.71 GHz 135.98 GHz 0.4 0.5 0.6 0.7 0.8 0.9 Fig. 3 Frequency (GHz) 135 133 130 128 0.5 0.6 0.7 0.8 Resonant Field (T) Fig. 2 Equilibrium magnetization as a function of field with and without the presence of radiation, as indicated. The dips in magnetization occur when the applied radiation frequency matches the transitions between states with the indicated magnetic quantum numbers. The applied radiation power and cavity Q were 12 17 mw and 1500 2500, respectively, depending on the frequency. The abrupt jumps in the data, most conspicuous just below 1 T, are due to instrumental artifacts. Fig. 3 Magnetization dips for the 10-to-9 transition at several radiation frequencies, as indicated. The radiation power was between 2.5 and 3.7 mw, depending on frequency, and the cavity Q value was several hundred. The inset shows the transition frequency vs. the field positions of the dips. A straight line fit allowed us to extract a value for the anisotropy parameter D of 0.2944 ± 0.0005 K. matched the energy difference between the ground and first-excited states. Here we increased the radiation amplitude both by using a high-power backward-wave oscillator source and a rectangular cavity with a Q as high as 2500. The largest cavity radiation field is 0.5 G, sufficient to create substantial radiative transition rates between energy levels [12]. We mounted an Fe 8 single crystal inside the cavity. One end of the cavity had a small opening to allow the coupling of radiation fromthe attached waveguide. A Hall-bar detector was mounted outside the cavity next to the sample position. We mounted the sample with its a-axis parallel to the applied field direction, which results in the easy axis being tilted 16 fromthe field direction [17]. Results and discussion. Figure 2 shows the measured magnetization of the sample both in the absence of radiation and when high-power (12 17 mw) radiation of various frequencies has been applied. The Q of the cavity is between 1500 and 2500, depending on the mode excited. At each frequency, the radiation induces a large dip in the sample s magnetization at certain values of magnetic field. At these fields, the frequency of the radiation matches the energy difference between two neighboring levels in, say, the right well in fig. 1, resulting in the absorption of radiation and a corresponding change in magnetization. The observed magnetization changes are so large that they cannot be attributed to simply moving population between the two states involved in the resonance. Instead, much of the population must be transferred to the opposite well, through thermal and/or quantum processes. Our results are consistent with numerous spectroscopic studies [14 16,18] on this material. To confirm the accuracy of our measurement technique as a spectroscopic tool, careful data

M. Bal et al.: Non-equilibrium magnetization dynamics etc. 113 2.5 2.199 Magnetization (a. u.) 2.4 2.3 2.2 2.1 2.196 2.193 2.190 2.187 Cavity Temperature (K) 0.5 0.55 0.6 0.65 0.7 0.75 0.8 Fig. 4 Simultaneously measured sample magnetization and cavity temperature as the magnetic field is swept through resonance for the 9-to-8 transition for the applied radiation. The applied radiation power was 3.2 mw and the cavity Q was 2400. The cavity temperature was measured with a thermometer mounted outside the cavity, while the sample was inside. The observed 10 mk temperature rise reflects a much larger change in the sample temperature. was taken at somewhat lower power (and with a lower cavity Q of a few hundred) of the 10- to-9 transition at several different frequencies, as shown in fig. 3. The inset shows the applied radiation frequency as a function of the measured field positions of the peaks. The straightline fit confirms the expected Zeeman dependence, the last term in eq. (1). The intercept of the fit yields a value for the anisotropy parameter D of 0.2944 ± 0.0005 K, in good agreement with the accepted values for this material. One surprising feature of the data in fig. 2 is the observation of transitions between excited states. The nominal temperature of our cryostat is 2 K during these experiments. When radiation is not resonantly absorbed by the sample, it nevertheless heats the cavity (which in turn warms the sample) to 4 K, as evidenced by the fact that the data in fig. 2 taken at 4 K in the absence of radiation approximately coincides with the data taken in the presence of radiation when the field is off resonance. Even at this elevated temperature, transitions between excited states should be weak, if observable at all, since the thermal populations of the excited states are small. We believe that a positive-feedback mechanism is at play that creates excess sample heating when the radiation is resonant with transitions between levels. To wit, even with a small population in an excited state the sample will absorb radiation when on resonance and some of this population will be put into a higher-energy state. In decaying fromthis state, a spin will release a phonon rather than a photon due to the much higher phonon density of states. These phonons rapidly thermalize, raising the temperature of the sample. This increases the population of the excited states, and, in turn, the number of photons absorbed, fuelling the feedback mechanismand driving the systemto a much higher temperature than its environment. To confirm that the sample heats additionally when on resonance, a thermometer was attached to the outside of our cavity about a centimeter away fromthe sample position and the temperature was monitored as the field was swept through one of the resonances (9 to 8). The thermometer registered a clear temperature rise of about 10 mk when the field tuned through the resonance, as shown in fig. 4. The specific heats of the cavity and sample are within an order of magnitude of each other at the experimental temperatures [19, 20]. However, the

114 EUROPHYSICS LETTERS 0.0 Off (a) 8 (b) On 7 M-M eq (a.u.) -0.10-0.20-0.30 117.89 GHz 0.01 T 0.05 T 0.09 T 0.13 T 0.17 T 0.23 T 0.29 T Relaxation Rate (khz) 6 5 4 3 2 117.89 GHz 0 1 2 3 4 Time (ms) 0.05 0.1 0.15 0.2 0.25 0.3 Fig. 5 (a) Magnetization change as a function of time during and after a 0.2 ms pulse of 117.89 GHz radiation at several values of magnetic field, as indicated, and a cryostat temperature of 1.8 K. The decrease in magnetization after the radiation has been turned off indicates that the spin system is out of equilibrium with the lattice. The abrupt jumps in the data at the beginning and end of the radiation pulse are instrumental artifacts. (b) Relaxation rate as a function of magnetic field for the data in (a) and similar data taken at. The relaxation rate was determined by fitting the decay after the radiation pulse to an exponential. cavity mass ( 10 g) is many orders of magnitude larger than the sample s ( 0.5 µg). Therefore, the observed cavity temperature rise probably corresponds to a several-kelvin increase in the sample s temperature. To better understand how this heating takes place, we performed a time-resolved study of the magnetization by applying a single 0.2 ms pulse [21] of 117.89 GHz radiation into a cavity with a low Q ( 100 200). The change in magnetization from its equilibrium value is plotted as a function of time at several fixed magnetic fields in fig. 5a. The abrupt jumps that occur when the radiation is turned on or off are instrumental artifacts. Ignoring these, the magnetization of the sample decreases during the radiation pulse due to photon-induced transitions from the ground state to the first-excited state as well as heating. However, after the radiation pulse is turned off, the magnetization continues to decrease for an additional fraction of a millisecond and reaches a minimum [22]. The decrease is largest at 0.13 T, the field at which the 10-to-9 transition occurs in fig. 2 for this frequency. Since energy is no longer being supplied to the system, one might expect the magnetization to increase as the system cools. In contrast, it appears that the spins absorb radiation during the radiation pulse, releasing phonons, which heat the lattice as they thermalize. The magnetization then continues to decrease as the spin system comes to equilibrium with the lattice in the (much longer) characteristic time required for the spins to cross the barrier (fig. 1). Thus, although photons are being absorbed by the spins, thermalization occurs first in the lattice and then in the spin system. We checked this hypothesis by fitting the region between the end of the pulse and the magnetization minimum to an exponential. The relaxation rate (fig. 5b) determined this way shows two peaks: one at a field of 0.13 T, the 10-to-9 transition resonance where the amount of heating is greatest, and another at 0.23 T, at which the first tunneling resonance beyond zero field occurs for this system(illustrated in fig. 1). Figure 5b also shows the results of a similar analysis of data taken at, where the field at which the 10-to-9 radiation

M. Bal et al.: Non-equilibrium magnetization dynamics etc. 115 0 On Off 4.4 K M-M eq (a.u.) -0.05-0.1 3.6 K 3K 2.4 K 1.8 K -0.15 0 1 2 3 4 Time (ms) Fig. 6 Magnetization change induced by radiation pulses as a function of time at several different cryostat temperatures when the magnetic field is fixed at 0.17 T. The induced magnetization change after the pulse becomes smaller and the response faster as the temperature is raised, consistent with the faster relaxation rate of the spin system at higher temperatures. transition nearly coincides with that for resonant tunneling, causing the two peaks to merge into one. These data indicate that the magnetization decrease after the radiation pulse appears to be controlled by the relaxation rate associated with the spin system, which is increased either by radiatively raising the temperature or by tunneling. This supports the notion that the decrease in magnetization after the pulse results fromthe the spin systemreturning to equilibriumwith the lattice, which has been indirectly heated by the radiation. These results may have some bearing on the slow recovery times measured by del Barco et al. when the Ni 4 molecular magnet is irradiated by microwaves at low temperatures [8]. The change in magnetization due to a radiation pulse at different temperatures is plotted in fig. 6. The decrease in magnetization after the pulse diminishes for increasing temperature up to 3.6 K, above which there is no longer a decrease in magnetization immediately after the pulse. This is consistent with our picture since at higher temperatures the spins relax faster (in accordance with an Arrhenius law) and thereby equilibriate with the lattice more quickly. In addition, the temperature increase of the lattice during the radiation pulse is expected to be smaller at higher temperatures because of the lattice s larger specific heat, reducing the magnitude of the observed effect. Conclusions. We find that resonant high-power microwave radiation can induce significant changes in the equilibrium magnetization of the molecular magnet Fe 8 when the radiation is on resonance with transitions between states with different magnetic quantum numbers. We observe transitions between excited states even when the thermal populations of these states are very small, indicating a substantial amount of sample heating when on resonance, as we have confirmed by monitoring the cavity temperature. When pulses of radiation are applied, the spin system and the lattice are driven out of thermal equilibrium, with recovery determined by the spin relaxation time. Any experiment probing photon-induced magnetization dynamics between only two levels of the spin systemneeds to be done at much shorter time scales.

116 EUROPHYSICS LETTERS We thank M. P. Sarachik, K. M. Mertes, J. Tu, E. M. Chudnovsky and J. Tejada for useful conversations. We are indebted to D. Krause, P. Grant and R. Chaudhuri for their technical contributions. Support for this work was provided by the US National Science Foundation under grant number CCF- 0218469, the Alfred P. Sloan Foundation, the Center of Excellence of the Israel Science Foundation, and the Amherst College Dean of Faculty. REFERENCES [1] Friedman J. R., Sarachik M. P., Tejada J. and Ziolo R., Phys. Rev. Lett., 76 (1996) 3830. [2] Friedman J. R., inexploring the Quantum/Classical Frontier: Recent Advances in Macroscopic Quantum Phenomena, edited by Friedman J. R. and Han S. (Nova Science, Hauppauge, NY) 2003, p. 179. [3] Thomas L., Lionti F., Ballou R., Gatteschi D., Sessoli R. and Barbara B., Nature, 383 (1996) 145. [4] Sangregorio C., Ohm T., Paulsen C., Sessoli R. and Gatteschi D., Phys. Rev. Lett., 78 (1997) 4645. [5] Amigó R., Hernandez J. M., García-Santiago A. and Tejada J., Phys. Rev. B, 67 (2003) 220402. [6] Sorace L., Wernsdorfer W., Thirion C., Barra A.-L., Pacchioni M., Mailly D. and Barbara B., Phys. Rev. B, 68 (2003) 220407. [7] Wernsdorfer W., Mueller A., Mailly D. and Barbara B., Europhys. Lett., 66 (2004) 861. [8] del Barco E., Kent A. D., Yang E. C. and Hendrickson D. N., Phys. Rev. Lett., 93 (2004) 157202. [9] Bal M., Friedman J. R., Suzuki Y., Mertes K., Rumberger E. M., Hendrickson D. N., Myasoedov Y., Shtrikman H., Avraham N. and Zeldov E., Phys. Rev. B, 70 (2004) 100408(R). [10] Petukhov K., Wernsdorfer W. and Barra A.-L., cond-mat/0502175 (2005). [11] Cage Brant, Russek Stephen E., Zipse David and Dalal Naresh S., J. Appl. Phys., 97 (2005) 10M507. [12] Leuenberger M. N. and Loss D., Nature, 410 (2001) 789. [13] Wernsdorfer W. and Sessoli R., Science, 284 (1999) 133. [14] Caciuffo R., Amoretti C., Murani A., Sessoli R., Caneschi A. and Gatteschi D., Phys. Rev. Lett., 81 (1998) 4744. [15] Barra A. L., Gatteschi D. and Sessoli R., Chem. Eur. J., 6 (2000) 1608. [16] Mukhin A. A., Gorshunov B., Dressel M., Sangregorio C. and Gatteschi D., Phys. Rev. B, 63 (2001) 214411. [17] Ueda M., Maegawa S., Miyasaka H. and Kitagawa S., J. Phys. Soc. Jpn., 70 (2001) 3084. [18] Park Kyungwha, Novotny M. A., Dalal N. S., Hill S. and Rikvold P. A., Phys. Rev. B, 66 (2002) 144409 and references therein. [19] Douglas L. M., Phys. Rev., 141 (1966) 576. [20] Mettes F. L., Luis F. and de Jongh L. J., Phys. Rev. B, 64 (2001) 174411. [21] We also did experiments with pulses ranging in duration from 1 µs to 2 ms and obtained qualitatively similar results. When pulse durations were short, the induced magnetization change was barely measurable. In these experiments we were restricted to time scales 1 µs, the response time of our Hall detector. Experiments at shorter time scales, at which coherent radiationinduced dynamics may be observable [12], are in progress. [22] Interestingly, Petukhov et al. [10], working on the same system, do not report any decrease in the magnetization after the radiation is turned off.