07.01.2014 Views

PHYS07200604007 Manas Kumar Dala - Homi Bhabha National ...

PHYS07200604007 Manas Kumar Dala - Homi Bhabha National ...

PHYS07200604007 Manas Kumar Dala - Homi Bhabha National ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

SPECTROSCOPIC INVESTIGATIONS OF THE<br />

ELECTRONIC STRUCTURE OF Pr 1−x Ca x MnO 3<br />

SYSTEM<br />

A THESIS SUBMITTED TO THE<br />

HOMI BHABHA NATIONAL INSTITUTE<br />

FOR THE DEGREE OF<br />

DOCTOR OF PHILOSOPHY IN SCIENCE<br />

(PHYSICS)<br />

MAY, 2009<br />

By<br />

MANAS KUMAR DALAI<br />

INSTITUTE OF PHYSICS,<br />

BHUBANESWAR 751 005,<br />

INDIA


CERTIFICATE<br />

This is to certify that the thesis entitled Spectroscopic investigations of the<br />

electronic structure of Pr 1−x Ca x MnO 3 system, which is being submitted by<br />

<strong>Manas</strong> <strong>Kumar</strong> <strong>Dala</strong>i in partial fulfillment of the degree of Doctor of Philosophy<br />

in Science (Physics) of <strong>Homi</strong> <strong>Bhabha</strong> <strong>National</strong> Institute is a record of his own<br />

research work carried out by him. He has carried out his investigations for the last<br />

six years on the subject matter of the thesis under my supervision at Institute of<br />

Physics, Bhubaneswar, India. To the best of my knowledge, the matter embodied<br />

in the thesis has not been submitted for the award of any other degree.<br />

Signature of the Candidate<br />

<strong>Manas</strong> <strong>Kumar</strong> <strong>Dala</strong>i<br />

Institute of Physics<br />

Bhubaneswar<br />

Date :<br />

Signature of the supervisor<br />

Dr. B. R. Sekhar<br />

Associate Professor<br />

Institute of Physics<br />

Bhubaneswar<br />

ii


iii<br />

To My Parents


Acknowledgements<br />

I owe a debt of gratitude to many people who made this thesis possible due to their<br />

support, suggestions, and encouragements.<br />

First of all, it is a great pleasure for me to express my sincere gratitude and indebtedness<br />

to my thesis supervisor Prof. B. R. Sekhar for his invaluable guidance<br />

and constant encouragement to complete this work. His excellent training in handling<br />

the intricacies of various experiments and making me acquainted with the data<br />

analysis have been an inspiration for me to pursue my career as an experimental<br />

condensed matter physicist. I wish to thank him for giving me the complete freedom<br />

both academically and socially during the period of my work.<br />

I would like to thank my group members Prabir and Rupali for their untiring<br />

support during the experiments and for many interesting discussions. I would also<br />

like to thank our technical staff Mr. Chittaranjan Swain for his help during the<br />

experiments. My sincere thanks go to Prof. C. Martin (Laboratoire CRISMAT,<br />

France) for providing good quality samples for this thesis work.<br />

I would like to thank Dr. S. R. Barman, Dr. S. Banik, Dr. A. K. Shukla and Dr.<br />

R. Dhaka (UGC-DAE CSR, Indore) for their help and discussions for inverse photoemission<br />

measurements. I would also like to thank Prof. Stefan Schuppler, Dr. Peter<br />

Nagel and Dr. Michael Merz for their help and cooperation during the high resolution<br />

photoemission and x-ray absorption measurements at WERA beamline of ANKA<br />

synchrotron, Karlsruhe, Germany. I would like to thank Dr. Federica Bondino, Dr.<br />

Marco Zangrando, Dr. Elena Magnano and Dr. Bryan Doyle of ELETTRA synchrotron,<br />

Trieste for their help and technical discussions for XAS experiments.<br />

It is a great pleasure for me to thank Prof. N. L. Saini (Departimento di fisica,<br />

Universita di Roma, Rome, Italy) for his fruitfull discussion at different stages of<br />

my Ph. D. career. I also thank Prof. Anil <strong>Kumar</strong> and Prof K. K. Nanda of IISc<br />

Bangalore for discussion and suggestions during their short visit to IOP.<br />

I would like to thank all my teachers, who had taught me at various stages of<br />

my educational career, especially my M. Sc teachers Profs. N. Barik, L. P. Singh, L.<br />

iv


Maharana, P. Khare, D. K. Basa, K. Maharana, S. Mohapatra and D. Behera and<br />

B. Sc teacher Mr. Dipak Singh for their encouragement and excellent teaching. It<br />

is my great pleasure to thank Profs. A. Khare, A. M. Jayannavar, S. G. Mishra, D.<br />

P. Mohapatra, K. Kundu, P. V. Satyam, S. Varma, S. Mukherji, J. Maharana, G. V.<br />

Raviprasad, A. M. Srivastava, S. M. Bhattacharjee, A. <strong>Kumar</strong>, P. Agrawal, T. Som,<br />

G. Tripathy, S. K. Patra for their teaching, suggestions and encouragements. I thank<br />

our director Dr. Y. P. Viyogi and ex-director Dr. R. K. Choudhury for their support<br />

and encouragements in all respect.<br />

I would like to thank Mr. Santosh Choudhury, Mr. Anup Behera, Mrs. Ramarani<br />

Das, Mr. Arun Dash, Mr. Khirod Patra, and Dr. S. N. Sarangi for providing technical<br />

help when needed. I would also like to thank the library staff Rama, Rabaneswar,<br />

Duryodhan and Kissan for their cooperation at every stage. I thank Dipankar, Srijesh<br />

and Dhiren for their help and cooperation at every computer related problems. I wish<br />

to thank the hostel mess staff for their wonderful services.<br />

I am thankful to our registrar Mr. C. B. Mishra for his imense help and cooperation<br />

in all administrative problems. I also thank Mr. Sk. Kefaytulla and Rajesh<br />

Mohapatra for their administrative help.<br />

I take this oppurtunity to thank my M.Sc friends Rajesh, Tapas, Swayambhu, Hitu<br />

and Manmohan and village friend Suda for their constant support and encouragement<br />

in all stage of my life. A special thank to Ananta for his constructive criticism, support<br />

and company for both academic and social activities in my Ph.D life.<br />

It is my pleasure to thank Zashmir bhai, Satya bhai, Jhasa bhai, Satchi bhai,<br />

Bedanga bhai, Pal bhai, Kamal bhai, Pradeep bhai, Bishnu bhai, Birendra bhai,<br />

Sudhira bhauja and Raj babu for their encouragement and cooperation during my<br />

stay at IOP. I would like to thank Millind, Sinu, Boby, Chandra, Hara, Amulya, Ajay,<br />

Trilochan, Smruti, Chitrasen, Jatis, Ashutosh, Kuntala, Sadhana, Mamata, Ranjita,<br />

Binata, Jaya, Poulomi, Saumia, Pramita, Probodh, Jay, Dipak, Sumalay, Bharat,<br />

Anupam, Umanand, Dilip, Suchi, Srikumar, Sankha, Nabyendu, Somnath, Sandeep,<br />

Sourabh, Subrat, Sanjay, Raghavendra, Vanaraj, and Tanmoy for making my stay at<br />

the Institute a memorable one. I also thank Dr. V. R. R. Medicherla for his valuable<br />

discussions.<br />

I would like to give my special thanks to Dr. Arun <strong>Kumar</strong> Pati and Dr. Biswajit<br />

Pradhan for making my stay in the beautiful campus of IOP an exciting, memorable<br />

and pleasing one.<br />

v


I am extremely grateful to Mrs. Ranjana Sekhar for making my stay homely<br />

during the course of my work.<br />

Gratefully I would like to thank my younger sister Tuin and younger brother Tulia<br />

for their unlimited love, affection, encouragements, assistance and support at every<br />

stage of my life. I do not find proper words to express my sense of indebtedness to my<br />

wife <strong>Manas</strong>mita, whose unlimited love, affection, encouragements, and support have<br />

become a special inspiration to me during my Ph. D life. I also thank my brother<br />

in-law Munu for his love, affection and support.<br />

Finally, I wish to express my deep sense of gratitude to my parents and late grand<br />

parents for their immense patience, love, affection, encouragements and support. I am<br />

indebted to them for giving me the freedom to choose my career and for constantly<br />

providing me the moral and emotional support at every stage of my life. It is only<br />

because of their kind blessings that I could complete this work.<br />

Date: 22.05.2009<br />

(<strong>Manas</strong> <strong>Kumar</strong> <strong>Dala</strong>i)<br />

vi


List of Publications/Preprints<br />

[1] Valence band electronic structure of Pr 1−x Sr x MnO 3 from photoemission studies,<br />

P. Pal, M. K. <strong>Dala</strong>i, B. R. Sekhar, S. N. Jha, S. V. N. Bhaskara Rao, N. C.<br />

Das, C. Martin, and F. Studer; J. Phys. : Condens. Matter 17, 2993<br />

(2005).<br />

*[2] Electronic structure of Pr 0.67 Ca 0.33 MnO 3 near the Fermi level studied by ultraviolet<br />

photoelectron and x-ray absorption spectroscopy, M. K. <strong>Dala</strong>i, P. Pal,<br />

B. R. Sekhar, N. L. Saini, R. K. Singhal, K. B. Garg, B. Doyle, S. Nannarone,<br />

C. Martin, and F. Studer; Phys. Rev. B 74, 165119 (2006).<br />

[3] Near Fermi level electronic structure of Pr 1−x Sr x MnO 3 : Photoemission study, P.<br />

Pal, M. K. <strong>Dala</strong>i, R. Kundu, M. Chakraborty, B. R. Sekhar, and C. Martin;<br />

Phys. Rev. B 76, 195120 (2007).<br />

[4] Pseudogap behavior of phase-separated Sm 1−x Ca x MnO 3 : A comparative photoemission<br />

study with double exchange, P. Pal, M. K. <strong>Dala</strong>i, R. Kundu, B. R.<br />

Sekhar, and C. Martin; Phys. Rev. B 77, 184405 (2008).<br />

*[5] Electronic structure of Pr 1−x Ca x MnO 3 from photoemission and inverse photoemission<br />

spectroscopy, M. K. <strong>Dala</strong>i, P. Pal, R. Kundu, B. R. Sekhar, S. Banik,<br />

A. K. Shukla, S. R. Barman, and C. Martin; Submitted to Physica B.<br />

*[6] Spectroscopic investigations of the electron-doped Ca 0.86 Pr 0.14 MnO 3 , M. K.<br />

<strong>Dala</strong>i, P. Pal, R. Kundu, B. R. Sekhar, M. Mertz, P. Nagel, S. Schuppler,<br />

and C. Martin; Submitted to Phys. Rev. B.<br />

*[7] Core level electronic structure of Pr 1−x Ca x MnO 3 using x-ray photoelextron spectroscopy,<br />

M. K. <strong>Dala</strong>i, P. Pal, R. Kundu, B. R. Sekhar, and C. Martin; To<br />

be submitted.<br />

vii


[8] Electron spectroscopic investigations of the paramagnetic insulator to antiferromagnetic<br />

insulator transition in Pr 0.5 Sr 0.5 MnO 3 , P. Pal, M. K. <strong>Dala</strong>i, R.<br />

Kundu, B. R. Sekhar, S. Schuppler, M. Merz, P. Nagel, M. Channabasappa and<br />

A. Sundaresan; Submitted to Phys. Rev. B.<br />

(*) indicates papers on which this thesis is based.<br />

viii


Synopsis<br />

The perovskite type manganites having general formula R 1−x A x MnO 3 (where R is a<br />

trivalent rare earth element like La, Pr, Nd. and A is a divalent alkaline earth element<br />

like Sr, Ca, Ba.) have been one of the interesting compounds among the transition<br />

metal oxide systems due to their important physical property such as colossal magnetoresistance<br />

(CMR) [1, 2, 3, 4]. CMR is defined as the colossal decrease of resistance<br />

by the application of magnetic field. These materials are insulators at high temperatures<br />

and poor metals at low temperatures. Accompanying this insulator-metal<br />

transition is a magnetic transition from a high temperature paramagnetic phase to a<br />

low temperature ferromagnetic phase. These transitions occurs around the same transition<br />

temperature T C . Applying a magnetic field around T C will greatly reduce the<br />

resistivity, i.e negative magnetoresistance effect. These materials were first described<br />

in 1950 by Jonker and van Santen [5] in perovskite type of manganites. One year later<br />

(1951), Zener [6, 7] explained this unusual correlation between magnetism and transport<br />

properties by introducing a novel concept called ”Double exchange” mechanism<br />

(DE). The pioneering work of Zener was followed by Anderson and Hasegawa [8] in<br />

1955 and de Gennes [9] in 1960. Double exchange theory argues that the electron<br />

hoping is related to the relative spin orientations of the neighbouring sites. Although<br />

it qualitatively explains the connection between the insulator-metal and the magnetic<br />

transitions, it was noticed latter by A. J. Millis et. al. [10] that it fails to explain<br />

the quantitative change in resistivity and the very low T C . Thus it is revealed that<br />

the double-exchange mechanism is not enough to explain the CMR effect. Several<br />

mechanisms [11, 12, 13, 14], based on electron-phonon interactions (the polaronic<br />

effect), electronic phase separation, charge and/or orbital ordering etc. have been<br />

proposed to account for the discrepancy. Although a detailed picture is still unclear,<br />

there seems to be a growing consensus that one or more of these additional effects<br />

probably exists, with the various phenomena both competing and cooperating with<br />

each other. This leads to a very rich and exotic behaviour, as observed in these<br />

ix


manganites. This mixed valence manganites have been studied for more than five<br />

decades but are still considered modern materials because of their wide potential for<br />

technological application.<br />

Experimental electronic structure studies [15, 16, 17, 18] on these materials have<br />

contributed substantially to the understanding of their unusual behavior which could<br />

not be fully accomodated within the framework of the DE model. The temperature<br />

and doping dependent metal-insulator transitions in these materials are found to be<br />

closely related to the unique electronic structure derived from the Mn 3d and O 2p<br />

hybridized orbitals. Most of their electrical and magnetic properties originate from<br />

the Jahn-Teller distortion in the MnO 6 octahedra around the Mn 3+ ions. The strong<br />

interplay between spin and orbital ordering originating from the single occupancy of<br />

the doubly degenerate e g orbitals of Mn 3+ (t 3 2ge 1 g) ions make the phase diagram of these<br />

compounds rich with physics. Many of these compounds, for example L 1−x Ca x MnO 3<br />

(L = La, Pr) show strong charge and/or orbital ordered (CO/OO) insulating nature<br />

at low temperatures, especially when x equals commensurate value.<br />

Electron spectroscopy is a very powerful experimental tool for probing the electronic<br />

structure of these materials. For this thesis work, the electronic structure of<br />

Pr 1−x Ca x MnO 3 series have been investigated using ultra-violet photoelectron spectroscopy<br />

(UPS), X-ray photoelectron spectroscopy (XPS), Inverse photoemission spectroscopy<br />

(IPES) and X-ray absorption spectroscopy (XAS). Pr 1−x Ca x MnO 3 is one of<br />

the interesting among these materials due to their insulating phase at all temperatures<br />

and great variety of ordered phases, that are very sensitive to the cation/anion<br />

doping. For 0.3 ≤ x ≤ 0.8, a charge ordering of Mn 3+ and Mn 4+ was found and an<br />

antiferromagnetic (AF) ordering can be observed with neel temperature ranging from<br />

100 to 170 K. For x ≤ 0.25, a ferromagnetic insulating state (FMI) is observed and<br />

with no CO, whatever the temperature. In this region the metallic state is never realized<br />

even upon the application of magnetic field. But the charge ordered insulating<br />

state can be melted into ferromagnetic metallic state (FMM) upon application of a<br />

magnetic field [4]. Interestingly the x = 0.33 doping shows the coexistence of FM and<br />

AFM phases. Around x = 0.9 a metallic cluster glass domain [19] has been observed.<br />

Using valence band photoemission and O K edge x-ray absorption, we studied<br />

the temperature dependent finer changes in the near E F electronic structure of<br />

Pr 0.67 Ca 0.33 MnO 3 , which is regarded as a prototype for the electronic phase separation<br />

models in CMR systems. With decrease in temperature the O 2p contributions<br />

x


to the t 2g and e g spin-up states in the valence band were found to increase until T c .<br />

Below T c , the density of states with e g spin-up symmetry increased while those with<br />

t 2g symmetry decreased, possibly due to the change in the orbital degrees of freedom<br />

associated with the Mn 3dO 2p hybridization in the pseudo-CE-type charge or orbital<br />

ordering. These changes in the density of states could well be connected to the electronic<br />

phase separation reported earlier. By combining the UPS and XAS spectra,<br />

we derived a quantitative estimate of the charge transfer energy E CT (2.6±0.1 eV),<br />

which is large compared to the earlier reported values in other CMR systems. Such<br />

a large charge transfer energy was found to support the phase separation model.<br />

For a complete understanding of the changes in the occupied and unoccupied<br />

electronic states we performed a detailed study of the Pr 1−x Ca x MnO 3 system across<br />

their ferromagnetic - antiferromagnetic phase boundary using UPS and IPES. We<br />

used three compositions x = 0.2, 0.33 and 0.4. Our photoemission studies showed<br />

that the pseudogap formation in these compositions occur over an energy scale of<br />

0.48±0.02 eV. Here again, we have estimated the charge transfer energy in these<br />

compositions to be of the order of ∼ 2.8±0.2 eV. We have also undertaken a detailed<br />

study of the core level electronic structure of Pr 1−x Ca x MnO 3 ((x = 0.2, 0.33, 0.4 and<br />

0.84) using X-ray photoelectron spectroscopy. These studies have been performed at<br />

the Mn 2p, Ca 2p and Pr 4d levels. We have made systematic measurments of the<br />

changes in the binding energy positions of these levels as a function of the charge<br />

carrier concentration. We also have analyzed the line shapes of these core levels.<br />

To a large extent the traditional models employing the charge-spin coupling, have<br />

been able to explain the CMR in many of the hole doped compositions. But, much<br />

less is known about the electron-doped versions of these materials in which the charge,<br />

orbital and spin ordering add more complexity to the ferromagnetic (FM) double exchange<br />

and the superexchange interactions. These Mn(IV) rich compositions exhibit<br />

marked differences in the electronic and magnetic properties from their Mn(III) rich<br />

counterparts. It has been shown that the substitution of Ca by a trivalent cation<br />

in the G-type antiferromagnet, CaMnO 3 leads to the formation of a FM component<br />

with a maximum for optimal electron doping. Depending on the nature of the trivalent<br />

cation, the x opt was found to lie between 0.135 and 0.16. This FM component is<br />

correlated to the distortion of the MnO 6 octahedra and thereby the e g band width.<br />

With low filling of the narrow e g band, these systems have strong electron-electron<br />

interactions and consequent charge localizations. Using high resolution photoelectron<br />

xi


spectroscopy and O K edge x-ray absorption spectroscopy we have studied the temperature<br />

dependence of the near Fermi level electronic structure in the electron doped<br />

CMR, Ca 0.86 Pr 0.14 MnO 3 . We found that the temperature dependent changes in the<br />

electronic structure is consistent with the conductivity behavior of the electron doped<br />

systems in general. As the temperature is lowered from 293 K a feature due to the e g<br />

states starts appearing near the E F . This increase in the DOS is a signature of the<br />

semi-metallic behavior of the sample in the range 110 - 300 K. Surprisingly, though<br />

the resistivity measurement on this sample showed an insulating behavior below its<br />

transition temperature T c (110 K), the intensity of this feature further increases below<br />

the T c . Further, we have found an increase in the e g band width (W) also at<br />

low temperatures. It was also interesting to note that the increase in the near E F<br />

DOS in this compound is apparently different in nature from that associated with the<br />

pseudogap formation in the other systems like Pr 1−x Sr x MnO 3 . We have discussed our<br />

results from the point of view of phase sepation models considering the temperature<br />

dependent structural changes and consequent localization of the electrons in a narrow<br />

e g band.<br />

xii


Bibliography<br />

[1] R. von Helmolt, J. Wecker, B. Holzapfel, L. Schultz, and K. Samwer; Phys. Rev.<br />

Lett. 71, 2331 (1993).<br />

[2] Y. Tokura, A. Urushibara, Y. Moritomo, T. Arima, A. Asamitsu, G. Kido and<br />

N. Furukawa; J. Phys. Soc. Jpn. 63, 3931 (1994).<br />

[3] A. Urushibara, Y. Moritomo, T. Arima, A. Asamitsu, G. Kido, and Y. Tokura;<br />

Phys. Rev.B 51, 14103 (1995).<br />

[4] Y. Tomioka, A. Asamitsu, Y. Moritomo, and Y. Tokura, J. Phys. Soc. Jpn 64,<br />

3626 (1995).<br />

[5] G. H. Jonker and J. H. Van Santen Physica 16, 337 (1950).<br />

[6] C. Zener, Phys. Rev. 81, 440 (1951).<br />

[7] C. Zener, Phys. Rev. 82, 403 (1951).<br />

[8] P. W. Anderson and H. Hasegawa, Phys. Rev. 100, 675 (1955).<br />

[9] P. G. de Gennes, Phys. Rev. 118, 141 (1960).<br />

[10] A. J. Millis, B. I. Shraiman, and P. B. Lttlewood, Phys. Rev. Lett. 74, 5144<br />

(1995).<br />

[11] A. J. Millis, Nature 392, 147 (1998).<br />

[12] A. Moreo, S. Yunoki, and E. Dagatto, Science 283 2034 (1999).<br />

[13] A. Moreo, M. Mayr, A. Feiguin, S. Yunoki, and E. Dagotto, Phys. Rev. Lett. 84<br />

5568 (2000).<br />

xiii


[14] H. Kuwahara, Y. Tomioka, A. Asamitsu, Y. Moritomo, Y. Tokura; Science 270,<br />

961 (1995).<br />

[15] T. Saitoh, A. E. Bocquet, T. Mizokawa, H. Namatame, A. Fujimori, M. Abbate,<br />

Y. Takeda, and M. Takano; Phys. Rev. B 51, 13942 (1995).<br />

[16] D. D. Sarma, N. Shanthi, S. R. Krishnakumar, T. Saitoh, T. Mizokawa, A.<br />

Sekiyama, K. Kobayashi, A. Fujimori, E. Weschke, R. Meier, G. Kaindl, Y.<br />

Takeda, and M. Takano; Phys. Rev. B 53, 6873 (1996).<br />

[17] J.-H. Park, C. T. Chen, S-W. Cheong, W. Bao, G. Meigs, V. Chakarian, and Y.<br />

U. Idzerda; Phys. Rev. Lett. 76, 4215 (1996).<br />

[18] A. Chainani, H. Kumigashira, T. Takahashi, Y. Tomioka, H. Kuwahara, and Y.<br />

Tokura; Phys. Rev. B 56, R15513 (1997).<br />

[19] C. Martin, A. Maignan, M. Hervieu, and B. Raveau, Phys. Rev. B 60 12191<br />

(1999).<br />

xiv


Contents<br />

CERTIFICATE<br />

Acknowledgements<br />

List of Publications/Preprints<br />

Synopsis<br />

ii<br />

iv<br />

vii<br />

ix<br />

1 Introduction 1<br />

1.1 An overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

1.2 Colossal magnetoresistance . . . . . . . . . . . . . . . . . . . . . 3<br />

1.3 Crystal structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.4 The basic electronic structure . . . . . . . . . . . . . . . . . . . . 5<br />

1.5 Double exchange mechanism . . . . . . . . . . . . . . . . . . . . 7<br />

1.6 Jahn-Teller effect in CMR manganites . . . . . . . . . . . . . . 8<br />

1.7 Phase separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

1.8 Spin, Charge and Orbital ordering in manganites . . . . . . . 11<br />

1.9 Pseudogap behavior in manganites . . . . . . . . . . . . . . . . 13<br />

1.10 Basic properties, phase diagram and CMR effect in Pr 1−x Ca x MnO 3<br />

system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2 Experimental Techniques 23<br />

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.2 Photoemission Spectroscopy . . . . . . . . . . . . . . . . . . . . . 24<br />

2.2.1 Basic principle . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.2.2 Theoretical interpretation of photoemission . . . . . . . . . . . 25<br />

2.2.3 The electron escape depth . . . . . . . . . . . . . . . . . . . . 29<br />

2.2.4 Experimental set-up . . . . . . . . . . . . . . . . . . . . . . . 30<br />

xv


2.3 Inverse Photoemission Spectroscopy . . . . . . . . . . . . . . . . 38<br />

2.3.1 Historical background . . . . . . . . . . . . . . . . . . . . . . 38<br />

2.3.2 Basic principle . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

2.3.3 Basic interpretation of inverse photoemission . . . . . . . . . . 39<br />

2.3.4 Experimental Set-up . . . . . . . . . . . . . . . . . . . . . . . 41<br />

2.4 X-ray Absorption Spectroscopy . . . . . . . . . . . . . . . . . . . 45<br />

2.4.1 Basic principle . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

2.4.2 Detection techniques . . . . . . . . . . . . . . . . . . . . . . . 47<br />

2.5 Synchrotron Radiation . . . . . . . . . . . . . . . . . . . . . . . . 49<br />

3 Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 59<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

3.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

3.3.1 The UPS study of Pr 0.67 Ca 0.33 MnO 3 . . . . . . . . . . . . . . 62<br />

3.3.2 The XAS study of Pr 0.67 Ca 0.33 MnO 3 . . . . . . . . . . . . . . 67<br />

3.3.3 The Charge Transfer Energy (E CT ) . . . . . . . . . . . . . . . 69<br />

3.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

4 Electronic Structure of Pr 1−x Ca x MnO 3 77<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

4.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

4.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

4.3.1 The UPS study of Pr 1−x Ca x MnO 3 , x = 0.2, 0.33 and 0.4 . . . 80<br />

4.3.2 The IPES study of Pr 1−x Ca x MnO 3 , x = 0.2, 0.33, 0.4 and 1 . 84<br />

4.3.3 The core levels of Pr 1−x Ca x MnO 3 , x = 0.2, 0.33, 0.4, and 0.84 87<br />

4.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

5 Electronic Structure of Ca 0.86 Pr 0.14 MnO 3 96<br />

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97<br />

5.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

5.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

5.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

6 Summary and Conclusions 108<br />

xvi


List of Figures<br />

1.1 Temperature dependence of the resistivity of La 1−x Sr x MnO 3 (x =<br />

0.175) under various magnetic fields [2]. . . . . . . . . . . . . . . . . . 4<br />

1.2 Schematic view of the cubic perovskite structure. . . . . . . . . . . . 5<br />

1.3 Phase diagram of temperature versus tolerance factor for the system<br />

R 0.7 A 0.3 MnO 3 , where R is a trivalent rare earth ion and A is a divalent<br />

alkaline earth ion [12]. . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

1.4 A schematic of the electronic structure of Mn 3+ (a) and Mn 4+ (b). . . 7<br />

1.5 Schematic features of the double exchange mechanism. . . . . . . . . 8<br />

1.6 Schematic view of the Jahn-Teller effect for the d 4 ion. . . . . . . . . 9<br />

1.7 Schematic representation of a macroscopic phase separated state (a),<br />

as well as possible charge inhomogeneous states stabilized by the longrange<br />

Coulomb interaction (spherical droplets in (b), stripes in (c))<br />

[20]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

1.8 Different types of spin arrangements. . . . . . . . . . . . . . . . . . . 11<br />

1.9 A Schematic picture of antiferromagnetic CE-type structure with the<br />

charge and e g orbital ordering. . . . . . . . . . . . . . . . . . . . . . . 12<br />

1.10 A schematic picture of the antiferromagnetic pseudo-CE-type structure<br />

in the case of Pr 1−x Ca x MnO 3 (x = 0.4) [38]. . . . . . . . . . . . . . . 13<br />

xvii


1.11 (a) DOS of the one-orbital model on a 10×10 cluster at J H = ∞ and<br />

temperature T = 1/130 (hopping t = 1). The four lines from the top<br />

correspond to densities 0.90, 0.92, 0.94, and 0.97. The inset has results<br />

at 〈n〉 = 0.86, a marginally stable density at T = 0. (b) DOS of the<br />

two-orbital model on a 20-site chain, working at 〈n〉 = 0.7, J H = 8, and<br />

λ = 1.5. Starting from the top at ω − µ = 0, the three lines represent<br />

temperatures 1/5, 1/10 and 1/20 respectively. Both (a) and (b) are<br />

taken from Ref. [40]. (c) DOS using a 20-site chain of the one-orbital<br />

model at T = 1/175, J H = 8, 〈n〉 = 0.87, and at a chemical potential<br />

such that the system is phase separated in the absence of disorder. W<br />

regulates the strength of the disorder, as explained in Moreo et. al.<br />

[41] from where this figure was taken. . . . . . . . . . . . . . . . . . . 14<br />

1.12 Schematic explanation of pseudogap formation at low electron density<br />

[40]. In (a) A typical Monte Carlo configuration of t 2g -spin is sketched.<br />

In (b) the corresponding e g density is shown, with electrons mostly<br />

located in the FM regions. In (c), the effective potential felt by the<br />

electrons is presented. A populated cluster band is formed (thick line).<br />

In (d) the resulting DOS is shown. . . . . . . . . . . . . . . . . . . . 15<br />

1.13 Phase diagram of Pr 1−x Ca x MnO 3 . Figure taken from ref. [44] . . . . 17<br />

1.14 (a) Temperature dependence of the resistivity of Pr 1−x Ca x MnO 3 with<br />

x = 0.3 at various magnetic fields. The inset is the phase diagram<br />

in the temperature magnetic field plane. The hatched area indicates<br />

the hysteretic region [45]. (b) The temperature dependence of the<br />

resistivity of Pr 1−x Ca x MnO 3 (x = 0.5, 0.4, 0.35 and 0.3) at magnetic<br />

fields of 0, 6 and 12 T [44]. . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.1 Principle of a modern photoemission experiment. The photon source is<br />

either UV light or X-ray, which hit the sample surface under an angle<br />

ψ with respect to the surface normal. The kinetic energy E kin of the<br />

photoelectrons can be analysed by use of electrostatic analysers. . . . 24<br />

xviii


2.2 Schematic view of the energetics of the photoemission process [3], which<br />

gives the relation between the energy levels in a solid and the electron<br />

energy distribution produced by photons of energy hν. The electron<br />

energy distribution which can be measured by the analyser as a function<br />

of kinetic energy (E kin ) is more conveniently expressed in terms of<br />

the binding energy E B when one refers to the density of states inside<br />

the solid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

2.3 ”Universal curve” for electron inelastic mean free path λ. . . . . . . . 30<br />

2.4 Photoemission spectrometer at Institute of Physics, Bhubaneswar. . . 31<br />

2.5 Schematic diagram of the VUV source. . . . . . . . . . . . . . . . . . 32<br />

2.6 Dimensional outline of the VUV source HIS 13. The viewport can be<br />

used for inspection of the discharge and for adjustment of the light<br />

spot position on the sample. The dimensions are in mm. . . . . . . . 32<br />

2.7 Schematic of the X-ray source in operation . . . . . . . . . . . . . . . 33<br />

2.8 Schematic diagram of the EA 125 hemispherical analyser. . . . . . . . 34<br />

2.9 The principle of operation of the concentric hemispherical analyser,<br />

schematic diagram. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

2.10 Schematic diagram of inverse photoemission spectroscopy (IPES). . . 39<br />

2.11 (a) IPES spectrometer at UGC-DAE consortium for scientific research,<br />

Indore, (b) External view of photon detector attached to the IPES<br />

chamber, and (c) Inside view of the chamber showing a rotatable (left)<br />

and fixed (right) electron gun and the sample holder . . . . . . . . . . 42<br />

2.12 A schematic diagram of the electron gun [30]. . . . . . . . . . . . . . 43<br />

2.13 Schematic cross section of the photon detector showing the window<br />

(1), O-ring (2), cylindrical cap (3), detector tube (4), teflon spacer (5),<br />

anode (6), double sided DN40 CF flanges (7, 10), gas inlet and outlet<br />

(8, 9), barrel connector (11), teflon sleeve (12), DN40 CF flange (13)<br />

and SHV connector (14). . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

2.14 Schematic diagram of x-ray absorption spectroscopy . . . . . . . . . . 46<br />

2.15 The x-ray loses its intensity via interactions with material. The absorption<br />

coefficient decreases smoothly with higher energy, except for<br />

special photon energies. When the photon energy reaches a critical<br />

value for a core electron transition, the absorption coefficient increases<br />

abruptly. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

xix


2.16 Schematic of the Si-photodiode. . . . . . . . . . . . . . . . . . . . . . 49<br />

2.17 Schematic of the synchrotron radiation. . . . . . . . . . . . . . . . . . 50<br />

2.18 Schematic representation of an insertion device. . . . . . . . . . . . . 51<br />

2.19 Optical layout of the WERA soft x-ray beam line, ANKA. . . . . . . 52<br />

2.20 Optical layout of the BEAR beam line. . . . . . . . . . . . . . . . . . 53<br />

2.21 Optical layout of the BACH beam line. . . . . . . . . . . . . . . . . . 53<br />

2.22 The variable included angle monochromator : the combined movement<br />

of PM1 (plane mirror) and one of the four spherical gratings SG1(-4)<br />

leads to the monochromatization of the light keeping both the entrance<br />

and exit slit fixed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

3.1 Valence band photoemission spectra of Pr 0.67 Ca 0.33 MnO 3 taken using<br />

He I photon energy (21.2 eV) at 77 (solid line), 110 (dotted), 150<br />

(dashed), 220 (dot-dashed), and 300 K (double-dot dashed). All the<br />

spectra have been normalized and shifted along y-axis by a constant<br />

for clarity. The subbands around -1.2 (e g↑ ), -3.5 (t 2g↑ ), and -5.6 eV (e g↑<br />

+ t 2g↑ ) are marked as A, B, and C respectively. . . . . . . . . . . . . 63<br />

3.2 High-resolution photoemission spectra of the near-E F region of the<br />

valence band of Pr 0.67 Ca 0.33 MnO 3 . In (a) the spectrum taken below T c<br />

(red) is compared with the normal state (300 K) spectrum (blue). The<br />

difference spectrum (green) obtained by subtracting the spectra at 300<br />

K from 77 K and multiplied by 5 is also shown in the panel. Similarly,<br />

in (b), (c), and (d) the near-E F spectra taken at 110, 150, and 220 K<br />

are compared with that taken at 300 K. The feature in the difference<br />

spectra corresponds to the tail feature A (e g↑ ) in Fig. 3.1. . . . . . . . 64<br />

3.3 Temperature dependence of the area of the three valence band features<br />

obtained from fitting the spectra with Lorentzian line shapes using a<br />

χ 2 iterative program. We have used an integral background, which was<br />

kept the same for all the spectra. The energy positions and FWHMs<br />

were determined by finding the best fit common to all the spectra by the<br />

iterative program. The final fit for all spectra at different temperatures<br />

were obtained with the same energy positions and FWHMs. (a), (b),<br />

and (c) correspond to the features A, B, and C positioned at -1.19,<br />

-3.49, and -5.63 eV with FWHMs 2.56, 1.93, and 1.37 eV respectively. 65<br />

xx


3.4 O K edge x-ray absorption spectra of Pr 0.67 Ca 0.33 MnO 3 taken at 300<br />

(solid line), 150 (dashed line), and 95 K (dotted). The pre-edge feature<br />

centered around 529.5 eV (marked by a box) is where most interest lies.<br />

Since this feature consists of two peaks (a main line and a shoulder on<br />

the low-energy side), this part of the spectrum was fitted with two<br />

components of Lorentzian line shapes. Results of the curve fit are<br />

given in Table 3.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

3.5 (a) Combined spectra from valence band photoemission (occupied energy<br />

states) and pre-peak in O K edge x-ray absorption spectra (unoccupied<br />

energy states) of Pr 0.67 Ca 0.33 MnO 3 taken at room temperature<br />

(above T c ). The three features corresponding to the subbands in the<br />

valence region are marked A, B, and C.The two peaks of the fitted O<br />

K edge pre-peak are marked A ′ and B ′ . We used integral background<br />

for both sides of E F . Details of the fitting are mentioned in the text.<br />

(b) Combined spectra of Pr 0.67 Ca 0.33 MnO 3 below T c . The valence band<br />

spectrum was taken at 77 K and O K edge XAS was taken at 95 K. . 70<br />

3.6 Schematic diagram of the near-E F energy levels of Pr 0.67 Ca 0.33 MnO 3<br />

in the occupied and unoccupied parts. The diagram is not drawn to<br />

scale. Some of the e g↑ states are occupied (at the Mn 3+ sites) and some<br />

are unoccupied (at the Mn 4+ sites). Hence the last occupied and first<br />

unoccupied states are marked with dotted lines. . . . . . . . . . . . . 71<br />

4.1 Phase diagram of the Pr 1−x Ca x MnO 3 system. In this study we have<br />

used the x = 0.2, x = 0.33 and x = 0.4 compositions. (FMI = ferromagnetic<br />

insulating, CO-AFMI = charge ordered antiferromagnetic<br />

insulating, CG = cluster glass, T N = Neel Temperature, T C = Critical<br />

Temperature, T CO = charge ordering temperature). . . . . . . . . . . 79<br />

4.2 The angle integrated valence band photoemission spectra of the Pr 1−x Ca x MnO 3<br />

(x = 0.0, 0.2, x = 0.33, x = 0.4 and x = 1.0) samples taken at room<br />

temperature and 77K using He I photons (21.2 eV). . . . . . . . . . . 81<br />

xxi


4.3 The near E F region of the valence band spectra (panel(a): 300 K,<br />

panel(b): 77 K) of the three samples plotted against that of CaMnO 3 .<br />

The difference spectra (lower panel) is obtained by subtracting the<br />

spectra of CaMnO 3 from those of the three Pr 1−x Ca x MnO 3 compositions<br />

and multiplied by 5. The difference spectra corresponds to the<br />

density of e g states. The red and blue peaks are A’ and A” respectively.<br />

The green line shows the spectral background used in the fitting. . . . 83<br />

4.4 The IPES spectra of the Pr 1−x Ca x MnO 3 samples (x = 0.2, 0.33 and<br />

0.4) and the CaMnO 3 sample taken at room temperature. The spectra<br />

represents the unoccupied electron states of the samples. . . . . . . . 85<br />

4.5 Panel(a) shows the IPES spectra of the near E F unoccupied states.<br />

Panel(b) shows the difference spectra obtained by subtracting the spectra<br />

corresponding to the x = 0.2 from those of other compositions. . . 86<br />

4.6 The Mn 2p core-level XPS (Mg Kα) spectra of Pr 1−x Ca x MnO 3 taken<br />

at room temperature (a) and 77 K (b) . . . . . . . . . . . . . . . . . . 88<br />

4.7 The Ca 2p core-level XPS (Mg Kα) spectra of Pr 1−x Ca x MnO 3 taken<br />

at room temperature (a) and 77 K (b). . . . . . . . . . . . . . . . . . 89<br />

4.8 The Pr 4d core-level XPS (Mg Kα) spectra of Pr 1−x Ca x MnO 3 taken<br />

at room temperature (a) and 77 K (b). . . . . . . . . . . . . . . . . . 90<br />

5.1 The angle integrated valence band photoemission spectrum of Ca 0.86 Pr 0.14 MnO 3<br />

taken using 124 eV photons. The spectra taken at different temperatures<br />

(30 K, 70 K, 110 K, 220 K and 293 K) are normalized for their<br />

intensities and shifted along Y-axis by a constant for clarity of presentation.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

5.2 The high resolution spectra of the near E F region of the valence band<br />

of Ca 0.86 Pr 0.14 MnO 3 for 30 K, 70 K, 110 K, 220 K and 300 K. This<br />

feature refers to the e g band. . . . . . . . . . . . . . . . . . . . . . . . 100<br />

5.3 The O K-edge x-ray absorption spectra of the Ca 0.86 Pr 0.14 MnO 3 taken<br />

at 30 K, 80 K, 150 K, 200 K and 300 K. The pre-edge feature appears<br />

between 527 - 533 eV is mostly due to the strong hybridization between<br />

Mn 3d and O 2p orbitals, where most interest lies. Origin of other<br />

features are given in the text. . . . . . . . . . . . . . . . . . . . . . . 101<br />

5.4 The pre-edge peak of the O K -edge XAS spectra of Ca 0.86 Pr 0.14 MnO 3<br />

for temperatures 30 K, 80 K, 150 K, 200 K and 300 K. . . . . . . . . 102<br />

xxii


Chapter 1<br />

Introduction<br />

1


Introduction 2<br />

1.1 An overview<br />

The perovskite type manganites having general formula R 1−x A x MnO 3 (where R is a<br />

trivalent rare earth element like La, Pr, Nd. and A is a divalent alkaline earth element<br />

like Sr, Ca, Ba.) have been one of the interesting compounds among the transition<br />

metal oxide systems due to their intriguing physical property such as colossal magnetoresistance<br />

(CMR) [1, 2, 3]. CMR is defined as the colossal decrease of resistance by<br />

the application of magnetic field. These materials are insulators at high temperatures<br />

and poor metals at low temperatures. Accompanying this insulator-metal transition<br />

is a magnetic transition from a high temperature paramagnetic phase to a low temperature<br />

ferromagnetic phase. These transitions occurs around the same transition<br />

temperature T C . Applying a magnetic field around T C will greatly reduce the resistivity,<br />

i.e negative magnetoresistance effect. These materials were first described in<br />

1950 by Jonker and van Santen [4] in perovskite type manganites.<br />

One year later (1951), Zener [5, 6] explained this unusual correlation between<br />

magnetism and transport properties by introducing a novel concept, so-called ”Double<br />

exchange” mechanism (DE). The pioneering work of Zener was followed by Anderson<br />

and Hasegawa [7] in 1955 and de Gennes [8] in 1960. Double exchange theory<br />

argues that the electron hopping is related to the relative spin orientations of the<br />

neighbouring sites. Although it qualitatively explains the connection between the<br />

insulator metal and the magnetic transitions, it was noticed latter by A. J. Millis et.<br />

al. [9] that it fails to explain the quantitative change in resistivity and the very low<br />

T C . Thus it is believed that double-exchange mechanism is not enough to explain<br />

the CMR effect. Several mechanisms, such as electron-phonon interactions (the polaronic<br />

effect), electronic phase separation, charge and/or orbital ordering etc. have<br />

been proposed to account for the discrepancy. Although the detailed picture is still<br />

unclear, there seems to be a growing consensus that one or more of these additional<br />

effects probably exists, with the various phenomena both competing and cooperating<br />

with each other. This can lead to very rich and exotic behaviour, as observed in the<br />

manganites. This mixed valence manganites have been studied for more than five<br />

decades but are still considered modern materials because of their wide potential for<br />

technological applications. Potential applications of the CMR effect in these manganites<br />

include magnetic sensors, magnetoresistive read heads, magnetoresistive random<br />

access memory (MRAM) etc.


Introduction 3<br />

The temperature and doping dependent metal-insulator transitions in these materials<br />

are found to be closely related to the unique electronic structure derived from<br />

the Mn 3d and O 2p hybridized orbitals of MnO 6 octahedra. The 3d orbitals of Mn<br />

in the MnO 6 octahedra, which are split by the crystal field into t 2g and e g states,<br />

are further split by the Jahn-Teller distortion. With charge carrier doping, many of<br />

them show well defined charge ordering (CO) and/or orbital ordering (OO) at low<br />

temperatures, especially when x equals a commensurate value (e.g x = 0.5). The CO<br />

state where the Mn 3+ (t 3 2ge 1 g) and Mn 4+ (t 3 2g) ions arranged like in a checkerboard was<br />

found to bear a strong influence on the one electron band width (W) of the e g band<br />

and the transfer interaction of the e g holes (electrons).<br />

The work presented in this thesis consists of the investigations of the electronic<br />

structure of Pr 1−x Ca x MnO 3 using different spectroscopic techniques as photoemission<br />

spectroscopy, inverse photoemision spectroscopy and x-ray absorption spectroscopy<br />

etc. The thesis is structured as follows : Following this overview, the Chapter 1 gives<br />

a general idea about the physical properties (structural, electronic and magnetic) and<br />

the mechanisms associated with the CMR manganites. Chapter 2 describes the principles<br />

and operations of the experimental techniques used for this thesis work. Chapter<br />

3 presents the electronic structure of Pr 0.67 Ca 0.33 MnO 3 at different temperatures<br />

using ultra-violet photoelectron and x-ray absorption spectroscopy. The electronic<br />

structure of the Pr 1−x Ca x MnO 3 series at different temperatures using ultra-violet<br />

photoemission and inverse photoemission spectroscopy are discussed in Chapter 4.<br />

In Chapter 5, the electronic structure of Ca 0.86 Pr 0.14 MnO 3 at different temperatures<br />

using high resolution photoemission and x-ray absorption spectroscopy are discussed.<br />

Chapter 6 gives the summary and conclusions of this thesis.<br />

1.2 Colossal magnetoresistance<br />

Colossal magnetoresistance is a property of some materials, which is the gigantic<br />

decrease of resistance by the application of magnetic field. The magnetoresistance<br />

can be defined as,<br />

MR =<br />

R(H) − R(0)<br />

, (1.1)<br />

R(0)<br />

where R(H) and R(0) are the resistances with and without magnetic field H respectively.<br />

Expressing the results as a percentage (i.e multiplying by an additional factor<br />

100), it has been shown by Jin et. al. [10] that the MR value as large as -100,000


Introduction 4<br />

Figure 1.1: Temperature dependence of the resistivity of La 1−x Sr x MnO 3 (x = 0.175)<br />

under various magnetic fields [2].<br />

% near 77 K. Which corresponds to thousand-fold magnetoresistance. The temperature<br />

dependent resistivity of La 1−x Sr x MnO 3 (x = 0.175) single crystal under several<br />

magnetic fields are shown in the Fig. 1.1.<br />

1.3 Crystal structure<br />

The CMR samples used in this thesis work are of perovskite type (ABO 3 ) structure.<br />

The general formula for perovskite type manganites is R 1−x A x MnO 3 , where R is the<br />

trivalent rare earth element such as La, Pr, Nd, and Sm etc. and A is the divalent<br />

alkaline earth ion such as Sr, Ca, and Ba etc. An ideal cubic perovskite structure is<br />

shown in the Fig. 1.2, which contains a MnO 6 octahedron, with one Mn ion at the<br />

center and oxygen ions at each of the six corners. The trivalent or divalent ions (R<br />

or A) are situated at the corners of the cube. The stability of the cubic perovskite<br />

structure depends strongly on the size of the cations (R, A or Mn). If there is a<br />

size mismatch occurs between the cations, then the cubic perovskite structure gets


Introduction 5<br />

Figure 1.2: Schematic view of the cubic perovskite structure.<br />

distorted. This size mismatch is governed by a factor called tolerance factor (t), and<br />

is defined by Goldschmidt [11] as,<br />

t = (r A + r O )<br />

√<br />

2(rMn + r O ) , (1.2)<br />

where r A , r Mn and r O are the ionic radii of the A-site(R or A), Mn- and O- atoms<br />

respectively. For cubic perovskite structure t = 1, where the Mn-O-Mn bond angles<br />

are relatively straight. Due to the ionic size mismatch the Mn-O-Mn bond angles<br />

deviate from 180 o and lead to distorted perovskite structure. When t decreases (0.96 <<br />

t < 1), the cubic structure transforms to rhombohedral and then to the orthorhombic<br />

structure (t < 0.96). The properties of manganites strongly depend on the tolerance<br />

factor. In 1995 Hwang et. al. [12] has reported the structure-property relationship<br />

as a function of tolerance factor for 30 % dopant concentration with a variety of<br />

rare earth and alkaline earth ions, shown in the Fig. 1.3. It shows the presence<br />

of three regimes: a paramagnetic insulator at high temperature, a low-temperature<br />

ferromagnetic metal at large tolerance factor and a low-temperature ferromagnetic<br />

insulator at small tolerance factor.<br />

1.4 The basic electronic structure<br />

In manganites the electronically active orbitals are the Mn 3d orbitals. In the cubic<br />

lattice, the five-fold degenerate 3d-orbitals of an isolated Mn atom or ion are split into


Introduction 6<br />

Figure 1.3: Phase diagram of temperature versus tolerance factor for the system<br />

R 0.7 A 0.3 MnO 3 , where R is a trivalent rare earth ion and A is a divalent alkaline earth<br />

ion [12].<br />

a triple degenerate (lower energy levels) d xy , d yz , and d zx orbitals, called t 2g states<br />

and a double degenerate (higher energy levels) d x 2 −y 2 and d 3z 2 −r 2 orbitals, called e g<br />

states [13]. Energy separation between the t 2g states and e g states is 10Dq due to the<br />

crystal field splitting of MnO 6 octahedra. The schematic of the electronic structure<br />

from ionic point of view are shown in the Fig. 1.4. Since the e g orbitals point towards<br />

the six O atoms, they are expected to have a larger hybridization with the O 2p<br />

states (p x , p y , and p z ) and hence are more delocalized than the t 2g states. So the e g<br />

state electrons play an important role of conduction electrons in this type of crystals.<br />

In Mn 3+ ions three electrons occupy the t 2g states and one electron occupies the e g<br />

states. The strong Hund’s coupling (J H ) in these systems favors the alignment of<br />

the e g electron spins with the core-like t 2g spins. When the manganites are doped<br />

with holes through chemical substitution, the Mn 4+ ions are formed. In Mn 4+ ions,<br />

there are only three electrons occupying the t 2g orbitals. When the excess electron is<br />

present in case of Mn 3+ ion, a further distortion known as the Jahn-teller distortion<br />

may lower the crystal symmetry and break the degeneracy of the e g states. This can<br />

be energetically favorable as the excess electron will fill the lower energy orbital while<br />

the upper energy orbital remains empty.


Introduction 7<br />

Figure 1.4: A schematic of the electronic structure of Mn 3+ (a) and Mn 4+ (b).<br />

1.5 Double exchange mechanism<br />

The starting point for understanding the properties of manganites is the so-called<br />

”double” exchange (DE) mechanism and was first proposed by C. Zener [5, 6] in<br />

1951, which says the simultaneous hopping of e g elctrons from Mn 3+ sites to O 2−<br />

sites and then from O 2− to the empty e g orbitals of Mn 4+ sites. This transfer of e g<br />

electron from Mn 3+ to Mn 4+ by DE is the basic mechanism of electrical conduction<br />

in manganites. In 1955, the DE theory was expanded by P. W. Anderson, and H.<br />

Hasegawa [7], where the hopping probability (t ij ) of the e g electron from Mn 3+ to<br />

neighbouring Mn 4+ is,<br />

t ij = t cos( θ ij<br />

), (1.3)<br />

2<br />

where θ ij is the angle between the Mn spins in the case of strong Hund’s coupling.<br />

The concept of double exchange mechanism is illustrated in Fig. 1.5. For parallel spin<br />

configuration (θ ij = 0) the hopping probability is maximum and for antiparallel spin<br />

(θ ij = 180 o ), the hopping probability is zero, that means the hopping cancels. Thus it<br />

predicts maximum conductivity at ferromagnetic state where as insulating behaviour<br />

at antiferromagnetic state. This provides the qualitative explanation for the close<br />

connection between the insulator-metal transition and the magnetic transition. It<br />

also explains why the external magnetic field will increase the conductivity since θ ij<br />

will be reduced as the external magnetic field polarizes the core (t 2g ) spins of the<br />

Mn ions. Using mean-field approximations, P. -G. degennes [8] in 1960 suggested<br />

that the interpolation between the antiferromagnetic state of the undoped limit and<br />

the ferromagnetic state at finite hole density, where the DE mechanism works, occurs


Introduction 8<br />

Figure 1.5: Schematic features of the double exchange mechanism.<br />

through a ”canted” state, similar as the state produced by a magnetic field acting over<br />

an antiferromagnetic state. Although DE mechanism gives an intuitive explanation<br />

for the coupling between the spin and charge degrees of freedom, it fails to predict<br />

the correct magnitude of the change in resistivity. For paramagnetic case (T > T C ),<br />

θ ij = 90 o , and so the hopping probability will be reduced to cos ( 90o ) = 0.7 times<br />

2<br />

that in the ferromagnetic case. While experimentally it is seen that the conductivity<br />

decreases across the ferro - para transition may be many orders of magnitude. Which<br />

says that some other mechanisms beyond DE are necessary to explain the conductivity<br />

change across the transition.<br />

1.6 Jahn-Teller effect in CMR manganites<br />

A. J. Millis et. al. [9] in 1995 pointed out that, the DE model alone is not enough to<br />

explain the CMR effect in manganites, where he argued that DE produces wrong T C<br />

by a large factor and the resistivity that grows with reducing temperature (insulating<br />

behaviour) even below T C . By taking these reasons Millis et. al. [9] concluded that<br />

the model based only on a large J H is not adequate for the manganites and claimed<br />

that the necessity of Jahn-Teller phonons to explain the change in curvature of the<br />

resistivity close to T C . The e g electrons on a given MnO 6 octahedron are strongly<br />

coupled to the distortions of the O 6 octahedron. In the pseudocubic materials such<br />

as La 1−x Ca x MnO 3 the strong coupling is due in part to the Jahn-Teller effect, the


Introduction 9<br />

Figure 1.6: Schematic view of the Jahn-Teller effect for the d 4 ion.<br />

essence of which is that a local distortion in which some Mn-O bonds get shorter and<br />

other gets longer breaks the local cubic symmetry and therefore splits the degeneracy<br />

of the e g level on that site (Fig. 1.6). If only one electron is present then it will in<br />

the absence of hybridization reside in the lower level and gain energy. For example,<br />

a distortion in which the Mn-O bond get longer in the ±z direction and get shorter<br />

in the ±x,y directions raise the energy of the d x 2 −y2 orbital and lowers the energy of<br />

the d 3z 2 −r 2 orbital (Q 3 mode). Millis et. al. [14, 15, 17, 18] argued that the physics of<br />

manganites is dominated by the interplay between a strong electron-phonon coupling<br />

and the large Hund’s coupling effect that optimizes the electronic kinetic energy by the<br />

generation of a FM-phase. The large value of electron-phonon coupling in manganites<br />

is clear in the regime below x = 0.2, where a static JT distortion plays a key role in the<br />

physics of the material and a dynamical JT effect may persists at higher hole densities<br />

[16], without leading to long-range order but producing important fluctuations that<br />

localize electrons by splitting the degenerate e g levels at a given MnO 6 octahedron.<br />

Millis et. al [14] argued that the ratio λ eff = E JT<br />

t eff<br />

(coupling strength) dominates<br />

the physics of the problem. Here E JT is the static trapping energy at a given octahedron,<br />

and t eff is an effective hopping that is temperature dependent following the<br />

standard DE discussion. In this context it was conjectured that when the temperature<br />

is larger than T C the effective coupling λ eff could be above the critical value that<br />

leads to the insulating behavior due to electron localization, while it becomes smaller


Introduction 10<br />

Figure 1.7: Schematic representation of a macroscopic phase separated state (a), as<br />

well as possible charge inhomogeneous states stabilized by the long-range Coulomb<br />

interaction (spherical droplets in (b), stripes in (c)) [20].<br />

than the critical value below T C , thus inducing metallic behavior. The calculations<br />

were carried out using classical phonons and t 2g spins. The results of Millis et. al.<br />

[14] for T C and the resistivity at a fixed density n = 1 when plotted as a function of<br />

λ eff had formal similarities with experimental results (which are produced as a function<br />

of density). In particular, if λ eff is tuned to be very close to the metal-insulator<br />

transition, the resistivity naturally strongly depends on even small external magnetic<br />

fields.<br />

1.7 Phase separation<br />

Another possibly complementary route to understand the CMR effect in manganites<br />

is the phase separation. It has been recently predicted from computational studies<br />

of realistic models that an electronic phase separation can occur in manganites in a<br />

certain range of doping [19]. In particular at low doping level and at low temperature,<br />

a phase separation between hole-poor antiferromagnetic (AFM) regions and hole-rich<br />

ferromagnetic (FM) regions is energetically more favourable than the homogeneous<br />

canted AFM phase. The energy of charge carrier is minimal for FM ordering. With<br />

a density insufficient for establishing the FM ordering in the entire sample, the carriers<br />

concentrate into droplets or stripes which become FM inside insulating AFM<br />

matrix [20](Fig. 1.7). The fluctuations length scales range from nm to µm. This<br />

point towards the conduction in metallic phase and the CMR effect is argued as a


Introduction 11<br />

Figure 1.8: Different types of spin arrangements.<br />

result of percolation [21, 22, 23]. On increasing the doping level the FM regions become<br />

connected to each other and a low electrical resistance is achieved above the<br />

percolation threshold. There are many reports about the coexistence of FM clusters<br />

in the CO AFM insulating phase of La 1−x Ca x MnO 3 using different experimental<br />

techniques [23, 24, 25, 26, 27]. And very recently the neutron diffraction and the<br />

inelastic neutron scattering studies of Pr 1−x Ca x MnO 3 have shown the coexistence of<br />

ferromagnetic fluctuations in antiferromagnetic domains [28, 29, 30, 31].<br />

1.8 Spin, Charge and Orbital ordering in manganites<br />

The unusual properties of manganites challenge the current topic of research in the<br />

field of strongly correlated electrons system due to their strong spin, charge and<br />

orbital degrees of freedom. The strong interplay between spin, charge and orbital<br />

ordering originating from the single occupancy of the doubly degenerate e g orbitals<br />

of Mn 3+ (t 3 2g e1 g ) ions make the phase diagram of these compounds rich with physics.<br />

With charge carrier doping, some of them for example, La 1−x Ca x Mon 3 system show<br />

strong charge and/or orbital ordered (CO/OO) insulating nature at low temperatures,<br />

especially when x equals a commensurate value. The CO state where the Mn 3+ (t 3 2ge 1 g)<br />

and Mn 4+ (t 3 2g ) are arranged like a checkerboard influences strongly the one electron<br />

bandwidth of the e g band [32] and the transfer interaction of the e g holes (electrons).<br />

There are various types of ordering occurs across the phase diagram of manganites,<br />

which are described below.<br />

The lattice distortion in manganites is due to the J-T interaction acting on the<br />

Mn 3+ ion, which lifts the degenerate e g orbitals in the cubic electronic field. The J-T


Introduction 12<br />

Figure 1.9: A Schematic picture of antiferromagnetic CE-type structure with the<br />

charge and e g orbital ordering.<br />

distorted Mn 3+ O 6 in LaMnO 3 form an orthorhombic structure (space group P bnm )<br />

where the e g charge occupies alternating 3d 3x 2 −r 2 and 3d 3y 2 −r2 orbitals. The magnetic<br />

structure of LaMnO 3 was found to be of A-type antiferromagnetic, where all Mn 3+<br />

moments lying in c-plane are aligned ferromagnetically and the the FM layer stacks<br />

antiferromagnetically along the c-axis by the weaker superexchange interaction along<br />

this axis. The A-type AFM ordering exists below 140 K [33]. Different types of spin<br />

arrangements are shown in the Fig. 1.8. Upon substituting divalent cations such<br />

as Ca, Sr, Pb etc. to the parent compound like LaMnO 3 , the magnetic structure<br />

changes from A-type AFM to the FM with the rotation of the spin direction from<br />

the b-axis in the c-plane towards the c-axis associated with the metallic conduction,<br />

which was interpreted successfully in terms of DE mechanism at an early stage of the<br />

investigation. Further increase of doping, the charge ordered (CO) state near x = 0.5<br />

is observed. For example in La 1−x Ca x MnO 3 (x = 0.5), the ordering becomes C-E type<br />

(Fig. 1.9), where the Mn 3+ and Mn 4+ line up forming alternate stripes in the C plane<br />

and stack also along the c-axis. This phenomenon in manganites was first observed<br />

by Wollan and Koehler [34] in the neutron powder diffraction of La 0.5 Ca 0.5 MnO 3 .<br />

Later these charge-ordering phases are observed in the low band width compound<br />

Pr 1−x Ca x MnO 3 with a broad doping region 0.3 < x < 0.75 [28, 35, 36, 37, 38].<br />

The CO state for Pr 0.6 Ca 0.4 MnO 3 is called pseudo-CE type ordering [38] is shown


Introduction 13<br />

Figure 1.10: A schematic picture of the antiferromagnetic pseudo-CE-type structure<br />

in the case of Pr 1−x Ca x MnO 3 (x = 0.4) [38].<br />

in Fig. 1.10, where the arrangement of spin is canted along the c-axis while the<br />

CE-type ordering is kept within the orthorhombic ab-plane. For the ground state<br />

of R 1−x A x MnO 3 , i.e AMnO 3 (x = 1), the magnetic structure is called G-type AFM<br />

[34, 39] ordering in which the spins on all Mn 4+ sites are antiparallel to their nearest<br />

neighbours.<br />

1.9 Pseudogap behavior in manganites<br />

The conducting properties of the materials depend on the finer changes associated<br />

with the density of states (DOS) near the Fermi level. If there are no allowed states<br />

at the Fermi level a gap forms by making the system insulating and finite density of<br />

states makes the system metallic. A Pseudogap exists in the mixed-phase regimes of<br />

manganites, which suggests that the prominent depletion of spectral weight (density<br />

of states) at the chemical potential, but not exactly zero (few states associated with<br />

it). This feature is similar to that extensively discussed in high T c superconductor.<br />

The calculations in the pseudogap context in manganites was carried out by Moreo<br />

et. al. [40, 41] using both one- and two-orbital models (Fig. 1.11). The density of<br />

states of the one-orbital model on 2D cluster varying the electronic density slightly<br />

below 〈n〉 = 1.0 is shown in the Fig. 1.11(a). At zero temperature, this density<br />

regime is unstable due to phase separation, but at the temperature of simulation<br />

those densities still correspond to stable states, but with a dynamical mixture of


Introduction 14<br />

Figure 1.11: (a) DOS of the one-orbital model on a 10×10 cluster at J H = ∞ and<br />

temperature T = 1/130 (hopping t = 1). The four lines from the top correspond to<br />

densities 0.90, 0.92, 0.94, and 0.97. The inset has results at 〈n〉 = 0.86, a marginally<br />

stable density at T = 0. (b) DOS of the two-orbital model on a 20-site chain, working<br />

at 〈n〉 = 0.7, J H = 8, and λ = 1.5. Starting from the top at ω − µ = 0, the three<br />

lines represent temperatures 1/5, 1/10 and 1/20 respectively. Both (a) and (b) are<br />

taken from Ref. [40]. (c) DOS using a 20-site chain of the one-orbital model at<br />

T = 1/175, J H = 8, 〈n〉 = 0.87, and at a chemical potential such that the system is<br />

phase separated in the absence of disorder. W regulates the strength of the disorder,<br />

as explained in Moreo et. al. [41] from where this figure was taken.


Introduction 15<br />

Figure 1.12: Schematic explanation of pseudogap formation at low electron density<br />

[40]. In (a) A typical Monte Carlo configuration of t 2g -spin is sketched. In (b) the<br />

corresponding e g density is shown, with electrons mostly located in the FM regions.<br />

In (c), the effective potential felt by the electrons is presented. A populated cluster<br />

band is formed (thick line). In (d) the resulting DOS is shown.<br />

AFM and FM features. A clear minimum in the density of states at the chemical<br />

potential can be observed. Very similar results also appear in 1D simulations [40].<br />

Fig. 1.11(b) shows the results for two-orbitals and a large electron-phonon coupling,<br />

at a fixed density and changing temperature. Clearly a pseudogap develop in the<br />

system as a precursor of the phase separation that is reached as the temperature is<br />

further reduced. Similar results have been obtained in other parts of parameter space,<br />

as long as the system is near unstable phase separated regimes. A pseudogap appears<br />

also in cases where disorder is added to the system. Fig. 1.11(c) shows the results<br />

for the case where a random on-site energy is added to the one orbital model.<br />

Moreo et. al. [40] explained this phenomenon tentatively for the case of ”without<br />

disorder”. A schematic of explanation is shown in the Fig. 1.12. In Fig. 1.12(a) a<br />

typical mixed phase FM-AFM state is sketched. In the FM regions, the e g electrons<br />

improve their kinetic energy, and thus they prefer to be located in those regions as<br />

shown in Fig. 1.12(b). The FM domains act as effective attractive potentials for<br />

electrons, as shown in Fig. 1.12(c). When other electrons are added FM clusters<br />

are created and new occupied levels appear below the chemical potential, creating<br />

a pseudogap [Fig. 1.12(d)]. These results are compatible with the photoemission


Introduction 16<br />

experiments by Dessau et. al. [42] for bilayer manganites (La 2−2x Sr 1+2x Mn 2 O 7 , x =<br />

0.4). In this experiment it was observed that the low-temperature ferromagnetic<br />

state was very different from a prototypical metal. Its resistivity is unusually high,<br />

the width of ARPES features are anomalously broad, and they don’t sharpen as they<br />

approach the Fermi momentum. In addition, the centroid of the near E F peak never<br />

approach close than approximately 0.65 eV to the Fermi level. This implies that<br />

even in the expected metallic regime, the density of states at the Fermi level is very<br />

small. Dessau et. al. [42] refers to these results as the formation of a pseudogap. This<br />

effect is present both in ferromagnetic and paramagnetic regimes. Scanning tunneling<br />

microscopy data of single crystals and thin films of hole doped manganites by Biswas<br />

et. al. [43] showed a rapid variation in the density of states for the temperatures<br />

near the Curie temperature, such that below T c a finite density of states is observed<br />

at the Fermi level while above T c a hard gap opens up. This result suggests that<br />

the presence of a gap or pseudogap is not just a feature of bilayers, but it appears in<br />

other manganites as well.<br />

1.10 Basic properties, phase diagram and CMR<br />

effect in Pr 1−x Ca x MnO 3 system<br />

The Pr 1−x Ca x MnO 3 is a low band width manganite. The band width W is smaller<br />

than in other manganites like La 1−x Sr x MnO 3 and La 1−x Ca x MnO 3 system. In the lowbandwidth<br />

compounds, a charge-ordered (CO) state is stabilized in the vicinity of x =<br />

0.5, while manganite with large W (La 1−x Sr x MnO 3 as example) present a metallic<br />

phase at this concentration. The Pr 1−x Ca x MnO 3 presents a stable CO state in a<br />

broad density region between x = 0.3 and 0.75, as showed by Jirak et. al [38]. Part of<br />

the phase diagram is shown in the Fig. 1.13 [44]. A ferromagnetic insulating (FMI)<br />

state exists in the range from x = 0.15 to 0.30. For x≥0.30, an antiferromagnetic CO<br />

state is stabilized. Neutron diffraction studies by Jirak et. al. [38] showed that at<br />

all densities between 0.30 and 0.75, the arrangement of charge/spin/orbital order of<br />

this state is similar to the CE-state (Fig. 1.9). However, certainly the hole density is<br />

changing with x, and as consequence the CE-state can not be perfect at all densities<br />

but electrons have to be added or removed from the structure. Jirak et. al. [38]<br />

discussed a ”pseudo”-CE-type structure for x = 0.4 that has the proper density.<br />

The effect of magnetic fields on the CO-state of Pr 1−x Ca x MnO 3 is remarkable. The


Introduction 17<br />

Figure 1.13: Phase diagram of Pr 1−x Ca x MnO 3 . Figure taken from ref. [44]<br />

resistivity vs. temperature graph of Pr 1−x Ca x MnO 3 (x = 0.3) under various magnetic<br />

fields [45] is shown in the Fig. 1.14(a). At low temperatures, changes in resistivity<br />

by several orders of magnitude can be observed. The stabilization of metallic state is<br />

realized upon the application of magnetic field. This state is ferromagnetic according<br />

to magnetization measurements, and thus it is curious to observe that a state not<br />

present at zero field in the phase diagram, is nevertheless stabilized at finite fields.<br />

The magnetic field effects on the other compositions of Pr 1−x CaxMnO 3 (x = 0.35,<br />

0.4 and 0.5) are also shown in the Fig. 1.14(b). The shape of the curves in Fig. 1.14<br />

present a large magnetoresistance effect, and a possible origin based on percolation<br />

between the CO- and FM-phase. Tomioka et. al. [46] showed that, as x grows away<br />

from x = 0.3, larger fields are needed to destabilize the charge-ordered state at low<br />

temperatures (e.g., 27 T at x = 0.50 compared with 4 T at x = 0.30).<br />

This thesis is based on the study of electronic structure of Pr 1−x Ca x MnO 3 series<br />

using different spectroscopic techniques.


Introduction 18<br />

Figure 1.14: (a) Temperature dependence of the resistivity of Pr 1−x Ca x MnO 3 with<br />

x = 0.3 at various magnetic fields. The inset is the phase diagram in the temperature<br />

magnetic field plane. The hatched area indicates the hysteretic region [45]. (b) The<br />

temperature dependence of the resistivity of Pr 1−x Ca x MnO 3 (x = 0.5, 0.4, 0.35 and<br />

0.3) at magnetic fields of 0, 6 and 12 T [44].


Bibliography<br />

[1] R. von Helmolt, J. Wecker, B. Holzapfel, L. Schultz, and K. Samwer; Phys. Rev.<br />

Lett. 71, 2331 (1993).<br />

[2] Y. Tokura, A. Urushibara, Y. Moritomo, T. Arima, A. Asamitsu, G. Kido, and<br />

N. Furukawa, J. Phys. Soc. Jpn. 63, 3931 (1994)<br />

[3] A. Urushibara, Y. Moritomo, T. Arima, A. Asamitsu, G. Kido, and Y. Tokura;<br />

Phys. Rev.B 51, 14103 (1995).<br />

[4] G. H. Jonker and J. H. Van Santen Physica 16, 337 (1950)<br />

[5] C. Zener, Phys. Rev. 81, 440 (1951)<br />

[6] C. Zener, Phys. Rev. 82, 403 (1951)<br />

[7] P. W. Anderson and H. Hasegawa, Phys. Rev. 100, 675 (1955)<br />

[8] P. G. de Gennes, Phys. Rev. 118, 141 (1960)<br />

[9] A. J. Millis, B. I. Shraiman, and P. B. Lttlewood, Phys. Rev. Lett. 74, 5144<br />

(1995)<br />

[10] S. Jin, T. H. Tiefel, M. McCormack, R. A. Fastnacht, R. Ramesh, and L. H.<br />

Chen, Science 264, 413 (1994)<br />

[11] V. Goldschmidt, Geochemistry, Oxford University Press (1958).<br />

[12] H. Y. Hwang, S.-W. Cheong, P. G. Radaelli, M. Marezio, and B. Batlogg, Phys.<br />

Rev. Lett. 75, 914 (1995)<br />

[13] J. M. D. Coey, M. Viret, L. Ranno, and K. Ounadjela, Phys. Rv. Lett. 75, 3910<br />

(1995)<br />

19


Introduction 20<br />

[14] A. J. Millis, B. I. Shraiman, and R. Mueller, Phys. Rev. Lett. 77, 175 (1996)<br />

[15] A. J. Millis, B. I. Shraiman, and R. Mueller, Phys. Rev. B 54, 5389 (1996)<br />

[16] A. J. Millis, R. Mueller, and B. I. Shraiman, Phys. Rev. B 54, 5405 (1996)<br />

[17] A. J. Millis, Nature 392, 147 (1998)<br />

[18] A. J. Millis, Phys. Rev. Lett. 80, 4358 (1998)<br />

[19] S. Yunoki, J. Hu, A. L. Malvezzi, A. Moreo, N. Furukawa, and E. Dagotto, Phys.<br />

Rev. Lett. 80, 845 (1998)<br />

[20] A. Moreo, S. Yunoki, and E. Dagotto, Science 283, 2034 (1999)<br />

[21] M. Fath, S. Freisem, A. A. Menovsky, T. Tomioka, J. Aarts, and J. A. Mydosh,<br />

Science 285, 1540 (1999)<br />

[22] M. Uehara, S. Mori, C. H. Chen, and S. -W. Cheong, Nature 399, 560 (1999)<br />

[23] S. Mori, C. H. Chen, and S. -W. Cheong, Nature 392, 473 (1998)<br />

[24] C. H. Chen and S. -W. Cheong, Phys. Rev. Lett. 76, 4042 (1996)<br />

[25] P. G. Radaelli, D. E. Cox, M. Marezio, and S. W. Cheong, Phys. Rev. B 55,<br />

3015 (1997)<br />

[26] G. Papavassiliou, M. Fardis, M. Belesi, M. Pissas, I. Panagiotopoulos, G. Kallias,<br />

D. Niarchos, C. Dimitropoulos, and J. Dolinsek, Phys. Rev. B 59, 6390 (1999)<br />

[27] G. Allodi, R. De Renzi, F. Licci, and M. W. Pieper, Phys. Rev. Lett. 81, 4736<br />

(1998)<br />

[28] R. Kajimoto, T. Kakeshita, Y. Oohara, H. Yoshizawa, Y. Tomioka, and Y.<br />

Tokura, Phys. Rev. B 58, R11837 (1998)<br />

[29] P. G. Radaelli, R. M. Ibberson, D. N. Argyriou, H. Casalta, K. H. Andersen, S.<br />

-W. Cheong, and J. F. Mitchell, Phys. Rev. B 63, 172419 (2001)<br />

[30] S. Mercone, V. Hardy, C. Martin, Ch. Simon, D. Saurel, and A. Brulet, Phys.<br />

Rev. B 68, (2003)


Introduction 21<br />

[31] Damien Saurel, Annie Brulet, Andre Heinemann, Christine Martin, Silvana Mercone,<br />

and Charles Simon, Phys. Rev. B 73, 094438 (2006)<br />

[32] Y. Tomioka, A. Asamitsu, Y. Moritomo, H. Kuwahara and Y. Tokura, Phys.<br />

Rev. Lett. 74, 5108 (1995)<br />

[33] Y. Murakami, J. P. Hill, D. Gibbs, M. Blume, I. Koyama, M. Tanaka, H. Kawata,<br />

T. Arima, Y. Tokura, K. Hirota, and Y. Endoh, Phys. Rev. Lett. 81, 582 (1998)<br />

[34] E. O. Wollan, W. C. Koehler, Phys. Rev. 100, 545 (1955)<br />

[35] M. V. Zimmermann, C. S. Nelson, J. P. Hill, D. Gibbs, M. Blume, D. Casa, B.<br />

Keimer, Y. Murakami, Y. Tomioka, and Y. Tokura, Phys. Rev. Lett. 83, 4872<br />

(1999)<br />

[36] R. Kajimoto, H. Yoshizawa, H. Kawano, H. Kuwahara, Y. Tokura, K. Ohoyama,<br />

and M. Ohasi, Phys. Rev. B 60, 9506 (1999)<br />

[37] D. E. Cox, P. G. Radaelli, M. Marezio, and S. -W. Cheong, Phys. Rev. B 57,<br />

3305 (1998)<br />

[38] Z. Jirak, S. Krupicka, Z. Simsa, M. Dlouha, and Z. Vratislav, J. Magn. Magn.<br />

Mater. 53, 153 (1985)<br />

[39] O. Chmaissem, B. Dabrowski, S. Kolesnik, J. Mais, D. E. Brown, R. Kruk, P.<br />

Prior, B. Pyles, and J. D. Jorgensen, Phys. Rev. B 64, 134412 (2001)<br />

[40] A. Moreo, S. Yunoki, and E. Dagotto, Phys. Rev. Lett. 83, 2773 (1999)<br />

[41] A. Moreo, M. Mayr, A. Feiguin, S. Yunoki, and E. Dagotto, Phys. Rev. Lett. 84,<br />

5568 (2000)<br />

[42] D. S. Dessau, T. Saitoh, C. -H. Park, Z. -X. Shen, P. Villella, N. Hamada, Y.<br />

Moritomo, and Y. Tokura, Phys. Rev. Lett. 81, 192 (1998)<br />

[43] A. Biswas, S. Elizabeth, A. K. Raychaudhuri, H. L. Bhat, Phys. Rev. B 59, 5368<br />

(1999)<br />

[44] Y. Tomioka, A. Asamitsu, H. Kuwahara, Y. Moritomo, and Y. Tokura, Phys.<br />

Rev. B 53, R1689 (1996)


Introduction 22<br />

[45] Y. Tomioka, A. Asamitsu, Y. Moritomo, and Y. Tokura, J. Phys. Soc. Jpn 64,<br />

3626 (1995)<br />

[46] Y. Tomioka and Y. Tokura, Science 288, 462 (2000)


Chapter 2<br />

Experimental Techniques<br />

23


Experimental Techniques 24<br />

2.1 Introduction<br />

Spectroscopic methods are some of the most powerful techniques for studying the<br />

electronic structure of solids. This chapter is intended to describe the details of the<br />

different spectroscopic techniques which have been used in this thesis work.<br />

2.2 Photoemission Spectroscopy<br />

2.2.1 Basic principle<br />

The principle of photoemission spectroscopy is based on the principle of photoelectric<br />

effect. The photoelectric effect was first discovered by the German Physicist Heinrich<br />

Hertz [1] in 1887 and latter explained by Albert Einstein [2] in 1905 as a manifestation<br />

of the quantum nature of light.<br />

Figure 2.1: Principle of a modern photoemission experiment. The photon source is<br />

either UV light or X-ray, which hit the sample surface under an angle ψ with respect<br />

to the surface normal. The kinetic energy E kin of the photoelectrons can be analysed<br />

by use of electrostatic analysers.<br />

When light is incident on a sample, electron can absorb the photon and escape<br />

from the sample with a maximum kinetic energy as shown in the Fig. 2.1. If a photon<br />

having energy hν is absorbed by the electron of binding energy E B , then the kinetic<br />

energy E kin of the electron is given by,<br />

E kin = hν − E B − φ, (2.1)<br />

where φ is the work function of the material. The resulting photoelectron spectra<br />

is a plot of the number of emitted electrons versus their kinetic energy, which<br />

gives the information about the electronic structure of the material. Fig. 2.2 shows


Experimental Techniques 25<br />

Figure 2.2: Schematic view of the energetics of the photoemission process [3], which<br />

gives the relation between the energy levels in a solid and the electron energy distribution<br />

produced by photons of energy hν. The electron energy distribution which<br />

can be measured by the analyser as a function of kinetic energy (E kin ) is more conveniently<br />

expressed in terms of the binding energy E B when one refers to the density<br />

of states inside the solid.<br />

schematically how the energy level diagram and the energy distribution of photoemitted<br />

electrons relate to each other.<br />

2.2.2 Theoretical interpretation of photoemission<br />

There have been many important theoretical studies to describe and analyse the spectral<br />

features in photoemission process [3, 4, 5, 6]. The most general and widely applied<br />

theoretical description of the photoemission spectrum is based on using Fermi’s golden<br />

rule as a result of perturbation theory in first order. The photo current produced in<br />

the photoemission spectroscopy experiment is due to the excitation of electrons from<br />

the initial states i with wave function ψ i to the final states f with wave function ψ f<br />

by the photon field having vector potential A. According to the Fermi’s Golden Rule,


Experimental Techniques 26<br />

the transition probability can be written as,<br />

w ∝ 2π¯h |〈ψ f|r|ψ i 〉| 2 δ(E f − E i − hν) = 2π¯h m fiδ(E f − E i − hν) (2.2)<br />

where m fi is the square of the transition matrix element. To discuss about the<br />

transition matrix element, one has to take certain assumptions of the wave functions<br />

contained in it. In the simplest approximation one can take a one electron view for<br />

the initial and final state wave functions. The final state is in addition with a free<br />

electron having kinetic energy E kin . The initial state wave function is then written as<br />

a product of the orbital φ k from which the electron is excited and the wave function<br />

of the remaining electrons ψi,R(N k −1), assuming that system under consideration has<br />

N electrons, where the index k indicates that the electron k has been left out and R<br />

refers to the remaining. Now one can write,<br />

ψ i (N) = Cφ i,k ψ k i,R(N − 1) (2.3)<br />

where C is the operator that antisymmetrizes the wave function properly.<br />

Similarly the final state can be written as a product of the wave function of the<br />

photoemitted electron φ f,Ekin and that of the remaining N −1 electrons ψf,R k (N −1),<br />

ψ f (N) = Cφ f,Ekin ψf,R k (N − 1) (2.4)<br />

So the transition matrix element can be written as,<br />

〈ψ f |r|ψ i 〉 = 〈φ f,Ekin |r|φ i,k 〉〈ψf,R(N k − 1)|ψi,R(N k − 1)〉 (2.5)<br />

Thus the matrix element is the product of one electon matrix element and the<br />

(N − 1) electron overlap integral. Let us assume that the remaining orbitals (called<br />

the passive orbitals) are the same in the final state as they were in the initial state,<br />

meaning ψf,R(N k − 1) = ψi,R(N k − 1), which renders the overlap integral unity and<br />

the transition matrix element is just the one-electron matrix element. Under this<br />

assumption [7] the PES experiment measures the negative Hartree-Fock orbital energy<br />

of the orbital k meaning,<br />

which is sometimes called Koopman’s binding energy.<br />

E B,k ≃ − ǫ k , (2.6)


Experimental Techniques 27<br />

Let us now assume that the final state with N − 1 electrons has s excited states<br />

with the wave function ψf,s k and energy E s (N − 1).<br />

The transition matrix element can be calculated by summing over all the possible<br />

excited final states yielding,<br />

〈ψ f |r|ψ i 〉 = 〈φ f,Ekin |r|φ i,k 〉 ∑ s<br />

c s (2.7)<br />

with<br />

c s = 〈ψf,s k (N − 1)|ψk i,R (N − 1)〉 (2.8)<br />

Here |c s | 2 is the probability that removal of electron from orbital φ k of the N<br />

electron ground state leaves the system in the excited state s of the N − 1 electron<br />

system. For strongly correlated systems many of the c s ’s are nonzero. In terms of PE<br />

spectrum this means that for s = k one has the so-called main line and for the other<br />

non-zero c s additional so-called satellite lines occur. If the correlations are weak, then<br />

ψ k f,s(N − 1)≃ψ k i,R(N − 1) (2.9)<br />

meaning that c s 2 ≃ 1 for s = k and c s 2 ≃ 0 for s ≠ k, i.e one has only one peak.<br />

Now the photocurrent I detected in a PE experiment as,<br />

I∝ ∑ |〈φ f,Ekin |r|φ i,k 〉| 2∑<br />

f,i,k<br />

s<br />

|c s | 2 δ(E f,kin + E s (N − 1) − E 0 (N) − hν, (2.10)<br />

where E 0 (N) is the ground state energy of the N-electron system. The photo<br />

current thus consists of the lines created by photoionizing the various orbitals k where<br />

each line can be accompanied by satellites according to the number of excited states<br />

s created in the photoexcitation of that particular orbital k.<br />

The above discussion is very convenient for atoms and molecules [4]. However,<br />

it is not always easy to apply for solids. For solids, a slightly different formalism<br />

has been developed [3, 8, 9] and hence the above equation can be rewritten in the<br />

following form<br />

I∝ ∑ f,i,k|〈φ f,Ekin |r|φ i,k 〉| 2 A(k, E) (2.11)<br />

where A(k, E) is the so-called spectral function for the wave number k and energy<br />

E. The spectral function can be related to the single particle Green’s function by


Experimental Techniques 28<br />

A(k, E) = π −1 Im{G(k, E)} (2.12)<br />

Here A(k, E) describes the probability of removing (below the E F ) or adding<br />

(above the E F ) an electron with the energy E and the wave vector k from (to) the<br />

N-electron system (Interacting).<br />

For a non interacting system with the one electron energy Ek,<br />

0<br />

G 0 (K, E) =<br />

where ǫ is very small<br />

and<br />

1<br />

(2.13)<br />

E − Ek 0 − iǫ,<br />

A 0 (k, E) = δ(E − E 0 k) (2.14)<br />

which means the spectral function is a δ-function at E = Ek 0 . This result is<br />

identical to that of equation (2.6) as in a noninteracting system the binding energy<br />

is by definition the Koopman’s binding energy.<br />

In an interacting electron system the electron energy gets renormalized by the<br />

so-called self energy Σ(k, E) = Re{Σ(k, E)} + iIm{Σ(k, E)}, yielding<br />

and<br />

G(K, E) =<br />

1<br />

E − E 0 k − Σ(k, E) (2.15)<br />

A(k, E) = 1 π<br />

Im{Σ(k, E)}<br />

[E − E 0 k − Re{Σ(k, E)}]2 + [Im{Σ(k, E)}] 2 (2.16)<br />

If we assume that the spectrum of the interacting system is not too different<br />

from that of the noninteracting system (Im{Σ(k, E)}), then the poles in the Green’s<br />

function of the interacting one occur at<br />

with the condition<br />

E 1 k = Re{E1 k } + iIm{E1 k } (2.17)<br />

E 1 k − E0 k − Σ(k, E1 k ) = 0 (2.18)<br />

This allows to separate the Greens function into two parts, a pole part and a<br />

non-singular term G inc , i.e,


Experimental Techniques 29<br />

and<br />

G(k, E) =<br />

Z k<br />

E − (Re{E 1 k } + iIm{E1 k }) + G inc (2.19)<br />

A(k, E) = 1 π<br />

Z k Im{E 1 k }<br />

(E − Re{E 1 k })2 + (Im{E 1 k })2 + A inc (2.20)<br />

where Z k is a number smaller than one. The first term in (2.19) is similar to<br />

the result for the noninteracting system, and one can therefore say that it describes<br />

a slightly modified electron (with renormalized mass) which is generally called a<br />

quasiparticle. The first term in (2.19) is called the coherent part of the Green’s<br />

function, it gives rise to the coherent part of the spectrum. And the second term in<br />

(2.19) is called the incoherent part of the Green’s function, which gives rise to the<br />

incoherent part of the spectrum. In a very approximate way one can identify the<br />

coherent part of the spectrum with the main line and the incoherent one with the<br />

satellites in equation (2.20) [3].<br />

2.2.3 The electron escape depth<br />

In photoemission spectroscopy, when electrons are impinged on, or emitted from a<br />

sample surface, the fundamental phenomenon occurs is inelastic scattering of electrons.<br />

The escape depth of electron from the sample surface is depend on the inelastic<br />

mean free path (IMFP) λ, which is the average distance between the scattering<br />

events. λ is small for low energy electrons in solids and hence gives rise to surface<br />

sensitivity. So ultra - high vacuum (UHV) condition is necessary to study the solid<br />

surfaces [3, 10, 11]. Because a monolayer is adsorbed on the surface of the sample<br />

in about 2.5 Langmuirs (1L = 10 −6 torr.s). At a pressure of 10 −9 torr it therefore<br />

takes about 1000 seconds, or less than an hour, until a coverage of one monolayer is<br />

achieved. Thus in order to get spectra of clean surfaces, vacuum of even better than<br />

10 −10 torr is mandatory. The IMFP of an electron in a solid is a strong function of<br />

its kinetic energy [3, 12, 13], and the relationship termed as the ”universal curve” as<br />

shown in Fig. 2.3.<br />

The electron kinetic energy in the range of interest between about 10 and 2000<br />

eV. The escape depth is only on the order of a few Å. From the universal curve, one<br />

can see the IMFP has a minimum value of 5 to 10 Å for the kinetic energy range


Experimental Techniques 30<br />

Figure 2.3: ”Universal curve” for electron inelastic mean free path λ.<br />

of 20 to 200 eV. And hence the surface sensitivity of photoemission spectroscopy is<br />

highest in this energy range.<br />

2.2.4 Experimental set-up<br />

The OMICRON photoemission system at Institute of Physics consists of two connected<br />

UHV chambers, a sample preparation chamber and an analysis chamber. The<br />

preparation chamber is equipped with a precision manipulator, sample heater, evaporator<br />

and a diamond file. The analysis chamber is made up of µ metal to protect low<br />

energy electrons from earth’s magnetic field and stray magnetic fields. It is equipped<br />

with a 125 mm hemispherical analyser [10] for angle integrated photoemission and a<br />

movable 65 mm hemispherical analyser mounted on a 2-axis goniometer is also there<br />

for angle resolved photoemission. Two photon sources [10, 14] (ultra-violet and x-ray)<br />

are used for photoexcitations. The analysis chamber is also equipped with a 4-axis<br />

sample manipulator-cum cryostat which can go down to 20 K. Facility for performing<br />

low energy electron diffraction (LEED) is also available in the analysis chamber. The<br />

external view of the spectrometer is shown in the Fig. 2.4. Some of the important<br />

components of the spectrometer are described below.


Experimental Techniques 31<br />

Figure 2.4: Photoemission spectrometer at Institute of Physics, Bhubaneswar.


Experimental Techniques 32<br />

Figure 2.5: Schematic diagram of the VUV source.<br />

Figure 2.6: Dimensional outline of the VUV source HIS 13. The viewport can be<br />

used for inspection of the discharge and for adjustment of the light spot position on<br />

the sample. The dimensions are in mm.<br />

Vacuum ultra-violet photon source (VUV source HIS 13)<br />

The high intensity VUV Source [14] HIS 13 can be operated with various discharge<br />

gases, such as helium, neon, argon, krypton, xenon or hydrogen. But in our laboratory<br />

we use only helium and neon gases. The operation of the lamp is based on the principle<br />

of a cold cathode capillary discharge [15]. If an increasing potential is applied between<br />

the ends of an insulating tube filled with gas at a pressure of typically 1 mbar, a


Experimental Techniques 33<br />

point will be reached when spontaneous breakthrough occurs leading to a continuous<br />

discharge. The ignition potential is about an order of magnitude higher than the<br />

operating potential necessary to maintain a continuous discharge. The schematic of<br />

the VUV source is shown in the Fig. 2.5. There is a windowless direct-sight connection<br />

between the discharge area and the target. The discharge current is electronically<br />

stabilised. The lamp is water cooled in order to allow for high discharge currents<br />

(up to 300 mA) and to reduce electrode degradation resulting in prolonged service<br />

intervals. The dimensional outline of the VUV source is shown in the Fig. 2.6<br />

X-ray source (DAR 400)<br />

The X-ray source we used is DAR 400 from OMICRON designed specifically for X-ray<br />

photoelectron spectroscopy [10]. It has twin anodes that allow either Mg Kα (1253.6<br />

eV) or Al Kα (1486.6 eV) radiation to be selected. The schematic of the X-ray source<br />

is shown in Fig. 2.7.<br />

Figure 2.7: Schematic of the X-ray source in operation<br />

Electrons are extracted from a heated filament to bombard the selected surface of<br />

an anode at high positive potential. The focus ring and the shape of the nose cone<br />

ensure that the electrons hit the anode in the right area. By switching the filament


Experimental Techniques 34<br />

used, the second face of the anode can be excited. The anode is water cooled to<br />

prevent the aluminium or magnesium surfaces from evaporating.<br />

X-rays generated in the surface of the anode pass through a thin aluminium window<br />

to the sample under analysis. The aluminium window forms a partial vacuum<br />

barrier between the source and the sample region.<br />

The hemispherical electron energy analyser (EA 125)<br />

The EA 125 energy analyser based on a 125 mm mean radius electrostatic hemispherical<br />

deflection analyser composed of two concentric hemispheres [10, 16]. The inner<br />

and outer spheres are biased positive and negative with respect to the pass energy of<br />

the analyser. The analyser disperses electrons according to the energy across the exit<br />

plane (between the two hemispheres) and focuses them in the angular dimension from<br />

the entrance to the exit plane. Slits which are variable are located at the entrance<br />

and exit of the analyser. A schematic of the EA 125 analyser is shown in Fig. 2.8.<br />

Figure 2.8: Schematic diagram of the EA 125 hemispherical analyser.<br />

The principle of operation of concentric hemispherical analyser is shown schematically<br />

in Fig. 2.9. It consists of two concentric hemispheres of radii R 1 (inner hemisphere)<br />

and R 2 (outer hemisphere) with mean radius R 0 mounted with a common<br />

centre O. A potential V is applied between the two hemispheres so that the outer is


Experimental Techniques 35<br />

negative and the inner positive with respect to V 0 which is the median equipotential<br />

surface between the hemispheres.<br />

Figure 2.9: The principle of operation of the concentric hemispherical analyser,<br />

schematic diagram.<br />

If E p is the kinetic energy of an electron travelling in an orbit of radius R 0 , then<br />

the relationship between E p and V 0 is given by the expression,<br />

eV 0 = E p ·<br />

(<br />

R2<br />

− R )<br />

1<br />

R 1 R 2<br />

(2.21)<br />

The voltages on the inner and outer hemispheres are V 1 and V 2 respectively and<br />

these are given by,<br />

V 1 = E p ·<br />

V 2 = E p ·<br />

(a) Modes of analyser operation<br />

(<br />

2 R )<br />

0<br />

− 1<br />

R 1<br />

(<br />

2 R )<br />

0<br />

− 1<br />

R 2<br />

(2.22)<br />

(2.23)<br />

The kinetic energy can be scanned either by varying the retardation ratio whilst<br />

holding the analyser pass energy constant known as constant analyzer energy (CAE)


Experimental Techniques 36<br />

mode or by varying the the pass energy E p whilst holding the retardation ratio constant<br />

known as the constant retard ratio (CRR) mode [10].<br />

i. Constant retard ratio (CRR) mode<br />

In the CRR mode electrons entering the analyser system from the sample are<br />

retarded by the lens stack by a constant proportion of their kinetic energy so that<br />

the ratio of electron kinetic energy to analyser pass energy is kept constant during a<br />

spectrum. The retard ratio k is defined as,<br />

K = E k − φ a<br />

E p<br />

∼ E k<br />

E p<br />

(2.24)<br />

where φ a is the analyser work function. Throughout the scan range the pass energy<br />

of the analyser is continuously varied to maintain a constant retard ratio. Sensitivity<br />

and resolution are proportional to the pass energy and therefore the kinetic energy<br />

in this mode. In this mode the analysed sample area and the emission angle remain<br />

almost constant throughout the whole kinetic energy range.<br />

ii. Constant analyser energy (CAE) mode<br />

In the CAE mode, the analyser pass energy is held constant, and the retarding<br />

voltage is changed thus scanning the kinetic energy of detected electrons. The resolution<br />

obtained in CAE is constant throughout the whole kinetic energy range. The<br />

sensitivity, however is inversely proportional to the kinetic energy and at low kinetic<br />

energies is improved over that of CRR. In the CAE mode, the analysed sample area<br />

and the emission angle may vary slightly with the kinetic energy, but this is dependent<br />

on the lens design.<br />

(b) Analyser resolution<br />

The analyser is a band pass energy filter for electrons at a specific energy E p<br />

and has a finite energy resolution ∆E which is dependent on the chosen mode of<br />

operation and specific operating conditions. The energy resolution of the analyser is<br />

given approximately by,


Experimental Techniques 37<br />

∆E = E p<br />

( d<br />

2R 0<br />

+ α 2 )<br />

(2.25)<br />

where d is the slit width, R 0 is the mean radius of the hemispheres and α is the<br />

half angle of electrons entering the analyser at the entrance slit.<br />

(c) The electrostatic input lens<br />

The input lens collects the electrons from the source and focuses them onto the<br />

entrance aperture of the analyser whilst simultaneously adjusting their kinetic energy<br />

to match the pass energy of the analyser. The lens is also designed to define the analysis<br />

area and angular acceptance of electrons which pass through the hemispherical<br />

analyser. The lens design employs double lens concept where two lenses are stacked<br />

one above the other.<br />

The first lens selects the analysis area (spot size) and angular acceptance. This is<br />

an Einzel lens, i.e it does not change the energy of the electrons and therefore has a<br />

constant magnification throughout the entire energy range. This lens can be operated<br />

in three discrete magnification modes: high, medium and low. In high magnification<br />

mode, the focal plane is nearer to the sample and the lens accept a wide angle of<br />

electron beams from a small region. In low magnification mode, the focal plane is<br />

farther from the sample and the lens accept only a small angle of beams but from a<br />

larger area. The medium magnification is in-between the two.<br />

The second lens retards or accelerates the electrons to match the pass energy of<br />

the analyser and uses the zoom lens function to ensure that the focal point remains<br />

on the analyser entrance aperture. The magnification of this lens varies with retard<br />

ratio as a result of the law of Helmholtz-Lagrange.<br />

The analysis area is defined by the combination of the selected analyser entrance<br />

aperture and the magnification of the entire lens. The magnification of the entire lens<br />

is a product of the magnifications of the two discrete lenses.<br />

(d) The detector<br />

A single channel electron multiplier (Channeltron) is placed across the exit plane<br />

of the analyser. The channeltron amplifies the current of a single electron by a factor<br />

of about 10 8 . The small current pulse present at the output of the channeltron is


Experimental Techniques 38<br />

passed through a vacuum feedthrough and then directly into the preamplifier. And<br />

then the signal is passed onto a pulse counter for processing and production of an<br />

electron energy spectrum.<br />

2.3 Inverse Photoemission Spectroscopy<br />

2.3.1 Historical background<br />

The inverse photoemission spectroscopy (IPES) [17, 18, 19] is a powerful technique<br />

to study the unoccupied electronic states of solids. The IPES experiment can be<br />

carried out in two different photon energy regimes. One is at higher energy range<br />

(detected photons are in x-ray range) called Bremsstrahlung isochromat spectroscopy<br />

(BIS) and the other one is at low energy region (detected photons are in the ultraviolet<br />

range) which is called UV-IPES or simply IPES. The history of IPES technique<br />

was developed in 1915 by Duane and Hunt [20] when an inverse relationship is observed<br />

between the short wave-length cutoff of emitted x-rays and incident electron<br />

energy when a solid is bombarded with electrons. Then pursued by Ohlin [21] and<br />

Nijboer [22] in 1940s and was resumed in the 1950s by Ulmer and Vernickel [23]. The<br />

IPES work in the ultra-violet region is of relatively recent origin. The experimental<br />

work was started by Dose [24] and the theoretical work by Pendry [25, 26]. The<br />

experimental advances since then have been rapid.<br />

2.3.2 Basic principle<br />

When a beam of electrons is incident on a solid surface, after entering the system,<br />

the electrons decay either radiatively or non-radiatively to states at lower energy. If<br />

the decay is radiative the emitted photon can be detected, and gives the information<br />

about the unoccupied states of the system. So this is complementary to direct photoemission<br />

spectroscopy (PES) [17, 18, 19, 27]. A schematic of the inverse photoemission<br />

spectroscopy is shown in the Fig. 2.10<br />

The energy of the photons (hν) emitted when a beam of electrons having kinetic<br />

energy E k is incident on a solid surface and then relax to a lower energy unoccupied<br />

state (E f ), is given by the conservation of energy as,<br />

E k = hν + E f − φ, (2.26)


Experimental Techniques 39<br />

Figure 2.10: Schematic diagram of inverse photoemission spectroscopy (IPES).<br />

where φ is the work function of the material.<br />

There are two modes [18, 19, 27] used for this measurement . One is isochromat<br />

mode and the other one is spectrograph mode. In isochromat mode, the incident<br />

electron energy is ramped and the emitted photons are detected at a fixed energy.<br />

In the spectrograph mode, the energy of the incident electron remains fixed and the<br />

emitted photons are detected over a range of photon energies.<br />

2.3.3 Basic interpretation of inverse photoemission<br />

The Inverse photoemission process can be described [28] by a comparison with photoemission.<br />

Both photoemission and inverse photoemission involve the interaction<br />

between photons and electrons. Now the interaction Hamiltonian associated with the<br />

electromagnetic field can be written as,


Experimental Techniques 40<br />

H ′ =<br />

e (A.p + p.A) (2.27)<br />

2mc<br />

where A is the vector potential of the electromagnetic radiation. For photoemission<br />

the field A can be treated as a classical time-dependent perturbation, and<br />

results will be the same as a proper quantum treatment had been made. For inverse<br />

photoemission, however, this is no longer true. Since photons are created it becomes<br />

necessary to quantize the electromagnetic field. In such a treatment the classical<br />

vector potential is replaced by the field operator A(x, t) defined by [29]<br />

A(x, t) =<br />

1 ∑ ∑ 2π¯hc<br />

(V p ) 1/2 q α ω [a q,α(t)ˆε (α) e iq·x + a † q,α(t)ˆε (α) e −iq·x ], (2.28)<br />

where ˆε α is the linear polarization vector, a real unit vector whose direction depends<br />

on the photon propagation direction q. The two operators a † q,α and a q,α either<br />

create or destroy a photon in the state q, α respectively. V p is the normalization<br />

volume for the photon. For inverse photoemission the initial state consists of an electron<br />

in a continuum state |i〉≡ψ i (r) with no photons and the final state consists of<br />

an electron in a bound state |f〉≡ψ f (r) and a photon with wave vector q. Then the<br />

transition rate is given by the above two equations using the first order perturbation<br />

theory,<br />

R =<br />

2π¯h<br />

c22π¯h<br />

ω<br />

where ρ p is the photon density of states<br />

and dΩ is the solid angle of emission. Then<br />

1<br />

|〈f|ˆε.p|i〉| 2 ρ p (2.29)<br />

V p<br />

V p<br />

ω 2<br />

ρ p =<br />

(2π) 3 ¯hc 3dΩ<br />

(2.30)<br />

R = 1 e 2 ω<br />

2π ¯hc m 2 c 2 |〈f|ˆε.p|i〉|2 dΩ, (2.31)<br />

where e2<br />

¯hc = α≈ 1<br />

137<br />

To obtain a cross section, we devide by the incident electron flux<br />

j e = ¯hk 1<br />

(2.32)<br />

m V e<br />

We further assume that the continuum electron state is normalized to a box of<br />

volume V e . Then


Experimental Techniques 41<br />

( ) dσ<br />

= α ω 1<br />

dΩ 2π mc<br />

IPES 2 ¯hk |〈f|ˆε.p|i〉|2 (2.33)<br />

It is interesting to compare this cross section with that of the photoemission.<br />

In photoemission the final states are those of electrons, The density of states factor<br />

becomes<br />

ρ e = V e mk<br />

8π 3 ¯h 2 (2.34)<br />

and the incident flux is the photon flux<br />

Then<br />

j p =<br />

ω 1<br />

(2.35)<br />

8π¯hc V p<br />

( ) dσ<br />

= α k 1<br />

dΩ 2π m ¯hω |〈i|ˆε.p|f〉|2 (2.36)<br />

PES<br />

To relate the cross sections of the two processes (PES and IPES), let us suppose<br />

the matrix element in PES is same as that of the IPES. Then the ratio of two cross<br />

sections reflects only the different phase space available for photon creation or electron<br />

creation, and is given by<br />

or<br />

r =<br />

( ) dσ<br />

/<br />

dΩ<br />

IPES<br />

( ) dσ<br />

= ω2<br />

dΩ c<br />

PES<br />

2 k = q2<br />

(2.37)<br />

2 k 2<br />

r =<br />

(<br />

λe<br />

λ p<br />

) 2<br />

(2.38)<br />

where λ e and λ p are the electron and photon wavelengths respectively.<br />

And the equation (2.38) says the ratio of the inverse photoemission cross section to<br />

that of the phtoemission is the ratio of the square of the wave length of electron to that<br />

of photon. At an energy of 10 eV, characteristics of the vacuum ultraviolet region, the<br />

ratio r∼10 −5 . Therefore the signal levels in IPES is much lower comparable to PES.<br />

At higher energies ∼1000 eV, characteristics of the x-ray range, we have r∼10 −3 .<br />

2.3.4 Experimental Set-up<br />

The external view of IPES spectrometer at UGC-DAE consortium for scientific research,<br />

Indore, is shown in the Fig. 2.11.


Experimental Techniques 42<br />

Figure 2.11: (a) IPES spectrometer at UGC-DAE consortium for scientific research,<br />

Indore, (b) External view of photon detector attached to the IPES chamber, and (c)<br />

Inside view of the chamber showing a rotatable (left) and fixed (right) electron gun<br />

and the sample holder


Experimental Techniques 43<br />

The experimental UHV chamber is made of µ metal which provides effective shielding<br />

for the low energy electrons from earth’s magnetic field and stray magnetic fields.<br />

A photon detector capable of detecting fixed energy photons with well defined narrow<br />

bandpass and an electron gun producing nearly monoenergetic low energy electrons<br />

with well defined angular spread are equipped in the UHV chamber. A sample manipulator<br />

is mounted vertically which allow to move the sample both along axial and<br />

azimuthal direction.<br />

The low energy electron gun (e-gun)<br />

The e-gun used for the experiment is of Stoffel Johnson type [30] is shown schematically<br />

in Fig. 2.12. This is chosen because it has less space-charge effect and high<br />

beam current.<br />

Figure 2.12: A schematic diagram of the electron gun [30].<br />

In this type of e-gun the minimum beam size obtained is 1.1 mm with full angular<br />

convergence of 5 ◦ to 7 ◦ . It consists of a filament in a mounting block followed by<br />

three element refocusing lens. All the lens elements are with cylindrical cross-section.<br />

In the design (Fig.) L = D = 16mm, P = 1.3D and Q = 2.5D. Ratio of voltages<br />

applied on the cathode and the subsequent electrostatic lens elements is V C : V A :<br />

V F : V O = −1 : +5 : −0.9 : 0. The Kinetic energy of emitting electrons is governed by<br />

V C . The whole surface of the lenses is gold plated to obtain a uniform work function.<br />

BaO dispenser cathode is used as the electron emitter. The e-gun operates in the<br />

range of 5 − 40 eV.


Experimental Techniques 44<br />

The photon detector<br />

Gas filled band-pass Geiger-Muller (GM) type counters are used for photon detection<br />

in IPES because of their high efficiency, low cost and simple design [31, 32, 33, 34,<br />

35, 36, 37, 38]. In our IPES experiment the acetone/CaF 2 detector [39] was used.<br />

The design of the photon detector is based on the following operating principle<br />

: the high energy cut-off of the acetone/CaF 2 detector is due to the CaF 2 window<br />

that does not transmit photons with energy > 10.2, while the threshold for the<br />

photoionization of acetone at 9.7 eV sets the low energy cut-off [40]. This determines<br />

the band pass function and results [33] in a mean photon detection energy of 9.9<br />

eV with a FWHM of 0.4 eV. Thus, 9.9±0.2 eV photons can enter the detector to<br />

photoionize acetone.<br />

Figure 2.13: Schematic cross section of the photon detector showing the window (1),<br />

O-ring (2), cylindrical cap (3), detector tube (4), teflon spacer (5), anode (6), double<br />

sided DN40 CF flanges (7, 10), gas inlet and outlet (8, 9), barrel connector (11),<br />

teflon sleeve (12), DN40 CF flange (13) and SHV connector (14).<br />

The detector assembly [41] (2.13) has a modular design consisting of a 250 mm<br />

long stainless steel cylindrical tube with inner diameter 17.8 mm and outer diameter<br />

21.2 mm (Fig), welded to a double-sided DN40 CF flange (7) with a through hole equal<br />

to the inner diameter of the tube (4). Another double-sided DN40 CF flange (10)<br />

with a similar through hole has two additional 6 mm diameter through holes drilled<br />

in the radial direction. Two 6 mm diameter of stainless steel tubes of length 25 mm


Experimental Techniques 45<br />

are welded to these radial holes, and on the other end, DN16 CF flanges (8, 9) are<br />

welded to these tubes. These are used as gas inlet (8) and outlet (9). The anode (6)<br />

is attached to a SHV feedthrough (14) mounted on another DB40 CF flange (13) by<br />

a barrel connector (11). The three DN40 CF flanges (7, 10, 13) are mounted together<br />

on a linear translator. The anode is a 1.6 mm diameter electropolished stainless steel<br />

wire and is held by two perforated teflon spacers (5) in the detector tube. High voltage<br />

is applied to the anode. A 25 mm diameter, 2 mm thick CaF 2 window (1) sits on a<br />

viton O-ring (2) and is tightly pressed by a stainless steel cap (3) at the entrance of<br />

the detector. This isolates the detector from the UHV chamber. Use of viton O-ring<br />

ensures easy exchange of windows at the testing stage of the detector. The count rate<br />

increases with higher solid angle of detection. So the detector is brought as close as<br />

possible to the sample using the linear translator.<br />

2.4 X-ray Absorption Spectroscopy<br />

X-ray absorption spectroscopy (XAS) is another technique to study the unoccupied<br />

electronic states as well as the geometrical arrangements of any materials around an<br />

atom [42, 43]. But in this thesis XAS technique is used only to study the electronic<br />

structure of the materials.<br />

2.4.1 Basic principle<br />

When a beam of monochromatic x-ray enters a solid, either it will be scattered by<br />

the electrons or absorbed and excite the core electrons. In the absorption process the<br />

core electron is excited to an empty state [21, 45]. The schematic of XAS is shown in<br />

Fig. 2.14.<br />

If the energy of the x-ray is much larger than the binding energy of the core state<br />

the excited electron behaves like a free electron in the solid. If however the x-ray<br />

energy is just enough to excite a core electron, it will occupy a lowest available empty<br />

state.<br />

When a monochromatic x-ray beam of intensity I 0 goes through a material of<br />

thickness d, its intensity I will get reduced as,<br />

I = I 0 e −ρµd (2.39)


Experimental Techniques 46<br />

Figure 2.14: Schematic diagram of x-ray absorption spectroscopy<br />

where µ is the absorption coefficient of the material, which depends on the types<br />

of atoms and density ρ of the material. The absorption coefficient decreases smoothly<br />

with higher energy, except for certain photon energies where the absorption increases<br />

drastically, and gives rise to an absorption edge. Each such edge occurs when the<br />

energy of the incident photons is just sufficient to excite a core electron to an unoccupied<br />

state or into the continuum state (leave the atom). After each absorption edge<br />

the absorption coefficient continues to decrease with increasing photon energy (Fig.<br />

2.15).<br />

There are two distinguishable parts in x-ray absorption process : The low energy,<br />

near edge x-ray absorption fine structure (NEXAFS) [46, 47] or x-ray absorption<br />

near edge structure (XANES) region and the extended x-ray absorption fine structure<br />

(EXAFS) [48, 49], where the ejected electrons have high kinetic energy with single<br />

scattering. In this thesis work, only the XANES (or NEXAFS) region has been taken<br />

to study the unoccupied states.


Experimental Techniques 47<br />

Figure 2.15: The x-ray loses its intensity via interactions with material. The absorption<br />

coefficient decreases smoothly with higher energy, except for special photon<br />

energies. When the photon energy reaches a critical value for a core electron transition,<br />

the absorption coefficient increases abruptly.<br />

2.4.2 Detection techniques<br />

The normal way to measure the x-ray absorption is in the transmission mode [21]. The<br />

intensity of the x-ray is measured before and after the sample and the percentage of<br />

transmitted x-rays is determined. Such transmission mode of experiment is standard<br />

for hard x-rays. But for soft x-rays they are difficult to perform because of the strong<br />

interaction of soft x-rays with matter and requires a very thin sample of 0.1 µm to<br />

obtain a detectable signal. An alternative to the tansmission mode experiment is the<br />

”Yield mode”, where the decay product of the core hole (created in x-ray absorption)<br />

will be measured. The decay of the core hole gives rise to an avalanche of electrons,<br />

photons and ions escaping from the sample surface. There are four types of detection<br />

methods [21] in yield mode XAS measurement such as Auger electron yield (or partial<br />

yield), fluorescence yield, ion yield and total electron yield.<br />

In Auger electron yield method, the electrons of a certain energy range (i.e electrons<br />

of a specific Auger decay channel of the core hole) are detected. It is found<br />

that the mean free path of a 500 eV Auger electron is of the order of 20 Å. Hence


Experimental Techniques 48<br />

the number of Auger electrons emitted is equal to the number of core holes which<br />

are created in the first 20 Å from the surface. The Auger electron yield method is a<br />

measurement of only 20 Å of material, hence the Auger electron yield is rather surface<br />

sensitive. In fluorescence yield XAS mode, the fluorescence decay of the core hole is<br />

measured instead of Auger decay electrons. The intensity of the fluorescence yield<br />

is small compared to Auger electron yield mode for the elements of atomic number<br />

Z ≤ 50. The main advantage of fluorescence yield is that the created photon has a<br />

mean free path of the same order of magnitude as the x-ray (∼ 1000 Å) and hence it<br />

probes the bulk. In ion yield mode, only the surface ions created by Auger decay are<br />

measured. So it is highly surface sensitive and is used for special surface study. The<br />

most applied detection technique for soft XAS is the total electron yield. In total<br />

electron yield mode, all escaping electrons are counted regardless of their energy. The<br />

count rate is very large in this case. This thesis contains the XAS spectra measured<br />

in total electron yield and fluorescence yield mode. In TEY mode, the electrons are<br />

detected by the hemispherical analyser, which is discussed in the last section. In the<br />

FY mode the radiative photons are detected by the Si-photodiode.<br />

Si - photodiode<br />

Silicon photodiode detectors are widely used over a broad wavelength range extending<br />

from the x-ray to the visible regions. The detector used for the XAS measurement<br />

involved in this thesis are AXUV and SXUV photodiodes [50, 51]. The photodiodes<br />

are fabricated by an ULSI (Ultra large Scale Integrated circuit) compatible process.<br />

Fig. 2.16 shows the cross section structure of the AXUV and SXUV photodiodes.<br />

The AXUV photodiode has a thin nitrided SiO 2 surface layer (approximately 7 nm<br />

thick) and 100 % internal quantum efficiency. The SiO 2 layer is known to have a<br />

fixed positive charge which repels minority carrier holes generated at the surface by<br />

strongly absorbing extreme ultra-violet (EUV) radiation. The SXUV photodiode has<br />

metal silicide surface layer which replaces the oxide layer on the AXUV photodiode<br />

and improves the radiation hardness. However, the absence of the oxide layer causes<br />

significant surface recombination in the SXUV photo diodes and reduces the internal<br />

quantum efficiency below 100 % value, that is characteristics of AXUV photodiodes.<br />

When the silicon photodiode is exposed to photons with energy greater than 1.12<br />

eV (wavelength less than 1100 nm), electron-hole pairs (carriers) are created. These<br />

photogenerated carriers are separated by the p-n junction electric field and a current


Experimental Techniques 49<br />

Figure 2.16: Schematic of the Si-photodiode.<br />

proportional to the number of electron hole pairs created flows through an external<br />

circuit. Several unique properties of the AXUV photodiodes have resulted in previously<br />

unattained stability and 100 % carrier collection efficiencies giving rise to near<br />

theoretical quantum efficiencies. The first property is the absence of a surface dead<br />

region i.e no photogenerated carrier recombination occurs in the doped n-type region<br />

and at the silicon-silicon dioxide interface. As the absorption depths for the majority<br />

of UV/EUV photons are less than 1 micrometer in silicon, the absence of a dead region<br />

yields complete collection of photogenerated carriers by an external circuit resulting<br />

into 100 % collection efficiency. The second unique property of the AXUV diodes is<br />

their extremely thin (4 to 8 nm), radiation-hard silicon dioxide junction passivating,<br />

protective entrance window. Owing to their 100 % collection efficiency and the thin<br />

entrance window, the quantum efficiency of the AXUV diodes can be approximately<br />

predicted in most of the XUV region by the theoretical expression E ph /3.65, where<br />

E ph is the photon energy in electron volts.<br />

2.5 Synchrotron Radiation<br />

The unique properties of synchrotron radiation are its continuous spectrum, high<br />

flux and brightness, and high coherence, which make it an indispensable tool in the


Experimental Techniques 50<br />

exploration of matter. Synchrotron radiation are produced when charge particles<br />

Figure 2.17: Schematic of the synchrotron radiation.<br />

moving with relativistic velocity are deflected in a magnetic field. The synchrotron<br />

radiation had been developed by taking the effect of Heinrich Hetz [1] as the electrons<br />

fly around the curves through the magnetic field emit electromagnetic radiation. In<br />

SR facilities the accelerated electrons fly in a closed orbit inside a stainless steel<br />

tube, which consists of large electromagnets at regular intervals to guide the electrons<br />

around the curves and keep them focussed in the center of the tube called the storage<br />

ring. The electrons are generated in an injection system consisting of an electron<br />

source and an accelerator, either a synchrotron or a linear accelerator. When the<br />

electrons are accelerated to the operating energy of the storage ring, they are injected<br />

into the storage ring. The bending magnets are used to deflect the electron beam.<br />

In order to compensate the energy loss of the electrons during the SR emission, the<br />

electrons are accelerated each time as they pass through the RF cavity installed inside<br />

the storage ring. A schematic of SR light source is shown in the Fig. 2.17.<br />

In the developed synchrotrons the intensity of the emitted radiation is enhanced<br />

by insertion devices (undulator or wiggler). The insertion devices are periodic arrays<br />

of permanent magnets. When electrons traverse in the periodic array of magnets are<br />

forced to undergo oscillations and radiate (Fig. 2.18 shows the schematic representation).The<br />

static magnetic field is alternating along the length of the insertion devices<br />

with a wave length λ 0 . The radiation produced by insertion devices is very intense and<br />

is guided through beam lines for various experiments. An important dimensionless<br />

parameter K which characterizes the nature of electron motion is defined as,


Experimental Techniques 51<br />

Figure 2.18: Schematic representation of an insertion device.<br />

K = eB 0λ 0<br />

(2.40)<br />

2πmc<br />

where e is the electronic charge, B 0 is the magnetic field, m is the electron rest<br />

mass and c is the speed of light.<br />

For undulator K > 1, the oscillation amplitude is bigger and the radiation contributions<br />

from each field period sum up independently leading to a broad energy spectrum.<br />

Insertion devices can provide several orders of magnitude higher flux than a simple<br />

bending magnet and are in high demand at synchrotron radiation facilities. For an<br />

undulator with N periods, the brightness can be up to N 2 more than a bending<br />

magnet.<br />

The experiments presented in this thesis were performed at different synchrotrons<br />

using monochromators of various type. The HRPES and XAS experiments were<br />

performed at the WERA soft x-ray beamline of ANKA synchrotron, Germany. At<br />

BEAR and BACH beamlines of ELETTRA synchrotron, Italy, the XAS experiments<br />

were performed.<br />

At WERA soft x-ray beamline of ANKA synchrotron, a spherical grating monochromator<br />

(DRAGON type) [21, 52] was used. The layout of the WERA soft x-ray beamline<br />

is shown in the Fig. 2.19. It consists of a pair of mirrors 1 (horizontal) and


Experimental Techniques 52<br />

Figure 2.19: Optical layout of the WERA soft x-ray beam line, ANKA.<br />

2 (vertical) to focus the incoming beam into the entrance slit of the grating box<br />

and also a pair of bendable focusing mirrors after the exit slit for optimum focus at<br />

sample position. Both entrance and exit slits are movable. The energy resolution<br />

∆E/E < 10 −4 . The photon energy range is 100 - 1500 eV. The experimental facilities<br />

associated with the WERA soft x-ray beamline are PES, PEEM, NEXAFS and<br />

SXMCD.<br />

The BEAR beamline at ELETTRA is dedicated to the study of optical, electronic<br />

and magnetic properties of materials under UHV condition. The lay out of the BEAR<br />

beam line is shown in the Fig. 2.20. The beamline optics is based on a couple of<br />

parabolic mirrors. The radiation is collimated by the first parabolic mirror (P1). The<br />

parallel beam produced by P1 reaches the monochromator stage based on the plane<br />

grating/plane mirror scheme according with the Naletto-Tondelo scheme [53]. There<br />

are three gratings (G-NIM, G-1200 and G1800) associated with the monochromator.<br />

Three gratings cover 5 - 1600 eV photon energy range. The G-NIM (1200 1/mm)<br />

works at near normal incidence to cover the 5 - 45 eV region. The G1200 (1200 1/mm)<br />

and G1800 (1800 1/mm) work in the 45 - 1600 eV energy range. The monochromatic<br />

radiation is focused onto the exit slit by a second parabolic mirror (P2). The elliptical<br />

re-focusing mirror (REFO) brings the beam at the sample position. The experimental


Experimental Techniques 53<br />

Figure 2.20: Optical layout of the BEAR beam line.<br />

facilities available at BEAR beamline are ARPES, XAS, photoelectron diffraction and<br />

specular and diffuse reflectivity.<br />

The BACH beam line is designed to deliver a photon beam with high intensity<br />

and brilliance in the soft x-rays energy range 35 -1600 eV with full control of the<br />

polarization of the light. The goal of this beam line is to perform the electron and<br />

photon spectroscopy experiment with polarization dependence. The optical layout of<br />

the BACH beam line [54, 55] is shown in the Fig. 2.21.<br />

Figure 2.21: Optical layout of the BACH beam line.


Experimental Techniques 54<br />

Figure 2.22: The variable included angle monochromator : the combined movement<br />

of PM1 (plane mirror) and one of the four spherical gratings SG1(-4) leads to the<br />

monochromatization of the light keeping both the entrance and exit slit fixed.<br />

The radiation source is based on two APPLE-II elliptical undulators that are<br />

used alternatively to optimize the flux. The photon dispersion system is based on the<br />

Padmore ”variable angle spherical grating monochromator (VASGM)” scheme and it<br />

includes four different interchangeable spherical gratings (Fig. 2.22). The first three<br />

gratings works in the energy ranges 35 - 200 eV, 200 - 500 eV and 500 - 1600 eV<br />

respectively. The fourth grating operates in the 400 - 1600 eV range providing higher<br />

flux allowing fluorescence and x-ray scattering experiments. Two refocusing sections,<br />

based on plane elliptical mirrors in Kirkpatrick - Baez configuration provide nearly<br />

aberration free spots on the samples. A third branch line after the monochromator will<br />

host a microscope based on Fresnel zone plate focusing. The following experimental<br />

facilities are associated with the BACH beam line ; XPS, UPS, XAS, XMCD, XMLD,<br />

SXES and RIXS.


Bibliography<br />

[1] H. Hertz, Ann. Physik 31, 983 (1887).<br />

[2] A. Einstein, Ann. Physik 17, 132 (1905).<br />

[3] Photoelectron Spectroscopy : Principles and Applications by S. Hüfner (Second<br />

Edition, Springer-Verlag Berlin 1995) and references therein.<br />

[4] C. S. Fadley, In Electron Spectroscopy : Theory, Techniques and Applications<br />

II, edited by C. R. Brundle, and A. D. Baker (Academic, New York 1978).<br />

[5] G. D. Mahan, Phys. Rev. B 2, 4334 (1970).<br />

[6] P. J. Feibelman and D. E. Eastman, Phys. Rev. B 10, 4932 (1974).<br />

[7] T. Koopmans, Physica 1, 104 (1933).<br />

[8] W. Bardyszewski, L. Hedin, Phys. Scripta 32, 439 (1985).<br />

[9] L. Hedin, S. Lundquist, Solid State Physics 23, 1 (Academic, New york 1969).<br />

[10] Practical Surface Analysis Volume I : Auger and X-ray Photoelectron Spectroscopy,<br />

edited by D. Briggs and M. P. Seah (Second Edition, John Wiley and<br />

Sons Ltd. 1990).<br />

[11] Photoemission in Solids I : General Principles by M. Cardona and L. Ley<br />

(Springer-Verlag Berlin 1978.<br />

[12] David R. Penn, Phys. Rev. B 13, 5248 (1976).<br />

[13] H. Gant and W. Mönch, Surface Sci. 105, 217 (1981) and references therein.<br />

[14] Techniques of Vacuum Ultra Violet Spectroscopy by J. A. R. Samson, (John<br />

Wiley and Sons, Inc 1967).<br />

55


Experimental Techniques 56<br />

[15] G. Schönhense and U. Heinzmann, J. Phys. E.: Sci. Instrum. 16, 74 (1983).<br />

[16] Bulding Scientific Apparatus by J. H. Moore, C. Davis, and M. A. Coplan,<br />

(Addison-Wesley Publishing Compan, Inc 1989).<br />

[17] V. Dose, Prog. Surf. Sci. 13, 225 (1983).<br />

[18] N. V. Smith, Rep. Prog. Phys. 51, 1227 (1998).<br />

[19] Topics in Applied Physics : Unoccupied Electronic States, edited by J. C. Fuggle<br />

and J. E. Inglesfield (Springer-Verlag Berlin, 1992).<br />

[20] W. Duane and F. L. Hunt, Phys. Rev. 6, 166 (1915).<br />

[21] P. Ohlin, Ark. Mat. Astron. Fys. A 29, 1 (1942).<br />

[22] B. R. A. Nijboer,Physica 12, 461 (1946).<br />

[23] K. Ulmer and H. Vernickel Z. Phys. 153, 149 (1958).<br />

[24] V. Dose, Appl. Phys. 14, 117 (1977).<br />

[25] J. B. Pendry, Phys. Rev. Lett. 45, 1356 (1980).<br />

[26] J. B. Pendry, J. Phys. C : Solid State Phys. 14, 1381 (1981).<br />

[27] P. D. Johnson and S. L. Hulbert, Rev. Sci. Instrum. 61, 2277 (1990).<br />

[28] P. D. Johnson and J. W. Davenport, Phys. Rev. B 31, 7521 (1985).<br />

[29] Advanced Quantum Mechanics by J. Sakurai, (Addiso-Wesley, New york, (1967).<br />

[30] N. G. Stoffel and P. D. Johnson, Nucl. Instrum. and Meth. in Phys. Res. A 234,<br />

230 (1985).<br />

[31] G. Debbinger, V. Dose, and H. Scheidt, Appl. Phys. A 18, 375 (1979).<br />

[32] P. M. G. Allen, P. J, Dobson, and R. G. Edgel, Solid State Comm. 55, 701<br />

(1985).<br />

[33] D. Funnemann and H. Merz, J. Phys. E : Sci. Instrum. 19, 554 (1985).<br />

[34] V. Dose, Th. Fauster, and R. Schneider, Appl. Phys. A 40, 203 (1986) ; A.<br />

Goldmann, M. Donath, W. Altmann, and V. Dose, Phys. Rev. B 32, 837 (1985).


Experimental Techniques 57<br />

[35] K. C. Prince, Rev. Sci. Instrum. 59, 741 (1988).<br />

[36] I. G. Hill and A. B. McLean, Rev. Sci. Instrum. 69, 261 (1998).<br />

[37] J. A. Lifton-Duffin, A. G. Mark, and A. B. MacLean, Rev. Sci. Instrum. 73,<br />

3149 (2002).<br />

[38] J. A. Lifton-Duffin, A. G. Mark, G. K. Mullins, G. E. Contant, and A. B.<br />

MacLean, Rev. Sci. Instrum. 75, 445 (2004).<br />

[39] S. Banik, A. K. Shukla, and S. R. Barman, Rev. Sci. Instrum. 76, 066102 (2005).<br />

[40] W. M. Trott, N. C. Blais, and E. A. Walters, J. Phys. Chem. 69, 3150 (1978).<br />

[41] A. K. Shukla, S. Banik, and S. R. Barman, Current Science 90, 490 (1978).<br />

[42] C. T. Chen, Y. Ma, and F. Sette, Phys. Rev. A 40, 6737 (1989).<br />

[43] X-ray Absorption Fine structure VI, (Proc. of the 6th Inter. XAFS Conf., York,<br />

1990), Ed. S. S. Hasnain, (Ellis Horwood, Chichester, 1991) ;and references<br />

therein.<br />

[44] F. M. F. de Groot, Ph. D. Thesis : X-Ray Absorption of Transition Metal Oxides,<br />

Katholike Universiteit Nijmegan, Netherlans (1991)<br />

[45] F. A. Gianturco, C. Guidotti, and U. Lamanna, J. Chem. Phys. 57, 840 (1972).<br />

[46] D. Norman, J. Stöhr, R. Jaeger, P. J. Durham, and J. B. Pendry, Phys. Rev.<br />

Lett. 51, 2052 (1983).<br />

[47] D. D. Vvedensky, J. B. Pendry, U. Döbler, and K. Baberschke, Phys. Rev. B 35,<br />

7756 (1987).<br />

[48] C. A. Ashley and S. Doniach, Phys. Rev. B 11, 1279 (1975).<br />

[49] P. A. Lee and J. B. Pendry, Phys. Rev. B 11, 2795 (1975).<br />

[50] Benjawan Kjornrattanawanich, Raj Korde, Craig N. Boyer, Glenn E. Holland,<br />

and John F. Seely, IEEE Trans. Elect. Dev. 53, 218 (2006) and references there<br />

in.<br />

[51] R. Korde, C. Prince, D. Cunningham, R. E. Vest, and E. Gullikson, Metrologia<br />

40, S145 (2003) and references therein.


Experimental Techniques 58<br />

[52] S. Schuppler, E. Pellegrin, P. Nagel, C. Pinta, B. Scheerer, and D. Fuchs, The<br />

soft x-ray beamline WERA at ANKA : International Conference for Synchrotron<br />

Radiation Instrumentation 2006, Daegu, South Korea, May 28 - June 2, 2006<br />

[53] G. Naletto, M.G. Pelizzo, G. Tondello, S. Nannarone, and A. Giglia, SPIE Proc.<br />

4145, 105 (2001).<br />

[54] M. Zangrando, M. Finazzi, F. Parmigiani, G. Paolucci, and D. Cocco, SPIE<br />

Proc. 4146, 132 (2000).<br />

[55] M. Zangrando, M. Finazzi, G. Paolucci, G. Comelli, B. Diviacco, R. P. Walker,<br />

D. Cocco, and F. Parmigiani, Rev. Sci. Instrum. 72, 1313 (2001).


Chapter 3<br />

Electronic Structure of<br />

Pr 0.67 Ca 0.33 MnO 3<br />

59


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 60<br />

3.1 Introduction<br />

Recently, a lot of attention have been focused on the charge-ordered compositions of<br />

Pr 1−x Ca x MnO 3 due to their importance as possible prototypes for the electronic phase<br />

separation (PS) models [1, 2, 3, 4, 5, 6, 7, 8, 9] proposed to explain the phenomenon<br />

of colossal magnetoresistance (CMR). The PS models are qualitatively different from<br />

the double-exchange model [10, 11, 12] or those based on strong Jahn-Teller polarons<br />

[13, 14]. According to the PS model the ground state of CMR materials is comprised<br />

of coexisting nanosize clusters of metallic ferromagnetic and insulating antiferromagnetic<br />

nature [6]. The insulator-metal transition in this scenario is through current<br />

percolation. Though there have been a number of experimental studies showing the<br />

existence of phase separation, their size varies from nano- to mesoscopic scales [8, 9].<br />

Radaelli et al. have shown the origin of mesoscopic phase separation to be the intergranular<br />

strain [9] rather than the electronic nature as in PS models. Nanosized<br />

stripes of a ferromagnetic phase [15] were reported in Pr 0.67 Ca 0.33 MnO 3 . This compound<br />

also attracted much attention earlier due to the existence of a nearly degenerate<br />

ferromagnetic metallic state and a charge-ordered antiferromagnetic insulating state<br />

with a field-induced phase transition possible between them [16]. With slight variations<br />

in the Ca doping the Pr 1−x Ca x MnO 3 system turns ferromagnetic (x=0.2) or<br />

antiferromagnetic (x=0.4) at low temperatures [17, 15]. The composition x=0.33<br />

shows a coexistence of ferromagnetic and antiferromagnetic phases [18]. This makes<br />

the charge- or orbital-ordered Pr 0.67 Ca 0.33 MnO 3 a prototype for the PS scenario.<br />

One of the requisites for the existence of an electronic phase separation in manganites<br />

is a strong-coupling interaction affecting the hopping of the itinerant e g electrons.<br />

The PS is expected to occur [5] when J H /t ≫ 1, where J H is the Hund’s coupling<br />

contribution between the localized t 2g and the e g electrons and t is the hopping amplitude<br />

of the e g electrons. The PS also favors a strong electron-phonon coupling,<br />

like the influence of a strong Jahn-Teller (JT) polaron arising from the Q 2 and Q 3<br />

JT modes. Essentially, one expects a strong localization of charge carriers to accompany<br />

the electronic separation of phases. Apart from charge, the orbital degrees of<br />

freedom also play an important role in this scenario [5]. The itinerant electron hopping<br />

term is strongly influenced by the symmetry of the orbitals (d x 2 −y 2 and d 3z 2 −r 2)<br />

hybridized with the O 2p orbitals of the MnO 6 octahedra. In comparison with the<br />

most popular CMR material La 1−x Sr x MnO 3 , the charge-ordered Pr system has an<br />

inherently reduced e g bandwidth W due to the smaller ionic radius of Pr, which also


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 61<br />

enhances the tendency for carrier localization. All these charge and orbital interactions<br />

are reflected in the near-E F electronic structure of these materials and could<br />

be probed using electron spectroscopic techniques. Valence band photoemission and<br />

O K x-ray absorption are two such proven tools sensitive to the changes in the lowenergy<br />

states crucial to these interactions. Although, there are many reports on<br />

the near-E F electronic structure of other CMR compounds using these techniques<br />

[19, 20, 21, 22, 23, 24, 25, 26, 27, 28] only a few studies have been reported on their<br />

charge-ordered compositions [8, 29]. One of the key energy terms these spectroscopies<br />

could show earlier was the charge transfer energy E CT which is intimately related to<br />

the electron-electron and electron-lattice interactions [20].<br />

In this study we have used ultraviolet photoelectron spectroscopy and x-ray absorption<br />

spectroscopy (XAS) in order to probe the electronic structure of the occupied<br />

and unoccupied states on a well-characterized, high-quality single crystal of<br />

Pr 0.67 Ca 0.33 MnO 3 . Another part of this single crystal had earlier been used for a detailed<br />

neutron scattering experiment that indicated the possible existence of a phase<br />

separation of ferromagnetic and antiferromagnetic stripes [15]. In the present study<br />

we have analyzed the temperature-dependent changes in the near-E F electron energy<br />

states from the perspective of a phase separation.<br />

3.2 Experimental<br />

The single-crystal sample of Pr 0.67 Ca 0.33 MnO 3 was grown by the floating zone method<br />

in a mirror furnace. The compositional homogeneity of the crystal was confirmed using<br />

energy-dispersive spectroscopic analysis. Magnetization and transport measurements<br />

on this crystal showed the transition temperatures T c and T N to be 100 and 110<br />

K, respectively. Details of the sample preparation, magnetization, and transport measurements<br />

and structural studies are published elsewhere [15, 30]. Angle-integrated<br />

ultraviolet photoemission measurements were performed using an Omicron µ-metal<br />

UHV system equipped with a high-intensity vacuum-ultraviolet source (HIS 13) and<br />

a hemispherical electron energy analyzer (EA 125 HR). At the He I (21.2 eV) line,<br />

the photon flux was of the order of 10 16 photons/s/sr with a beam spot of 2.5 mm<br />

diameter. Fermi energies for all measurements were calibrated using a freshly evaporated<br />

Ag film on a sample holder. The total energy resolution, estimated from the


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 62<br />

width of the Fermi edge, was about 80 meV. The single-crystalline samples were repeatedly<br />

scraped using a diamond file inside the chamber with a base vacuum of ∼<br />

1.0 × 10 −10 mbar. Scraping was repeated until negligible intensity was found for<br />

the bump around 9.5 eV, which is a signature of surface contamination [19]. For<br />

the temperature-dependent measurements, the sample was cooled by pumping liquid<br />

nitrogen through the sample manipulator fitted with a cryostat. Sample temperatures<br />

were measured using a silicon diode sensor touching the bottom of the sample<br />

holder. XAS measurements were performed using the BEAR [31] and BACH [32]<br />

beamlines associated with ELETTRA at Trieste, Italy. At the BEAR beamline we<br />

used monochromatized radiation from a bending magnet in order to record the O K<br />

edge spectra at room temperature and 150 K in fluorescence detection mode on a<br />

freshly scraped surface of the single crystal. The energy resolution was around 0.2<br />

eV in the case of these two spectra. The O K edge spectra at 95 K was recorded at<br />

the BACH beamline, using the total electron yield mode. Before the measurements,<br />

the sample surface was scraped inside the UHV chamber (∼ 1.0 × 10 −10 mbar) using<br />

a diamond file. At this beamline we used radiation from an undulator, monochromatized<br />

using a spherical grating. The resolution at the O Kedge for this measurement<br />

was better than 0.1 eV.<br />

3.3 Results and Discussion<br />

3.3.1 The UPS study of Pr 0.67 Ca 0.33 MnO 3<br />

The angle-integrated valence band photoemission spectra of Pr 0.67 Ca 0.33 MnO 3 taken<br />

at different temperatures below and above T c are shown in Fig. 3.1. Intensities of<br />

all the spectra were normalized and shifted along the y axis by a constant for clarity.<br />

The spectra, dominated by the states due to the Mn 3d-O 2p hybridized orbitals, look<br />

similar to those reported earlier on the La 1−x Sr x MnO 3 system [19, 20, 21, 22, 23]. The<br />

origin of the two prominent features, one at ∼ -3.5 eV (marked B) and another at<br />

∼ -5.6 eV (marked C) below E F , are now well known from earlier experiments and<br />

band structure calculations [23, 24, 33] on similar systems. While the feature at -3.5<br />

eV is mainly due to the t 2g↑ states of the MnO 6 octahedra, the -5.6 eV subband<br />

has contributions from both t 2g and e g states. More important contributors to the<br />

properties of these systems are the states nearer to E F , which appear as a tail at<br />

∼ -1.2 eV (marked A) from the chemical potential. The intensity of this feature


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 63<br />

C<br />

B<br />

Intensity (arb. units)<br />

A<br />

-10 -8 -6 -4 -2 0<br />

Energy Relative to E F<br />

(eV)<br />

Figure 3.1: Valence band photoemission spectra of Pr 0.67 Ca 0.33 MnO 3 taken using He<br />

I photon energy (21.2 eV) at 77 (solid line), 110 (dotted), 150 (dashed), 220 (dotdashed),<br />

and 300 K (double-dot dashed). All the spectra have been normalized and<br />

shifted along y-axis by a constant for clarity. The subbands around -1.2 (e g↑ ), -3.5<br />

(t 2g↑ ), and -5.6 eV (e g↑ + t 2g↑ ) are marked as A, B, and C respectively.


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 64<br />

300 K<br />

77 K<br />

5 X Difference<br />

(a)<br />

300 K<br />

110 K<br />

5 X Difference<br />

(b)<br />

Intensity (arb. units)<br />

-1 -0.5 0 0.5 1 1.5<br />

300 K<br />

220 K<br />

5 X Difference<br />

(d)<br />

-1 -0.5 0 0.5 1 1.5<br />

300 K<br />

150 K<br />

5 X Difference<br />

(c)<br />

-1 -0.5 0 0.5 1 1.5<br />

Binding Energy (eV)<br />

-1 -0.5 0 0.5 1 1.5<br />

Figure 3.2: High-resolution photoemission spectra of the near-E F region of the valence<br />

band of Pr 0.67 Ca 0.33 MnO 3 . In (a) the spectrum taken below T c (red) is compared with<br />

the normal state (300 K) spectrum (blue). The difference spectrum (green) obtained<br />

by subtracting the spectra at 300 K from 77 K and multiplied by 5 is also shown in<br />

the panel. Similarly, in (b), (c), and (d) the near-E F spectra taken at 110, 150, and<br />

220 K are compared with that taken at 300 K. The feature in the difference spectra<br />

corresponds to the tail feature A (e g↑ ) in Fig. 3.1.<br />

is quite small compared to B and C and is a signature of the insulating nature<br />

of this material. Earlier photoemission experiments on La 1−x Sr x MnO 3 also have<br />

shown that the intensity of this tail feature A is quite small [8, 20, 21, 22, 23].<br />

The presence of the feature A is clear from Fig. 3.2, where we have shown the<br />

near-E F region of the valence band spectra taken with a higher resolution. Here,<br />

the spectra taken at different temperatures have been compared with those at room<br />

temperature. The figure also shows the difference spectra obtained by subtracting<br />

the room-temperature spectra from spectra taken at low temperatures. The feature<br />

in the difference spectra corresponds to A, which originates from the e g↑ states [22].<br />

In order to estimate the relative changes in all the three features (A, B, and C)<br />

due to temperature, we have carefully fitted the whole valence band spectrum with


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 65<br />

0.014<br />

(a)<br />

0.012<br />

Normalized Intensity<br />

0.01<br />

0.6<br />

0.55<br />

0.5<br />

0.45<br />

(b)<br />

(c)<br />

0.4<br />

100 150 200 250 300<br />

Temperature (K)<br />

Figure 3.3: Temperature dependence of the area of the three valence band features<br />

obtained from fitting the spectra with Lorentzian line shapes using a χ 2 iterative<br />

program. We have used an integral background, which was kept the same for all the<br />

spectra. The energy positions and FWHMs were determined by finding the best fit<br />

common to all the spectra by the iterative program. The final fit for all spectra at<br />

different temperatures were obtained with the same energy positions and FWHMs.<br />

(a), (b), and (c) correspond to the features A, B, and C positioned at -1.19, -3.49,<br />

and -5.63 eV with FWHMs 2.56, 1.93, and 1.37 eV respectively.


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 66<br />

three components corresponding to the three subbands contributing to this region.<br />

We used a χ 2 iterative program for fitting the different spectra with Lorentzian line<br />

shapes for A, B, and C. Except for the area of the peaks, the energy positions,<br />

widths, background, and all other parameters of all the peaks were kept the same for<br />

different temperatures. The positions and full widths at half maximum (FWHMs)<br />

of the three peaks are given in the caption of Fig. 3.3. The fitted spectra for 77<br />

K (below T c ) and 300 K (above T c ) are shown on the left sides of Figs. 3.5(a) and<br />

3.5(b) below. One can see from Figs. 3.3(a) - (c) and that the area of all the peaks<br />

keeps increasing as we go down in temperature T c , but below this temperature the<br />

intensities of both peaks C and B decrease. On the other hand, the intensity of A<br />

does not show any decrease across the transition [Fig. 3.3(a)].<br />

Pr 0.67 Ca 0.33 MnO 3 shows a transition to a charge-ordered state below 220 K (T co )<br />

[30, 37, 38]. Neutron diffraction studies [15] on another part of this single crystal<br />

have shown that this charge ordering turns into an antiferromagnetically structured<br />

pseudo-CE type charge or orbital ordering below 110 K. Also, the coexistence of<br />

ferromagnetic and antiferromagnetic phases below this transition temperature has<br />

been shown on this single crystal. In the pseudo-CE-type ordering the Mn 3d z 2 −r 2<br />

orbitals, where the e g state is occupied, are aligned with the O 2p orbitals. Such an<br />

alignment can increase the hybridization between Mn 3d z 2 −r2 and O 2p. Furthermore,<br />

a simultaneous decrease in the hybridization strength could also be expected for the<br />

core-like in-plane t 2g and O 2p states. The results of the curve fitting (Fig. 3) of<br />

our valence band spectra reflects these changes in hybridization with temperature.<br />

The charge ordering following the decrease in temperature causes an increase in the<br />

O 2p contribution to both the t 2g and e g spin-up subbands in the valence region and<br />

hence the intensities of A, B, and C go up. As mentioned earlier, the pseudo-CE-type<br />

charge or orbital ordering below 110 K, results not only in a stronger hybridization of<br />

O 2p with the e gz 2 −r 2 ↑ states compared to that with the t 2g states but a simultaneous<br />

decrease in the latter also. This is reflected in the increase in intensity of A and<br />

decrease in the intensities of both C and B below 110 K. Though in Pr 0.67 Ca 0.33 MnO 3<br />

there is no temperature-dependent insulator-metal transition, the change in intensities<br />

of these features across T c appears similar to the shifting of spectral weight found in<br />

the La 1−x Sr x MnO 3 system [8].


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 67<br />

Normalized Absorption (arb. units)<br />

530 540 550 560<br />

Photon Energy (eV)<br />

Figure 3.4: O K edge x-ray absorption spectra of Pr 0.67 Ca 0.33 MnO 3 taken at 300<br />

(solid line), 150 (dashed line), and 95 K (dotted). The pre-edge feature centered<br />

around 529.5 eV (marked by a box) is where most interest lies. Since this feature<br />

consists of two peaks (a main line and a shoulder on the low-energy side), this part<br />

of the spectrum was fitted with two components of Lorentzian line shapes. Results<br />

of the curve fit are given in Table 3.1.<br />

3.3.2 The XAS study of Pr 0.67 Ca 0.33 MnO 3<br />

In a reasonable approximation, the near-E F region of the O K x-ray absorption<br />

spectra could well represent the density of unoccupied states in many of the transition<br />

metal oxide compounds [20, 34]. In order to probe the electronic structure of the<br />

unoccupied states we have performed XAS on our Pr 0.67 Ca 0.33 MnO 3 single crystal.<br />

The O K edge XAS spectra taken at room temperature and 150 and 95 K (below<br />

T c ), shown in Fig. 3.4, have been normalized in intensity all along the region starting<br />

from 550 eV. The pre-edge feature in the O K spectra (centered around 529.5 eV) is<br />

due to the strong hybridization between Mn 3d and O 2p orbitals. The broad feature<br />

around 536.5 eV is due to the bands from hybridized Pr 5d and Ca 3d orbitals, while<br />

the structure above 540 eV is due to states like Mn 4sp and Pr 6sp, etc. These<br />

states are known to contribute least to the near-E F electronic structure of these


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 68<br />

Table 3.1: χ 2 iterative fitting parameters for the pre-edge peak in O K XAS. The<br />

intensities of B ′ and A ′ at low temperature are normalized with respect to the total<br />

intensity of the pre-peak at room temperature. We have used an integral background,<br />

which was kept the same for all the spectra. The energy positions and FWHMs were<br />

determined by finding the best fit common to the three spectra by a χ 2 iterative<br />

program. The final fit for all the spectra from different temperatures were obtained<br />

with the same energy positions and FWHMs.<br />

Temperature (K)<br />

O K edge x-ray absorption spectra<br />

B ′ A ′<br />

Position FWHM Normalized Position FWHM Normalized<br />

95 (eV) (eV) area (eV) (eV) area<br />

2.11 1.52 0.72 1.48 0.97 0.29<br />

150 2.11 1.52 0.72 1.48 0.97 0.26<br />

300 2.12 1.52 0.73 1.47 0.97 0.26<br />

transition metal-oxide compounds. The assignments of the features mentioned above<br />

are consistent with the band structure calculations on similar systems [34, 35].<br />

The pre-edge feature in the O K edge spectra carries a substantial amount of<br />

physics involved in the properties of these materials. It has earlier been shown that<br />

the pre-peak in the O K edge spectra of different CMR materials consists of two lines,<br />

one of which appears as a shoulder on the low-energy side of the other [21, 25, 27, 28].<br />

The intensity of this shoulder was found to increase as the material goes across the<br />

insulator-metal transition [27, 28]. Since our Pr 0.67 Ca 0.33 MnO 3 sample is an insulator<br />

at all temperatures the pre-peak in our O K edge spectra (Fig. 3.4) does not show any<br />

splitting, though the presence of the shoulder is visible as an asymmetry on the lowenergy<br />

side of this peak. This shoulder feature is due to the first available unoccupied<br />

states and in XAS corresponds to the addition of an electron to the e g↑ state of the<br />

crystal-field-split MnO 6 octahedra. The main feature in the pre-peak arises from the<br />

t 2g↓ states [21]. For a quantitative estimate of the temperature-dependent changes in<br />

the intensities of the two features we have fitted the pre-peak using two components of<br />

Lorentzian line shapes. Here also the data fitting was done using χ 2 iterative method,<br />

keeping the energy positions and widths the same for all temperatures. Results of the<br />

data fit are shown in Table 3.1. Fitted spectra for temperatures above (300 K) and<br />

below Tc (95 K) are shown on the right side of Figs 3.5(a) and 3.5(b). Though very<br />

small, the intensity of the feature A ′ shows a slight increase below T c . The best fit


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 69<br />

gives a 0.6±0.1 eV difference between the energy positions of A ′ and B ′ . Following a<br />

different assignment of origin to the pre-peak features Dessau et al. have interpreted<br />

this energy difference as a measure of the Jahn-Teller distortion (4E JT ) [26].<br />

3.3.3 The Charge Transfer Energy (E CT )<br />

Having shown the temperature-dependent changes in the UPS and XAS results in<br />

Figs. 3.2 and 3.4, we present a combined picture of them in order to derive some insights<br />

into the charge transfer energy involved in the properties of the Pr 0.67 Ca 0.33 MnO 3<br />

sample. The combined spectra of the valence band from photoemission and the preedge<br />

region of the O K edge XAS, presented in Figs. 3.5(a) and 3.5(b) give the density<br />

of both occupied and unoccupied states above and below the Curie temperature (T c ).<br />

A schematic of these energy levels is shown in Fig. 3.6. The hole or electron doping<br />

(value of x) causes the symmetry of the last occupied and first unoccupied bands to<br />

be e gz 2 −r 2 ↑. In Fig. 3.5 these states are marked by A and A ′ . In order to scale the left<br />

and right sides of E F , the O K edge spectra have been shifted such that the energy<br />

of the first unoccupied state (1.4 eV from E F ) found from inverse photoemission [36]<br />

coincides with that of A ′ . We have used integral backgrounds for both sides of E F<br />

and, as mentioned earlier, all the fitting parameters were kept the same including<br />

the energy positions for the high and low temperatures. The important parameter<br />

which can be derived from this combined spectrum is the charge transfer energy E CT ,<br />

which is the energy required for an e g electron to hop between the Mn 3+ (t 3 2ge 1 g) and<br />

Mn 4+ (t 3 2g) sites. In our spectra (Fig. 3.5) E CT is the energy difference between A<br />

and A ′ which corresponds to the last occupied and first unoccupied states. The value<br />

of E CT is 2.6±0.1 eV from our spectra, which is higher than the value (1.5±0.4 eV)<br />

determined indirectly by Park et al. for the La 1−x Ca x MnO 3 system [20]. The larger<br />

value of E CT shows the existence of a stronger charge localization in the charge- or<br />

orbital-ordered Pr 0.67 Ca 0.33 MnO 3 compound.<br />

The value of E CT indicating a strong charge localization, found in our Pr 0.67 Ca 0.33<br />

MnO 3 has significant implications for the models proposed to explain the CMR effect<br />

in manganites. Charge localization can result from strong-coupling interactions<br />

of the e g electron with the core-like t 2g or the Q 2 and Q 3 Jahn-Teller modes. Both<br />

these interactions influence the electron hopping term (t). The former interaction,<br />

ferromagnetic in nature, is the Hund’s coupling J H . A large value of J H favors<br />

the electronic separation of phases. A strong electron-lattice coupling results from


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 70<br />

E CT<br />

(a)<br />

Intensity (arb. units)<br />

C<br />

B<br />

A’<br />

B’<br />

A<br />

-5 0 5 10<br />

Energy Relative to E F<br />

(eV)<br />

(b)<br />

B<br />

E CT<br />

B’<br />

Intensity (arb. units)<br />

C<br />

A’<br />

A<br />

-5 0 5 10<br />

Energy Relative to E F<br />

(eV)<br />

Figure 3.5: (a) Combined spectra from valence band photoemission (occupied energy<br />

states) and pre-peak in O K edge x-ray absorption spectra (unoccupied energy states)<br />

of Pr 0.67 Ca 0.33 MnO 3 taken at room temperature (above T c ). The three features corresponding<br />

to the subbands in the valence region are marked A, B, and C.The two<br />

peaks of the fitted O K edge pre-peak are marked A ′ and B ′ . We used integral background<br />

for both sides of E F . Details of the fitting are mentioned in the text. (b)<br />

Combined spectra of Pr 0.67 Ca 0.33 MnO 3 below T c . The valence band spectrum was<br />

taken at 77 K and O K edge XAS was taken at 95 K.


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 71<br />

2 2<br />

e g x − y<br />

J ex<br />

2 2<br />

e g z − r<br />

e g x 2 − y 2<br />

t 2g<br />

e z 2 2<br />

g − r<br />

2 2<br />

e g z − r<br />

E F<br />

t 2g<br />

Figure 3.6: Schematic diagram of the near-E F energy levels of Pr 0.67 Ca 0.33 MnO 3 in<br />

the occupied and unoccupied parts. The diagram is not drawn to scale. Some of the<br />

e g↑ states are occupied (at the Mn 3+ sites) and some are unoccupied (at the Mn 4+<br />

sites). Hence the last occupied and first unoccupied states are marked with dotted<br />

lines.<br />

the cooperative Jahn-Teller distortions of the MnO 6 octahedra. These distortions,<br />

particularly the Q 2 and Q 3 vibrational modes of the oxygen ions are stronger in<br />

the case of Pr-containing manganites compared to La-containing ones. Park et al.<br />

[20] have reported a smaller charge transfer energy (E CT = 1.5 eV) in the case of<br />

La 1−x Ca x MnO 3 and have attributed it to small polarons (Anderson localization) induced<br />

from the ionic size difference between Mn 3+ and Mn 4+ . A small-polaronic<br />

model may not be able to explain the high value of E CT found in Pr 0.67 Ca 0.33 MnO 3 .<br />

On the other hand, a strong charge localization is expected in the case of the PS<br />

model in which the ground state is described as a mixture of ferromagnetic and antiferromagnetic<br />

regions. It is also possible that this large charge localization is due to<br />

the Zener polaron proposed by Aladine et al [39]. In the Zener polaron picture, the<br />

Mn ions of adjacent MnO 6 octahedra form a dimer with variations in the Mn-O-Mn<br />

bond angle. The pseudo-CE-type charge ordering in Pr 1−x Ca x MnO 3 favors such regular<br />

distortions in the MnO 6 octahedra. Further studies using electron spectroscopic<br />

techniques on samples in the vicinity of the charge-ordered compositions may be able<br />

to differentiate between these possible driving mechanisms behind the strong charge


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 72<br />

localization.<br />

3.4 Conclusions<br />

Using valence band photoemission and O K edge x-ray absorption, we have probed<br />

the electronic structure of Pr 0.67 Ca 0.33 MnO 3 , which is regarded as a prototype for the<br />

electronic phase separation models in CMR systems. With decrease in temperature<br />

the O 2p contributions to the t 2g and e g spin-up states in the valence band were<br />

found to increase until T c . Below T c , the density of states with e g spin-up symmetry<br />

increased while those with t 2g symmetry decreased, possibly due to the change in<br />

the orbital degrees of freedom associated with the Mn 3d - O 2p hybridization in the<br />

pseudo-CE-type charge or orbital ordering. These changes in the density of states<br />

could well be connected to the electronic phase separation reported earlier. Our<br />

quantitative estimate of the charge transfer energy (E CT ) is 2.6±0.1 eV, which is<br />

large compared to the earlier reported values in other CMR systems. Such a large<br />

charge transfer energy may support the phase separation model.


Bibliography<br />

[1] I. G. Deac, J. F. Mitchell and P. Schiffer Phys. Rev. B 63, 172408 (2001).<br />

[2] L. M. Fisher, A. V. Kalinov, I. F. Voloshin, N. A. Babushkina, K. I. Kugel, and<br />

D. I. Khomskii Phys. Rev. B 68, 174403 (2003).<br />

[3] R. Kajimoto, H. Mochizuki, H. Yoshizawa, S. Okamoto, and S. Ishihara, Phys.<br />

Rev. B 69, 54433 (2004).<br />

[4] K. S. Nagapriya, A. K. Raychoudhuri, B. Bansal, V. Venkataraman, S. Parashar,<br />

and C. N. R. Rao, Phys. Rev. B 71, 024426 (2005).<br />

[5] A. Moreo, S. Yunoki, and E. Dagatto, Science 283, 2034 (1999).<br />

[6] A. Moreo, M. Mayr, A. Feiguin, S. Yunoki, and E. Dagotto, Phys. Rev. Lett. 84,<br />

5568 (2000).<br />

[7] S. Yunoki, J. Hu, A. L. Malvezzi, A. Moreo, N. Furukawa, and E. Dagotto, Phys.<br />

Rev. Lett. 80, 845 (1998).<br />

[8] D. D. Sarma, D. Topwal, U. Manju, S. R. Krishnakumar, M. Bertolo, S. La Rosa,<br />

G. Cautero, T. Y. Koo, P. A. Sharma, S.-W. Cheong, and A. Fujimori, Phys.<br />

Rev. Lett. 93, 97202 (2004).<br />

[9] P. G. Radaelli, R. M. Ibberson, D. N. Argyriou, H. Casalta, K. H. Andersen, S.<br />

-W. Cheong and J. F. Mitchell, Phys. Rev. B 63, 172419 (2001)<br />

[10] C. Zener, Phys. Rev. 82, 403 (1951).<br />

[11] P. W. Anderson and H. Hasegawa, Phys. Rev. 100, 675 (1955).<br />

[12] P. -G. de Gennes, Phys. Rev. 118, 141 (1960).<br />

73


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 74<br />

[13] A. J. Millis, P. B. Littlewood, and B. I. Shraiman, Phys. Rev. Lett. 74, 5144<br />

(1995).<br />

[14] A. J. Millis, B. I. Shraiman, and R. Mueller Phys. Rev. Lett. 77, 175 (1996).<br />

[15] Ch. Simon, S. Mercone, N. Guiblin, C. Martin, A. Brulet, and G. Andre, Phys.<br />

Rev. Lett. 89, 207202 (2002).<br />

[16] K. Miyano, T. Tanaka,Y. Tomioka, and Y. Tokura, Phys. Rev. Lett. 78, 4257<br />

(1997).<br />

[17] E. O. Wollan, and W. C. Koehler, Phys. Rev. 100, 545 (1955)<br />

[18] Z. Jirak, S. Krupicka, Z. Simsa, M. Dlouha, and S. Vratislav, J. Magn. Magn.<br />

Mater. 53, 153 (1985)<br />

[19] D. D. Sarma, N. Shanthi, S. R. Krishnakumar, T. Saitoh, T. Mizokawa, A.<br />

Sekiyama, K. Kobayashi, A. Fujimori, E. Weschke, R. Meier, G. Kaindl, Y.<br />

Takeda, and M. Takano, Phys. Rev. B 53, 6873 (1996).<br />

[20] J.-H. Park, C. T. Chen, S. W. Cheong, W. Bao, G. Meigs, V. Chakarian, and Y.<br />

U. Idzerda, Phys. Rev. Lett. 76, 4215 (1996)<br />

[21] J.-H. Park, T. Kimura, and Y. Tokura, Phys.Rev. B 58, R13330 (1998).<br />

[22] T. Saitoh, A. E. Bocquet, T. Mizokawa, H. Namatame, A. Fujimori, M. Abbate,<br />

Y. Takeda, and M. Takano, Phys. Rev. B 51, 13942 (1995).<br />

[23] T. Saitoh, A. Sekiyama, K. Kobayashi, T. Mizokawa, A. Fujimori, D. D. Sarma,<br />

Y. Takeda, and M. Takano, Phys. Rev. B 56, 8836 (1997).<br />

[24] P. Pal, M. K. <strong>Dala</strong>i, B. R. Sekhar, S. N. Jha, S. V. N. Bhaskara Rao, N. C. Das,<br />

C. Martin, and F. Studer, J. Phys.: Condens. Matter 17, 2993 (2005).<br />

[25] D. S. Dessau, Y. D. Chuang, A. Gromko, T. Saitoh, T. Kimura, and Y. Tokura,<br />

J. Electron Spectrosc. and Relat. Phenom. 117-117, 265 (2001).<br />

[26] D. S. Dessau, T. Saitoh, C.-H. Park, Z.-X. Shen, Y. Moritomo, and Y. Tokura, in<br />

Science and Technology of Magnetic Oxides, edited by M. Hundley, J. Nickel,<br />

R. Ramesh, and Y. Tokura, MRS Symposia Proceedings No. 494 (Materials<br />

Research Society, Pittsburgh, 1998),p. 181.


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 75<br />

[27] N. Mannella, A. Rosenhahn, M. Watanabe, B. Sell, A. Nambu, S. Ritchey, E.<br />

Arenholz, A. Young, Y. Tomioka, and C. S. Fadley, Phys. Rev. B 71, 125117<br />

(2005).<br />

[28] O. Toulemonde, F. Millange, F. Studer, B. Raveau, J. H. Park, and C. T. Chen,<br />

J. Phys. Condens. Matter 11, 109 (1999).<br />

[29] J.-H. Park, C. T. Chen, S.-W.Cheong, W. Bao,and G. Meigs, J. Appl. Phys. 79,<br />

4558 (1996).<br />

[30] S. Mercone, V. Hardy, C. Martin, Ch. Simon, D. Saurel,and A. Brulet, Phys.<br />

Rev. B 68, 094422 (2003).<br />

[31] S. Nannarone, F. Borgatti, A. DeLuisa, B. P. Doyle, G. C. Gazzadi, A. Giglia,<br />

P. Finetti, N. Mahne, L. Pasquali, M. Pedio, G. Selvaggi, G. Naletto, M. G.<br />

Pelizzo, and G. Tondello, in SYNCHROTRON RADIATION INSTRUMENTA-<br />

TION: Eighth international conference on synchrotron radiation instrumentation,<br />

edited by T. Warwick, J. Stohr, H. A. Padmore, and J. Arthur, AIP Conf.<br />

Proc. No. 705 (AIP, San Franscisco, California, 2003), pp. 450-453.<br />

[32] M. Zangrando, M. Finazzi, G. Paolucci, G.Comelli, B. Diviacco, R. P. Walker,<br />

D. Cocco, and F. Parmigiani, Rev. Sci. Instrum. 72, 1313 (2001).<br />

[33] E. Z. Kurmaev, M. A. Korotin, V. R. Galakhov, L. D. Finkelstein, E. I.<br />

Zabolotzky, N. N. Efremova, N. I. Lobachevskaya, S. Stadler, D. L. Ederer, T.<br />

A. Callcott, L. Zhou, A. Moewes, S. Bartkowski, M. Neumann, J. Matsuno, T.<br />

Mizokawa, A. Fujimori, and J. Mitchell, Phys. Rev. B 59, 12799 (1999).<br />

[34] F. M. F. de Groot, M. Grioni, J. C. Fuggle, J. Ghijsen, G. A. Sawatzky, and H.<br />

Petersen, Phys. Rev. B 40, 5715 (1989).<br />

[35] H. Kurata, E. Lefevre, C. Colliex, and R. Brydson, Phys. Rev. B 47, 13763<br />

(1993).<br />

[36] A. Chainani, M. Mathew, and D. D. Sarma, Phys. Rev. B 47, 15397 (1993).<br />

[37] C. Martin, A. Maignan, M. Hervieu, and B. Raveau, Phys. Rev. B 60, 12191<br />

(1999).<br />

[38] V. Hardy, A. Wahl, and C. Martin, Phys. Rev. B 64, 64402 (2001)


Electronic Structure of Pr 0.67 Ca 0.33 MnO 3 76<br />

[39] A. Daoud-Aladine, J. Rodriguez-Carvajal, L. Pinsard-Gaudart, M. T. Fernandez-<br />

Diaz, and A. Revcolevschi, Phys. Rev. Lett. 89, 097205 (2002).


Chapter 4<br />

Electronic Structure of<br />

Pr 1−x Ca x MnO 3<br />

77


Electronic Structure of Pr 1−x Ca x MnO 3 78<br />

4.1 Introduction<br />

Mixed phase manganite systems with coexisting ferromagnetic (FM) and antiferromagnetic<br />

(AFM) domains have been attracting considerable attention recently due to<br />

their importance in understanding the role of spin, charge, and orbital ordering in the<br />

phenomena of colossal magneto resistance (CMR). Theoretical models, particularly<br />

those [1, 2, 3] based on electronic phase separation (PS) exploits this character of the<br />

CMR compounds in explaining the insulator-metal (IM) transitions found in these<br />

systems. The electronic PS scenario is phenomenologically different from the double<br />

exchange (DE) mechanism [4, 5, 6] or those based on strong Jahn-Teller polarons<br />

[7, 8]. In this scenario, the IM transition is driven by current percolation through<br />

FM metallic domains embedded in an AFM insulating background [2]. Though, there<br />

have been many structural studies [9, 10] showing the co-existence of such magnetic<br />

domains, the reported sizes of those vary from nano - to mesoscopic scales [11, 12]. In<br />

manganites, the electronic PS is expected to occur in systems in which the itinerant<br />

e g electrons have strong Hund’s coupling interaction with the localized t 2g electrons<br />

of the MnO 6 octahedra in their structure.<br />

Some of the charge ordered (CO) compositions of the Pr 1−x Ca x MnO 3 system are<br />

considered to be the prototypes for the PS models owing to their inherently reduced<br />

e g bandwidth W. Neutron diffraction studies have shown the co-existence of nanosized<br />

FM and AFM stripes [10, 12] in Pr 0.67 Ca 0.33 MnO 3 . It was also found that a fieldinduced<br />

transition possible between their nearly degenerate FM metallic and chargeordered<br />

AFM insulating phases [12, 13]. With slight variations in the Ca doping<br />

the Pr 1−x Ca x MnO 3 system turns ferromagnetic (x = 0.2) or antiferromagnetic (x =<br />

0.4) at low temperatures [10, 14]. The composition x = 0.33 shows a co-existence of<br />

ferromagnetic and antiferromagnetic phases [15]. The energy scales involved in the<br />

IM transitions in this FM-AFM system are of importance for the understanding of<br />

the physics behind the phenomena of CMR.<br />

In this chapter we have compared the binding energy positions of some of the near<br />

Fermi level (E F ) occupied and unoccupied states, crucial to the electrical and magnetic<br />

transitions of the Pr 1−x Ca x MnO 3 system using the complementary techniques<br />

of direct and inverse photoelectron spectroscopies. Electron spectroscopic techniques<br />

have been quite successful in probing the near E F low energy states involved in the<br />

charge and orbital interactions of CMR materials. Although, there are many reports<br />

on the electronic structure of other CMR compounds using some of these techniques


Electronic Structure of Pr 1−x Ca x MnO 3 79<br />

Figure 4.1: Phase diagram of the Pr 1−x Ca x MnO 3 system. In this study we have used<br />

the x = 0.2, x = 0.33 and x = 0.4 compositions. (FMI = ferromagnetic insulating,<br />

CO-AFMI = charge ordered antiferromagnetic insulating, CG = cluster glass, T N =<br />

Neel Temperature, T C = Critical Temperature, T CO = charge ordering temperature).<br />

[16, 17, 18, 19, 20, 21, 22] only a few studies have dealt with the energy scales of the<br />

electron interactions in charge ordered compositions [23, 24]. One of the key terms,<br />

these spectroscopies could show earlier, was the charge transfer energy E CT which is<br />

intimately related to the electron-electron and electron-lattice interactions [23, 24].<br />

Here we have analyzed the temperature dependent changes shown by three compositions<br />

across the charge ordered FM-AFM phase boundary of the Pr 1−x Ca x MnO 3<br />

system using ultraviolet photoelectron spectroscopy (UPS) and inverse photoelectron<br />

spectroscopy (IPES). Also we have studied different core levels of Pr 1−x Ca x MnO 3<br />

using x-ray photoelectron spectroscopy (XPS).<br />

4.2 Experimental<br />

Single crystals of Pr 1−x Ca x MnO 3 with x = 0.2, 0.33 and 0.4 were grown by the floating<br />

zone method in a mirror furnace. The compositional homogeneity of the crystals<br />

had been confirmed using energy dispersive spectroscopic (EDS) analysis. The<br />

crystals were characterized by using magnetization, electrical transport and neutron<br />

diffraction measurements. Details of the sample preparation and characterization


Electronic Structure of Pr 1−x Ca x MnO 3 80<br />

measurements are published elsewhere [10, 25]. There are also three polycrystalline<br />

samples (PrMnO 3 , CaMnO 3 and Pr 0.16 Ca 0.14 MnO 3 ) used for different studies. Which<br />

were prepared by the standard solid state reaction method. The details of the sample<br />

preparation is published elsewhere [26]. Consolidated results of these studies are<br />

presented in the phase diagram shown in Fig. 4.1 [27]. Angle integrated ultraviolet<br />

photoemission measurements were performed by using an Omicron mu-metal ultra<br />

high vacuum system equipped with a high intensity vacuum ultraviolet source (HIS<br />

13) and a hemispherical electron energy analyzer (EA 125 HR). At the He I (21.2<br />

eV) line, the photon flux was of the order of 10 16 photons/sec/steradian with a beam<br />

spot of 2.5 mm diameter. The Fermi energies (E F ) for all measurements were calibrated<br />

by using a freshly evaporated Ag film on a sample holder. The total energy<br />

resolution, estimated from the width of the Fermi edge, was about 80 meV. The XPS<br />

measurements were performed using a twin anode Mg/Al x-ray source (DAR 400) in<br />

the same chamber as of UPS. In our measurements we have used the Mg Kα line with<br />

photon energy 1253.6 eV and the resolution was about 1.0 eV. The samples were repeatedly<br />

scraped by using a diamond file inside the chamber under a base vacuum of<br />

∼ 1.0 x 10 −10 mbar. The negligible intensity found for the ∼ 9.5 eV bump [16] in the<br />

UPS spectra ensures the cleanliness of our sample surface. For the temperature dependent<br />

measurements, the samples were cooled by pumping liquid nitrogen through<br />

the sample manipulator fitted with a cryostat. The sample temperatures were measured<br />

by using a silicon diode sensor touching the bottom of the sample holder. The<br />

IPES measurements were performed by using an electrostatically focused electron gun<br />

of Stoffel Johnson design and an acetone gas filled band pass Geiger Muller counter<br />

with a CaF 2 window [28, 29]. These measurements were carried out in the isochromat<br />

mode in which the photon energy is fixed at 9.9 eV and kinetic energy of the incident<br />

electrons is varied in 0.05 eV steps. Overall resolution of the IPES measurements<br />

were estimated [29] to be approximately 0.55 eV.<br />

4.3 Results and Discussion<br />

4.3.1 The UPS study of Pr 1−x Ca x MnO 3 , x = 0.2, 0.33 and 0.4<br />

Neutron diffraction studies [27] on Pr 1−x Ca x MnO 3 have shown that the compositions<br />

with x = 0.2 and x = 0.4 are FM and AFM respectively (Fig. 4.1) while the x = 0.33<br />

has co-existing FM and AFM phases [14, 15]. The ground state of x = 0.33 has been


Electronic Structure of Pr 1−x Ca x MnO 3 81<br />

identified to be an inhomogeneous state below T CO . The x = 0.4 is above the critical<br />

doping (x c ) found in the Pr 1−x Ca x MnO 3 system and is homogeneous in nature.<br />

(a) 300 K<br />

C<br />

B<br />

(b) 77 K<br />

C<br />

B<br />

Intensity (arb. units)<br />

A<br />

1.0<br />

0.4<br />

0.33<br />

Intenisty (arb. units)<br />

A<br />

1.0<br />

0.4<br />

0.33<br />

0.2<br />

0.2<br />

0.0<br />

0.0<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Binding Energy (eV)<br />

Binding Energy (eV)<br />

Figure 4.2: The angle integrated valence band photoemission spectra of the<br />

Pr 1−x Ca x MnO 3 (x = 0.0, 0.2, x = 0.33, x = 0.4 and x = 1.0) samples taken at<br />

room temperature and 77K using He I photons (21.2 eV).<br />

Figure 4.2 shows the angle integrated valence band photoemission spectra of the<br />

Pr 1−x Ca x MnO 3 (x = 0.2, x = 0.33 and x = 0.4) samples taken at room temperature<br />

and 77K. The figure also shows the spectra of the end compositions of this system (x<br />

= 0.0 and x = 1.0). It may be noted from the phase diagram that the compositions<br />

with x = 0.2, x = 0.33 and x = 0.4 are PM insulators at 300K and FM/AFM insulators<br />

at 77K. Intensities of all the spectra shown in both the panels were normalized<br />

and shifted along the ordinate axis by a constant value for clarity. All the features<br />

seen in the spectra are dominated by the states due to the Mn 3d - O 2p hybridized<br />

orbitals. The two prominent features, one at ∼ 3.5 eV (marked B) and another at ∼<br />

5.6 eV (marked C), are primarily due to the O 2p nonbonding states and the Mn 3d<br />

- O 2p bonding states [30]. The feature closest to the E F , which appear as a tail at


Electronic Structure of Pr 1−x Ca x MnO 3 82<br />

∼ 1 eV (marked A) is due to the e g states of the MnO 6 octahedra [18, 24, 31, 32].<br />

The intensity of this feature is low due to the insulating nature of all our samples.<br />

Presence of the feature A is clear from Fig. 4.3(a) and 4.3(b) where we have shown<br />

the photoemission spectra from the near E F region of the x = 0.2, x = 0.33 and x = 0.4<br />

samples along with that of CaMnO 3 , taken at temperatures above and below the T CO .<br />

The lower panels of the figures show the difference spectra obtained by subtracting the<br />

spectra of CaMnO 3 from those of the different Pr 1−x Ca x MnO 3 compositions. Since,<br />

CaMnO 3 does not have any electrons in its e g levels, the difference spectra should<br />

correspond to the density of the e g electrons in the Pr 1−x Ca x MnO 3 samples. The<br />

shape of the difference spectra shows that it consists of two features, similar to those<br />

reported by Ebata et al. [33]. We have fitted all the difference spectra corresponding<br />

to the three compositions at two different temperatures with a mixture of Lorentzian<br />

and Gaussian line shapes using a χ 2 iterative programme. We used a fixed width<br />

for all the fitted lines marked A’ for both the temperatures in order to see the finer<br />

changes associated with the A”. The energy positions of the A’ and A” and the<br />

width of the A” were determined by the iterative programme. Table 4.2 shows the<br />

results of the fitting. At room temperature, the changes across the three compositions<br />

are negligible except that the sample with x = 0.33 has a larger width for its near<br />

E F feature marked A”. At 77K, both A’ and A” have been found to be shifted<br />

slightly towards the E F . Further, at this temperature also, the x = 0.33 shows a<br />

wider A”. The width of the near E F feature is coupled to the conductivity of these<br />

samples. Earlier experiments using valence band photoemission have shown that<br />

a finite number of states get transferred from A” to A’ when some CMR systems<br />

go across the temperature - composition phase diagram [31, 33, 34] resulting in the<br />

formation of pseudogaps at the E F . The energy scales involved in such shifts of DOS<br />

are of importance to the theoretical models, particularly for those based on electronic<br />

phase separation. Formation of pseudogaps are a generic feature of mixed phase<br />

CMR systems [31, 35] with FM electron rich phases embedded in an insulating AFM<br />

matrix. The feature A’ in our spectra could possibly be ascribed to the e g electrons<br />

in the AFM phases while the A” to those in the FM phases. The difference in the<br />

binding energy positions of A’ and A” (0.47±0.01 eV) gives an approximate measure<br />

of the energy difference between the e g electronic states in these two phases. In other<br />

words, this gives the energy scale over which the pseudogap forms. The increased<br />

width of the feature A” of x = 0.33 at temperatures above and below the T CO may


Electronic Structure of Pr 1−x Ca x MnO 3 83<br />

(a)<br />

300 K 300 K 300 K<br />

Intensity (arb. units)<br />

A’<br />

A’<br />

A’<br />

A’’ A’’ A’’<br />

0.2 - CaMnO3 0.33 - CaMnO3 0.4 - CaMnO3<br />

1.5 1 0.5 0 1.5 1 0.5 0<br />

1.5<br />

1<br />

0.5<br />

0<br />

Binding Energy (eV)<br />

(b)<br />

77 K 77 K<br />

77 K<br />

Intensity (arb. units)<br />

A’<br />

A’<br />

A’<br />

A’’<br />

A’’<br />

A’’<br />

0.2 - CaMnO3 0.33 - CaMNO3 0.4 - CaMnO3<br />

1.5<br />

1<br />

0.5<br />

0<br />

1.5<br />

1<br />

0.5<br />

0<br />

1.5<br />

1<br />

0.5<br />

0<br />

Binding Energy (eV)<br />

Figure 4.3: The near E F region of the valence band spectra (panel(a): 300 K,<br />

panel(b): 77 K) of the three samples plotted against that of CaMnO 3 . The difference<br />

spectra (lower panel) is obtained by subtracting the spectra of CaMnO 3 from<br />

those of the three Pr 1−x Ca x MnO 3 compositions and multiplied by 5. The difference<br />

spectra corresponds to the density of e g states. The red and blue peaks are A’ and<br />

A” respectively. The green line shows the spectral background used in the fitting.


Electronic Structure of Pr 1−x Ca x MnO 3 84<br />

Table 4.2: Results of the fitting of the difference spectra using a χ 2 iterative programme.<br />

We have used a mixture of Lorentzian and Gaussian line shapes for fitting<br />

the various difference spectra corresponding to the three compositions at 300K and<br />

77K. In order to see the finer changes in the near E F feature A”, we used a fixed<br />

width for all the fitted lines marked A’. The energy positions of the A’ and A” and the<br />

width of the A” were determined by the iterative programme. At both the temperatures<br />

the sample with x = 0.33 shows a larger full width at half maximum (FWHM)<br />

for its A”. At 77K, both A’ and A” have been found to be shifted slightly towards<br />

the E F .<br />

Temperature (K)<br />

300<br />

77<br />

x<br />

A’ A”<br />

Position (eV) FWHM (eV) Position (eV) FWHM (eV)<br />

0.2 0.97 0.5 0.49 0.5<br />

0.33 0.96 0.5 0.49 0.65<br />

0.4 0.96 0.5 0.49 0.59<br />

0.2 0.92 0.5 0.45 0.49<br />

0.33 0.93 0.5 0.45 0.53<br />

0.4 0.92 0.5 0.46 0.5<br />

indicate the formation of FM clusters in the AFM insulating background. From the<br />

fitting of the difference spectra we feel that the increase of the width of A” is largely<br />

due to the formation of a tail towards the E F , which makes it difficult to estimate<br />

the energy difference between A’ and A” accurately.<br />

4.3.2 The IPES study of Pr 1−x Ca x MnO 3 , x = 0.2, 0.33, 0.4<br />

and 1<br />

Fig. 4.4 shows the IPES spectra of the Pr 1−x Ca x MnO 3 samples (x = 0.2, 0.33 and<br />

0.4) and the CaMnO 3 sample taken at room temperature. The spectra have been<br />

normalized for their intensities and shifted along the vertical axis by a constant for<br />

clarity. The spectra shows two broad features; one between 5 to 12 eV (marked Q)<br />

and another at around 2.4 eV (marked P). For the x = 0.2, 0.33, and 0.4, the feature<br />

appearing at higher energy is mostly due to Pr 5d and Ca 3d orbitals, whereas for<br />

the CaMnO 3 it is due only to the Ca 3d, 4s orbitals. The feature P is due to the<br />

hybridized Mn 3d - O 2p states. These assignments of the features are consistent<br />

with band structure calculations and other studies on similar systems [24, 36, 37].<br />

Similar energy positions were reported earlier for La 1−x Sr x MnO 3 by Chainani et al.


Electronic Structure of Pr 1−x Ca x MnO 3 85<br />

Q<br />

x=1<br />

P<br />

x=0.4<br />

Intensity (arb. units)<br />

x=0.33<br />

x=0.2<br />

0 2.5 5 7.5 10 12.5 15<br />

Energy Relative to Fermi Energy (eV)<br />

Figure 4.4: The IPES spectra of the Pr 1−x Ca x MnO 3 samples (x = 0.2, 0.33 and 0.4)<br />

and the CaMnO 3 sample taken at room temperature. The spectra represents the<br />

unoccupied electron states of the samples.<br />

[38], though another study [39] showed the feature P to be more closer to the E F .<br />

Again, in the IPES spectra also, the feature appearing near to E F is important to<br />

properties of these compounds. In Fig. 4.5 we have shown the spectra of the near E F<br />

region containing this feature along with the difference spectra (Fig. 4.5(b)). One can<br />

see clearly that the intensity of P increases with the Ca concentration x due to the<br />

increase of unoccupied states in the Mn 3d orbitals with hole doping. From a linear<br />

extrapolation of the rising band edge in the spectra shown in the Fig. 4.5(a) we have<br />

estimated the band gap to be ∼ 0.7 eV for the three compositions (0.2, 0.33, and 0.4).<br />

At this temperature, the value of the gap does not seem to vary with the charge carrier<br />

concentration, whereas the parent compound (Ca x MnO 3 ) shows a smaller value (∼<br />

0.67 eV). In case of La 1−x Sr x MnO 3 , Chainani et al. [38] has shown the value of the<br />

band gap to be ∼ 0.6 eV. The larger values shown by our Pr 1−x Ca x MnO 3 samples<br />

could be attributed to their insulating nature. In an alternate approach, one can also<br />

estimate the band gap from the difference spectra shown in the Fig. 4.5(b). This<br />

gives a some what larger value; i.e ∼ 0.86 eV for the x = 0.4 and 0.33 and ∼ 0.82 for


Electronic Structure of Pr 1−x Ca x MnO 3 86<br />

(a)<br />

(b)<br />

Intensity (arb. units)<br />

x=1<br />

x=0.4<br />

1 - 0.2<br />

0.4 - 0.2<br />

x=0.33<br />

x=0.2<br />

0.33 - 0.2<br />

-1 0 1 2 3 4<br />

0 1 2 3 4<br />

Energy Relative to Fermi Energy (eV)<br />

Figure 4.5: Panel(a) shows the IPES spectra of the near E F unoccupied states.<br />

Panel(b) shows the difference spectra obtained by subtracting the spectra corresponding<br />

to the x = 0.2 from those of other compositions.<br />

the Ca x MnO 3 .<br />

Band structure calculations on similar CMR systems [12, 32] have unambiguously<br />

shown that both the highest occupied and lowest unoccupied states in these systems<br />

are of Mn 3d e g spin-up symmetry. As mentioned earlier, within this frame work the<br />

features marked A’ and A” in Fig. 4.3 and the feature P in Fig. 4.4 correspond to the<br />

e g spin-up states. Combining the results presented in Table I and these two figures<br />

one can estimate the energy required by an e g electron to hop from a Mn 3+ site to the<br />

neighboring Mn 4+ site. If we consider the centroids of the A’ in Fig. 4.3 and the P in<br />

Fig. 4.4 we arrive at a value of ∼ 3.3±0.2 eV, which is large compared to the earlier<br />

reported value [24]. With the A” of Fig. 4.3 and P of Fig. 4.4 we get ∼ 2.8±0.2<br />

eV which is close to the earlier report (2.6±0.1 eV). Due to the large widths and<br />

long tails of the A” and P towards the E F we expect that this value of the energy be<br />

close to its upper bound. Energy required for the transfer of e g electrons between the<br />

Mn sites again is an important parameter for the phase separation models to work in


Electronic Structure of Pr 1−x Ca x MnO 3 87<br />

these systems.<br />

4.3.3 The core levels of Pr 1−x Ca x MnO 3 , x = 0.2, 0.33, 0.4,<br />

and 0.84<br />

The Mn2p core level spectra of Pr 1−x Ca x MnO 3 (x = 0.2, 0.33, 0.4, and 0.84) taken<br />

at room temperature as well as at 77 K are shown in Fig. 4.6. All the spectra were<br />

normalized and shifted along the Y axis by a constant for clarity. The two prominent<br />

peaks at around ∼ 641.5 and ∼ 653.5 eV are due to the spin orbit split 2p 3/2 and 2p 1/2 .<br />

These spectra look similar to those reported earlier on the similar system [17, 33]. In<br />

Fig. 4.6, the Mn 2p peak positions shift towards the higher binding energy for x =<br />

0.4 and 0.84 at room temperature (a) as well as at 77 K (b). But the binding energy<br />

shift is higher incase of room temperature spectra as compared to that of 77 K. The<br />

Mn LMV Auger peak is observed on the higher binding energy side of 2p 1/2 peak<br />

for all four concentrations at room temperature, whereas at 77 K this Auger peak is<br />

absent for x = 0.2, and 0.3. This Mn LMV was earlier observed in other hole doped<br />

manganites [17].<br />

The Ca 2p spectra of Pr 1−x Ca x MnO 3 (x = 0.2, 0.33, 0.4 and 0.84) both at room<br />

temperature as well as 77 K are shown in the Fig. 4.7. The Ca 2p 3/2 and Ca 2p 1/2 ,<br />

due to spin orbit splitting, are clearly distinguishable at around ∼ 345.5 and ∼ 349<br />

eV respectively. It is clear from the Fig. 4.7 (a) that, at room temperature the peak<br />

positions shift towards the lower binding energy when we move from x = 0.2 to x =<br />

0.84. Also the same kind of shift is observed for 77 K spectra (Fig. 4.7 (b)) with a<br />

lesser magnitude.<br />

The Pr 4d spectra of Pr 1−x Ca x MnO 3 (x = 0.2, 0.33 and 0.4) taken at room<br />

temperature as well as 77 K are shown in Fig. 4.8. The Pr 4d peaks come around<br />

116 eV binding energy. The peak positions shift towards the lower binding energy<br />

at room temperature (Fig. 4.8 (a)) when we move from x = 0.2 to 0.4. Here also<br />

the similar kind of shift is observed for 77 K spectra (Fig. 4.8 (b)) with a lesser<br />

magnitude.<br />

As mentioned above the binding energy shifts with respect to the value of x have<br />

been observed at the Ca 2p, Pr 4d and Mn 2p core levels. Incase of Mn 2p level, the<br />

binding energy shift is to the higher energy side while the other two levels shift to lower<br />

energy. The magnitude of the binding energy shifts are found to be higher at room<br />

temperature compared to 77 K. The shift to higher binding energy observed in Mn


Electronic Structure of Pr 1−x Ca x MnO 3 88<br />

Mn LMV<br />

Temp = 300 K<br />

2p 3/2<br />

Intensity (arb. units)<br />

0.84<br />

0.4<br />

0.33<br />

Mn 2p<br />

0.2<br />

670<br />

660<br />

2p 1/2<br />

640<br />

650<br />

640<br />

Binding energy (eV)<br />

(a)<br />

Mn LMV<br />

Temp = 77 K<br />

2p 2p 1/2 3/2<br />

Intensity (arb. units)<br />

0.84<br />

0.4<br />

0.33<br />

Mn 2p<br />

0.2<br />

670<br />

660<br />

650<br />

Binding Energy (eV)<br />

(b)<br />

Figure 4.6: The Mn 2p core-level XPS (Mg Kα) spectra of Pr 1−x Ca x MnO 3 taken at<br />

room temperature (a) and 77 K (b) .


Electronic Structure of Pr 1−x Ca x MnO 3 89<br />

Ca 2p<br />

Temp = 300 K<br />

2p 3/2<br />

2p<br />

1/2<br />

Intensity (arb. units)<br />

0.84<br />

0.4<br />

0.33<br />

0.2<br />

352<br />

348 344<br />

Binding Energy (eV)<br />

(a)<br />

340<br />

Ca 2p<br />

2p 1/2<br />

2p 3/2<br />

Temp = 77 K<br />

Intensity (arb. units)<br />

0.84<br />

0.4<br />

0.33<br />

0.2<br />

352<br />

348 344<br />

Binding Energy (eV)<br />

(b)<br />

340<br />

Figure 4.7: The Ca 2p core-level XPS (Mg Kα) spectra of Pr 1−x Ca x MnO 3 taken at<br />

room temperature (a) and 77 K (b).


Electronic Structure of Pr 1−x Ca x MnO 3 90<br />

Pr 4d<br />

Temp = 300 K<br />

Intensity (arb. unis)<br />

0.4<br />

0.3<br />

0.2<br />

130<br />

120 110<br />

Binding Energy (eV)<br />

(a)<br />

100<br />

Pr 4d<br />

Temp = 77 K<br />

Intensity (arb. units)<br />

0.4<br />

0.33<br />

0.2<br />

130<br />

120 110<br />

Binding Energy (eV)<br />

(b)<br />

100<br />

Figure 4.8: The Pr 4d core-level XPS (Mg Kα) spectra of Pr 1−x Ca x MnO 3 taken at<br />

room temperature (a) and 77 K (b).


Electronic Structure of Pr 1−x Ca x MnO 3 91<br />

2p could be due to the increase in the effective Mn valence state from Mn 3+ towards<br />

Mn 4+ . The binding energy shift incase of Ca 2p and Pr 4d core levels could possibly<br />

due to the shift in the chemical potential [33, 40]. The low temperature data shows<br />

a smaller value for this chemical potential shift. Such a suppression of the chemical<br />

potential shift have been observed earlier in other transition metal oxides [41, 42] is<br />

mostly due to the formation of stripes. According to the phase separation model, a<br />

pinning of the chemical potential is consequence of the mixed phase nature of these<br />

materials. The numerical studies [1, 3] have indeed shown the pinning of chemical<br />

potential in the composition range, where the phase separation is realized. These<br />

results also support the model of phase separation in the Pr 1−x Ca x MnO 3 system.<br />

4.4 Conclusions<br />

Using UV photoelectron spectroscopy and inverse photoelectron spectroscopy we have<br />

deduced some of the important energy scales in the Pr 1−x Ca x MnO 3 system. Our<br />

photoemission studies across the FM-AFM phase boundary show that the pseudogap<br />

formation in these compounds occur over an energy scale of 0.47±0.01 eV. From a<br />

combined analysis of the photoemission and inverse photoemission spectra we also<br />

have found the energy required for the transfer of Mn 3d e g electrons between the adjacent<br />

Mn sites to be of the order of ∼ 2.8±0.2 eV. Though an accurate measurement<br />

of these energies are not possible due to the inherent limitations of the compounds<br />

and the technique, our results give an upper bound of them. Further, we have discussed<br />

our results from a possible phase separation scenario. From the core level<br />

study of Pr 1−x Ca x MnO 3 , we observed a suppression of chemical potential shift for 77<br />

K spectra compared to that of room temperature. These results support the model<br />

of phase separation in Pr 1−x Ca x MnO 3 system.


Bibliography<br />

[1] A. Moreo, S. Yunoki, and E. Dagatto, Science 283, 2034 (1999).<br />

[2] A. Moreo, M. Mayr, A. Feiguin, S. Yunoki, and E. Dagotto, Phys. Rev. Lett. 84,<br />

5568 (2000).<br />

[3] S. Yunoki, J. Hu, A. L. Malvezzi, A. Moreo, N. Furukawa, and E. Dagotto, Phys.<br />

Rev. Lett. 80, 845 (1998).<br />

[4] C. Zener, Phys. Rev. 82, 403 (1951).<br />

[5] P. W. Anderson and H. Hasegawa, Phys. Rev. 100, 675 (1955).<br />

[6] P. -G. de Gennes, Phys. Rev. 118, 141 (1960).<br />

[7] A. J. Millis, P. B. Littlewood, and B. I. Shraiman, Phys. Rev. Lett. 74, 5144<br />

(1995).<br />

[8] A. J. Millis, B. I. Shraiman, and R. Mueller Phys. Rev. Lett. 77, 175 (1996).<br />

[9] C. Martin, A. Maignan, M. Hervieu, B. Raveau, Z. Jirak, M. M. Savosta, A.<br />

Kurbakov, V. Trounov, G. Andre, and F. Bouree, Phys. Rev. B 62, 6442 (2000).<br />

[10] Ch. Simon, S. Mercone, N. Guiblin, C. Martin, A. Brulet, and G. Andre, Phys.<br />

Rev. Lett. 89, 207202 (2002).<br />

[11] D. D. Sarma, D. Topwal, U., M., S. R. Krishnakumar, M. Bertolo, S. La Rosa,<br />

G. Cautero, T. Y. Koo, P. A. Sharma, S.-W. Cheong, and A. Fujimori, Phys.<br />

Rev. Lett. 93, 97202 (2004).<br />

[12] P. G. Radaelli, R. M. Ibberson, D. N. Argyriou, H. Casalta, K. H. Andersen, S.<br />

-W. Cheong and J. F. Mitchell, Phys. Rev. B 63, 172419 (2001).<br />

92


Electronic Structure of Pr 1−x Ca x MnO 3 93<br />

[13] Y. Tomioka, A. Asamitsu, H. Kuwahara, Y. Moritomo, Y. Tokura, Phys. Rev.<br />

B 53, R1689 (1996).<br />

[14] E. O. Wollan, and W. C. Koehler, Phys. Rev. 100, 545 (1955).<br />

[15] Z. Jirak, S. Krupicka, Z. Simsa, M. Dlouha, and S. Vratislav, J. Magn. Magn.<br />

Mater. 53, 153 (1985).<br />

[16] D. D. Sarma, N. Shanthi, S. R. Krishnakumar, T. Saitoh, T. Mizokawa, A.<br />

Sekiyama, K. Kobayashi, A. Fujimori, E. Weschke, R. Meier, G. Kaindl, Y.<br />

Takeda, and M. Takano, Phys. Rev. B 53, 6873 (1996).<br />

[17] T. Saitoh, A. E. Bocquet, T. Mizokawa, H. Namatame, A. Fujimori, M. Abbate,<br />

Y. Takeda, and M. Takano, Phys. Rev. B 51, 13942 (1995).<br />

[18] P. Pal, M. K. <strong>Dala</strong>i, R. Kundu, M. Chakraborty, B. R. Sekhar, and C. Martin,<br />

Phys. Rev. B 76, 195120 (2007).<br />

[19] D. S. Dessau, Y. D. Chuang, A. Gromko, T. Saitoh, T. Kimura, and Y. Tokura,<br />

J. Electron Spectrosc. and Relat. Phenom. 117-117, 265 (2001).<br />

[20] D. S. Dessau, T. Saitoh, C.-H. Park, Z.-X. Shen, Y. Moritomo, and Y. Tokura, in<br />

Science and Technology of Magnetic Oxides, edited by M. Hundley, J. Nickel,<br />

R. Ramesh, and Y. Tokura, MRS Symposia Proceedings No. 494 (Materials<br />

Research Society, Pittsburgh, 1998), p. 181.<br />

[21] H. Kajueter, G. Kotliar, D. D. Sarma and S. R. Barman, Int. J. Mod. Phys. B,<br />

11, 3849, (1997).<br />

[22] N. Mannella, A. Rosenhahn, M. Watanabe, B. Sell, A. Nambu, S. Ritchey, E.<br />

Arenholz, A. Young, Y. Tomioka, and C. S. Fadley, Phys. Rev. B 71, 125117<br />

(2005).<br />

[23] J.-H. Park, C. T. Chen, S. W. Cheong, W. Bao, G. Meigs, V. Chakarian, and Y.<br />

U. Idzerda, Phys. Rev. Lett. 76, 4215 (1996).<br />

[24] M. K. <strong>Dala</strong>i, P. Pal, B. R. Sekhar, N. L. Saini, R. K. Singhal, K. B. Garg, B.<br />

Doyle, S. Nannarone, C. Martin, and F. Studer, Phys. Rev. B 74, 165119 (2006).<br />

[25] S. Mercone, V. Hardy, C. Martin, Ch. Simon, D. Saurel,and A. Brulet, Phys.<br />

Rev. B 68, 094422 (2003).


Electronic Structure of Pr 1−x Ca x MnO 3 94<br />

[26] A. Maignan, C. Martin, F. Damay, and B. Raveau, Chem. Mater 10, 950 (1998)<br />

[27] C. Martin, A. Maignan, M. Hervieu, and B. Raveau, Phys. Rev. B 60, 12191<br />

(1999).<br />

[28] S. Banik, A. K. Shukla, and S. R. Barman, Rev. Sci. Intrum. 76, 066102 (2005).<br />

[29] A. K. Shukla, S. Banik, and S. R. Barman, Curr. Sci. 90, 490 (2006).<br />

[30] T. Saitoh, A. Sekiyama, K. Kobayashi, T. Mizokawa, A. Fujimori, D. D. Sarma,<br />

Y. Takeda, and M. Takano, Phys. Rev. B 56, 8836 (1997).<br />

[31] P. Pal, M. K. <strong>Dala</strong>i, R. Kundu, B. R. Sekhar, and C. Martin, Phys. Rev. B 77,<br />

184405 (2008).<br />

[32] M. Chakraborty, P. Pal, and B. R. sekhar, Solid State Commun. 145, 197 (2008).<br />

[33] K. Ebata, H. Wadati, M. Takizawa, A. Fujimori, A. Chikamatsu, H. Kumigashira,<br />

M. Oshima, Y. Tomioka, and Y. Tokura, Phys. Rev. B 74, 064419<br />

(2006).<br />

[34] K. Ebata, M. Hashimoto, K. Tanaka, A. Fujimori, Y. Tomioka, and Y. Tokura,<br />

Phys. Rev. B 76, 174418 (2007).<br />

[35] A. Moreo, S. Yunoki, and E. Dagotto, Phys. Rev. Lett. 83, 2773 (1999).<br />

[36] F. M. F. de Groot, M. Grioni, J. C. Fuggle, J. Ghijsen, G. A. Sawatzky, and H.<br />

Petersen, Phys. Rev. B 40 5715 (1989).<br />

[37] H. Kurata, E. Lefevre, C. Colliex, and R. Brydson, Phys. Rev. B 47, 13763<br />

(1993).<br />

[38] A. Chainani, M. Mathew, and D. D. Sarma, Phys. Rev. B 47, 15397 (1993).<br />

[39] Y. Kuwata, S. Suga, S. Imada, A. sekiyama, S. Ueda, T. Iwasaki, H. Harada,<br />

T. Muro, T. Fukawa, K. Ashida, H. Yoshioka, T. Terauchi, J. Sameshima, H.<br />

Kuwahara, Y. Moritomo, and Y. Tokura, J. Electron Spectrosc. Relat. Phenom.<br />

88-91, 281 (1998).<br />

[40] J. Matsuno, A. Fujimori, Y. Takeda, and M. Takano, Europhys. Lett. 59, 252<br />

(2002).


Electronic Structure of Pr 1−x Ca x MnO 3 95<br />

[41] A. Ino, T. Mizokawa, A. Fujimori, K. Tamasaku, H. Eisaki, S. Uchida, T. Kimura,<br />

T. Sasagawa, and K. Kishio, Phys. Rev. Lett. 79, 2101 (1997).<br />

[42] M. Satake, K. Kobayashi, T. Mizokawa, A. Fujimori, T. Tanabe, T. Katsufuji,<br />

and Y. Tokura, Phys. Rev. B 61, 15515 (2000).


Chapter 5<br />

Electronic Structure of<br />

Ca 0.86 Pr 0.14 MnO 3<br />

96


Electronic Structure of Ca 0.86 Pr 0.14 MnO 3 97<br />

5.1 Introduction<br />

The strong interplay among charge, spin, lattice and orbital degrees of freedom of electrons<br />

form the basis of the phenomena of colossal magnetoresistance (CMR) shown<br />

by manganites. To a large extent the traditional models [1, 2, 3] employing the<br />

charge-spin coupling, have been able to explain the CMR in many of the hole doped<br />

compositions (R 1−x A x MnO 3 with x


Electronic Structure of Ca 0.86 Pr 0.14 MnO 3 98<br />

5.2 Experimental<br />

The polycrystalline sample of Ca 0.86 Pr 0.14 MnO 3 was prepared by standard solid state<br />

reactions. Details of the sample preparation technique is published elsewhere [4]. Purity,<br />

homogeneity and composition of the sample were checked by electron diffraction<br />

and energy dispersive spectroscopic analysis. The electrical and magnetic properties<br />

of this composition has also been published earlier [4, 15]. This sample shows<br />

a metallic behavior down to ∼ 200K and then the resistivity slightly increases till<br />

the T c (110 K). Below the T C the behavior is insulating and the resistivity goes up<br />

by many orders of magnitude. Generally, the conductivity behavior of this sample is<br />

similar to that of the other electron doped systems [4, 8, 15].<br />

High resolution photoemission and x - ray absorption spectroscopy measurements<br />

were performed by using the light from the soft x - ray beam line WERA of ANKA<br />

synchrotron, Karlsruhe, Germany. We used the monochromatized radiation from<br />

an undulator. The HRPES measurments were carried out using the photon energy<br />

124 eV. The photoelectrons were collected using a Scienta SES 2002 electron energy<br />

analyzer. The energy resolution was 25 meV in case of the narrow scan (-2 - 2.5 eV)<br />

and it was 125 meV incase of wide scan (-2 - 15 eV) of the valence band. The O K XAS<br />

spectra has been taken in the fluorescence detection mode and the energy resolution<br />

was set to 150 meV. The sample surface was cut using a diamond knife under a base<br />

vacuum of ∼ 1.0 x 10 −10 mbar before taking the data for each temperature. In case of<br />

the photoemission measurements, scraping was repeated until negligible intensity was<br />

found for the bump around 9 - 10 eV, which is a signature of surface contamination<br />

[16]. For the temperature dependent measurements, the samples were cooled by<br />

pumping liquid helium through the sample manipulator fitted with a cryostat. The<br />

sample temperatures were measured by using a silicon diode sensor touching the<br />

bottom of the sample holder.<br />

5.3 Results and Discussion<br />

Figure 5.1 shows the high resolution photoemission spectra of the Ca 0.86 Pr 0.14 MnO 3 .<br />

Varius spectra, taken at 30 K, 70 K, 110 K, 220 K and 293 K show the temperature<br />

dependent changes in the valence band electronic structure. Intensities of all the<br />

spectra were normalized and shifted along the ordinate axis by a constant value for<br />

the clarity of presentation. All the features seen in the spectra are dominated by the


Electronic Structure of Ca 0.86 Pr 0.14 MnO 3 99<br />

B<br />

D<br />

C<br />

Intensity (arb. units)<br />

300 K<br />

220 K<br />

110 K<br />

A<br />

70 K<br />

30 K<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Binding Energy (eV)<br />

Figure 5.1: The angle integrated valence band photoemission spectrum of<br />

Ca 0.86 Pr 0.14 MnO 3 taken using 124 eV photons. The spectra taken at different temperatures<br />

(30 K, 70 K, 110 K, 220 K and 293 K) are normalized for their intensities<br />

and shifted along Y-axis by a constant for clarity of presentation.


Electronic Structure of Ca 0.86 Pr 0.14 MnO 3 100<br />

Intensity (arb. units)<br />

A<br />

30 K<br />

70 K<br />

110 K<br />

220 K<br />

300 K<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

Binding Energy (eV)<br />

Figure 5.2: The high resolution spectra of the near E F region of the valence band of<br />

Ca 0.86 Pr 0.14 MnO 3 for 30 K, 70 K, 110 K, 220 K and 300 K. This feature refers to the<br />

e g band.<br />

states due to the Mn 3d - O 2p hybridization. The broad feature appearing at ∼ 6 eV<br />

(marked D) is due to the bonding states of this hybridization while its non-bonding<br />

states appear as a shoulder at ∼ 3.2 eV (marked C). The peak at ∼ 2 eV (marked B)<br />

is due to the Mn 3d t 2g states of the MnO 6 octahedra [17]. The Mn 3d e g↑ states of<br />

the octahedra appear around ∼ 0.3 eV (marked A) in the spectra. It should be noted<br />

that the intensity of the peak B decreases with decrease of temperature while that of<br />

the feature A increases. These spectral weight shifts are characteristic of many CMR<br />

systems. The e g states lying close to the E F are important for the conductivity of this<br />

material. A high resolution spectra of the near E F region is presented in figure 5.2.<br />

As the temperature is lowered from 293 K the feature A starts appearing. Further,<br />

below the transition temperature T c (110 K), the intensity of A steeply increases and<br />

goes through a maximum. The 30 K spectrum again shows a smaller intensity for<br />

this feature compared to the one at 70 k. Interestingly, below the T c , the leading<br />

edges of the different spectra move towards the E F as we go down in temperature.


Electronic Structure of Ca 0.86 Pr 0.14 MnO 3 101<br />

Mn 3d + O 2p<br />

Pr 5d, Ca 3d<br />

Mn 4sp, Pr 6sp<br />

Normalized Absorption (arb. units)<br />

300 K<br />

200 K<br />

150 K<br />

80 K<br />

30 K<br />

530 540 550 560<br />

Photon Energy (eV)<br />

Figure 5.3: The O K-edge x-ray absorption spectra of the Ca 0.86 Pr 0.14 MnO 3 taken at<br />

30 K, 80 K, 150 K, 200 K and 300 K. The pre-edge feature appears between 527 -<br />

533 eV is mostly due to the strong hybridization between Mn 3d and O 2p orbitals,<br />

where most interest lies. Origin of other features are given in the text.<br />

This shifting of the valence band edges is a clear indication of increase in the width<br />

(W) of the e g band below the T c .<br />

In order to see the corresponding changes in the near E F unoccupied states we have<br />

taken the O K edge x-ray absorption spectra (XAS) of the Ca 0.86 Pr 0.14 MnO 3 sample<br />

at different temperatures (Fig. 5.3). Here, all the spectra have been normalized to<br />

the incoming intensity and absolute cross sections and are also corrected for selfabsorption<br />

and saturation effects. The spectra shows three broad features. The one<br />

between 527 eV to 533 eV, called the pre-edge feature in the O K edge spectra, is<br />

mostly due to the strong hybridization between Mn 3d and O 2p orbitals. The feature<br />

around 536.5 eV is due to the bands from hybridized Pr 5d and Ca 3d orbitals, while


Electronic Structure of Ca 0.86 Pr 0.14 MnO 3 102<br />

Normalized Absorption (arb. units)<br />

A’<br />

B’<br />

C’<br />

30 K<br />

80 K<br />

150 K<br />

200 K<br />

300 K<br />

527 528 529 530 531 532 533 534<br />

Photon Energy (eV)<br />

Figure 5.4: The pre-edge peak of the O K -edge XAS spectra of Ca 0.86 Pr 0.14 MnO 3 for<br />

temperatures 30 K, 80 K, 150 K, 200 K and 300 K.<br />

the structure above 540 eV originates from the states like Mn 4sp and Pr 6sp, etc.<br />

These assignments of the features are consistent with the earlier experiments and<br />

band structure calculations on similar systems [21, 22, 23]. The pre-edge region of<br />

the O K XAS has earlier been shown to represent the near E F unoccupied states<br />

in these kind of compounds [21, 22, 23]. Figure 5.4 shows the pre-edge region of<br />

the XAS spectra of taken at different temperatures. The feature extending from<br />

∼ 527 to 531 eV consists of two lines. The first one is due to the O 2p orbitals<br />

hybridized with the Mn 3d e g↑ (marked A’) while the second arises from the O 2p -<br />

Mn 3d t 2g↓ hybridization (marked B’). The weak feature centered around ∼ 531.5 eV<br />

corresponds to the e g↓ (marked C’) states [23, 24, 25, 26, 27]. All the features show<br />

some temperature dependence. As the temperature is lowered from 300 K to 30 K<br />

the intensity of A’ decreases while that of B’ increases. This again shows a spectral<br />

weight shift complementary to the one observed in the photoemission spectra (Fig.


Electronic Structure of Ca 0.86 Pr 0.14 MnO 3 103<br />

5.1). Here, the spectral weight get shifted from near E F positions to higher energy<br />

as we go down in temperature. One can also see a shift in the energy position of the<br />

feature originating from the e g↓ states.<br />

Structural studies [19] have shown that the Ca 0.86 Pr 0.14 MnO 3 has an orthorhombic<br />

perovskite structure at 293 K. Below 110 K its structure changes to elongated pseudotetragonal<br />

with four short a-b plane and two long out-of-plane Mn-O-Mn bonds.<br />

This structural change can lead to an increase in the in-plane Mn 3d - O 2p hybridization<br />

strength and a consequent increase in the W. The broadening of feature A below<br />

T c in the photoemission spectra (Fig. 2) is due to this increase in W. The parent<br />

compound of our sample, CaMnO 3 has an empty e g band. In the Ca 0.86 Pr 0.14 MnO 3 ,<br />

the e g band could well be considered as rarely occupied. The spin-spin interaction of<br />

the e g electrons in this narrow band can lead to a localization of them resulting in<br />

the formation of FM clusters. Earlier Martin et. al [20] have shown the existence of<br />

such FM clusters embedded in an AFM matrix in electron-doped compositions. The<br />

near E F electronic spectra of such phase separated systems are known to display a<br />

pseudogap behavior [18]. The spectral weight shifts observed in the photoemission<br />

and O K XAS spectra of Ca 0.86 Pr 0.14 MnO 3 corroborate this behavior. The enhanced<br />

shift of electronic states from B to A observed in the photoemission spectra taken below<br />

the T c could be due to the structural changes across the transition. The changes<br />

in the topology of the MnO 6 octahedra below the T c mentioned earlier, might result<br />

in the shift of electrons from the t 2g levels to the near E F positions because of<br />

the increased strength in the in-plane Mn 3d -O 2p hybridization. The unoccupied<br />

states represented in pre-edge region of the O K XAS spectra, on the other hand,<br />

shows complementary shifts of states from the near E F to higher energy positions<br />

with decrease of temperature.<br />

5.4 Conclusions<br />

Using high resolution photoelectron spectroscopy and O K edge XAS, we have studied<br />

the temperature dependent changes in the near E F electronic structure of the electron<br />

doped CMR, Ca 0.86 Pr 0.14 MnO 3 . High resolution photoemission study show changes<br />

in the e g electron band width and a spectral weight shift across the transition which<br />

could be ascribed to the structural changes and consequent changes in the Mn 3d - O<br />

2p hybridization strengths. On the other hand, a complementary shifts of states from


Electronic Structure of Ca 0.86 Pr 0.14 MnO 3 104<br />

the near E F to higher energy positions were observed with decrease of temperature<br />

in the unoccupied states presented in the pre-edge region of O K XAS spectra.


Bibliography<br />

[1] C. Zener, Phys. Rev. 82, 403 (1951).<br />

[2] P. W. Anderson and H. Hasegawa, Phys. Rev. 100, 675 (1955).<br />

[3] P. -G. de Gennes, Phys. Rev. 118, 141 (1960).<br />

[4] A. Maignan, C. Martin, F. Damay and B. Raveau, Chem. Mater 10, 950 (1998).<br />

[5] W. Bao, J. D. Axe, C. H. Chen and S. W. Cheong, Phys. Rev. Lett. 78, 543<br />

(1997).<br />

[6] A. Maignan, C. Martin, F. Damay, B. Raveau and J. Hejtmanek, Phys. Rev. B<br />

58, 2758 (1998).<br />

[7] Shun-Qing Shen and Z. D. Wang, Phys. Rev. B 58, R8877(1998).<br />

[8] M. M. Savosta, P. Novak, M. Marysko, Z. Jirak, J. Hejtmanek, J. Englich, J.<br />

Kohout, C. Martin and B. Raveau, Phys. Rev. B 62, 9532 (2000).<br />

[9] J. J. Neumeier and J. L. Cohn, Phys. Rev. B 61, 14319 (2000).<br />

[10] C. martin, A. Maignan, F. Damay and B. Raveau, Solid State Chem. 134, 198<br />

(1997).<br />

[11] A. Moreo, S. Yunoki and E. Dagatto, Science 283, 2034 (1999).<br />

[12] A. Moreo, M. Mayr, A. Feiguin, S. Yunoki and E. Dagotto, Phys. Rev. Lett. 84,<br />

5568 (2000).<br />

[13] S. Yunoki, J. Hu, A. L. Malvezzi, A. Moreo, N. Furukawa and E. Dagotto, Phys.<br />

Rev. Lett. 80, 845 (1998).<br />

105


Electronic Structure of Ca 0.86 Pr 0.14 MnO 3 106<br />

[14] M. Uehara, S. Mori, C. H. Chen and S. W. Cheong, Nature (London) 399, 560<br />

(1999).<br />

[15] C. Martin, A. Maignan, M. Harvieu, B. Raveau, Z. Jirak, M. M. Savosta, A.<br />

Kurbakov, V. Trounov, G. Andre, and F. Bouree, Phys. Rev. B 62, 6442 (2000)<br />

[16] D. D. Sarma, N. Shanthi, S. R. Krishnakumar, T. Saitoh, T. Mizokawa, A.<br />

Sekiyama, K. Kobayashi, A. Fujimori, E. Weschke, R. Meier, G. Kaindl, Y.<br />

Takeda, and M. Takano, Phys. Rev. B 53, 6873 (1996).<br />

[17] T. Saitoh, A. E. Bocquet, T. Mizokawa, H. Namatame, A. Fujimori, M. Abbate,<br />

Y. Takeda, and M. Takano, Phys. Rev. B 51, 13942 (1995).<br />

[18] P. Pal, M. K. <strong>Dala</strong>i, R. Kundu, B. R. Sekhar and C. Martin, Phys. Rev. B 76,<br />

184405 (2008).<br />

[19] Z. Jirak, S. Krupicka, Z. Simsa, M. Dlouha and S. Vratislav, J. Magn. Magn.<br />

Mater. 53, 153 (1985).<br />

[20] C. Martin, A. Maignan, M. Hervieu and B. Raveau, Phys. Rev. B 60, 12191<br />

(1999).<br />

[21] F. M. F. de Groot, M. Grioni, J. C. Fuggle, J. Ghijsen, G. A. Sawatzky and H.<br />

Petersen, Phys. Rev. B 40 5715 (1989).<br />

[22] H. Kurata, E. Lefevre, C. Colliex and R. Brydson, Phys. Rev. B 47, 13763 (1993).<br />

[23] M. K. <strong>Dala</strong>i, P. Pal, B. R. Sekhar, N. L. Saini, R. K. Singhal, K. B. Garg, B.<br />

Doyle, S. Nannarone, C. Martin, and F. Studer, Phys. Rev. B 74, 165119 (2006).<br />

[24] M. Abbate, F. M. F. de Groot, J. C. Fuggle, A. Fujimori, O. Strebel, F. Lopez,<br />

M. Domke,, G. Kaindl, G. A. Sawatzky, M. Takano, Y. Takeda, H. Eisaki and<br />

S. Uchida, Phys. Rev. B 46, 4511 (1992).<br />

[25] J.-H. Park, C. T. Chen, S. W. Cheong, W. Bao, G. Meigs, V. Chakarian and Y.<br />

U. Idzerda, Phys. Rev. Lett. 76, 4215 (1996).<br />

[26] O. Toulemonde, F. Millange, F. Studer, B. Raveau, J. H. Park and C. T. Chen,<br />

J. Phys. Condens. Matter 11, 109 (1999).


Electronic Structure of Ca 0.86 Pr 0.14 MnO 3 107<br />

[27] N. Mannella, A. Rosenhahn, M. Watanabe, B. Sell, A. Nambu, S. Ritchey, E.<br />

Arenholz, A. Young, Y. Tomioka, and C. S. Fadley, Phys. Rev. B 71, 125117<br />

(2005).


Chapter 6<br />

Summary and Conclusions<br />

108


Summary and Conclusions 109<br />

In this thesis the electronic structure of Pr 1−x Ca x MnO 3 series have been investigated<br />

using photoemission, inverse photoemission and x-ray absorption spectroscopy.<br />

Using photoemission (UPS and XPS), we have studied the valence band and core levels<br />

of Pr 1−x Ca x MnO 3 . The unoccupied states of Pr 1−x Ca x MnO 3 have been studied<br />

by inverse photoemission spectroscopy and x-ray absorption spectroscopy.<br />

A brief overview of the structural, transport and magnetic properties associated<br />

with the CMR manganites have been described in the introduction of this thesis.<br />

Details of the experimental techniques used to study the electronic structure of<br />

Pr 1−x Ca x MnO 3 have been described in the chapter 2.<br />

Using valence band photoemission and O K edge x-ray absorption, we studied<br />

the temperature dependent finer changes in the near E F electronic structure of<br />

Pr 0.67 Ca 0.33 MnO 3 , which is regarded as a prototype for the electronic phase separation<br />

models in CMR systems. With decrease in temperature the O 2p contributions<br />

to the t 2g and e g spin-up states in the valence band were found to increase until T c .<br />

Below T c , the density of states with e g spin-up symmetry increased while those with<br />

t 2g symmetry decreased, possibly due to the change in the orbital degrees of freedom<br />

associated with the Mn 3dO 2p hybridization in the pseudo-CE-type charge or orbital<br />

ordering. These changes in the density of states could well be connected to the electronic<br />

phase separation reported earlier. By combining the UPS and XAS spectra,<br />

we derived a quantitative estimate of the charge transfer energy E CT (2.6±0.1 eV),<br />

which is large compared to the earlier reported values in other CMR systems. Such<br />

a large charge transfer energy was found to support the phase separation model.<br />

For a complete understanding of the changes in the occupied and unoccupied<br />

electronic states we performed a detailed study of the Pr 1−x Ca x MnO 3 system across<br />

their ferromagnetic - antiferromagnetic phase boundary using UPS and IPES. We<br />

used three compositions x = 0.2, 0.33 and 0.4. Our photoemission studies showed<br />

that the pseudogap formation in these compositions occur over an energy scale of<br />

0.48±0.02 eV. Here again, we have estimated the charge transfer energy in these<br />

compositions to be of the order of ∼ 2.8±0.2 eV. The core level electronic structure<br />

of Pr 1−x Ca x MnO 3 (x = 0.2, 0.33, 0.4 and 0.84) also have been studied using X-ray<br />

photoelectron spectroscopy. These studies have been performed at the Mn 2p, Ca 2p<br />

and Pr 4d levels. From these core level studies we observed a suppression of chemical<br />

potential shift for 77 K spectra compared to that of room temperature. These results<br />

support the model of phase separation in Pr 1−x Ca x MnO 3 system.


Summary and Conclusions 110<br />

Using high resolution photoelectron spectroscopy and O K edge x-ray absorption<br />

spectroscopy we have studied the temperature dependence of the near Fermi level<br />

electronic structure in the electron doped CMR, Ca 0.86 Pr 0.14 MnO 3 . We found that<br />

the temperature dependent changes in the electronic structure is consistent with the<br />

conductivity behavior of the electron doped systems in general. As the temperature<br />

is lowered from 293 K a feature due to the e g↑ states starts appearing near the E F .<br />

The intensity of this feature further increases below the T c with a shift of spectral<br />

weight from higher binding energy to the near E F states. Further, we have found an<br />

increase in the e g↑ band width (W) also at low temperatures. On the other hand,<br />

a complementary shifts of states from the near E F to higher energy positions were<br />

observed with decrease of temperature in the unoccupied states presented in the preedge<br />

region of O K edge x-ray absorption spectra. We have discussed our results from<br />

the point of view of phase separation models considering the temperature dependent<br />

structural changes and consequent localization of the electrons in a narrow e g band.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!